Buschow KHJ Handbook of Magnetic Materials Volume 09

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 651

Handbook of Magnetic Materials,

Volume 9
Elsevier, 1995

Edited by: K.H.J. Buschow


ISBN: 978-0-444-82232-1

by kmno4
PREFACE TO VOLUME 9

The Handbook series Magnetic Materials is a continuation of the Handbook series


Ferromagnetic Materials. The original aim of Peter Wohlfarth when he started the
latter series was to combine new developments in magnetism with the achievements
of earlier compilations of monographs, producing a worthy successor to Bozorth's
classical and monumental book Ferromagnetism. This is the main reason that Fer-
romagnetic Materials was initially chosen as title for the Handbook series, although
the latter aimed at giving a more complete cross-section of magnetism than Bozorth's
book.
Magnetism has seen an enormous expansion into a variety of different areas of
research in the last few years, comprising the magnetism of several classes of novel
materials that share with truly ferromagnetic materials only the presence of magnetic
moments. For this reason the Editor and Publisher of this Handbook series have
carefully reconsidered the title of the Handbook series and changed it into Magnetic
Materials. It is with much pleasure that I can introduce to you now Volume 9 of
this Handbook series.
The magnetism of the majority of metallic systems can adequately be described
by the well known concepts of localised or itinerant moment magnetism. These
traditional concepts are, however, not able to describe the magnetism of a fairly large
class of materials generally indicated as heavy-fermion systems. The magnetism of
these strongly correlated charge-carrier systems has developed from two different
sources, the Kondo-impurity concept and the intermediate-valence concept. The last
decade has seen a strong proliferation in experimental and theoretical studies of such
systems. Progress made in this field by means of inelastic neutron scattering was
described already in Chapter 6 of Volume 7 of the Handbook. A more general
account of the magnetism of heavy-fermion systems is presented in Chapter 1 of the
present Volume.
Towers of strengths to the understanding of the physics of magnetism are theory
and experiment. In Volume 7 of the Handbook two different chapters were devoted
to the former, emphasising results of electronic band structure calculations and their
beneficial influence on the understanding of magnetism in many materials. As a
counterweight, two novel experimental techniques will be described in the present
Volume. The first one, Chapter 2, deals with muon spin rotation, the second one,
Chapter 5, gives an account of the possibilities offered by photon beam spectroscopy.
In both chapters it is shown how these sophisticated experimental methods can be
vi PREFACE TO VOLUME 9

used to obtain experimental information not easily obtainable by conventional ex-


perimental methods.
Interstitially modified intermetallic compounds of rare earth and 3d elements are
described in Chapter 3. These materials can be obtained from the pure intermetallics
by filling some of the available interstitial hole sites in their crystal structure with
carbon, nitrogen or hydrogen atoms. Though the drastic changes of magnetocrys-
talline anisotropy and magnetic couplings are of substantial fundamental interest, a
large part of the Chapter is devoted to practical consequences as found in modern
permanent magnet technology.
In one of the preceding volumes, Vol. 7, a major updating of the experimental
results was presented for intermetallics in which rare earths are combined with 3d
transition metals, while progress in ferrite research was presented in Vol. 8. Both
groups of materials are fairly extensive, as are the many experimental results that
have accumulated over the years. Of particular interest in these two groups of ma-
terials is the occurrence of field-induced phase transitions. These phase transitions
are commonly treated in a rather phenomenological way, and at best, are described
in terms of anisotropy and moment couplings. The last chapter of the present vol-
ume deals with the thermodynamic approach and shows how the understanding and
description of these magnetic phase transitions can be considerably enriched.
Volume 9 of the Handbook on the Properties of Magnetic Materials, as the preced-
ing volumes, has a dual purpose. As a textbook it is intended to be of assistance to
those who wish to be introduced to a given topic in the field of magnetism without
the need to read the vast amount of literature published. As a work of reference it is
intended for scientists active in magnetism research. To this dual purpose, Volume 9
of the Handbook is composed of topical review articles written by leading authori-
ties. In each of these articles an extensive description is given in graphical as well as
in tabular form, much emphasis being placed on the discussion of the experimental
material in the framework of physics, chemistry and material science.
The task to provide the readership with novel trends and achievements in mag-
netism would have been extremely difficult without the professionalism of the North-
Holland Physics Division of Elsevier Science B.V., and I wish to thank Joep Ver-
heggen and Wim Spaans for their great help and expertise.

K.H.J. Buschow
Van der Waals-Zeeman Institute
University of Amsterdam
CONTENTS

Preface to V o l u m e 9 . . . . . . . . . . . . . . . . . . . . v

Contents . . . . . . . . . . . . . . . . . . . . . . . . vii

Contents o f Volumes 1-8 . . . . . . . . . . . . . . . . . . ix

List o f Contributors . . . . . . . . . . . . . . . . . . . . xi

1. H e a v y F e r m i o n s and R e l a t e d C o m p o u n d s
G.J. N I E U W E N H U Y S . . . . . . . . . . . . . . . . . 1
2. M a g n e t i c M a t e r i a l s Studied b y M u o n Spin Rotation S p e c t r o s c o p y
A. S C H E N C K and E N . G Y G A X . . . . . . . . . . . . . . 57
3. Interstitially M o d i f i e d Intermetallics o f Rare Earth and 3d Elements
H. F U J I I and H. S U N . . . . . . . . . . . . . . . . . . 303
4. F i e l d I n d u c e d P h a s e Transitions in F e r r i m a g n e t s
A.K. ZVEZDIN . . . . . . . . . . . . . . . . . . . . 405
5. Photon B e a m Studies o f M a g n e t i c Materials
S.W. L O V E S E Y . . . . . . . . . . . . . . . . . . . 545

Author Index . . . . . . . . . . . . . . . . . . . . . . 631

Subject Index . . . . . . . . . . . . . . . . . . . . . . 679

Materials Index . . . . . . . . . . . . . . . . . . . . . 689

vii
CONTENTS OF VOLUMES 1-8

Volume 1
1. Iron, Cobalt and Nickel, by E.P. Woh!farth . . . . . . . . . . . . . . . . 1
2. Dilute Transition Metal Alloys: Spin Glasses, by J.A. Mydosh and G.J. Nieuwenhuys 71
3. Rare Earth Metals and Alloys, by S. Legvold . . . . . . . . . . . . . . . . 183
4. Rare Earth Compounds, by K. H.J. Buschow . . . . . . . . . . . . . . . . 297
5. Actinide Elements and Compounds, by W. Trzebiatowski . . . . . . . . . . . . 415
6. Amorphous Ferromagnets, by E E . Luborsky . . . . . . . . . . . . . . . . 451
7. Magnetostrictive Rare Earth-Fe2 Compounds, by A. E. Clark . . . . . . . . . . 531

Volume 2
1. Ferromagnetic Insulators: Garnets, by M.A. Gilleo . . . . . . . . . . . . . . 1
2. Soft Magnetic Metallic Materials, by G. Y Chin and J. 14. Wernick . . . . . . . . . 55
3. Ferrites for Non-Microwave Applications, by P L Slick . . . . . . . . . . . . 189
4. Microwave Ferrites, by J. Nicolas . . . . . . . . . . . . . . . . . . . . 243
5. Crystalline Films for Bubbles, by A.H. Eschenfelder . . . . . . . . . . . . . 297
6. Amorphous Films for Bubbles, by A.H. Eschenfelder . . . . . . . . . . . . . 345
7. Recording Materials, by G. Bate . . . . . . . . . . . . . . . . . . . . 381
8. Ferromagnetic Liquids, by S. W. Charles and J. Popplewell . . . . . . . . . . . 509

Volume 3
1. Magnetism and Magnetic Materials: Historical Developments and Present Role in Industry
and Technology, by U. Enz . . . . . . . . . . . . . . . . . . . . . . 1
2. Permanent Magnets; Theory, by 11. Zifistra . . . . . . . . . . . . . . . . 37
3. The Structure and Properties of Alnico Permanent Magnet Alloys, by R.A. McCurrie . . 107
4. Oxide Spinels, by S. Krupi&a and P Novdk . . . . . . . . . . . . . . . . 189
5. Fundamental Properties of Hexagonal Ferrites with Magnetoplumbite Structure,
by 14. Kojima . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
6. Properties of Ferroxplana-Type Hexagonal Ferrites, by M. Sugimoto . . . . . . . . 393
7. Hard Ferrites and Plastoferrites, by H. St~blein . . . . . . . . . . . . . . . 441
8. Sulphospinels, by R. P van S t @ d e . . . . . . . . . . . . . . . . . . . 603
9. Transport Properties of Ferromagnets, by I.A. Campbell and A. Fert . . . . . . . . 747

ix
x CONTENTS OF VOLUMES 1-8

Volume 4
1. Permanent Magnet Materials Based on 3d-rich Ternary Compounds, by K. H. J. Buschow 1
2. Rare Earth-Cobalt Permanent Magnets, by K.J. Strnat . . . . . . . . . . . . . 131
3. Ferromagnetic Transition Metal Intermetallic Compounds, by J. G. Booth . . . . . . 211
4. Intermetallic Compounds of Actinides, by V. Sechovsky and L. Havela . . . . . . . 309
5. Magneto-optical Properties of Alloys and Intermetallic Compounds, by K. H.J. Buschow 493

Volume 5
1. Quadrupolar Interactions and Magneto-elastic Effects in Rare-earth Intermetallic
Compounds, by P. Morin and D. Schrnitt . . . . . . . . . . . . . . . . . 1
2. Magneto-optical Spectroscopy of f-electron Systems, by W. Reim and J. Schoenes . . 133
3. INVAR: Moment-volume Instabilities in Transition Metals and Alloys, by E. E Wasserman 237
4. Strongly Enhanced Itinerant Intermetallics and Alloys, by P E. Brommer and J. J.M. Franse 323
5. First-order Magnetic Processes, by G. Asti . . . . . . . . . . . . . . . . . 397
6. Magnetic Superconductors, by ~. Fischer . . . . . . . . . . . . . . . . . 465

Volume 6
1. Magnetic Properties of Ternary Rare-earth Transition-metal Compounds, by H.-S. Li and
J . M . D . Coey . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2. Magnetic Properties of Ternary Intermetallic Rare-earth Compounds, by A. Szytula 85
3. Compounds of Transition Elements with Nonmetals, by O. Beckman and L. Lundgren 181
4. Magnetic Amorphous Alloys, by P. Hansen . . . . . . . . . . . . . . . . 289
5. Magnetism and Quasicrystals, by R.C. O'Handley, R.A. Dunlap and M.E. McHenry 453
6. Magnetism of Hydrides, by G. Wiesinger and G. Hilscher . . . . . . . . . . . 511

Volume 7
1. Magnetism in Ultrathin Transition Metal Films, by U. Gradmann . . . . . . . . . 1
2. Energy Band Theory of Metallic Magnetism in the Elements, by V.L. Moruzzi and
P.M. Marcus . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
3. Density Functional Theory of the Ground State Magnetic Properties of Rare Earths and
Actinides, by 34. S. S. Brooks and B. Johansson . . . . . . . . . . . . . . . 139
4. Diluted Magnetic Semiconductors, by J. Kossut and W. Dobrowolski . . . . . . . . 231
5. Magnetic Properties of Binary Rare-earth 3d-transition-metal Intermetallic Compounds,
by J.J.M. Franse and R.J. Radwarlski . . . . . . . . . . . . . . . . . . 307
6. Neutron Scattering on Heavy Fermion and Valence Fluctuation 4f-systems,
by M. Loewenhaupt and K.H. Fischer . . . . . . . . . . . . . . . . . . 503

Volume 8
1. Magnetism in Artificial Metallic Superlattices of Rare Earth Metals, by J.J. Rhyne and
R. W. Erwin . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2. Thermal Expansion Anomalies and Spontaneous Magnetostriction in Rare-Earth Intermetallics
with Cobalt and Iron, by A. V. Andreev . . . . . . . . . . . . . . . . . . 59
3. Progress in Spinel Ferrite Research, by V.A.M. Brabers . . . . . . . . . . . . 189
4. Anisotropy in Iron-Based Soft Magnetic Materials, by M. Soinski and A.J. Moses 325
5. Magnetic Properties of Rare Earth-Cu 2 Compounds, by Nguyen Hoang Luong and
J. J.M. Franse . . . . . . . . . . . . . . . . . . . . . . . . . . 415
chapter 1

HEAVY FERMIONS AND RELATED


COMPOUNDS

G.J. NIEUWENHUYS
Kamerlingh Onnes Laborato~ Leiden University
RO. Box 9506, 2300 RA Leiden
The Netherlands

Handbook of Magnetic Materials, Vol. 9


Edited by K. H.J. Buschow
©1995 Elsevier Science B.V. All rights reserved
CONTENTS

1. I n t r o d u c t i o n ................................................................. 3
1.1. S c o p e o f this c h a p t e r ..................................................... 3
1.2. The picture ............................................................. 3
1.3. Archetypal heavy fermions ................................................ 5
1.4. The terms .............................................................. 6
1.5. Other reviews ........................................................... 10

2. E x p e r i m e n t a l results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1. 122-compounds ......................................................... 10
2.2. 111-compounds ......................................................... 21
2.3. U2T2X-compounds ....................................................... 26
2.4. U-compounds with CaCu5 structure and related borides ......................... 27
2.5. 334-compounds ......................................................... 34
2.6. Gaps .................................................................. 36
2.7. Magnetism and superconductivity ........................................... 39
2.8. Miscellaneous compounds ................................................. 40
2.9. Yb-compounds .......................................................... 42
2.10. T h i n films .............................................................. 44

3. C o n c l u s i o n .................................................................. 45

4. A c k n o w l e d g e m e n t s ........................................................... 45
References ..................................................................... 45
1. Introduction

1.1. Scope of this chapter


As always, this review has to be restricted for many reasons, few of them being the
interest and knowledge of the author. In this chapter some magnetic properties of
heavy fermions and related compounds will be emphasized and presented mainly in
tabular form. For other aspects of these very interesting new materials the reader
will be referred to other reviews at the end of this introduction.

1.2. The picture


Heavy fermions have been intensively studied during the last decade. The name
'heavy fermion' stems from the fact that these materials exhibit an anomalous specific
heat. At temperatures much lower than the Debye temperature (and much lower then
the Fermi temperature) the specific heat of a conductor can be described by

Cp = 3`T +/3T 3, (1)

where 3`T is the so-called linear term due to the excitations of the itinerant electrons
and ¢3T3 is the low temperature approximation of the specific heat of the lattice. For
normal conductors, 3' is of order 1 ... 10 mJ/(moleK2). However, a new class of
materials - the heavy fermions - showed 3"-values up to 1000 mI/(moleK2). The
expression for 3' in terms of the effective mass of the itinerant electrons reads

3" = (m*kF/Trzhz)(k2/3), (2)

where kF is the Fermi wavevector and hB is Boltzmann's constant. The large value
for 3' is then ascribed to a large value for ra*, hence the name heavy fermion. On
the other hand, if the specific heat, Cv(T), is described in terms of the number of
possible excitations at temperature T, Cp(T) c< nhB(T/To), then the characteristic
temperature, To, is estimated at 10 to 100 K, much lower than the Fermi temperature
of ordinary metals. Heavy fermions are typically found amoung the Ce and U
containing compounds. The magnetic susceptibility of these compounds follows
a Curie-Weiss law at high temperatures, with an effective moment approximately
given by Hund's rule, but at low temperatures the magnetic susceptibility flattens
off (sometimes after attaining a maximum) and becomes constant with decreasing
temperature. This constant, T --+ 0, value is much larger than the Pauli susceptibility
4 G.J. NIEUWENHUYS

of normal metals. In fact, for those heavy fermions where no magnetic ordering is
found, the enhancement of the magnetic susceptibility for T --+ 0 is as large as that of
the specific heat. The crystal structure of most of the heavy fermions is tetragonal or
hexagonal. A large anisotropy in the magnetic susceptibility is generally observed,
e.g., in tetragonal URuzSi2 the magnetic susceptibility measured with the magnetic
field directed along the c-axis shows the features mentioned above, with the field
along the a-axis no temperature dependence at all is found for T ~< 100 K. The
same is roughly true for hexagonal UPdzA13, however with a change in the role of
the a- and c-axis. In both cases the susceptibility can be very well described by
a simple crystalline electric field, CEF, model. A direct consequence of the large
contribution of the itinerant electrons to the (linear term) of the specific heat is that
at a temperature of only a few Kelvin the entropy gain of the itinerant electrons
is already R ln(2), R being the gas constant, a value only expected for localized
magnetic moments. Therefore, in spite of the linear term in the specific heat the
question arises whether the f-electrons of the Ce or the U should be described in an
itinerant model. Moreover, some heavy fermions are superconducting and the jump
in the specific heat at Tc equals about 3'To indicating that the heavy electrons are
involved in the superconductivity. Still the large entropy gain at low temperatures
has to be explained as well, within the same model for heavy fermions. A priori
it is not clear whether charge or spin degrees of freedom cause the entropy (Kagan
et al. 1992). The same ambiguity between itinerancy and localization governs the
role of the crystalline electric fields, CEF. Whereas inelastic neutron scattering can
observe CEF levels in a number of Ce based compounds, the non-observation in
U-based compounds is almost a rule rather than an exception. On the other hand, as
mentioned above the strong magnetic anisotropy can easily be explained by CEF as
is the case also for a number of other macroscopic properties such as specific heat
and magnetic susceptibility as functions of temperature and external magnetic field.
The electrical resistance of heavy fermions is anomalous too. At low temperatures
it can be described by

p(T) = p(O) + A T 2, (3)

where A was experimentally found by Kadowaki and Woods (1986) to be


10-53 '2 #~cm/(moleKZ/mJ)2, and where 3' is the coefficient of the linear term in
the specific heat. In other words, A is (ra*/m) 2 larger than in normal metals. On
increasing the temperature a maximum is attained followed by a logarithmic de-
crease towards higher temperatures. A number of different ground states has been
found in heavy fermions. Some stay paramagnetic down to the lowest temperatures
(20 mK), others order antiferromagnetically (ferromagnetism is seldom found), and
a number of superconductors are observed. Finally, even a semiconducting ground
state is possible. The transition temperature for the superconductivity is rather low,
~< 2 K, and the superconductivity coexists with antiferromagnetism in almost all cases
(UBe13 seems to be the exception amoung the superconducting heavy fermions as
it is with respect to its crystal structure, being cubic). The type of superconducting
order parameter is not known with great certainty, but is not expected to be a simple
HEAVY FERMIONS AND RELATED COMPOUNDS 5

one. UNiaA13 and UPd2A13 have rather normal spontaneous moments as measured
by neutron scattering ( ~ I#B) and the ordering is of long range. URu2Si2 has an
ordered moment of only 0.02#B and the range of order parameter is restricted to
only about 100 A. UPt3 and CeCu2Si2 show only short range correlations of small
moments. These magnetic correlations have been observed, e.g., with magnetic X-
ray scattering (Isaacs et al. 1989, Mason et al. 1990 and Isaacs et al. 1994). In the
case of CeCu2Si2, there is even growing evidence that the superconductivity and the
antiferromagnetic order live in different parts of the sample, Feyerherm et al. (1995)
and Luke et al. (1994). Another general feature of heavy fermions is the large
sample dependence of the magnetic and superconducting properties. Low concentra-
tion replacements by other atoms or slight departures from the exact stoichiometry
induce large effects. For example, small changes in the Cu content in CeCu2Si2 can
alter the volume fractions mentioned above from zero to one hundred percent. Heat
treatments can change the properties also in strange ways, without any evidence for
a change the crystallographic structure. Neutron diffraction experiments (F5k et al.
1995) have shown that annealing URu2Si2 changes the temperature dependence of
the ordered moment drastically without changing the Ndel temperature or the zero-
temperature value of the ordered moment. This strong dependence on the samples
indicates that heavy fermions live close to an instability. It makes experimental re-
search in this field rather difficult and supports clearly the statement that the quality
of the experimental results ~ the quality of the sample. The need for good sample
analysis is evident. Maybe we are slowly reaching the point where we cannot im-
prove our single crystalline samples, since starting materials have limited purity and
thermodynamics tells us that the minimum of the free energy at final temperature
(normal annealing temperature) is obtained for a finite number of imperfections.

1.3. Archetypal heavy fermions

TABLE 1
Main features of archetypal heavy fermions.

Compound Tc TN "7 X(0) A Crystal- Ref.


(K) (K) (mJ/moleK2) (10-9m3/mole) (Iz~ cm/K2) structure
CeCuzSi2 0.65 - 1000 98 - tetrag. [1]
UBel3 0.9 - 1100 172 - cubic [2]
UPt3 0.5 5 450 95 - hexag. [3]
URuzSi2 1.5 17 180 67 - tetrag. [4]
UPdzA13 2.0 14 150 110 - hexag. [5]
U2Znl7 - 9.7 500 545 - rhomb. [6]
UCdl 1 - 5 840 468 - cubic [7]
CeA13 - - 1620 263 35 hexag. [8]
CeCu6 - - 1300 224 - orthor. [9]
YbCu4Ag - - 245 342 10.7 cubic [10]
UA12 - - 142 53 0.06 cubic [11]
6 G.J.N~UWENHUYS

References:
[1] Bredlet al. (1983) [7] Fisket al. (1984)
[2] Ott et al. (1983) [8] Andreset al. (1975)
[3] Stewart et al. (1984) [9] Fujitaet al. (1985)
[4] Palstra et al. (1985) [10] Rosselet al. (1987)
[5] Geibel et al. (1991a) [11] Stewart (1984)
[6] Ott et al. (1984)

1.4. The terms


This section bears on the very well readable reviews by Lee et al. (1986), and by
Bauer (1991).

1.4.1. Single impurity


Since f-electron ions play the major role in heavy fermions and Anderson (1961) has
invented a simple, nontrivial, successful and well understood model for an single
impurity in a metal, this should be the starting point for a theoretical description. In
its simplest form the three ingredients are the conduction electrons, an f-ion with a
single orbital and a hybridization matrix element Vfk that couples this orbital with
the conduction band. The Hamiltonian then reads

H: ~-~ 6(k)nks q- 6f ~ nfs ~- Unfsnf$ + ~ (Wfkf+Ck~+ VfkekJ~),


+ (4)
ks s ks

where U is the Coulomb correlation energy associated with double occupancy of the
f orbital. The width, F, of the f level is given by ~rN(ez)lVfkFI2, where N(eF) is
the single-spin density of states at the Fermi level. Ce has Nf degenerate orbitals
with maximum total occupancy, nf of 1, since U is estimated to be 5 eV. The
f orbitals have an energy of a few eV below the Fermi level. For Uranium the
relevant configurations are f2 (angular momentum, J = 4) and f3 ( j = 9/2), The
correlation energy is estimated to be 2 eV. The hybridization has often been taken
as an adjustable, constant, parameter, but ab initio band structure calculations have
been carried out, e.g., by Sheng et al. (1990). In order to obtain absolute information
on this parameter Endstra et al. (1993a) have used a simple model to estimate Vek to
explain trends in the magnetic ordering temperatures. The energy of the f level, cf,
with respect to the chemical potential and the bandwidth, F, govern the equilibrium
occupancy of the f levels, thus non-integer occupancies may occur, leading to so-
called valence fluctuations. Additionally, because of these fluctuations, f electrons
may exchange spin components with the conduction band, without an actual charge
transfer. In that case one speaks of spin fluctuations. Crystalline electric fields,
CEF, effects tend to lower the degeneracy, on the other hand, the hybridization can
renormalize the strength of the CEF downwards. During the last decade many exact
results have been obtained for the degenerate Anderson model taking into account
the effects due to crystalline electric fields, spin-orbit coupling and external magnetic
fields (Schlottmann 1989).
HEAVY FERMIONS AND RELATED COMPOUNDS 7

1.4.2. Kondo
Under strict conditions (nondegeneracy and integer occupation) the Hamiltonian in
eq. (4) can be transformed (Schrieffer and Wolff 1966) into

H K = -23S.s(0), (5)

which is exactly the Hamiltonian that Kondo (1964) used to describe the on site
interaction of an S -- 1/2 single impurity interacting with the electron density s(0).
Since 3 is negative, this interaction favours antiparallel alignment of the conduction
electron spins with respect to the impurity spin at low temperatures, finally leading
to a singlet state at T = 0. This moment compensation leads to a result for the
magnetic susceptibility which agrees with the observations in heavy fermions. The
compensation or screening also leads to a term in the electrical resistivity ~ ln(TK/T)
in the vicinity of the Kondo temperature, TK. In the limit for U --+ oo the interaction
3 = - v Z / @ f - EF), where Vfk is replaced by a constant V. As a consequence
of the singlet ground state of the impurity, the low-temperature properties of the
system can be described within a Fermi liquid picture. In the degenerate case, the
transformation from eq. (1) to eq. (2) can also be made for integer occupation (no
charge fluctuations) and then leads to the Coqblin-Schrieffer model (Coqblin and
Schrieffer 1969). Heavy fermions, consisting of a regular lattice of Ce or U atoms,
are called Kondo lattices, but one should bear in mind that the Kondo Hamiltonian
can only be obtained exactly in the single impurity case with integer occupation.

1.4.3. The compounds


In a heavy fermion compound, we are dealing with a regular array of magnetic 'im-
purities'. In that case the interactions between the Ce or U ions cannot be neglected
in general; the problem has acquired a translation symmetry and charge transfer be-
tween the f states and the conduction band can no longer be considered as a minor
perturbation. The hybridization causes on the one hand the screening as in the single
impurity model, but at the other hand induces an interaction between the magnetic
moments mediated by the conduction electrons. The screening picture itself has to
be reconsidered, since there are not enough conduction available for all magnetic
moments. Charge transfer now renormalizes the chemical potential when a finite
number of f electrons is promoted into the conduction band. As this process is tem-
perature dependent, the chemical potential becomes temperature dependent, because
of the conservation of the total number of electrons. On the one hand, this notion of
the 'lattice' aspect of the Kondo problem has led to an item like correlation, which
denotes the subtle antiferromagnetic correlations between the - partly - screened
magnetic moments thought to cause the rapid decrease in the electrical resistance
with decreasing temperature observed in heavy fermion compounds even without
long range magnetic order. On the other hand, Strong and Millis (1994) showed
that most properties (e.g., specific heat, magnetic susceptibility) of the archetypel
heavy fermions CeA13 and CeCu6 c a n be beautifully explaned using crystal field and
Kondo effects of single atoms only. The exception is the electrical resistance, where
the agreement between calculation and experiment is less satisfactory. Ab initio cal-
culations of the size of the magnetic moments and of the transition temperatures of
8 G.J. NIEUWENHUYS

magnetically ordering compounds have been carried out by Cooper et al. (Cooper
1992; Hu and Cooper 1993). In their calculation scheme, they have treated both
hybridization and Coulomb exchange effects simultaneously in the presence of inter-
configurational correlation effects. In this way, they are able to successfully compute
the magnetic moments and transition temperatures of a number of U- and Ce-based
compounds adopting the NaC1 structure. Band-structure calculations (Norman and
Koelling 1992) are not able to correctly predict these magnetic properties because
they include only those aspects of the valence fluctuations that can be captured by
time averaging. Therefore, the true character of the 4f or 5f states is missed and has
to be included by adding an additional on-site scattering term (Fulde et al. 1988) or
interconfigurational correlation effects. Due to computational limits such calculations
have not been carried out for the compounds we will discuss in this chapter. We will
therefore limit ourselves to rough estimates of the interactions involved and use that
as a guidline through the experimental results. To lowest order the strength of the
conduction electron mediated interaction between the magnetic atoms, the RKKY
interaction, is proportional to 32, and thus proportional to V 4. As long as this in-
teraction does not dominate, the screening process is still effective and the physical
properties of the concentrated systems can roughly be considered as a summation
of the single impurity effects. The most apparent exception is seen in the electrical
resistance of stoichiometric compounds, where coherent excitations of the f electrons
become important. In the single impurity limit the electrical resistance increases as
ln(TK/T) with decreasing temperature and saturates as T --+ 0. Coherency causes
the resistance to decrease again for T << TK, which leads to the AT 2 term mentioned
earlier. When the RKKY interaction dominates, the system will order magnetically,
although sometimes with very small ordered moments. The competition between the
'screening', with a characteristic strength

TK oc (1/N(cv) ) exp ( - 1/N(eF)3), (6)

and the RKKY interaction with a characteristic strength

TRKKY ~ 32N(eF), (7)

has lead to a phase diagram for the present type of compounds (Doniach 1977, Brandt
and Moshchalkov 1984) depicted schematically in fig. 1. This phase diagram draws
the boundaries between the magnetically ordering compounds, the 'Kondo-lattices'
and the paramagnetic systems. Although it does not provide quantitative information
for three dimensional systems, it has been shown by Endstra et al. (1993a) that it
serves very well as a guidline for experimentalists to order the compounds within a
given series based on a rude estimate of the hybridization parameter. Recognizing
that

:J ~x V 2 / ( e f - eF) (8)
and assuming that: (i) the hybridization is mainly due to the d-itinerant electrons,
so that 3 can be replaced by 3dr and (ii) that ef - eF is nearly constant over a given
HEAVY FERMIONS AND RELATED COMPOUNDS 9
2.0

1.5

i1.0 f
/r.
PARAMAGNET
KONDO
e METAL
0.5

0,5 1.0
d/W~
Fig. 1. The one-dimensional Kondo-necklace phase diagram as derived by Doniach (1977).

series of analogous compounds, 3 can by assumed to be proportional to V~. Using


an approach put forward by Harrison and Straub (Harrison 1969, 1980, 1983, 1987,
and Straub and Harrison 1985) the hybridization matrix elements can be written as

ggl, m =(771l,mj~2/~'~e)[(T21--1~'2/'-l)1/2/dl+l'+1] (9)

The input parameters are the atomic radii of the respective atoms (rl and rv), the
interatomic distance d, the angular momenta l, l' (l = 0, 1, 2, 3 for s, p, d, f orbitals),
and the symmetry of the bond m. Note that me in eq. (9) is the electron mass. r/u,,~
is given by (Harrison and Straub 1987)

(-1) z'+l (l + 1')!(2l)!(2l')!


?]ll'm -- - - X
6rr 2z+Z' l!l, !

× (-1)~ [ ( 2 / + 1)(2/' + 1) I 1/2


(1 + m ) ! ( l -- m ) ! ( l ' + m ) ! ( l ' -- m ) !

with m = 0, 1, 2, 3 for ~r, re, 5 and ~o bounds. This simple approach has enabled
Endstra et al. (1993a) to order compounds of several series and to predict the
behaviour of a number of pseudo binary compounds. This picture makes it clear
why some compounds do order magnetically, while others don't. The typical heavy
fermion compounds have to be placed near the right-hand side of the phase diagram,
which partly explains their sensitivity for the exact composition.
10 G.J. NIEUWENHUYS

1.5. Other reviews


Numerous reviews on several aspects of the heavy fermion phenomena have appeared
in the literature. Reviews with emphasis on the development of the theory can be
found in Lee et al. (1986), Gor'kov (1987), Fulde et al. (1988), Schlottmann
(1989), Grewe et al. (1990) and Hewson (1993). An experimental review of the
superconductivity is given in Fisk et al. (1986), while Fisk et al. (1988) deal mainly
with the Fermi-liquid character. Other experimental reviews are: Stewart (1984), Ott
(1987), Ott and Fisk (1987), Sechovsky and Havela (1988) (actinide compounds),
Sigrist and Ueda (1991) (unconventional superconductivity), Bauer (1991) (CeCu
and YbCu compounds), Sarma et al. (1992) (sound experiments), Aeppli and Fisk
(1992) (Kondo insulators), Endstra et al. (1993b) (122-compounds), Schenck (1993)
and Schenck and Gygax (this volume) (#+SR), Hess et al. (1993), Loewenhaupt and
Fisher (1994) (neutron scattering), Knetsch et al. (1994) (URuzSi2), Sauls (1994)
(phase diagram of UPt3), and Ott (1994) (superconductivity).

2. Experimental results

The number of papers on experimental results on heavy fermion materials and related
compounds amounts to a few thousands. Only in the last few years our data base
grew with 2000 titles. Ineventable, a selection had to be made for this review paper.
The consistency and readability of this chapter were the only criteria used. In view
of the large number of papers, the idea about a world-wide data base for storing
experimental data and there interpretation should be given some serious though,
bearing in mind that the technical infrastructure is available.

2.1. 122-compounds
There has been an extensive research on the MT2X 2 compounds, with M a mag-
netic (rare-earth or actinide) ion, T a transition metal and X being Si, Ge, Sn or Sb.
Hundreds of compounds exist which makes these series very suitable for comparison
along the periodic table. Almost all compounds crystallize in a tetragonal structure
with a e/a ratio of order of 2.3. However, two different allotropic derivatives ex-
ist: The body-centered ThCrzSi2 structure (I4/mmm) and the primitive CaBe2Ge2
one (P4/nmm). These structures are depicted in fig. 2. Note the structures differ
in the sequence of the atomic planes, being M-X-T-X-M-X-T-X-M for the body-
centered cell and M-T-X-T-M-X-T-X-M for the primitive cell. Since the extra X-ray
diffraction lines from the primitive structure can be rather weak, it is not always
clear from the literature which structure is the correct one for a given compound.
Some compounds (e.g., UCo2Ge2) can be stabilized in either of the two structures.
The magnetic properties appear to be strongly different in those cases. A general
rule about the electrical resistivity, p, has been found by Steeman et al. (1990b).
Compounds which crystallize in the ThCrzSi2 structure have Pt[c < P±c, while it is
the other around for the compounds with the CaBezGe2 structure. Up to now the
only exception is UNizSi2 (Ning et al. 1990). The family of 122-compounds has
HEAVY FERMIONS AND RELATEDCOMPOUNDS 11

(a) (b)

T I_ _T I

(~M ©T oX
Fig. 2. A schematic drawing of (a) the ThCrzSi2 and CaBezGe2 crystal structure. The origin of the
CaBe~Ge2 unit cell is shifted by (3/4, 3/4, --ZM)to facilitate a comparison with the ThCrzSi2 unit cell.
M denote the rare-earth atoms, T are the transition-metal atoms and X the metalloids Si, Ge, Sn, etc.
(after Endstra 1992).

two famous members: CeCu2Si2, being the first heavy fermion materials found to
become superconducting (Steglich et al. 1979), and URu2Si2 found to be an an-
tiferromagnetically ordering (TN = 17.5 K) superconductor with moderately heavy
electrons (Palstra et al. 1985). Extensive investigations have been devoted to these
two compounds since the discovery of their peculiar properties. In both cases the
physics is not straightforward and a large sample dependence has been found. In the
case of CeCuaSi2 superconductivity can only be found if the material is melted with
an excess amount of Cu (up to 10%). However, e.g., microprobe analysis shows the
compound to be of the correct stoichiometric composition. The superconductivity is
remarkable in the sense that the temperature derivative of the upper critical field is
enhanced as much as the effective mass of the electrons and that the 'jump' in the
specific heat at Tc divided by the now enhanced 3' is still of order 2. This means
that the heavy electrons (being quasi particles) attain the superconducting ground
state. Bearing in mind that the large effective mass of the quasi particles also means
that the Fermi temperature, TF, is exceptionally low (~ 100 K), superconductivity
is found now in a compound where TF is much smaller then the Debye tempera-
ture and that Tc is of order 1% of TF. Note that a similar situation exists in the
high temperature superconductors. Recently (Feyerherm et al. 1995 and Luke et al.
1994) magnetic order of a spin glass type has been found in CeCu2Si2. However,
12 G.J. NIEUWENHUYS

depending on the starting composition, the volume fraction of the magnetically or-
dering part of the sample and the volume fraction of the superconducting part are
both a function of temperature, in such a way that the total is 100%. Therefore,
superconductivity and magnetic order do not coexist in CeCu2Si2 on a microscopic
scale. We mention 'starting composition' on purpose, since microprobe analyse is
not able to distinguish between the samples obtained. The other famous member of
the family, URu2Si2, shows antiferromagnetic order at TN = 17.5 K, as evidenced by,
e.g., specific heat (Palstra et al. 1985) and neutron diffraction (Broholm et al. 1987
and 1991). The latter experiments also showed that the magnetic order persists into
the superconducting state below Tc = 1.3 K, and that the magnetic order must be of
a peculiar type, since the magnetic moment in the ordered state is only 0.03#B per
U atom. The smallness of the ordered moment seems not to be of any consequence,
neither for the size of the anomaly in the specific heat at TN (being large), nor on
the dependence of the magnetic properties in large external fields. For example, a
transition to ferromagnetic alignment has been observed (de Boer et al. 1986) at
B = 36 T, which would point, together with TN = 17.5 K, to a normal value of the
ordered moment. A number of attempts has been made the explain the magnetic
behaviour of URu2Si2 on the basis of CEF effects. The first one (Nieuwenhuys
1987) could describe the magnetic anisotropy and the general features of the specific
heat and the magnetic susceptibility, but certainly not the small ordered moment,
the last one (Santini and Amoretti 1994) ascribes the main features as being due
to quadrupolar ordering and considers the staggered magnetic moment as a weak
secondary effect. This, however, should assume that the quadrupolar ordering is not
observable via neutron diffraction, since the detailed experiments by Walker et al.
(1993) did not reveal any other order parameter than the magnetization along the
crystallographic c-axis. A further remarkable effect is that the electrical resistivity
first increases while cooling below TN, indicating the formation of a charge gap. We
will come back to this point later. Pressure has an anisotropic effect on the properties
of URu2Si2. Bakker et al. (1992) found for the magnetism dTN/dPa = 126 mK/kbar
and dTN/dPc = -41 mK/kbar. The pressure dependence of the superconductivity
can be described by dTc/dPa = - 3 5 mK/kbar and dTc/dPc = 25 mK/kbar. A
hydrostatic high pressure experiment revealed a value of 36 K for TN at 80 kbar
(Kagayama et al. 1994). The magnetic ordering is, in spite of the very small or-
dered moment, rather stable as a function of sample preparation. This in contrast to
the superconductivity, where a large sample dependence can be observed (Ramirez
et al. 1991). The large sample dependence has lead to conclusions pointing towards
a double transition to superconductivity as observed in UPt3, but, e.g., the detailed
specific heat investigation by Knetsch et al. (1992) has refuted this idea. Super-
conductivity in URu2Si2 is also not of a simple type. The power-law behaviour
of the specific heat below Tc already indicates that the energy gap due to super-
conductivity does not exists over the whole Fermi-surface. Measurements of the
penetration depth (Knetsch et al. 1992) support this conclusion. Very remarkable
are the results of #+SR experiments (Luke et al. 1990): for fields along the c-axis a
reasonable penetration depth (about 8000 A) is observed, however for fields directed
perpendicular to the c-axis no sign of superconductivity, neither in the Knight shift
HEAVY FERMIONS AND RELATED COMPOUNDS 13

nor in the relaxation rate can be discerned. That all in spite of the observation by
Knetsch et al. (1992) that Bcl is isotropic. For details on of the ~+SR results, the
reader is referred to the review by Schenck and Gygax in this volume, more detailed
information on URu2Si2 has been collected by Knetsch et al. (1994). The tetragonal
character of the structure of the 122-series almost immediately leads to anisotropic
magnetic behaviour. The investigation of single crystals is therefore sometimes a
condition for the research. The materials can be grown into single crystal form via
the tri-arc method (Menovsky et al. 1983). Quite a number of these systems show
magnetic ordering (mainly antiferromagnetic) 1. The exact valency of the magnetic
atom is of importance, particularly for Ce, since Ce 3+ has one 4f-electron and thus
a magnetic moment and C e 4+ is nonmagnetic. In the CeTzX2 series the valency
attains a maximum of 3.18 for T = Co and X = Si (Neifeld et al. 1985). The
valency approaches 3 going to 4d and 5d transition metals and/or to X = Ge or Sn.
Since magnetic C e 3+ is a Kramers ion with J = 5/2, the J-multiplet is split up by
the crystalline electric field into 3 doublets. The energy difference between the two
lowest doublets can be remarkably small (10-100 K) which is of the same order as
the magnetic ordering temperature or the characteristic energy of the Kondo effect
(TK). CEF effects in U based compounds are less clear, but they certainly play a
role, e.g., in determining the magnetic anisotropy. For U, the ground state can be a
non-magnetic singlet, assuming that U is in 4+ valence state.

2.1.1. CeTzX2
In table 2 the main structural and magnetic properties of the Ce based 122-compounds
are lis(ed. In this table (and in the following tables) the structure is denoted by I for
the inversion symmetric ThCr2Si2 tetragonal type and P for the primitive CaBezGe2
one. The lattice parameters are in ,~. The value of the linear specific heat coefficient,
7, is given in mJ/mole K: only if it is not disturbed by magnetic ordering and/or CEF
effects. The magnetic ordering is denoted by A for antiferromagnetism, F for ferro-
or ferrimagnetism, C for complex ordering and G for spin-glass type ordering. The
temperatures, TM and T~: (TM being the magnetic ordering temperature), and the
energy of the first excited state in the CEF level scheme A 1 divided by Boltz-
mann's constant are in Kelvin. The crystal structure of CePt2Si2 is a unique type
for REPt2Si2 compounds and has the P4/mm space group, and CePt2Ge2 crystallizes
in a monoclinic deformation of the CaBe2Ge2 with space group P21, b -- 4.402 and
fi = 90.83 °. CelrzSi2 can crystallize in two different structures, depending on the
heat treatment (see discussion below on UCo2Ge2). CeNizSn2 adopts a structure
which is slightly monoclinically distorted with respect to the CaBezGe2 structure
(Liang et al. 1990). Most Kondo temperatures and crystalline electric field splittings
have been obtained from the detailed neutron research by Severing et al. (1989a
and b) and Loidl et al. (1992), others from have been estimated on the basis of bulk
measurements such as specific heat and magnetic susceptibility. CeCuzGe: exhibits

1 We will not consider compounds with Mn, since Mn is the only transition metal carrying a magnetic
moment in these type of compounds.
14 G.J. NIEUWENHUYS

TABLE 2
Structural and magnetic properties of CeT2X 2. The structure is denoted by I for the inversion symmetric
ThCr2Si2 tetragonal type and P for the primitive CaBe2Ge 2 one. The lattice parameters are in ,~. The
specific heat coefficient, % is given in mJ/(mole) K 2 only if it is not disturbed by magnetic ordering
and/or CEF erects. The magnetic ordering is denoted by A for antiferromagnetism, F for ferro- or
ferrimagnetism, C for complex ordering and G for spin-glass type ordering. The temperatures, T M and
TK (TM being the magnetic ordering temperature), and the energy of the first excited state in the CEF
level scheme A 1 divided by Boltzmann's constant are in Kelvin.

Compound Struc. a c 7 Magn. TM TK A1 Ref.


CeAg2Ge 2 I 4.280 10.93 - A 6.3 3 0 [1-4]
CeAgzSi 2 I 4.250 10.66 - A 9.5 - 100 [5-8]
CeAlzGa2 - 4.203 10.95 80 F 9 - - [3]
CeAuzGe2 I 4.367 10.41 - A 15 - 127 [2, 4]
CeAu2Si 2 I 4.310 10.20 - A 10.1 2 200 [5-10]
CeCozGe2 I 4.099 10.24 - A 29.5 - 171 [11]
CeCozSi 2 I 3.953 9.78 - P - - - [5, 6]
CeCul.3 Sb2 P 4.341 10.26 - C 7.7 - 20 [12]
CeCuzGe2 I 4.140 10.16 100 A 4.15 6 190 [4, 13-16]
CeCu2Si2 I 4.105 9.93 1000 G 0.8 10 145 [7, 9, 17-20]
CeCuzSn2 P 4.433 10.35 - A 2.1 2 150 [21-23]
CeFe2Ge2 I 4.070 10.48 - P - - - [24]
CeFe2Si2 I 3.995 9.88 - P - - - [25]
CeIrzGe2 P 4.246 10.10 - P - - - [26]
CeIrzSi 2 P 4.143 9.85 - P - - - [27]
CelrzSi z I 4.085 10.16 - P - - - [27]
CeIr2Sn 2 P 4.499 10.04 - A 4.1 28 - [22, 23]
CeNizGe2 I 4.150 9.85 350 P - 30 - [28]
CeNi2Sb2 P 4.416 9.97 400 P - - 3 [12, 21]
CeNi2Si2 I 4.036 9.57 - P - - - [6]
CeNizSn 2 P 4.409 10.11 600 A 1.8 7 17 [21-23, 29]
CeOs2Si2 I 4.162 9.85 - P - - - [27, 30]
CePdzGe2 I 4.369 10.06 - A 5.1 - 110 [31, 32]
CePd2Si 2 I 4.237 9.89 - A 10 10 230 [7-9, 27, 31, 33]
CePdzSn2 P 4.554 10.48 - A 0.5 2 - [22, 23]
CePtzGe 2 * 4.400 9.80 - A 2.2 2 - [34, 35]
CePtzSi 2 * 4.252 9.80 89 P - 60 194 [10, 27, 36-39]
CePt2Sn2 P 4.581 10.39 - A 0.88 1 29 [22, 23, 40, 41]
CeRh2Ge2 I 4.161 10.45 - A 14 - - [42--44]
CeRh2Si2 I 4.089 10.18 - A 36 33 - [8, 9, 27, 45]
CeRh2Sn2 P 4.472 10.56 - A 0.47 2 - [22, 23]
CeRu2Ge2 I 4.256 10.00 - F 7.91 - 500 [2, 4, 31, 46, 47]
CeRu2Si2 I 4.197 9.80 - P - 15 220 [9, 27, 37, 48]

References:
[1] Gignoux et al. (1988) [10] Heeb et al. (1991)
[2] B6hm et al. (1988) [li] Fujii et al. (1988)
[3] Rauchschwalbe et al. (1985) [12] Mentink et al. (1994)
[4] Loidl et al. (1992) [13] de Boer et al. (1987)
[5] Palstra (1986) [14] Knopp et al. (1989)
[6] Palstra et al. (1986a) [15] Jaccard et al. (1992a)
[7] Severing et al (1989a) [16] Jaccard et al. (1992b)
[8] Grier et al. (1984) [17] Bredl et al. (1983)
[9] Severing et al. (1989b) [18] Luke et al. (1994)
HEAVY FERMIONS AND RELATED COMPOUNDS 15
[19] Jarlborg et al. (1983) [34] Sampathkumaranet al. (1991)
[20] Feyerherm et al. (1995) [35] Das et al. (1991)
[21] Kaczmarska et al. (1993) [36] Ayache et al. (1987)
[22] Beyermann et al. (1991) [37] Gignoux et al. (1991)
[23] Selsane et al. (1990) [38] Bhattacharjee et al. (1989)
[24] Felner et al. (1975) [39] Hiebl et al. (1985)
[25] Rogl (1984) [40] Mignot et al. (1993)
[26] Francois et al. (1985) [41] Shigeoka et al. (1992)
[27] Hiebl et al. (1986) [42] Thompson et al. (1994)
[28] Knopp et al. (1988) [43] Venturini et al. 1988a
[29] Takabatake et al. (1990) [44] Venturini et al. (1998b)
[30] Horvath et al. (1983) [45] Quezel et al. (1984)
[31] Besnus et al. (1992) [46] Besnus et al. (1991)
[32] Rossi et al. (1979) [47] Godart et al. (1987)
[33] Steeman et al. (1988) [48] Regnault et al. (1988)

an accidental quartet as the excited state, while CeAg2Ge2 attains such a quartet as
ground state. The calculated ordered moment obtained from CEF splittings is in
excellent agreement with the measured values for Si and Ge compounds, except for
CeCu2Ge2. This indicates that the ground state is not much influenced by the Kondo
effect, i.e. A1 is much larger than TK, as shown in the table. However, for the Sn
compounds these energies are of comparable magnitude. The ratio of the Kondo
temperature and the CEF-splitting can therefore be used to classify these magnetic
compounds. Small CEF splittings can easily lead to incorrect determinations of the
linear specific heat contribution, e.g., CeNi2Sn2 with 7 K and CeNi2Sb2 with 3.2 K,
have maxima in the low temperature specific heat at about 3 or 1.5 K, respectively,
which could appear as an enormous "~ when C/T is plotted versus T over a restricted
temperature range. CeCul.3Sb2 is a remarkable compound in the sense that it has the
correct crystal structure, but that part of the Cu sites are not occupied. The magnetism
in this compound is rather complex including field cooling effects, suggestive for
spin-glass or some random ordering (Mentink 1994a). Polycrystalline CePt2Sn2
does show the magnetic ordering at TN = 0.47 K, however, a single crystalline
sample exhibited at m a x i m u m only short range order (Mignot et al. 1993). We will
come back to this point in connection with a detailed studied example: CePd2A13.
Using eqs (8-10) and the structure parameters given in table 2 and in the references
mentioned therein, an estimate can be given for the strength of the hybridization and
thus of the magnetic ordering temperature, see section 1.4.3. Doing so, the CeT2Si2
and CeT2Ge2 can be placed in a phase diagram, as has been done by Endstra et
al. (1993a) and displayed in figs 3 and 4. As can be seen from these figures, the
Ce-based alloys can be classified within this magnetic phase-diagram. The Sn and
Sb compounds have a smaller hybridization then their Si and Ge counterparts. This
then would suggest that these compounds should be found at the left-hand side of the
phase diagram, and thus order magnetically as well localized systems. However, it
appears that the ordering temperatures are much lower or even zero, but at the same
time these compounds have a considerable electronic contribution to the specific
heat, which would point to the right-hand side of the diagram. On the other hand,
for all transition metals there is a clear trend when going from Si --+ Ge --+ Sn and
Sb, which suggest that a renormalization of the T/W axis would fit the Sn and
16 G.J. NIEUWENHUYS

I-- CeT2Si 2

Rh

Ru P t Ir Lr Ir~ Os

J._,/W
Fig. 3. Schematic phase diagram for the Kondo lattice with magnetic ordering temperature of the
CeT2Si2 compounds (after Endstra et al. 1993a).

I.- CeT2G %

Rh

ge rid_/W

Fig. 4. Schematic phase diagram for the Kondo lattice with magnetic ordering temperature of the
CeT2Ge2 compounds (after Endstra et al. 1993a).

Sb compounds without much difficulty. The fact that almost all above mentioned
materials crystallize in a similar crystal structure implies that probably also pseudo
binary mixtures can be made, which then opens the possibility to tune parameters
like the Kondo temperature and/or the hybridization via small changes in the lattice
parameters and the density of the d-electrons. For example, very rich phase diagrams
have been found for Ce(Cul_~Ni=)2Ge2 by Steglich et al. (1990a, b), while Heeb
et al. (1991) have shown that CePtAuSi2 forms a heavy fermion system between
the local moment antiferromagnet CeAu2Si2 and the typical Kondo-type compound
CePt2Si2. Another beautiful experiment was carried out by Jaccard et al. (1992a
and b) on CeCu2Ge2, which is on the right-hand side of the phase diagram in fig. 4.
By applying pressure the volume was decreased, thus the hybridization increased,
pushing the compound towards the right. Indeed the magnetic ordering temperature
decreased, a heavy fermion region was entered leading finally to superconductivity.
Other pressure experiments were performed by Thompson et al. (1986) on CeM2Si2,
HEAVY FERMIONS AND RELATED COMPOUNDS 17

M = Ag, Au, Pd and Rh, where TN was determined as a function of the volume of
the unit cell. Again these early experimental result are in good agreement with the
phase diagram shown above.

2.1.2. CeCuX3
As an intermezzo, we will discuss here the properties of two new 113-compounds,
which almost the same structure as the 122. CeCuA13 and CeCuGa3 have been
prepared and investigated by Mentink et al. (1993a). CeCuA13 crystallizes in a
disordered variant of the ThCr2Si2 structure, with a = 4.256 ,~ and e = 10.633 A,
while CeCuGa~ adopts the atommically ordered primitive tetragonal BaNiSn3 struc-
ture (space group I4/mm), depicted in fig. 5, with lattice parameters a = 4.266 ,~ and
c = 10.434 A. CeCuA13 orders antiferromagnetically at 3 K, while CeCuGa3 stays
paramagnetic down to 0.4 K in spite of the fact the former one has a site disordered
structure and the latter a atommically ordered lattice. From magnetic susceptibility
and specific heat measurements an estimate can be given for the crystalline electric
field splitting, being 13.1 K for the A1 compound and 3.7 K for the other one. The
linear terms in the specific heat are 25 and 72 mJ/(moleK2), respectively. Based
on the difference in the magnitude of the magnetisation a [ + 1/2) ground state is
inferred for CeCuGa3 and a I + 3/2) ground state for CeCuA13.

2.1.3. UT2X2
Magnetism in UT2X2 have been studied quite extensively, partly because interesting
phenomena are present, but of course also because of the presence the heavy fermion
antiferromagnetic superconductor URu2Si2 in this series. Except for the so-called
low-temperature phase of UIr2Ge2, again all compounds crystallize in a tetragonal

@M, 0 T, • Q X
Fig. 5. The BaNiSn3-type crystal structure (after Mentink et al. 1993).
18 G.J. NIEUWENHUYS

structure. The structural and main magnetic properties are given in table 3 based on
the review paper by Endstra et al. (1993b). In those cases where neutron diffraction
has been carried out, it was generally found that the ordered magnetic moment is
directed along the crystallographic c-axis. The antiferromagnetic ordered state in
general consists of ferromagnetic magnetic planes (being the crystallographic basal
planes) stacked antiferromagnetically along the c-axis. This stacking is not always
of the simple + - + - type, some compounds even change their stacking as a
function of temperature, e.g., UNi2Si2 (Lin et al. 1991). This ge.neral anisotropy
can be understood if we assume that the U ions are in the 4+ state. If we may
apply Hund's rule, then the magnetic quantum number J equals 4, which leads to
the possibility of non-magnetic singlet states. Konno (1993) has given algebraic
expressions for the eigenvalues and eigenstates for the J = 4 case in a tetragonal
surrounding. Assuming the CEF picture to be completely valid, the magnetism of the
tetragonal U-based compounds is of the induced type and since the transition-matrix
elements between the low-lying eigenstates are only non-zero for the Jz operator,
the moments are confined to the c-axis (Nieuwenhuys 1987). Although this simple
picture does explain major features, it should be noted that the calculated magnitude
of the ordered moment is generally too large. Also, inelastic neutron scattering
experiments have great difficulty to reveal the CEF levels, which of course might be
due to a large broadening by hybridization. Nevertheless, the crystalline electric field
plays a important role in determining the anisotropy of these compounds. In a number
of cases, table 3 denotes the magnetic ordering as C, meaning complex, implying
that more then one magnetic structure has been found depending on temperature. For
example, UCu2Ge2 orders ferrromagnetically at 100 K and changes its structure to a
+ - - + type of antiferromagnetic order below 50 K. UCo2Ge2 and UIr2Ge2 have
been mentioned twice in table 3 because of the different crystal structures. UCo2Ge2
forms at the melting point in the CaBe2Ge2 structure (P) and transforms at annealing
at 1300 K into the ThCr2Si2 structure (I), thereby increasing its c-axis considerably.
Details of this transformation have been studied by Drost (1995) via in situ neutron
diffraction experiments. Note that this structural transformation is accompanied by
a large change in the magnetism, the I structure orders antiferromagnetically at 175
K, while the P structure is a paramagnet down to 35 inK. UIr2Ge2 also adopts the
P structure at its melting point, but transforms upon annealin~ into orthorhombic
structure (space group Pmmm) with a = 4.054 A, b = 4.195 A and c = 10.25 A.
Here the Nrel temperature increases. There are not many UT2Ge2 compounds with
the 4d transition metal series. Only Rh and Pd exist, for Ru a compound U4RuTGe6
has been found. This is a ferromagnet with Tc = 6.8 K (Mentink et al. 1991, Endstra
et al. 1992b). The magnetic properties of the superconducting compound URu2Si2
have already been mentioned above. As in the case of the Ce-based compounds, the
trends of the magnetic ordering temperatures (including the absence of ordering) can
be described in a simple model based on a calculation of the hybridization strength
(Endstra et al. 1993a). In fig. 6 we display their results for the Si compounds, while
fig. 7 shows the Ge compounds. As can be seen from these figures, the model indeed
explains the trends, also, e.g., the absence of magnetic ordering in the P-phase of
UCo2Ge2. Ab initio self-consistent density-functional band structure calculations in
HEAVY FERMIONS AND RELATED COMPOUNDS 19

TABLE 3
Structural and magnetic properties of UT2X 2. The structure is denoted by I for the inversion symmetric
ThCr2Si2 tetragonal type and P for the primitive CaBe2Ge 2 one. The lattice parameters are in ,~. The
specific heat coefficient, "3', is given in mJ/(mole K 2) only if it is not disturbed by magnetic ordering
and/or CEF effects. The magnetic ordering is denoted by A for antiferromagnetism, F for fen'o- or
ferrimagnetism, C for complex ordering and G for spin-glass type ordering. The temperatures, TM and
TK (TM being the magnetic ordering temperature), and the energy of the first excited state in the CEF
level scheme A1 divided by Boltzmann's constant are in Kelvin.

Compound Struc. a e 3' Magn. TM TK A1 Ref.


UAu2Si 2 I 4.280 10.29 - C 48 - - [1-4]
UCo2Ge 2 I 4.010 9.88 - A 174 - - [5, 6]
UCozGe2 P 4.043 9.30 62 P - - - [7]
UCo2P2 P 3.955 8.97 - C 230 - - [8, 9]
UCozSi 2 I 3.917 9.61 - A 85 - - [3, 10]
UCr2Si2 I 3.911 10.50 - A 30 - - [3, 11]
UCu2Ge2 I 4.063 10.23 - C 100 - - [7, 12, 13]
UCu2Si2 I 3.984 9.95 - F 104 - - [4, 10, 11]
UFe2Ge2 I 4.024 9.96 24 P - - - [7, 12, 14]
UFezSi2 I 3.951 9.53 18 P - - - [14, 15]
UIr2Ge2 P 4.156 9.77 - A 19 - - [16]
UIrzGe~ * - 10.25 - A 33 - - [16]
UlrzSi 2 P 4.087 9.83 105 A 4.9 - - [3, 17]
UNizGe2 I 4.095 9.48 - A 77 - - [7, 10, 12]
UNizSi2 I 3.958 9.51 22 C 124 - - [4, 18, 19]
UOszSi 2 I 4.121 9.68 - P - - - [2]
UPdzGe 2 I 4.200 10.23 - A 140 - - [20]
UPd2Si 2 I 4.121 10.19 - A 97 - - [2, 20]
UPt2Ge2 P 4.330 9.75 - A 72 - - [21]
UPtzSi 2 P 4.197 9.69 32 A 35 - - [22]
URezSi 2 I - - - P - - - [2]
URhzGe 2 ? 4.154 9.76 305 P - - - [23]
URhzSi 2 I 4.012 10.06 - A 130 - - [2, 20]
URu2Si2 I 4.128 9.59 180 A 17.5 - 50 [24]

References:
[1] Rebelsky et al. (1991) [13] Kuznietz et al. (1990)
[2] Palstra et al. (1986a) [14] Szytula et al. (1988a)
[3] Buschow and de Mooij (1986) [15] Szytula et al. (1988b)
[4] Torikachvili et al. (1992) [16] Lloret et al. (1987)
[5] Kuznietz et al. (1989) [17] Dirkmaat et al. (1990a)
[6] Endstra et al. (1991) [18] Lin et al. (1991)
[7] Dirkmaat et al. (1990b) [19] Ning et al. (1991)
[8] Reehuis et al. (1991) [20] Ptasiewicz-Bak et al. (1981)
[9] Trod et al. (1993) [21] Endstra et al. (1992)
[10] Chelmieki et al. (1985) [22] Steeman et al. (1990)
[11] Hiebl et al. (1990) [23] Dirkmaat et al. (1990c)
[12] Endstra et al. (1990) [241 Palstra et al. (1985)

the local approximation ( t r e a t i n g t h e U 5 f s t a t e s as b a n d s t a t e s ) b y S a n d r a t s k i i a n d


Ktibler (1994) on the UT2Si2 compounds correctly predict the occurrence of mag-
n e t i s m i n t h e s e m a t e r i a l s . U R u 2 S i 2 f o r m s i n d e e d a n e x c e p t i o n i n t h e s e n s e t h a t it is
p r ed i c t e d as n o n - magnetic, whereas the experiment shows a very small magnetic mo-
20 G,J. NIEUWENHUYS

tI 1"

T K iI tI T~,Ky
£-- iI
I /
J
i
UT2Si 2
/ /
I/
I/
H
g

Ir OS jd_f/W

Fig. 6. Schematic phase diagram for the Kondo lattice together with magnetic ordering temperatures of
the UT2Si 2 compounds. The dashed curves indicate the Kondo and the RKKY temperature. The thick
curve indicates the effective magnetic ordering temperature (after Endstra et al. 1993a).

P UT2Ge 2

C o ~r

j Fe co-
Rh
Jo_/W
Fig. 7. Schematic phase diagram for the Kondo lattice together with magnetic ordering temperatures of
the UT2Gez compounds (after Endstra et al. 1993a).

ment. In the compound UCo2P2, the Co atoms also have a small magnetic moment
so that this compound does not fit into the total picture. Another P containing com-
pound is UCuzP2, which crystallizes in the hexagonal CaA12Si2 structure (Zolnierek
et al. 1987) was reported to be a ferromagnetic with the record breaking Tc of 216 K
(Kaczorowski and Tro6 1990). Magneto-optical spectroscopy has been carried out on
this compound by Schoenes et al. (1989) and Fumagalli et al. (1988). They found a
max~imum Kerr rotation of 3 ° and could characterize UCu2P2 as a semi-metal. Polar-
ized neutron diffraction experiments (Delapalme et al. 1994) suggest that U is in the
3+ state in this compound and that the U 6d-electrons play an important role. In the
course of the search for new boron-carbide superconductors, two new U compound
have been found by Takabatake et al. (1994b): UNi2B2C and URh2B2C. Both crys-
tallize in the ThCrzSi2 structure with a = 3.513 ,~, c -- 10.54 ,~ and a -- 3.782 ,~,
c = 10.214 ,~, respectively. UNizBzC is a antiferromagnetic with the remarkable
HEAVY FERMIONS AND RELATED COMPOUNDS 21

200 0.28
U(C°'-xNix;2Ge2 - 1
150
\ t 0.26 ¢
100
0.24 >
50

0
0.0
. . . . . . . . . .
0.5
1
1.0
0.22

x
Fig. 8. Magnetic ordering temperature (n) and calculated Vdf (o) versus x for U(Col_~Ni~)2Ge2
compounds (after Endstra et al. 1993c).

high TN of 218 K and URh2B2C orders ferromagnetically at also a relatively high


temperature of 185 K. The similarity between the crystal structures and the lattice
parameters enables, as mentioned before, the study of pseudobinary compounds.
This tuning of the various parameters makes comparison with Endstra's model more
challenging. The U(Col_xNix)2Ge2 system has been studied by Endstra (1992) and
Kuznietz et al. (1990b). The U(Col_xCux)2Ge2 system has been a subject of an-
other paper by Kuznietz et al. (1990a), while the compounds U(Nil_zCux)2Ge2 were
reported by Kuznietz et al. (1992a, b). Finally, the U(Co1_~Fe~)2Gez compounds
has been investigated by van Rossum (1993). The U(Col_zNix)zGe2 is particularly
interesting, since looking at fig. 7, one would expect the transition temperature
to vary monotonously between the parent compounds. However, a minimum was
found, which could be explained easily by the occurrence of a maximum in the hy-
bridization strength (both parent compounds are on the right-hand side of the phase
diagram). Figure 8 shows this result.

2.2. l l l-compounds
2.2.1. CeTX compounds
The CeTX compounds crystallize in a large variety of structures, resulting in a
broad spectrum of magnetic properties. It is therefore rather difficult to describe a
general line. However, one thing is clear, there is always a relation between the
volume of the unit cell and the valency of Ce. As pointed out above, the latter is
of essential importance for the magnetic behaviour. Crystalline electric field effects
do play a important role, they can be observed via inelastic neutron scattering.
No extremely small splittings between the ground state and the first excited state
have been found in this series. Remarkably, the influence of the CEF is mentioned
always in the literature, but it is seldom taken into account, e.g., when analyzing the
temperature dependence of the electrical resistance. Table 4 gives the main structural,
magnetic en electrical properties of the CeTX compounds. The entries have their
usual meaning; under the head 'conduc.' there is an M for metallic behaviour and
an S for semiconducting behaviour. The latter may occur only at low temperatures.
22 G.J. NIEUWENHUYS

TABLE 4
Structural and magnetic properties of Ce 111-compounds. 'Crys.' denotes the symmetry of the lattice,
'Struc.' the structure type. The lattice parameters are given in ,~. Under the heading 'Magn.' an A
means antiferromagnetism and F ferromagnetism. TM is the magnetic transition temperature in Kelvin.
An M under 'Conduc.' means metallic and S semiconducting.

Compound Crys. Struc. a b c Magn. TM Conduc. Ref.


CeA1Ga hex A1B 2 4.378 - 4.329 A 6 M [11
CeAuIn hex Fe2P 7.698 - 4.256 A 5.7 M [2]
CeCuGe hex Ni2 4.311 - 7.933 A 10.2 M [3, 4]
CeCuSi hex NizIn 4.238 - 7.988 A 14.9 M [3, 4]
CeCuSn hex NizIn 4.583 - 7.865 A 8 M [3-5]
CeIrGe o.rh TiNiSn 7.073 4.374 7.575 - - M [6]
CeNiAI hex ZrNiA1 6.978 - 4.020 - - M [3, 7-10]
CeNiGe o.rh TiNiSn 4.317 7.388 7.201 - - M [3, 7]
CeNiIn hex ZrNiA1 7.552 - 3.986 - - M [3, 7, 11-14]
CeNiSb hex A1B2 4.384 - 4.110 F 7 M [15, 16]
CeNiSi tetra LaPtSi 4.061 - 14.045- - M [3, 7]
CeNiSn o.rh E-TiNiSi 7.522 4.555 7.602- - S [3, 7, 17-24]
CePdAI hex ZrNiA1 7.221 - 4.233 A 2.7 M [25-27]
CePdGa o.rh TiNiSi - - - A 2.2 M [28-31]
CePdGe o.rh CeCu2 4.488 7.300 7.676 A 3.4 M [6, 31]
CePdln hex ZrNiA1 7.709 - 4.078 A 1.7 M [11, 12, 32-37]
CePdSb hex GaGeLi 4.935 - 7.890 F 17 M [38-40]
CePdSn o.rh c-TiNiSi 7.535 4.700 7.955 A 7.5 M [19, 20, 30, 31, 41, 42]
CePtGa o.rh TiNiSi - - - A 3.5 M [31, 43]
CePtGe o.rh CeCu2 4.45l 7.344 7.616 A 3.4 M [6, 31]
CePtIn hex FezP 7.657 - 4.069 - - - [11, 44]
CePtSb hex GaGeLi 4.533 - 8.058 F 4.5 M [45]
CePtSi tetra LaPtSi 4.202 - 14.484- - M [10, 46--49]
CePtSn o.rh ~-TiNiSi 7.463 4.628 8.016 A 7.5 M [21, 31, 50, 51]
CeRhGe o.rh TiNiSi 7.430 4.466 7.120 A 9 M [6]
CeRhIn hex Fe2P 7.547 - 4.050 - - M [32, 52]
CeRhSb o.rh CeCu2 4.609 7.416 7.846 - - S [53-55]

References:
[1] Fremy et al. (1989) [19] Nakamura et al. (1991)
[2] Pleger et al. (1987) [20] Koghi et al. (1992)
[3] Nakotte (1994) [21] Yamaguchi et al. (1990)
[4] Fuming et al. (1991) [22] Kyogatu et al. (1990)
[5] Sakurai et al. (1993) [23] Kalvius et al. (1993)
[6] Rogl. et al. (1989) [24] Nohara et al (1993)
[7] Kuang et al. (1992) [25] Hulliger (1993)
[8] Buschow (1979) [26] Schank et al. (1994)
[9] Singhal et al. (1993) [27] Kitazawa et al. (1994)
[10] Lee et al. (1987b) [28] Hovestreydt et al. (1972)
[11] Satoh et al. (1990) [29] Adroja et al. (1994)
[12] Fujii et al. (1987) [30] Rainford et al. (1994)
[13] Fujii et al. (1989) [31] Sakurai et al. (1989)
[14] Takabatake et al. (1990b) [32] Bruck et al. (1993)
[15] Pecharski et al. (1983) [33] Adroja et al. (1991)
[16] Skolozdra et al. (1994) [34] BrUck et al. (1988)
[17] Aliev et al. (1990) [35] Maeno et al. (1987)
[18] Takabatake et al. (1990) [36] Suzuki et al. (1990)
HEAVY FERMIONS AND RELATEDCOMPOUNDS 23

[37] Fujii et al. (1990) [47] Krimmel et al. (1992)


[38] Malik and Adroja (1991b) [48] Ott et al. (1992)
[39] Trovarelli et al. (1994) [49] Mielke et al. (1993)
[40] Riedi et al. (1994) [50] Adroja et al. (1988)
[41] Adroja et al. (1992) [51] Takabatake et al. (1993)
[42] Fujita et al. (1992) [52] Adroja et al. (1989)
[43] Uwatoko et aL (1995) [53] Malik and Adroja (1991a)
[44] Fujita et al. (1988) [54] Takabatake et al. (1994)
[45] Rainford and Adroja (1994) [55] Nishigori et al. (1994)
[46] Lee et al. (1987a)

Two compounds show the S for the conductance, CeNiSn and CeRhSb. The first
one shows semiconducting behaviour for temperatures below 6 K, signalled by a
rapid increase in the electrical resistance and a sharp drop in the specific heat.
The size of the corresponding energy gap is only 2.4, 5.5 and 5 K for the current
along the a-, b- and e-axis, respectively (Takabatake et al. 1990b). Also, nuclear
magnetic resonance (Kyogatu et al. 1990) reveals the gap via a more rapid decrease
of the relaxation rate, l/T1, as function of temperature. The Knight shift follows
the magnetic susceptibility, exhibiting a maximum at 12 K. The experimental results
could be described assuming that the structure of the gap is that of a pseudo gap, i.e.
the density of states varies linearly as function of energy in stead of making a jump.
Preliminary vacuum tunneling experiments (Aarts and Volodin 1995) confirm this
picture. Within the CeNiX compounds, CeNiSn is situated at the end of a monotonic
gradual trend, starting with A1, towards integer valency. CeNiA1 is a typical mixed
valence compound, while CeNiSn for temperatures above 6 K should be characterized
as a medium heavy fermion with 7 = 173 mJ/(moleK2). CeNiSn does not show
magnetic ordering. If Ni is replaced partly by Pd, as done by Kasaya et al. (1991),
then magnetic ordering starts at CeNi0.8Pd0.zSn; CePdSn is a ferromagnet with the
same crystal structure as CeNiSn. CeRhSb shows the semiconducting behaviour
at temperatures below 20 K, again a pseudo gap is found with a width of about
4 K. It should be noted, that in both cases, CeNiSn and CeRhSb, the corresponding
nonmagnetic compound (with La in stead of Ce) has a normal metallic behaviour.
We will come back to this point later. Also, 20% Ce replaced by La in CeRhSb
recovers the metallic conductance. CePdSb is a ferromagnet with Tc = 17 K as
found from magnetic measurements, which is a remarkable high temperature for
Ce compounds in this series. The corresponding Nd, Gd and Sm compounds have
ordering temperatures of similar values. Ce is here in the trivalent state, the CEF
excitations are found at 240 and 560 K, respectively. The absolute value for the
electrical resistance is quite high, 4 m~2cm. The specific heat, however, shows
no sharp feature at Tc, only a swallow maximum around 10 K has been found.
No clear explanation has been given yet. CePtIn shows no magnetic order. In
contrast the electronic contribution to the specific heat has a coefficient of 1100
mJ/(mole K 2) at T = 0 K due to an upturn with decreasing temperature starting at
0.5 K. This would characterize this compound as extremely heavy. However, an
external magnetic field of 5 T suppresses this -,/ value down to 500 mJ/(mole K2).
The overall temperature dependence can be described by a Kondo contribution with
TK = 11 K and a CEF splitting of the lowest two levels of A = 80 K. CePdIn,
24 G.J. N1EUWENHUYS

with 7 = 700 mJ/(moleK2), orders ferromagnetically at Tc = 1.7 K and CeRhln


has an unstable moment. Mixtures of these two latter compounds show a gradual
change of the Ce valency, accompanied by a related change of the volume of the
unit cell (Brtick et al. 1993). Going from Rh towards Pd the hybridization and
the mass enhancement (7) increase with a maximum for 3' at 800 mJ/(mole K 2) just
before the ordering sets in at 80% Pd. Replacing Pd by Pt (Fujita et al. 1992)
causes the magnetic order to disappear. CePtIn shows no magnetic order down
to 60 mK. CePtSb and CeA1Ga both order antiferromagnetically at 4.5 and 6 K,
respectively. In order to fit the data for the crystalline electric field, the addition
of a 043 term is necessary. For CePtSb this could be related via detailed neutron
diffraction to the GaGeLi structure, with is an atomically ordered version of the
Cain2 structure. Probably, this implies that CeA1Ga is also not exactly in the A1B2
structure. CePtSi can be characterized as a heavy fermion system: specific heat
measurements have shown a upturn in C/T starting at 10 K and resulting in a
zero temperature extrapolation of about 700 mJ/(mole K2). In this case, an external
magnetic field of 13 Tesla does not change essentially the results, ruling out short
range magnetic order as a cause of the specific heat increase. Via replacing Pt by Ni,
Lee et al. (1987b) have shown the gradual change from a heavy fermion system to a
typical mixed valence behaviour. The magnetic order in case of antiferromagnetism
is generally not of a simple type. For example, CePtSn (Sakurai et al. 1989) and
CeCuSn (Nakotte 1994) show two peaks in the specific heat. Both anomalies are
influenced by external magnetic fields in different ways. To explain these type of
phenomena, detailed microscopic investigations should be carried out.

2.2.2. UTX-compounds
The equiatomic UTX compounds have been studied quit extensively, starting with
the work of Dwight (1974), Buschow et al. (1985) and Palstra et al. (1986b, 1987a,
b). A review has been given by Sechovsky and Havela (1988) in vol. 4 of this book
series and by Sechovsky et al. (1990). Therefore, I will restrict here to some new
developments. As is the case with the Ce compounds, different crystal structures
are found in the UTX system. The structure seems to be governed by the number
of d-electrons of the transition metal atom (Nakotte 1994), starting with the hexag-
onal MgZn2 structure for Mn-, Tc- or Re-compounds, though the hexagonal Fe2P
or orthorhombic CeCu2 structures towards the cubic MgAgAs and hexagonal Cain2
structures for the compounds containing Cu, Ag or Au. Compounds with the cubic
MgAgAs structure are semiconductors, the other ones show metallic behaviour. Note
that the corresponding non-magnetic Th compounds with the MgAgAs structure are
also semiconductors, so that in these cases the magnetism is not a condition for
the semiconducting behaviour. UNiSn is a rather famous compound crystallizing in
the cubic MgAgAs structure. It undergoes an exotic phase transition from semicon-
ducting paramagnetism for T > TN (= 43 K) to a metallic antiferromagnetic phase
below TN (Palstra et al. 1986b, Fujii et al. 1989b). Neutron diffraction experiments
have shown UNiSn to order in a type I antiferromagnetic structure (Kawanaka et al.
1989), while indications for a lattice distortion were found from far-infrared mea-
surements (Kilibarda-Dalafave et al. 1993). The crystalline electric field scheme has
HEAVYFERMIONSAND RELATEDCOMPOUNDS 25

been deduced by Aoki et al. (1993) to consist for the U 4+ ion of a non-magnetic
doublet ground state and a first excited triplet state at 180 K. From these scheme it
can be shown that the phase transition in UNiSn should be of first order and includes
the spins as well as the lattice, which was nicely confirmed by Suzuki et al. (1994)
by powder X-ray and ultra-sonic measurements. The series URhIn, UPtIn and UPdIn
(Nakotte et al. 1992, Havela et al. 1994a, Sugiura et al. 1990) can be classified as
magnetic heavy fermions; the strength of the magnetism increases when going from
Rh towards Pd, i.e. the moment increases from 1.2 to 1.6#B and TN increases from
7 to 20.4 K. This trend can be explained be the decrease in the hybridization. From
the specific heat there are indications that the bandstructure changes at TN, optical
and tunneling experiments should clarify this point. UPdSn and UAuSn both order
antiferromagnetically at 37 and 40 K, respectively (De Boer 1992, Robinson et al.
1993a, 1994a, Nakotte et al. 1993a, b). From neutron diffraction it appears that both
compounds have a very rich magnetic phase diagram. In both cases the magnetic unit
cell of orthorhombic symmetry, while the crystallographic cell is hexagonal. This im-
plies that different domains are possible. UPdSn has two transitions in zero external
field. At both transition temperatures, 40 and 25 K, different Cartesian components
of the magnetic moments order. Similar effects have been seen in the pure rare-earth
metals. The electrical resistances of the two compounds is rather different, the one
of UPdSn decreases by a factor of 4 below the magnetic ordering temperature, the
resistivity of UAuSn increases gradually with decreasing temperature over the whole
range from room temperature to the lowest temperatures. UPdGe, UPtGe and UNiGe
also show quit complicated magnetic structures (Kawamata et al. 1992, Proke~ et al.
1994, Robinson et al. 1994a, b, Havela et al. 1994a and Sechovsky et al. 1994a, b).
The electronic properties of UNiA1 have been described by Brtick et al. (1994). The-
oretical or phenomenological predictions of this variety of magnetic structures are
not (yet) possible, due to the delicate balance between the different interactions. The
latter also implies that excellent sample quality is essential for this research. In all
cases there is a huge magnetocrystalline anisotropy due to several reasons: crystalline
electric fields, anisotropic interaction and/or anisotropic hybridization. As a general
trend, as far as single crystals have been investigated, compounds crystallizing in
the ZrNiA1 structure have their ordered moments directed along the c-axis, while the
moments in the TiNiSi derived compounds prefer the b-c plane. Other structures
reveal more complicated anisotropies. Evidence for anisotropic hybridization stems
from detailed neutron diffraction studies on URhA1, URuA1 and UPdSn, where also
a large orbital moment for the essentially itinerant 5f-electrons is found (Paixao et
al. 1993, Nakotte 1994). Measurements of this kind (U-formfactor) as a function
the external magnetic field - and thus for metamagnetic transitions as a function
of the magnetic structure - should be able to shed light on the intriguing questions
concerning the magnetic order in heavy fermion compounds (Wulff et al. 1990).
The electrical resistance of the 111-compounds (and of heavy fermions in general)
is rather high, several hundreds of # ~ cm. Also, it sometimes drastically changes
as a result of magnetic order, although, as mentioned above, there are exceptions
where the temperature dependence of the electrical resistance does not change at
all at the ordering temperature. This implies that a simple conduction electron vs.
26 G.J. NIEUWENHUYS
1.6

|.0
/
v
/
UNIGa
0.5 / T : 4.2 •
t

/
. t l v v . . . . v

15o,
m ~ x ~ UNiGa
100
~ T = 4.2K
E V A
50 ~ i // ~ // c

0,

0
0.0 0.5 1.0 1.5
B 03
Fig. 9. Field dependence of the magnetization for BIle and of the electrical resistivity of UNiGa at 4.2
K for i[[BlJc and i±Blle. The open symbols represent the result obtained with increasing field, while the
full symbols represent the one obtained with decreasing field (after Nakotte 1994).

local moment scattering picture is much too simple. There is in general a very strong
interaction between the conduction electrons and the magnetic degrees of freedom of
the 5f-electrons. The latter are also believed to take part in the electrical conductance.
Then, the magnetic structure is going to play an important role, since the conduction
electrons now also have to obey the symmetry of the magnetic unit cell. Conductivity
calculations should now be carried out on the basis of band-structure calculations
which include the various magnetic structures. We will come back to this point. For
the case of the UTX compounds it is interesting to note that very large magneto-
resistance effects have been observed, larger than in the multilayer materials, see,
e.g., Havela et al. (1994a), de Boer et al. (1994) and Nakotte et al. (1994a, b). A
typical example is UNiGa, which orders antiferromagnetically at 40 K (Havela et
al. 1991), the ground state is characterized by the sequence + + - - + - . At low
temperatures, an external magnetic field of 1 Tesla is sufficient to align the moments
along the c-axis, which leads to a reduction of the electrical resistivity to only 15%
of its zero-field value. In fig. 9 the magnetization and the resistivity are displayed.

2.3. UzT2X-compounds
A new family of heavy fermion compounds have been synthesized with the compo-
sition UzTzX (Peron et al. 1994, Nakotte et al. 1994b, Mirambet et al. 1993, 1994,
Nakotte 1994, Havela et al. 1994b). The compounds with T -= Co, Rh, Ni, Pd, Ir, Pt
HEAVYFERMIONSAND RELATEDCOMPOUNDS 27

C) U
OT
c
@X

Fig. 10. Schematicrepresentation of the unit cell of U2T2X crystallizingin the U3Si2 structure. The
two shortest inter-uraniumdistances are indicated(after Nakotte 1994).

and X = In, Sn all crystallize in the tetragonal U3Si2 structure, depicted in fig. 10.
The shortest distances between the uranium atoms in this series of compounds ranges
from 3.40 ,~ to 4.00 A, that is just around and above the famous Hill limit of 3.5 ,~
(Hill 1970). Therefore, these compounds are expected to cover the range between
non-magnetic behaviour via heavy-fermion behaviour towards magnetic ordering. In
table 5 the structure parameters, together with the magnetic ordering temperature
and the linear coefficient for the specific heat are given. All magnetic ordering is
antiferromagnetic. Note, that the value for 7 is still in mJ/(mole/.~. K2), but that
the formula unit now contains two uranium atoms. High-field magnetization exper-
iments (Nakotte et al. 1994b) have confirmed the antiferromagnetism by revealing
metamagnetic transitions and confirmed the paramagnetism of the other compounds
by showing that the magnetization is a linear function of the external field at higher
fields. Calculations of the strength of the hybridization shows that indeed the mag-
netically ordering compounds have a lower value than the paramagnetic ones, placing
this series of compounds to the right-hand side of the Doniach phase diagram.

2.4. U-compounds with CaCu5 structure and related borides


A number of rare-earth intermetallic borides are excellent magnet materials with
strong uniaxial magnetocrystalline anisotropy, see, e.g., Strnat (1988). Uranium
borides, unfortunately, have not so many useful properties. The main structural
and magnetic properties of a number of these compounds are given in table 6. The
heading 'Magn.' now shows AF for antiferromagnetism, PP for Pauli paramagnetism
and SF for spin fluctuator, meaning that the magnetic susceptibility is still rather
temperature independent, but that the electrical resistivity shows a heavy-fermion
like decrease for temperatures decreasing below ~ 150 K. Because of the similarity
of the structure, also some 123-compounds are listed in table 6, including the new
28 G.J. NIEUWENHUYS

TABLE 5
Structural and magnetic properties of U2T2X. The lattice parameters are in ,~. The
specific heat coefficient, "7, is given in mJ/(mole K 2) only if it is not disturbed by
magnetic ordering and/or CEF effects. TN is the N6el temperature in Kelvin.

Compound a c TN
U2Co2In 7.361 3.431 - 78
U2Ni2In 7.375 3.572 14.3 -
UzRhzIn 7.553 3.605 - 237
UzPdzIn 7.637 3.752 36.6 -
UzPtzIn 7.654 3.725 - 850
UzCo2Sn 7.654 3.725 - 261
U2Ni2Sn 7.263 3.691 26.0 -
U2Rh2Sn 7.524 3.630 24.4 -
U2Pd2Sn 7.603 3.785 40.6 -
U21r2Sn 7.566 3.601 - 98
U2Pt2Sn 7.670 3.698 - 305

TABLE 6
U-compounds with CaCu5-type structure and related borides. 'Crys.' denotes the symmetry of the
lattice, 'Struc.' the structure type. The lattice parameters are given in A. The specific heat coefficient,
7, is given in mJ/(mole K 2) only if it is not disturbed by magnetic ordering and/or CEF effects. Under
the heading 'Magn.' A F means antiferromagnetism, S F spinfluctuator and P P Pauli paramagnet. TM
is the magnetic transition temperature in Kelvin.

Compound Crys. Struc. a c 3' Magn. TM Ref.


U(Coo.25 Nio.75)4B hex CeCo4B 4.931 6.964 - AF 5 [1]
U(Coo.5 Nio.5)4B hex CeCo4B 4.924 6.954 73 SF - [1]
U(Coo.75 Ni0.25)4B hex CeConB 4.910 6.928 - SF - [11
U(Ruo.7Rho.3)4B 4 tetra LuRu4B 4 - - 114 PP - [2]
UCosB 2 hex CeCosB 2 4.947 3.065 48 SF - [3, 4]
UCo4B hex CeCo4B 4.895 6.933 - SF - [5]
UCo4B 4 tetra CeCo4B 4 5.027 7.030 52 PP - [2]
UFesB 2 hex CeCosB 2 5.047 2.997 32 PP - [3, 41
UlrsB 2 hex CeCosB 2 5.373 3.181 64 SF - [3, 4]
UNi2A13 hex PrNi2A13 5.204 4.018 - AF 5.2 [6-81
UNi4B hex CeCo4B 4.950 6.962 - AF 21 [1, 9]
UOssB2 hex CeCosB 2 5.511 2.971 37 PP - [3, 4]
UOs4B 4 tetra LuRu4B 4 - - 55 PP - [2]
UPd2A13 hex PrNi2A13 5.365 4.191 - AF 14 [6,7, 10, 11]
UPd2Ga 3 hex CeCo4B 5.295 8.505 - AF 13 [12]
URusB 2 hex CeCosB 2 5.473 2.965 21 PP - [3, 41
URu4B 4 tetra LuRu4B 4 7.460 15.000 88 SF - [2]

References:
[1] Mentink et al. (1994e) [7] Geibel et al. (1992)
[2] Mentink et al. (1992) [8] Scr6der et al. (1994)
[3] Yang et al. (1984) [9] Mentink et al. (1994b)
[4] Mentink (1994) [10] de Visser et al. (1992)
[5] Nakotte et al. (1993c) [11] de Visser et al. (1993)
[6] Geibel et al. (1991b) [12] Siillow (1994)
HEAVYFERMIONSAND RELATEDCOMPOUNDS 29

magnetic heavy fermion superconductors UPd2A13 and UNi2A13. Inspired by the


anomalous ferromagnetism in CeRh3B2 (Dhar et al. 1981) with Tc = 115 K, CeT3B2
and UT3B2 materials have been investigated by Yang et al. (1984, 1985), Ku et al.
(1980), Kasaya et al. (1990) and Mentink et al. (1994e). The compounds crystallize
in CeCo3B2 structure, see fig. 11. All the uranium compounds of this series are
either Pauli paramagnets or spin fluctuating materials with rather small temperature
independent 7 values. This can be understood from the small inter-uranium distances
(< 3.5 ]k) causing an overlap of the 5f-wavefunctions. UCo3B2 and UIr3B2 have
somewhat larger distances and therefore exhibit the spin fluctuation phenomena as
also found in typical heavy fermions. The 141-compounds order in the CeCo4B
structure, depicted in fig. 12. The formation of a magnetic moment on the U-site

c I 3g

la

a 2c

@U, © T, eB

Fig. 11. The CeCo3Bz-typestructureas adoptedby the UT3B2compounds(afterMentink1994).

c ] ~ l b ~- - ~ z~-~--~__-_.
- - = ~©
: 6i

la
~: a
u 2c

O U, © T, • B
Fig. 12. The CeCo4B-typestructureas adoptedby the UT4B compounds. Notethe presenceof two
differentU- and T-atomsites. The uraniumatomshave a triangulararrangementin the basalplane(after
Mentink 1994).
30 G.J. NIEUWENHUYS

d
',,o/",o , ' \ o ,' ',,o/",o , ' \
,i_o ok
/o,, O/o',,o ,,'o',,
o ,,'0',,
O/o',,o /
~ ..... ,~, .... ~ ,,

',, o .-' ,-, oo ,'o, -.,,o


, , ,' o'.,._ o . ' ,o , ° , , ' o \
,'t
'~' ~ ~ - - ~ d ....~,
/\7,,' ', o / " , , o / \ o ,,, ',, o f
,d o__k, o , /
Fig. 13. Magnetic structure of hexagonal UNi4B, projected on to the crystallographic basal plane.
Given the FM coupling along the e-axis, every layer at z = 0 and z = 1/2 exhibits the same moment
orientations. The thin solid line represents the magnetic unit cell. Note the presence of the 'free' uranium
moments with two different magnetic environments (after Mentink 1994).

in U(Co~Nil_~)4B can be understood since the d-band shifts downwards in energy


with decreasing z (more Ni-rich), thereby decreasing its density at EF, while also
the distance between the U-atom and the transition metal increases. Both effects
will decrease the hybridization and enable the formation of moments on the U-
site. At the end of this series, nature has placed a remarkable partially ordered
antiferromagnet: UNi4B. In the ordered state, the magnetic moments are confined to
the basal-plane of the crystal structure. According the neutron diffraction experiments
(Mentink et al. 1994b) two-thirds of the magnetic moments order in a vortex-like
structure in the basal plane, while being ferromagnetically aligned along the c-axis,
see fig. 13. The other one-third of the magnetic moments are free to rotate in
the basal plane. They are only coupled ferromagnetically along the c-axis, but a
one-dimensional ferromagnetic chain does not order at finite temperature. Note that
the nearest-neighbour coupling in the basal plane cancels for all magnetic moments,
while the next-nearest-neighbour ones cancel for the 'free' moments. The unique
magnetic structure is also reflected by the magnetic susceptibility, see fig. 14. The
low-temperature upturn in small external magnetic fields represents the response of
the ferromagnetic linear chains, which can easily be saturated by larger fields. A
similar extra contribution is found in the specific heat. Also the large anisotropy is
evident from these susceptibility results. This anisotropy of about a factor of ten
remains present in the paramagnetic region, it is therefore not caused by a spatial
anisotropy of the magnetic interaction between the moments. Probably, crystalline
HEAVY FERMIONS AND RELATED COMPOUNDS 31

| i
o m Ooooo
300

*~. °o o
"6" ~'~.o
O
200 "

O
100
X
U~m annjllmmIe ° ° ° ° ° °°e • • o ooe eoo ooo ooo ooql • • • •

, I n I ,
0
0 10 20 30

T (K)
Fig. 14. Magnetic susceptibility of a UNi4B single crystal, measured in fields of 0.3 mT (o), 0.5 T (+)
and 5 T (rn), perpendicular to the c-axis. The lower points (-) represent the result for a field of 0.5 T
directed parallel to the c-axis (after Mentink 1994).

electric field effect and anisotropic hybridization dictate the magnetic moments to lie
preferentially in the crystallographic basal plane. Small steps have been observed
in the magnetization versus field curves at low temperatures for fields parallel to
the a-axis and pm'allel to the b-axis. These could be explained by assuming a small
6-fold anisotropy in the basal plane (Mentink et al. 1994c). Remains the question
why one-third of the spins behave differently from the others. The same neutron
diffraction experiments also showed that UNi4B does not crystalize exactly in the
CeCo4B structure, but that it adopts a superstructure with a unit cell of size av/3 with
the axis rotated by 30 ° with respect to the CeCo4B cell. This superstructure persists
to temperatures above TN. Presently the atomic positions are being investigated by
single crystal X-ray diffraction. The new heavy fermion superconductors, UPd2A13
and UPd2A13 crystallize in the PrNi2A13 structure, which is depicted in fig. 15. This
structure is an ordered variant of the CaCu5 type, where 3 of the Cu atom are replaced
by A1. The existence of magnetic moments in these compounds is caused by the -
for the CaCu5 structure - relatively large volume of the unit cell, leading to inter-
uranium distances larger than Hill's limit. The Ni system has an ordered magnetic
moment of (0.24 + 0.1)#B (Schr6der et al. 1994), the transition temperature is 5.2 K
and the compound becomes superconducting below a temperature of about 1 K. Its
Pd counterpart has a larger ordered moment of 0.85#B (Krimmel et al. 1992b), an
higher TN of 14 K and a superconducting transition temperature of ~2 K. The exact
temperature of the transition to superconductivity depends on the sample preparation
(Sakon et al. 1993). We will come back to that point. The magnetic structure of
UPd2A13 consists of ferromagnetic layers in the crystallographic basal planes with
32 G.J. N1EUWENHUYS

c I 3g

la

> 2c
Et

U, © Pd, • A1
Fig. 15. The PrNi2A13-type structure adopted by UPd2AI3 and UPd2A13. The uranium-transition metal
layers are separated by an aluminium layer at z = 1/2 (after Mentink 1994).

250 i |
2.00

200
©

0
E 150

o 100
x
50

i I i I i I

0 100 200 300

T (K)
Fig. 16. Magnetic susceptibility of UPd2A13 along the a-axis (o) and c-axis (+). The solid lines
denote a fit to a CEF model with a singlet ground state described in the text. The inset shows the high-
field magnetization at T = 4.2 K parallel to the a-axis. Again, the solid line represents the calculated
magnetization (after de Visser et al. 1993 and Mentink 1994).

the m o m e n t s in the plane. The layers are stacked antiferromagnetically along the
c-axis. For UNi2A13 the magnetic structure is slightly different, the ordering wave
vector is n o w (½ + 3 , 0 , ½), with 5 = 0 . 1 1 0 + 0 . 0 0 3 . In both compounds the magnetic
order persists into the superconducting phase as evidenced by neutron diffraction
HEAVY FERMIONS AND RELATEDCOMPOUNDS 33

and #+SR experiments (see Krimmel et al. 1993 and Schenck and Gygax in this
volume). The magnetic susceptibility of UPd2A13 is strongly anisotropic as de-
picted in fig. 16. The experimental results were explained on the basis of a CEF
1(
model for the U 4+ ion with a singlet ground state, /"4 = ~ I + 3) - I - 3)), a
first excited state, F1 = 10) at 33 K, a third state, /"6 = I + 1) at 102 K, a fourth
state, F}2~ = a I q: 2) - bl + 4) at 152 K and other states at much higher energy
(Grauel et al. 1993, Mentink 1994). The magnetic moment of UPd2A13 then is
assumed to be due to an admixture of the first and second excited state as a con-
sequence of an external and/or internal magnetic field. Note that this admixture
will give rise to large moments in the basal plane. Any magnetic moment induced
along the c-axis must come from an admixture between the/"4 and/"3 states, hav-
ing an energy difference of about 560 K and is therefore relatively small. The
maximum in the susceptibility versus temperature curve is due to the depopulation
of the first excited state with decreasing temperature. The step in the magnetiza-
tion as function of external field (inset to fig. 16) can be understood as a jump
to ferromagnetic alignment of the moments due to the external field. Similar ex-
planations for the magnetic susceptibility and the magnetization as a function of
external fields were given in the case of URu2Si2 (see above). In that case (due
to the tetragonal crystal symmetry) the level scheme is such that a field along the
c-axis causes the largest admixture, and thus the moments are directed only along
the c-axis, i.e. the basal-plane component simply does not exist. Note, however,
that direct observations of the crystalline electric field levels by inelastic neutron
scattering have not been successful up to now. The value of the ordered magnetic
moment as well as the magnetic anisotropy have also been successfully explained
on the basis of self-consistent density-functional calculations by Sandratskii et al.
(1994), where the U 5f states are treated as band states. The magnetism in UPd2A13
is rather robust versus doping with other elements, this in contrast to the super-
conductivity (Geibel et al. 1994). In order to investigate the sample dependences
further, Mentink et al. (1994d) have studied the sister compound CePd2A13. This
compound does not become a superconductor so that the investigations can be fo-
cussed on the magnetism. In poly-crystalline form, after sufficient heat treatment,
CePdaA13 orders antiferromagnetically at 2.7 K. However, as-cast samples as well
as the best single crystals show at most only short range magnetic order. Such a
discrepancy has also been observed in CePt2Sn2. Detailed nuclear quadrupole res-
onance experiments have shown that, e.g., in as-cast polycrystalline samples 11%
of the A1 experience a different surrounding from the other 89%. In view of the
number of nearest neighbours of the A1, this would imply that 2% of the A1 atom
are wrongly placed. In the annealed polycrystal this number is about 1%. This
experiment shows in a quantitative way the extreme sensitivity of the magnetism
in heavy fermions compounds for the exact structure of the samples, and thus for
the sample preparation. In other words this experiment emphasizes the notion that
the ground state in heavy fermions is a result of a delicate balance between interac-
tions.
34 G.J. NIEUWENHUYS

2.5. 334-compounds
A remarkable set of compounds is formed by the U- and Ce-containing M 3 T 3 X 4
compounds, with T a transition metal and X is Sb, Sn or Bi. The existence of
the Sb compounds had already been reported by Dwight in 1979, but it took a few
years before the remarkable electronic and magnetic properties were studied in more
detail (Buschow et al. 1985, Takabatake et al. 1990c and Endstra et al. 1990b). All
compounds crystallize in the cubic Y3Au3Sb4 structure (space group 1 2~3d). The unit
cell, depicted in fig. 17, contains 4 formula units. The inter-uranium distances are
relatively large (~ 4.4 A), much larger than the distances between the magnetic atoms
and the transition metal atoms (~ 3.3 A). Therefore, reasonable magnetic moments
are expected, since no direct f - f overlap will occur. Most of the Sb compounds are
semiconductors. In the case of U3T3Sb4 the gap in the density of states is rather
large (,.~ 0.2 eV) in contrast to the tree Ce containing compounds mentioned, where
gaps are found of the order of meV. The U compounds are therefore considered as
'normal' semiconductors, while the Ce compounds are most probably hybridization
induced semiconductors. The normal semiconducting properties were confirmed
by bandstructure calculation (Takegahara et al. 1990). Table 7 summarizes the
structural, electronic and magnetic properties. It should be noted that in the case of
U 3 C u 3 S b 4 electron-microprobe analysis shows that the Cu content was less than 3,
so this is a Cu deficient material. From table 7 it is clear that the semiconducting
compounds have a small coefficient of the linear term in the specific heat, i.e. no
density of states at the Fermi level. Also, these compounds do not order magnetically,
indicating that the conduction electrons are essential for the magnetic interaction.
Again the trend in the magnetic properties in these system can be understood in terms
of the strength of the hybridization. For example, replacing Ni by Cu in the Sn series

Fig. 17. The Y3Au3Sb4 crystal structure type. Small open circles indicate the Y atoms, small patterned
circles are the Au atoms and the large open circles the Sb atom (after Takegahara et al. 1990).
HEAVY FERMIONS AND RELATED COMPOUNDS 35

TABLE 7
334-compounds. The lattice parameter is given in ,~. The specific heat coefficient, % is given in
mJ/(mole K2) only if it is not disturbed by magnetic ordering and/or CEF effects. An M under 'Conduc.'
means metallic and S semiconducting. Under the heading 'Magn.' an A means antiferromagnetism and
F ferromagnetism. TM is the magnetic transition temperature in Kelvin. The effective moment/Zeff is
given in/~B.

Compound a 3' Conduc. Magn. TM #eff Ref.


Ce3Au3Sb4 10.058 - S - - 2.45 [ 1]
Ce3Pt3Bi4 9.998 3 S - - 2.5 [2]
Ce3Pt3Sb4 9.820 0 S - - - [1]
U3Au3Sn4 9.824 280 M - - 3.58 [3]
U3Co3Sb4 9.284 69 M F 10 2.1 [4]
U3Cu3 Sb4 9.415 - M F 91 3.39 [3, 4]
U3Cu3Sn4 9.522 380 M A 12 3.34 [3]
U3Ni3Sb4 9.393 2 S - - 3.65 [3, 4]
U3Ni3Sn4 9.380 92 M - - 1.8 [3, 4]
U3Pd3Sb4 9.684 - S - - 3.58 [3]
U3Pt3Sb4 9.683 - S - - 3.68 [3]
UPt3Sn4 9.675 94 M - - 1.84 [3]
URh3Sb4 9.531 - - F 105 2.4 [4]

References:
[1] Kasaya et al. (1991b) [3] Takabatake et al. (1990c)
[2] Hundley et al. (1990) [4] Endstra et al. (1990)

will d e c r e a s e the f - d h y b r i d i z a t i o n b e c a u s e the d - b a n d shifts a w a y from the F e r m i


level and the lattice p a r a m e t e r increases. In U3Ni3Sn4 the h y b r i d i z a t i o n is too large
to f o r m o r d e r e d m a g n e t i s m and the c o m p o u n d is close to the h e a v y f e r m i o n regime,
while U3Cu3Sn4 is an antiferromagnet. In the s e m i c o n d u c t i n g U3T3Sb4 c o m p o u n d s a
r e p l a c e m e n t o f Ni b y Pd or Pt ( s a m e c o l u m n in the p e r i o d i c system) does not c h a n g e
m u c h in the g r o u n d state, e x c e p t for the size o f the gap. H o w e v e r , substituting Cu for
Ni (different c o l u m n ) drastically changes the features o f the c o m p o u n d s as was shown
in detail b y Fujii et al. (1992) b y investigating the transition f r o m s e m i c o n d u c t i v i t y
to f e r r o m a g n e t i s m in U3(Nil_xCux)3Sb4. The increasing n u m b e r o f d-electrons with
increasing first causes the b a n d gap to close and subsequently i m p r o v e s the m a g n e t i c
ordering via a R K K Y - t y p e interaction b e t w e e n the m a g n e t i c atoms. Substituting Sn
b y Sb in the series U3Ni3(Sbl_xSn~)4 has two effects: the b a n d gap is r e m o v e d
and the h y b r i d i z a t i o n is g r a d u a l l y increased, reflected b y a m o n o t o n e o u s reduction
o f the effective m o m e n t o b s e r v e d at high temperature (Endstra et al. 1992b). A
m a x i m u m for 7 is f o u n d for z = 0.5. The z e r o - t e m p e r a t u r e resistivity changes b y
a factor o f 105. A s m e n t i o n e d above, Ce3Pt3Bi4 has been c o n s i d e r e d as so-called
K o n d o insulator (see A e p p l i and F i s k 1992). Note that in this case, as for C e N i S n ,
the charge gap is e x t r e m e l y small, 0.01 e V for Ce3Pt3Bi4 and 0.001 e V for CeNiSn.
Interestingly, the charge gaps in Ce3Pt3Bi4 and in C e N i S n are a c c o m p a n i e d b y a
spin gap o f c o m p a r a b l e m a g n i t u d e as o b s e r v e d b y neutron scattering ( S e v e r i n g et
al. 1991, M a s o n et al. 1992). As m e n t i o n e d in the section on i l l - c o m p o u n d s ,
substituting L a for Ce b r e a k s the f-lattice s y m m e t r y and leads rapidly to a metallic
like resistivity ( T h o m p s o n et al. 1993a, b). Severing et al. (1994) have shown that
36 G.J. NIEUWENHUYS

the spin gap is less sensitive to doping with non-magnetic elements. Probably the
formation of a spin gap is a more local phenomenon. The charge gap in Ce3Pt3Bi4
has also been observed via reflectivity and optical conductance measurements by
Bucher et al. (1994). These experiments showed that for temperatures below 100
K a charge gap develops. The spectral weight depends on temperature, its decrease
scales with the decrease of the Ce 4f-moments, but the size of the gap is temperature
independent. Electronic band structure calculations in Ce3Pt3Sb4 and Ce3Pt3Bi4 by
Takegahara et al. (1993) show a gap of 300 K, while the experiment reveals 900 K
for the Sb compound. In the Bi compound the calculated gap is 3.5 larger than
the experimental one of 100 K. From these results the conclusion can be drawn that
Ce3Pt3Sb4 the term 'Kondo insulator' is not appropriate. If the small value of the gap
in Ce3Pt3Bi4 can be associated with impurity states, then the term Kondo insulator
is also in doubt for this compound. Other band structure calculations by the same
group (Takegahara et al. 1992, 1993) have shown that the non-magnetic compounds
Th3Ni3Sb4 and La3Au3Sb4 have a narrow gap at the Fermi level. Therefore, the
related compound U3Ni3Sb4 and Ce3Au3Sb4 are 'normal' semiconductors. The same
is true for the 111-compounds crystalizing in the MgAgAs structure, where the non-
magnetic equivalents are also found to be semiconducting (Palstra et al. 1987a and
b).

2.6. Gaps
Several types of energy gaps in a dispersion relation can occur in metallic (magnetic)
systems:
• The gap in the conduction band due to superconductivity.
• The gap in the conduction band, which leads the semiconducting behaviour.
This gap is found in a number of magnetic and non-magnetic compounds, see
above.
• The gap in the spin wave spectrum of magnetically ordered systems. The
resulting low-temperature electrical resistance has been described by Hessel
Anderson et al. (1979, 1980), who showed that for temperatures smaller than
the gap size the contribution to the resistivity should follow a behaviour given
by:

T
T(l +-~)e -zVT (11)

If the gap in the spin wave spectrum is zero than the normal contributions
proportional to the second and fifth power of the temperature remain. It is
rather difficult to distinguish the normal behaviour from the one described by
eq. (11).
• Gaps in the dispersion relation of the conduction electrons due to the lowering
of the symmetry of the crystal lattice as a consequence of the magnetic ordering.
Necessary conditions are that the symmetry is lowered (ferromagnets will not
exhibit this phenomena) and that the itinerant electrons "see" the lowering of the
HEAVYFERMIONSAND RELATEDCOMPOUNDS 37

symmetry, thus a band electron - magnetic moment interaction must be present.


Changes in the band structure can cause changes in the electrical conduction
through a change in the number of electrons available at the Fermi surface, even
sometimes in an anisotropic way. A textbook example is the c-axis resistivity
of pure Er due to a cone-like structure between 20 and 80 K, see Miya (1963),
Freeman (1972) and Coqblin (1977)
• Gaps in Kondo insulators as described above, probably there are charge gaps as
well as spin gaps in these materials.
Note that only the last type of gap is typical for heavy fermions, the other ones are
found in normal metals, semiconductors and magnets, and thus also in some heavy
fermions. Of course it is very interesting how these 'normal' gaps behave when they
are formed in a band of heavy quasi particles. Magnetic gaps in spin or in spin wave
excitation spectra can be observed via inelastic neutron scattering, see Broholm et
al. (1991), Severing et al. (1991), Mason et al. (1992). Charge gaps can be seen
in optical experiments and in vacuum tunneling measurements (Wolff 1985). The
latter type of experiments is rather new in the research on heavy fermions (Aarts and
Volodin 1995), although the tunneling spectra of the most famous ones have been
investigated. URu2Si2 does show a gap in the spin wave excitation spectrum, it shows
the exponential term in the resistivity (Palstra et al. 1986a) and in the specific heat
(Maple et al. 1986). URu2Si2 also shows an increase in the electrical resistivity when
the temperature is lowered through TN, indicative of a change in the band structure.
Indeed, optical data (Bonn et al. 1988) and vacuum tunneling experiments (Aarts
et al. 1994) confirmed this conclusion. The latter experiment clearly showed that
in the antiferromagnetic phase a gap exits for electrons tunneling in the basal plane,
which gap decreased on increasing temperature and disappeared at TN. No such gap
could be detected for electrons tunneling along the crystallographic c-axis. UPd2A13
has no gap in the spin wave spectrum (Mason 1994). Vacuum tunneling experiments
(Aarts et al. 1994) show the same type of gap as observed in URu2Si2. This could
be expected, bearing in mind that in both compounds the magnetic ordering consists
of ferromagnetic layers in the basal plane stacked antiferromagnetically along the
c-axis. Thus in both cases the symmetry of the crystal along the c-axis is broken
be the antiferromagnetic ordering. On the other hand, the electrical resistivity of
UPd2A13 shows no increase at TN, which can be explained by assuming that the
effects on the number of conduction electrons near the Fermi level is accidentally
too small to overcompensate the decrease in the electrical resistivity due to the
decrease in magnetic disorder scattering. However, also optical experiments do not
show a charge gap developing at TN (Degiorgi et al. 1994a) and that has still to be
explained. Gaps have also been observed via optical experiments in UNi2Si2 (Cat
et al. 1993) and UCus, but no gap was detected in U2Zn17 (Degiorgi et al. 1994b).
Note that the resistivity of UCu5 also increases for temperatures below TN (Ott et al.
1985a). Finally, I show two other examples where strong indications emerge form
electrical resistance that the antiferromagnetic order changes the band structure. The
first one is a series of pseudo binary compounds U(Col_xFe~)2Ge2 (van Rossum
1993). The resistivity is shown in fig. 18.
38 G.J. NIEUWENHUYS

1.50

re
q.
1 .OO
j,,¢.a~ oooo °° °
o. oo

U(Co 1-X F e x )2Ge2


0.50 I i I I

O 100 200 300

T ( K )

Fig. 18. Normalized resistivity versus temperature for U(CoI_~Fe=)2Ge2 after van Rossum (1993).

500
J ~t--8.XlS
250 -- i//b-axis -- i//c-axis

0 1.10
g. 5 o o .- ,~ 1.05
0
1.00
25O 095
0.90
i i i
0 0.85
4OO 35 40 45 50

T [K]
200

I I

0 20 40 60
T (~<)
Fig. 19. Temperature dependence of the electrical resistivity of UNiGe for current along the three
principal directions. The fits with an additional exponential term involving electron-magnon scattering
are shown by solid lines. The right hand side figure shows an enlargement of the area around 41.5 K
(after Proke~ et al. 1994).
HEAVYFERMIONSAND RELATEDCOMPOUNDS 39

Also U(Col_xNi~)2Ge2 (Endstra et al. 1993c) shows strong increases in the resis-
tance. The other example is UNiGe, which exhibits a transition from a commensurate
antiferromagnet to an incommensurate structure at T* = 41.5 K (Prokeg et al. 1994).
The electrical resistivity as a function of temperature is depicted in fig. 19, where
nothing can be seen when the current is directed along the a-axis, an increase at
T* when the current flows parallel with the b-axis and finally a sharp decrease for
currents along the c-axis. Clearly the change in the band structure has an anisotropic
influence on the conduction and also it is clear that this change can induce an increase
as well as a decrease in the number of available carriers.
The same experiment gives indications for a gap in the spin wave spectrum,
shown by fits to the expression given in eq. (11) (Proke~ et al. 1994). Another
remarkable signature of the change of the band structure was found from M/3ssbauer
experiments by Mulder (1994). He has determined the electric field gradient at the
site of the magnetic ion - in the case Gd - by measuring the quadrupolar splitting
of the 155Gd nucleus. The electric field gradient is governed by the symmetry
of the crystal lattice and in particular its size by the band structure. It appeared
that in some antiferromagnetic compounds, the gradient changes at the magnetic
transition temperature, whereas in ferromagnetic compounds no change could be
observed. Evidently, careful experiments is the area of spin and charge gaps will
reveal more about the mechanisms in effect in magnetic and semiconducting heavy
fermion materials.

2.7. Magnetism and superconductivity


The interplay between magnetism and superconductivity plays an important role
in heavy fermion physics. Almost all superconducting heavy fermions also show
magnetic ordering, the only exception seems to be the cubic UBe13. In good samples
of UPt3 two superconductivity transitions can be seen in the specific heat as a function
of temperature. This double transition is though to be caused by the symmetry
breaking effect of the antiferromagnetic order in the basal plane present below TN = 5
K (Aeppli et al. 1988). Extensive measurements as function of external magnetic
field and external pressure have revealed a rich phase diagram where the different
crossings of the phase separating lines merge in a tetracritical point at 2.6 kbar
(Boukhny et al. 1994a, b; see also van Dijk 1994 and references therein). Keller et
al. (1994) observed a clear dependence of the upper critical field, Bc2, on the angle
between the magnetic field and the crystalline axis. As mentioned above URuzSi2
and UPdzAI3 show anisotropic gaps at the Fermi surface due to the antiferromagnetic
order. The superconductivity therefore develops in this already partly gapped state
and is evidently influenced by its symmetry. It is difficult to obtain really hard facts
on this aspect, but, e.g., as mentioned earlier, the observation of superconductivity in
URuzSi2 by #+SR is also highly anisotropic in this compound in spite of the apparent
isotropic behaviour of the lower critical field. In the case of UPd2A13 it has been
argued that the superconductivity and the magnetism are carried by two different
f-electron systems (Caspary et al. 1993 and Feyerherm et al. 1994), being itinerant
for the superconductivity and more localized for the magnetism. When U in UBe13
40 G.J. NIEUWENHUYS

is partly (few atomic percent only) replaced by Th, a rather exotic phase diagram is
found, see, e.g., Ott (1994). Between 2 and 4 at.%, a second transition is observed
from a second maximum in the specific heat (Ott 1989). Heffner et al. (1989a, b;
1990) have shown that below the second transition temperature the superconductivity
coexists with a magnetic state, characterized at least by static randomly distributed
internal magnetic fields as seen via #+SR. In all cases it is rather clear that the
superconductivity cannot be described by an ordinary order parameter.

2.8. Miscellaneous compounds


2.8.1. CeCu6 and derivatives
CeCu6 has been studied for a long time. Onuki et al. (1985b) found the first
indication for its Kondo lattice or heavy fermion behaviour based on resistivity mea-
surements on a single crystal. Rapidly other measurements followed (a.o. Ott et al.
1985b, Amato et al. 1987) showing that no long range magnetic order nor super-
conductivity was present down to temperatures of 15 mK. See also Bauer (1991)
for a review. The compound CeCu6 crystalizes in an orthorhombic structure, space
group Pnma, having 4 formula units per unit cell and lattice parameters a = 8.109 A,
b = 5.098 A, and c = 10.172 A. Below a temperature of 220 K, the compound trans-
forms into a monoclinic structure, space group P21/c, with gradually increasing angle
/3 up to 91.36 ° and lattice parameters a = 5.080 A, b = 10.221 A and e = 8.067 A
(Asano et al. 1986 and Gratz et al. 1987). With its 3' value of 1600 mJ/(molK2)
it is now considered as an archetypal heavy fermion system. Neutron diffraction
experiments showed the development of antiferromagnetic correlations below 1 K
(Aeppli et al. 1986, Rossat-Mignod et al. 1988), while #+SR (Amato et al. 1993)
reveals that the moment associated with these correlations should be smaller then
0.01#B. This then places CeCu6 really at the borderline between magnetic ordering
and heavy fermion behaviour. Small amounts of Ag or Au (Gangopadhyay et al.
1992, Germann and von L6hneysen 1989) drive the compound into the long range
ordered region, most probably by increasing the volume of the unit cell and thereby
decreasing the Kondo temperature. Another beautiful experiment was carried out by
Stroka et al. (1993) on CeCu6_~Au~ compounds showing that the Schottky anomaly
due to the crystalline electric field splitting develops from a broadened maximum
into the well-known free ion behaviour with increasing x from 0 to 0.9. At x = 0.9
the long range magnetic order is observed at 2 K. Non-Fermi-liquid behaviour has
also been observed in Au doped CeCu6, at the composition CeCus.9Au0A by von
LShneysen et al. (1994). CeCu6 is a famous heavy fermion without magnetic or-
dering, by increasing x in CeCu6_~Aux antiferromagnetism is rapidly obtained for
small values of x, the transition is zero for the critical concentration x = 0.1.

2.8.2. UxYl-xPd3
UxYl-xPd3 forms a series of strongly correlated electron systems with a complex
and rich magnetic phase diagram (Seaman et al. 1991, 1992, Andraka and Tvselik
1991). For x ~< 0.2 the specific heat, magnetization and electrical properties indicate
non-Fermi liquid behaviour, while for 0.3 ~< x ~< 0.5 a spin-glass type ordering
HEAVYFERMIONSAND RELATEDCOMPOUNDS 41

is found. The lower concentration samples obey remarkably well the theoretical
predictions for the two-channel Kondo effect (Cox 1987). The T -+ 0 behaviour is
not completely clear, Andraka and Tvselik suggest a T = 0 second-order magnetic
phase transition, while Cox assumes a collective Jahn-Teller instability. Stillow
et al. (1994) have shown via detailed metallurgical analysis that compounds with
z ~ 0.2 are intrinsically inhomogeneous. For example, for z = 0.2, the measured
concentrations vary from 0.14 to 0.23. A long term heat treatment could only halve
the range of concentrations. Nevertheless, the magnetic and electrical properties did
not depend strongly on the homogeneity. Remains the question what the behaviour
of a homogeneous alloy will be.

2.8.3. CeCu5 and UCu5 and derivatives


CeCu5 crystallizes in the hexagonal CaCu5 structure, it orders antiferromagnetically
at 3.9 K, and has a '7-value of 100 mJ/(mole K 2) and a logarithmic contribution to the
electrical resistivity (Bauer 1991 and references therein). Inelastic neutron scattering
experiments (Alekseev et al. 1992) and magnetic measurements have shown that
the I + 1) is the ground state of the CEF levels and that the I + 23-}is at 17 meV
(200 K) the first excited state. At the ordering temperature only the ground state
plays a significant role. Remarkably, in spite of the [ 4- ½) ground state, from which
a large magnetic moment in the basal can be expected, neutron diffraction (Bauer
et al. 1994a) showed that the ordered moment is directed along the c-axis and has
1
a value of 0.36#B. The propagation vector of the magnetic structure is (0, 0, ~).
The measured moment is smaller than the one calculated from the CEF scheme
and also the entropy involved in the magnetic transition is smaller than R ln(2),
indicating that the Kondo effect is present. The unexpected direction of the ordered
moment points to a strongly anisotropic (in spin space) interaction between the Ce-
moments, i.e. the compound is Ising like (Bauer et al. 1994a). #+SR experiments
(Wiesinger et al. 1994) showed some evidence for inhomogeneous magnetic order.
These experiments also showed the strong anisotropic susceptibility from the Knight
shift, confirming the Ising like behaviour. The study of CeCu5 is of particular interest
because substituting A1 or Ga for Cu destroys the long range magnetic order (Bauer
1991, Bauer et al. 1990) and drastically increases "7 up to a value of 2.8 mJ/(mole K 2)
for CeCu4A1. Presumably this large "7 should be interpreted as being due to short
range magnetic correlations, since it appeared to be strongly dependent on external
magnetic fields (Andraka et al. 1991). The electrical resistivity of CeCu4Ga is not
pressure dependent (Eichler et al. 1994). UCu5 orders antiferromagnetically below
16 K. The crystal structure is of the cubic fcc AuBe5 type, and the magnetic structure
has been determined via neutron diffraction (Murasik et al. 1974 and Schenck et
al. 1990). The U sublattice orders ferromagnetically in the (111) planes, which
are stacked antiferromagnetically. The ordered moments are aligned along the (111)
direction and have a magnitude of ,.~ 1.55#B. The specific heat reveals a second
transition at 1 K (Ott et al. 1985a, b), which is hysteretic in temperature, suggesting
first order although no latent contribution could be observed. Schenck et al. (1990)
showed that this transition could not be observed in the neutron diffraction, and that
42 G.J. NIEUWENHUYS

for the #+SR only the relaxation, not the frequency of the spontaneous precession
changes at about 1 K. A possible explanation that remains is that besides the already
preset antiferromagnetic order a second ordering takes place of small moments like
in, e.g., UPt3 and URuzSi2 (Schenck 1993). If that interpretation is correct, then
this would be a strong indication for the existence of to different f-electron systems.
Nakamura et al. (1994) interpreted the 1 K transition as a spin reorientation on the
bases of the an analysis of the nuclear magnetic resonance data. When part of the
Cu atoms in UCu5 are replaced by Pd, the antiferromagnetism rapidly disappears.
UCus_~Pd~ is an antiferromagnet only for z < 0.75 and becomes a spin glass for
z > 2. For z = 1.5 the specific heat divided by temperature diverges logarithmically
for T --+ 0 (Andraka and Stewart 1993). No Fermi liquid behaviour could be detected
for this alloy down to 0.3 K, thereby characterizing this material as a non-Fermi liquid
one, like Uo.zY0.8Pd3. The pseudo binary compounds mentioned in this section give
the experimentalist in principle the opportunity to scan the zero-temperature transition
between heavy fermion behaviour and long rang magnetic order. Such ideas are
known for some time (Lawrence 1982) and elaborate theories are available (Millis
1993, 1994). Whether it will be possible to really investigate the critical region
depends on our ability to engineer our samples or the use other external variables
such as pressure or magnetic field sufficiently cleverly. See for a recent review:
Andraka (1994) and Abrahams (1994).

2.9. Yb-compounds
Not many Yb compounds can be considered as heavy fermions. Most of them are
typical valence fluctuators. Yb 3+ is in the 4f 13 state and thus magnetic since it is
off by 1 electron from the full 4f-shell. Yb 2+ is non magnetic. This should be con-
trasted with the Ce-compounds, where adding one f-electron leads to the magnetic
state. Therefore, e.g., pressure dependences of the properties of Yb-compounds are
just opposite from those in Ce-compounds. The counterpart of UBe13, YbBel3 is
a simple antiferromagnet with an ordering temperature of 1.115 K and the crystal
field splittings in this cubic material are rather small. The ground state is /"6, the
first excited state a/"8 at 37 K and the last state is a/"6 at 51 K. These observation
could be made by neutron scattering (Walter et al. 1985) and via specific heat mea-
surements (Ramirez et al. 1986), indicating that hybridization and/or Kondo effects
hardly play any role. YbCu5 acts as a parent compound for various substitution at
the Cu site, see Bauer (1991) and references therein. A few of the resulting doped
materials are heavy fermions. Two crystal structures seem to be possible for YCus,
the simple hexagonal CaCu5 (Iandelli and Palenzona 1971) or the MgSnCu4 struc-
ture, related to the cubic AuBe5 (Hornstra and Buschow 1971). Substitution by A1 or
Ga stabilize the CaCu5 and substitutions by Au, Ag, Pd or In stabilize the MgSnCu4
structure. Yb in YbCu5 is close to the divalent state and thus non magnetic. All
the substitutions mentioned above drive the Yb ion into the trivalent state, thereby
initiating magnetism. YbCu4Au orders antiferromagnetically below 1 K, the Kondo
temperature is of order of 0.3 K (Bonville et al. 1992). The crystal field levels
are /"7, /"8 and/"6 at 0, 46 and 81 K, respectively. YbCu4Pd behaves in the same
way. The only compound in this series with a large -,/-coefficient for the specific
HEAVYFERMIONSAND RELATEDCOMPOUNDS 43

heat is YbCu4Ag. Rossel et al. (1987) found 200 mJ/(mole K2). A value of 245 was
determined for the Pd sample, but this may be in error because of the magnetic order
and the crystal field states. The magnetic and thermodynamic properties of the sys-
tems mentioned can be remarkably well described by the Coqblin-Schrieffer model
(Besnus et al. 1990, Schlottmann 1993). The same can be said about YbCu2Si 2 (Ra-
sul and Schlottmann 1989) and YbPdzSi2 (Schlottmann 1992). The interpretation of
the neutron data obtained by Severing et al. (1990a) for YbCu4Ag is not unambi-
geous. The original authors could not find a consistent fit over the whole temperature
range. A reanalyzis by Polatsek and Bonville (1992) revealed a characteristic tem-
perature (approx. Kondo temperature) of 60 K and a crystal field level scheme such
as that for YbCu4Au. On the other hand, Schlottmann (1993) concludes that the
observed inelastic scattering is a result of transitions from the Kondo resonance to
the Fermi level. YbCu4In is a remarkable compound. It exhibits a first-order phase
transition at T~ = 41 K from the divalent (non-magnetic) state at low temperatures
to a magnetic trivalent one at higher temperatures. The transition is marked by a
sudden change in the magnetic, thermodynamic and transport properties (Felner and
Novik 1986, Felner et al. 1987). Of course also the signature of the crystal field
splittings in the inelastic neutron scattering (Severing et al. 1990b) disappears upon
cooling below T~. The CEF scheme in the magnetic phase is Fs, F6 and F7 at 0,
38 and 45 K, respectively, almost the same as that for the stable trivalent YbNi4In.
The pressure and magnetic field dependence of the electrical and thermal properties
of YbCu4Ag have been studied by Graf et al. (1994), Bauer et al. (1994b, 1995),
Thompson et al. (1994a) and Lacerda et al. (1995). In general the observations can
be described by a decrease of the Kondo temperature with increasing pressure. A
magnetic field of 8 T appears to be equivalent to 30 kbar. Thompson et al. (1994a)
show that the relation between the quadratic term in the electrical resistivity and the
linear term in the specific heat,

A c< 72 (12)

holds when the pressure is the implicit variable. The same is not true for the external
magnetic field (Lacerda et al. 1995). This is probably due to the small CEF splittings,
so that external fields can effectively change the ground state of the system. The
N6el temperature of YbCu4Au, being 0.6 K at ambient pressure, increases to more
than 2 K at 50 kbar. The pressure dependence of the resistivity was described by
a low Kondo temperature, consistent with the earlier observations (Bonville et al.
1992). A1 and Ga substitution also drive the YbCu5 system towards the magnetic
state of Yb (Bauer et al. 1992, 1995) and the full trivalent state is reached around 40
at.% A1 or Ga. YbPtBi has received lots of attention due to its very large value of the
extrapolation of Cp/T to zero temperature, being 8000 mJ/(mole K 2) (Canfield et al.
1991, Fisk et al. 1991). The specific heat exhibits a small maximum at 0.4 K, #+SR
experiments showed spatially inhomogeneous disordered static magnetism below
~0.5 K, with a small moment of ~ 0.1#B in about 50% of the total sample volume.
(Amato et al. 1992). In fact, the #+SR data suggest a spin-glass like behaviour for
YbPtBi, although also a incommensurate spin density wave is possible (Heffner et
44 G.J. NIEUWENHUYS

al. 1994). The latter is in accord with the resistivity measurements by Movshovich
et al. (1994a, b), who found an increase in p when cooling down below 0.4 K,
indicating an ordering such that the band structure changes and at least the number
of carriers at the Fermi level decreases. They also showed that a pressure of only
1 kbar is sufficient to suppress the low temperature state. The crystal structure has
been solved by Robinson et al. (1994b). YbPtBi crystallizes in the cubic MgAgAs
structure, Pt occupies the 4a sites, Bi the 4c, the 4b site is vacant and the 4d site is
occupied by Yb. The same neutron diffraction experiment showed that any ordered
moment should by smaller than 0.25#B. An inelastic neutron scattering experiment,
also made by Robinson et al. (1993b) gives information about the CEF splitting.
Most probably the order of the levels is FT, Fs and F6, the corresponding splittings
however are of order of 1 K and about 65 K. The near degeneracy of the two lowest
states is important. As Thompson et al. (1993b) pointed out, it is the fact that the
CEF energies, the ordering temperature and possibly the Kondo interaction are of
comparable strength which gives YbPtBi its peculiar properties. Remarkably, most of
the properties are not influenced by doping with non-magnetic elements like Lu and
Y (Lacerda et al. 1993): down to 50 at.% Yb the thermal and magnetic properties are
proportional to the Yb content, even the temperature of the maximum in the specific
heat does not depend on dilution of the Yb sub-lattice. Thompson et al. (1993b)
also showed that the isostructural YbPtSb orders normally at T = 0.35 K and shows
accordingly contributions from the nuclear specific heat at lower temperatures.

2.10. Thin films


Thin films of heavy fermion materials have already been made occasionally for a
number of years (e.g., Tedrow and Quateman 1986, Roessler and Tedrow 1990).
The general conclusion is that the typical properties of the heavy electron persist in
the thin film. Note, however, that these were thicker than any characteristic length
of the heavy fermion state. A systematic effort was undertaken by the Darmstadt
group: Huth et aI. (1993, 1994a, b, c, 1995), Hessert et al. (1995) (UPd2A13
films) and Lunkenheimer et al. (1994) (UCua+xA18_x films). The UPd2A13 films
are highly c-axis oriented, the films show the antiferromagnetic transition at 14 K
as well as the superconducting one at 1.8 K. The availability of these oriented
films enables detailed investigation of the Hall effect and of the critical current in
the superconducting state. The Hall coefficient exhibits a maximum around 55 K
and a minimum at 6 K and saturates quadratically towards zero temperature. The
temperature dependence can be understood as a superposition of skew scattering
(dominating at high temperatures) and coherent effects at low temperatures. Criti-
cal current measurements in the superconducting state showed this magnetic heavy
fermion superconductor to be able to carry 6.5 x 107 A]m2 at 0.32 K and 1 T external
field. The activation energy for flux greep appeared to be 25 K in zero external field
and to decrease with field along the c-axis to 5 K at 3 T. UCu4A18 crystallizes in the
ThMn~2 structure and orders antiferromagnetically below 30 K. When z is increased
in the series UCu4+xA18_x, then the antiferromagnetism decreases for z > 1.5 and a
heavy fermion state emerges resulting in 7 = 800 mJ/(mole K2) at z = 1.9 (Geibel
et al. 1990, K/3hler et al. 1990). Thermal and transport properties can be described
HEAVY FERMIONS AND RELATED COMPOUNDS 45

in a single ion Kondo effect model. Lunkenheimer et al. (1994) have prepared this
system in amorphous thin film form. In this state the onset of magnetic order is
suppressed with respect to the crystalline results at low z values and the coherent
heavy-fermion behaviour is suppressed at large z values. From dc- and ac-resistivity
measurements they conclude that all sample reveal a single ion Kondo behaviour,
but that significant deviations from dipolar Kondo effect are found for the lowest
temperatures.

3. Conclusion

Heavy fermions and other related magnetic compounds have attracted lots of attention
over the last decade. They still from the subject of many publications: the subject
is alive. In spite of this efforts a number of questions have remained unanswered up
to now, see, e.g., Coleman (1995) in the final talk of the Amsterdam conference on
Strongly Correlated Electron Systems. Future attention will probably be focussed on
the zero-temperature transition from the 'paramagnetic' state to magnetic order and
the accompanying non-Fermi liquid behaviour. Also the band structure will become
more important, via further ab initio calculations, and via spectroscopic experiments.

4. A c k n o w l e d g e m e n t s

During the past ten years I have had the pleasure to work with a number of PhD
students: T.T.M. Palstra, A.J. Dirkmaat, T. Endstra, R.A. Steeman E.A. Knetsch,
S.A.M. Mentink, A. Drost and S. Stillow. Without their efforts this chapter would
not have been written. I thank J. Aarts and J.A. Mydosh for many stimulating
discussions. Without the cooperation and friendship of many colleagues around the
world the research on heavy fermion would be dule, if not impossible.

References

Aarts, J. and A.R Volodin, 1995, Physica B Adroja, D.T., B.D. Rainford and S.K. Malik,
206-207, 43. 1994, Physica B 194-196, 169.
Aarts, J., A.R Volodin, A.A. Menovsky, Aeppli, G., H. Yoshizawa, Y. Endoh, E. Bucher,
G.J. Nieuwenhuys and J.A. Mydosh, 1994, J. Hufnagl, Y. Onuki and T. Komatsubara,
Europhys. Lett. 26, 203. 1986, Phys. Rev. Lett. 57, 122.
Abrahams, E., 1994, Physica B 197, 435. Aeppli, G., E. Bucher, C. Broholm, J.K. Kjems,
Adroja, D.T., S.K. Malik, B.D. Padalia J. Baumann and J. Hufnagl, 1988, Phys. Rev.
and R. Vijayaraghavan, 1988, Solid State Lett. 60, 615.
Commun. 66, 1201. Aeppli, G, and Z. Fisk, 1992, Comments Cond.
Adroja, D.T., S.K. Malik, B.D. Padalia and Mat. Phys. 16, 155.
R. Vijayaraghavan, 1989, Phys. Rev. B 39, Alekseev, RA., V.N. Lazukov, I.R Sadonov,
4831. M.N. Khlopkin, G.A. Takzei, Y.U.R Grebe-
Adroja, D.T. and S.K. Malik, 1991, J. Magn. nyuk and I.I. Sych, 1992, J. Magn. Magn.
Magn. Mater. 100, 126. Mater. 110, 119.
Adroja, D.T., B.D. Padalia, S.N. Bhatia and Aliev, EG., V.V. Moshchalkov, M.K. Zalyalyut-
S.K. Malik, 1992, Phys. Rev. B 45, 477. dinov, G.I. Pak, R.V. Skolozdra, RA. Alekseev,
46 G.J. NIEUWENHUYS

V.N. Lakuzov and I.P. Sadikov, 1990, Physica Bauer, E., E. Gratz, R. Hauser, Le Tuan,
B 163, 358. T. Kagayama and G. Oomi, 1995, Physica B
Amato, A., D. Jaccard, J. Flouquet, E Lapierre, 206-207, 352.
J.L. Tholence, R.A. Fisher, S.E. Lacy, Besnus, M.J., P. Haen, N. Hamdaoui, A. Herr
J.A. Olsen and N.E. Philips, 1987, J. Low and A. Meijer, 1990, Physica B 163, 571.
Temp. Phys. 68, 371. Besnus, M.J., A. Essaihi, N. Hamdaoui,
Amato, A., RC. Canfield, R. Feyerherm, Z. Fisk, G. Fisher, J.P. Kappler, A. Meyer, J. Pierre,
EN. Gygax, R.H. Heffner, D.E. MacLaughlin, P. Haen and P. Lejay, 1991, Physica 171, 350.
H.R. Ott, A. Schenck and J.D. Thompson, Besnus, M.J., A. Essaihi, G. Fisher, N. Hamdaoui
1992, Phys. Rev. B 46, 3151. and A. Meyer, 1992, J. Magn. Magn. Mater.
104-107, 1387.
Amato, A. et al. 1993, Physica B 186-188, 273.
Beyermann, W.P., M.E Hundley, P.C. Canfield,
Anderson, P.W., 1961, Phys. Rev. 124, 41. J.D. Thompson, M. Latroche, C. Godart,
Andraka, B. and A.M. Tvselik, 1991, Phys. Rev.
M. Selsane, Z. Fisk and J.L. Smith, 1991,
Lett. 67, 2886. Phys. Rev. B 43, 13130.
Andraka,B, J.S. Kim, G.R. Stewart and Z. Fisk, Bhattacharjee, A.K., B. Coqblin, M. Raki,
1991, Phys. Rev. B 44, 4371. L. Forro, C. Ayache and D. Schmitt, 1989, J.
Andraka, B. and G.R. Stewart, 1993, Phys. Rev. Phys. (Paris) 50, 2781.
B 47, 3208. B0hm, A., R. Caspary, U. Habel, L. Pawlak,
Andraka, B., 1994, Physica B 199-200, 239. A. Zuber, E Steglich and A. Loidl, 1988, J.
Andres, K., J.E. Graebner and H.R. Ott, 1975, Magn. Magn. Mater. 76-77, 150.
Phys. Rev. Lett. 35, 1779. Bonn, D.A., J.D. Garrett and T. Timusk, 1988,
Aoki, Y, T. Suzuki, T. Fujita, H. Kawanaka, Phys. Rev. Lett. 61, 1305.
T. Takabatake and H. Fujii, 1993, Phys. Rev. Bonville, E, B. Canoud, J. Hammann, A. Hodges,
B 47, 15060• P. Imbert, G. J6hanno, A. Severing and
Asano, H., M. Umino, Y. Onuki, T. Komatsubara, Z. Fisk, 1992, J. Phys. I (France) 2, 459.
E Izumi and N. Watanabe, 1986, J. Phys. Boukhny, M., G.L. Bullock, B.S. Shivaram and
Soc. Jpn 55, 454. D.G. Hicks, 1994a, Phys. Rev. Lett. 73,
Ayache, C., J. Beille, E. Bonjour, R. Calemczuk, 1707.
G. Greuzet, D. Gignoux, A. Najib, D. Schmitt, Boukhny, M., G.L. Bullock and B.S. Shivaram,
J. Voiron and M. Zerguine, 1987, J. Magn. 1994b, Phys. Rev. B 50, 8985.
Magn. Mater. 63-64, 329. Brandt, N.B. and V.V. Moshchalkov, 1984, Adv.
Bakker, K., A. de Visser, E. BrOck, A.A. Phys. 33, 373.
Menovsky and JJ.M. Franse, 1992, J. Magn. Bredl, C.D., H. Spille, U. Rauschwalbe,
Magn. Mater. 108, 63. W. Lieke, E Steglich, G. Gordier, W. Assmus,
M. Hermann and J. Aarts, 1983, J. Magn.
Bauer, E., E. Gratz, J. Kohlmann, K. Winzer,
Magn. Mater. 31-34, 373.
D. Gignoux and D. Schmitt, 1990, Z. Phys.
Broholm, C., J.K. Kjems, W.J.L. Buyers,
B 80, 263. P. Matthews, T.T.M. Palstra, A.A. Menovsky
Bauer, E., 1991, Adv. Phys. 40, 417. and J.A. Mydosh, 1987, Phys. Rev. Lett. 58,
Bauer, E., K. Payer, R. Hauser, E. Gratz, 1467.
D. Gignoux, D. Schmitt, N. Pillmayr and Broholm, C., H. Lin, P. Matthews, T.E. Mason,
G. Schaudy, 1992, J. Magn. Magn. Mater. W.J.L. Buyers, M.E Collins, A.A. Menovsky,
104-107, 651. J.A. Mydosh and J.K. Kjems, 1991, Phys.
Bauer, E., M. Rotter, L. Keller, P. Fischer, Rev. B 43, 12809.
M. Ellerby and K.A. McEwen, 1994a, J. Brtick, E., M. van Sprang, J.C.P. Klaasse and
Phys.: Condens. Matter 6, 5533. ER. de Boer, 1988, J. Appl. Phys. 63, 3417.
Bauer, E., R. Hauser, E. Gratz, M. Maikis, BrUck, E., H. Nakotte, K. Bakker, F.R. de Boer,
Le Tuan, A. Indinger, G. Oomi and P.E de Ch~tel, Li Jingyan, J.P. Kuang and
T. Kagayama, 1994b, Physica B 199-200, Y. Fuming, 1993, J. Alloys Comp. 200, 79.
527. Brack, E., H. Nakotte, P.F. de Ch~tel, H.P. van
Bauer, E., E. Gratz, R. Hauser, Le Tuan, der Meulen, J.J.M. Franse, A.A. Menovsky,
A. Galatam, A. Kottar, H. Michor, W. Perthold N.H. Kim-Ngan, L. Havela, V. Sechovsky,
and G. Hilscher, 1994c, Phys. Rev. B 50, J.A.A.J. Perenboom, N.C. Tuan and J. Sebek,
9300. 1994, Phys. Rev. B 49, 8852.
HEAVY FERMIONS AND RELATED COMPOUNDS 47

Bucher, B., Z. Schlesinger, RC. Canfield and huys and J.A. Mydosh, 1992, Physica 179,
Z. Fisk, 1994, Physica B 199-200, 489. 84.
Buschow, K.H.J., 1979, Rep. Prog. Phys. 42, de Visser, A., H. Nakotte, L.T. Lai, A.A.
1373. Menovsky, S.A.M. Mentink, G.J. Nieuwen-
Buschow, K.H.J., D.B. de Mooij, T.T.M. Palstra, huys and J.A. Mydosh, 1993, Physica
G.J. Nieuwenhuys and J.A. Mydosh, 1985, 186-188, 291.
Philips J. Res. 40, 313. Degiorgi, L., M. Dressel, G. Grtiner, R Wachter,
Buschow, K.HJ. and D.B. de Mooij, 1986, N. Sato and T. Komatsubara, 1994a,
Philips J. Res. 41, 55. Europhys. Lett. 25, 311.
Canfield, RC., J. Thompson, W.P. Beyermann, Degiorgi, L, H.R. Ott, M. Dressel, G. Grtiner
A. Lacerda, M.E Hundley, E. Peterson, and Z. Fisk, 1994b, Europhys. Lett. 26, 221.
Z. Fisk and H.R. Ott, 1991, J. Appl. Phys. Delapalme, A., Z. Zolnierek, R. Trod and
70, 5800. D. Kaczorowski, 1994, J. Phys.: Condens.
Cao, N., J.D. Garrett and T. Timusk, 1993, Matter. 6, 8877.
Physica B 191, 263. Dhar, S.K., S.K. Malik and R. Vijayaraghavan,
Caspary, R. et al. 1993, Phys. Rev. Lett. 71, 1981, J. Phys. C 14, L321.
Dirkmaat, A.J., T. Endstra, E.A. Knetsch,
2146.
G.J. Nieuwenhuys, J.A. Mydosh, A.A.
Chelmicki, L., J. Leciejewicz and A. Zygmunt,
Menovsky, ER. de Boer and Z. Tarnawski,
1985, J. Phys. Chem. Solids 46, 529.
1990a, Phys. Rev. B 41, 2589.
Coleman, R, 1995, Physica B 206-207, 352.
Dirkmaat, A.J., T. Endstra, E.A. Knetsch,
Cooper, B.R., Q.C. Cheng, S.R Lim, C. Sanchez-
A.A. Menovsky, G.J. Nieuwenhuys and
Castro, N. Kioussis and J.M. Wills, 1992, J.
J.A. Mydosh, 1990b, J. Magn. Magn. Mater.
Magn. Magn. Mater. 108, 10.
84, 143.
Coqblin, B., 1977, The Electronic Structure
Dirkmaat, A.J., T. Endstra, E.A.Knetsch,
of Rare-Earth Metals and Alloys: The A.A. Menovsky, G.J. Nieuwenhuys and
Magnetic Heavy Rare-Earths (Academic J.A. Mydosh, 1990c, Europhys. Lett. 11,
Press, London). 275.
Coqblin, B. and J.R. Schrieffer, 1969, Phys. Rev. Doniach, S., 1977, Physica B 91, 231.
185, 847. Drost, A., 1995, PhD Thesis, Leiden University,
Cox, D.L., 1987, Phys. Rev. Lett. 59, 1240. unpublised.
Das, I., E.V. Sampathkumaran, R. Nagarajan and Dwight, A.E., 1974, J. Less-Common Met. 34,
R. Vijayaraghavan, 1991, Phys. Rev. B 43, 278.
13159. Dwight, A.E., 1979, J. Nucl. Mat. 79, 417.
de Boer, ER., J.J.M. Franse, E. Louis, Dijk, N.H., 1994, PhD Thesis, University of
A.A. Menovsky, J.A. Mydosh, T.T.M. Palstra, Amsterdam, unpublished.
U. Rauchschwalbe, W. Schlabitz, E Steglich Eichler, A., C. Mehls and Ch. Sutter, 1994,
and A. de Visser, 1986, Physica B 138, 1. Physica B 199-200, 181.
de Boer, ER., J.C.P. Klaasse, RA. Veenhuizen, Endstra, T., 1992, PhD Thesis, Leiden University.
A. BOhm, C.D. Bredl, U. Gottwick, H.M. Endstra, T., A.J. Dirkmaat, S.A.M. Mentink,
Mayer, L. Pawlak, U. Rauchschwalbe, A.A. Menovsky, G.J. Nieuwenhuys and
H. Spille and E Steglich, 1987, J. Magn. J.A. Mydosh, 1990a, Physica B 163, 309.
Magn. Mater. 63-64, 91. Endstra, T., G.J. Nieuwenhuys, J.A. Mydosh and
de Boer, ER., E. Brt~ck, H. Nakotte, A.V. K.H.J. Buschow, 1990b, J. Magn. Magn.
Andreev, V. Sechovsky, L. Havela, R Nozar, Mater. 89, L273.
C.J.M. Denisen, K.H.J. Buschow, B. Varizi, Endstra, T., G.J. Nieuwenhuys, A.A. Menovsky
M. Meissner, H. Maletta and R Rogl, 1992, and J.A. Mydosh, 1991, J. Appl. Phys. 69,
Physica B 176, 275. 4816.
de Boer, ER., K. Proke~, H. Nakotte, E. Br~ick, Endstra, T., G.J. Nieuwenhuys, A.A. Menovsky
M. Hilber, R Svoboda, V. Sechowsky, and J.A. Mydosh, 1992a, J. Magn. Magn.
L. Havela and H. Maletta, 1994, Physica, in Mater. 108, 67.
press. Endstra, T., G.J. Nieuwenhuys, A.A. Menovsky
de Visser, A., H. Nakotte, L.T. Lai, A.A. and J.A. Mydosh, 1992b, J. Magn. Magn.
Menovsky, S.A.M. Mentink, G.J. Nieuwen- Mater. 108, 69.
48 G.J. NIEUWENHUYS

Endstra, T., G.J. Nieuwenhuys and J.A. Mydosh, and T. Okamoto, 1987, Jpn. J. Appl. Phys.
1993a, Phys. Rev. B 48, 9595. 26, 549.
Endstra, T., S.A.M. Mentink, G.J. Nieuwenhuys Fujii, H., E. Ueda, Y. Uwatoko and T. Shigeoka,
and J.A. Mydosh, 1993b, in: Frontiers 1988, J. Magn. Magn. Mater. 76-77, 179.
in Solid State Sciences, Vol. 2, eds. Fujii, H., T. Inoue, Y. Andoh, T. Takabatake,
L.C. Gupta and M.S. Multani (World K. Satoh, Y. Maeno, E Fujita, J. Sakurai and
Scientific, Singapore) p. 167. Y. Yamaguchi, 1989a, Phys. Rev. B 39,
Endstra, T., B.J. van Rossum, G.J. Nieuwenhuys 6840.
and J.A. Mydosh, 1993c, Physica B 186-188, Fujii, H., H. Kawanaka, T. Takabatake,
778. M. Kurisu, Y. Yamagucji, J. Sakurai,
F/ik, B., C. Vettier, J. Flouquet, R Lejay and H. Fujiwara, T. Fujita and I. Oguro, 1989b, J.
J.M. Mignot, 1995, Physica B 206-207, 415. Phys. Soc. Jpn 58, 2495.
Felner, I., I. Mayer, A. Grill and M. Schieber, Fujii, H., M. Nagasawa, H. Kawanaka, T. Inoue
and T. Takabatake, 1990, Physica B 165-166,
1975, Solid. State Commun. 16, 1005.
435.
Felner, I. and I. Novik, 1986, Phys. Rev. B 33,
Fujii, H., S. Miyata and T. Takabatake, 1992, J.
617. Magn. Magn. Mater. 104-107, 45.
Felner, I., I. Novik, D. Vaknin, U. Potzel, Fujita, T., K. Satoh, Y. Onuki and T. Komatsubara,
J. Moser, G.M. Kalvius, G. Wortmann, 1985, J. Magn. Magn. Mater. 47-48, 66.
G. Schmeister, G. Hilscher, E. Gratz, Fujita, T., K. Satoh, Y. Maeno, Y. Uwatoko and
C. Schmitzer, N. Pillmayer, K.G. Prasad, H. Fujii, 1988, J. Magn. Magn. Mater.
H. de Waard and H. Pinto, 1987, Phys. Rev. 76-77, 133.
B 35, 6956. Fujita, T., T. Suzuki, S. Nishigori, T. Takabatake,
Feyerherm, R., A. Amato, EN. Gygax, H. Fujii and J. Sakurai, 1992, J. Magn. Magn.
A. Schenck, C. Geibel, F. Steglich, N. Sato Mater. 108, 35.
and T. Komatsubara, 1994, Phys. Rev. Lett. Fulde, E, J. Keller and G. Zwicknagel, 1988,
73, 1849. in: Solid State Physics, Vol. 41, eds.
Feyerherm, R., A. Amato, C. Geibel, EN. Gygax, H. Ehrenreich and D. Turnbn11 (Academic
R Hellmann, R.H. Heffner, D.E. MacLaughlin, Press, New York) p. 1.
R. Mtiller-Reisener, GJ. Nieuwenhuys, Fumagalli, R, J. Schoenes, H. Ruegsegger and
A. Schenck and E Steglich, 1995, Physica D. Kaczorowski, 1988, Helv. Phys. Acta 61,
207-207, 470. 829.
Fisk, Z., G.R. Stewart, J.O. Willis, H.R. Ott and Fuming, Y., J.R Kuang, J. Li, E. BrOck,
E Hulliger, 1984, Phys. Rev. B 30, 6360. H. Nakotte, ER. de Boer, X. Wu, Z. Li and
Fisk, Z., H.R. Ott, T.M. Rice and J.L. Smith, Y. Wang, 1991, J. Appl. Phys. 69 4705.
1986, Nature 320, 124. Gangopadhyay, A.K., K. Lettau, C. Lettau,
Fisk, Z., D.W. Hess, CJ. Pethick, D. Pines, J.S. Schilling, E. Schuberth, W. Fuchs, K.
J.L. Smith, J.D. Thompson and J.O. Willis, Andres, M. Damento and K.A. Gschneidner
1988, Science 239, 33. Jr, 1992, J. Magn. Magn. Mater. 103, 267.
Fisk, Z., RC. Canfield, W.R Beyermann, Geibel, C., U. Ahlheim, A.L. Giorgi, G. Sparn,
J.D. Thompson, M.E Hundley, H.R. Ott, H. Spille, E Steglich and W. Suski, 1990,
Physica 163, 194.
E. Felder, M.B. Maple, M.A. Lopez de la
Geibel, C., U. Ahlheim, C.D. Bredl, J. Diehl,
Torte, E Visani and C. Seaman, 1991, Phys.
A. Grauel, R. Helffich, H. Kitazawa,
Rev. Lett. 67, 3310.
R. KOhler, R. Modler, M. Lang, C. Schank,
Francois, M., G. Venturini, J.E Mar~ch6,
S. Thies, E Steglich, N. Sato and T.
B. Malaman and B. Roques, 1985, J. Less- Komatsubara, 1991a, Physica C 185-189,
Common Met. 113, 231. 2651.
Freeman, A.J., 1972, in: Magnetic Properties of Geibel, C., S. Thies, D. Kaczprowski, A. Mehner,
Rare Earth Metals (Plenum, London) p. 245. A. Grauel, B. Seidel, U. Ahlheim, R. Helfrich,
Fremy, M.A., D. Gignoux, D. Schmidt and R Petersen, C.D. Bredl and E Steglich, 1991b,
A.Y. Takenchi, 1989, J. Magn. Magn. Mater. Z. Phys. B 83, 305.
82, 175. Geibel, C., C. Schank, S. Thies, H. Kitazawa,
Fujii, H., Y. Uwatoko, M. Akayama, K. Satoh, C.D. Bredl, A. Boehm, M. Rau, A. Grauel,
Y. Maeno, T. Fujita, J. Sakurai, H. Kamimura R. Caspary, R. Helfrich, U. Ahlheim,
HEAVY FERMIONS AND RELATED COMPOUNDS 49

G. Weber and E Steglich, 1992, Z. Phys. Heffner, R.H., D.W. Cooke, Z. Fisk, R.L.
84, 1. Hutson, M.E. Schillaci, H.D. Rempp, J.L.
Geibel, C., C. Schank, E Jahrling, B. Buschinger, Smith, J.O. Willis, D.E. MacLaughlin,
A. Grauel, T. Liahmann, P. Gegenwart, C. Boekema, R.L. Lichti, J. Oostensand and
R. Helfrich, R.H.P. Reinders and E Steglich, A.B. Denison, 1989a, Phys. Rev. B 39,
1994, Physica B 199-200, 128. 11345.
Gerrnann, A. and H. von LOhneysen, 1989, Heffner, R.H., J.O. Willis, J.L. Smith, P. Birrer,
Europhys. Lett. 9, 367. C. Baines, EN. Gygax, B. Hitti, E. Lippelt,
Gignoux, D., D. Schmitt, M. Zerguine, E. Bauer, H.R. Ott, A. Schenck and D.E. MacLaughlin,
N. Pillmayr, J..Y. Henry, V.N. Nguyen and 1989b, Phys. Rev. B 40, 806.
Heffner, R.H., J.L. Smith, J.O. Willis, P. Bitter,
J. Rossat-Mignod, 1988, J. Magn. Magn.
C. Baines, EN. Gygax, B. Hitti, H.R. Ott,
Mater. 74, 1.
A. Schenck, E.A. Knetsch, J.A. Mydosh and
Gignoux, D., A.P. Murani, D. Schmitt and
D.E. MacLaughlin, 1990, Phys. Rev. Lett.
M. Zerguine, 1991, J. Phys. I 1, 281.
65, 2816.
Godart, C., A.M. Unarji, L.C. Gupta and Heffner, R.H., A. Amato, RC. Canfield,
R. Vijayaraghavan, 1987, J. Magn. Magn. R. Feyerherm, Z. Fisk, EN. Gygax, D.E.
Mater. 63-64, 326. MacLaughlin, A. Schenck, J.D. Thompson
Gor'kov, L.P., 1987, Sov. Sci. Rev. A Phys. 9, and H.R. Ott, 1994, Physica B 199-200, 113.
1. Hess, D.W., P.S. Riseborough and J.L. Smith,
Graf, T., J.D. Thompson, P.C. Canfield, 1993, in: Encyclopedia of Applied Physics,
M.B. Maple and Z. Fisk, 1994, Physica Vol. 7, (VCH Publishers) p. 435.
199-200, 520. Hessel Anderson, N. and H. Smith, 1979, Phys.
Gratz, E., E. Bauer, H. Novotky, H. Mueller, Rev. B 19, 384.
S. Zeminli and B. Barbara, 1987, J. Magn. Hessel Anderson, N., 1980, in: Crystalline
Magn. Mater. 63-64, 312. Electric Field and Structural Effects in
Grauel, A., A. Btihm, C. Geibel, G. Weber and f-Electron Systems, eds J.E. Crow, R.P. Guer-
E Steglich, 1993, J. Appl. Phys. 73, 5421. tin and T.W. Mihalisin (Plenum Press, New
Grier, B.H., J.M. Lawrence, V. Murgai and York) p. 373.
R.D. Parks, 1984, Phys. Rev. B 29, 2664. Hessert, J., M. Huth, M. Jourdan and H. Adrian,
Grewe, N. and E Steglich, 1990, in: Handbook 1995, Physica B 206-207, 618.
on the Physics and Chemistry of the Rare Hewson, A.C., 1993, The Kondo Problem
Earths, Vol. 13, eds. K.A. Gschneidner and to Heavy Fermions (Cambridge University
L. Eyring (Elsevier, New York) p. 341. Press, Cambridge).
Harrison, W.A., 1969, Phys. Rev. 181, 1036. Hiebl, K. and P. Rogl, 1985, J. Magn. Magn.
Harrison, W.A. and S. Froyen, 1980, Phys. Rev. Mater. 50, 39.
Hiebl, K., C. Horvath and P. Rogl, 1986, J.
B 21, 3214.
Less-Common Met. 117, 375.
Harrison, W.A., 1983, Phys. Rev. B 28, 550.
Hiebl, K., C. Horvath, K. Remsching and
Harrison, W.A. and G.K. Straub, 1987, Phys. H. No~l, 1990, J. Appl. Phys. 67, 943.
Rev. B 36, 2695. Hill, H.H., 1970, in: Plutonium and Other
Havela, L., V. Sechovsky, L. Jirman, ER. de Boer Actinides, ed. W.N. Miller (AIME, New
and E. BrOck, 1991, J. Appl. Phys. 69, 4813. York) p. 2.
Havela, L., V. Sechovsky, H. Nakotte, E. Brfick Hornstra, J. and K.HJ. Buschow, 1971, J.
and ER. de Boer, 1994a, J. Alloys Comp. Less-Common Met. 27, 123.
213, 243. Horwath, C. and P. Rogl, 1983, Mat. Res. Bull.
Havela, L., V. Sechovsky, P. Svoboda, M. Divi~, 18, 443.
H. Nakotte, K. Proke~, ER. de Boer, Hovestreydt, E. et al., 1972, J. Less-Common
A. Purwanto, R.A. Robinson, A. Seret, Met. 85, 247.
J.M. Winand, J. Rebizant, J.C. Spirlet, Hu, G.J. and B.R. Cooper, 1993, Phys. Rev. B
M. Richter and H. Eschrig, 1994b, J. Appl. 48, 12743.
Phys. 76, 6214. Hulliger, F., 1993, J. Alloys Comp. 196, 225.
Heeb, E., H.R. Ott, E. Felder, E Hulliger, Hundley, M.E, P.C. Canfield, J.D. Thompson,
A. Schilling and Z. Fisk, 1991, Solid State Z. Fisk and J.M. Lawrence, 1990, Phys. Rev.
Commun. 77, 323. B 42, 6842.
50 G.J. NIEUWENHUYS

Huth, M., A. Kaldowski, J. Hessert, Th. Steinborn Kawanaka, H., H. Fujii, M. Nishi, T. Takabatake,
and H. Adrian, 1993, Solid State Commun. K. Motoya, Y. Uwatoko and Y. Ito, 1989, J.
87, 1133. Phys. Soc. Jpn 59, 3481.
Huth, M., A. Kaldowski, J. Hessert, C. Heske Keller, N., J.L. Tholence, A. Huxley and
and H. Adrian, 1994a, Physica B 199-200, J. Flouquet, 1994, Phys. Rev. Lett. 73, 2364.
116. Kilibarda-Dalafave, S., H.K. Ng, J.E. Crow,
Huth, M., J. Hessert, M. Jourdan, A. Kaldowski P. Pernambuco-Wise, T. Yuen and C.L. Lin,
and H. Adrian, 1994b, Phys. Rev. B 50, 1993, Physica 186-188, 989.
1309. Kitazawa, H., A. Matsushita, T. Matsumoto and
Huth, M., J. Hessert, M. Jourdan, A. Kaldowski
T. Suzuki, 1994, Physica B 199-200, 28.
and H. Adrian, 1994c, Physica C 235, 2439.
Huth, M., J. Hessert, M. Jourdan and H. Adrian, Knetsch, E.A., J.J. Petersen, A.A. Menovsky,
1995, Physica B 206-207, 615. M.W. Meisel, G.J. Nieuwenhuys and J.A.
Iandelli, A. and A. Palenzona, 1971, J. Mydosh, 1992, Euro. Phys. Lett. 19, 637.
Less-Common Met. 25, 333. Knetseh, E.A., A.A. Menovsky and J.A. Mydosh,
Isaacs, E.D., D.B. McWham, C. Peters, G.E. Ice, 1994, in: Frontiers in Solid State Sciences,
D.R Siddons, J.B. Hastings, C. Vettier and eds. L.C. Gupta and M.S. Multani (World
O. Vogt, 1989, Phys. Rev. Lett. 61, 1241. Scientific, Singapore) in press.
Isaacs, E.D., R Zschack, A.R Ramirez, Knopp, G., A. Loidl, R. Caspary, U. Gottwick,
C.S. Oglesby and E. Bucher, 1994, J. Appl. C.D. Bredl, H. Spille and E Steglich, 1988,
Phys. 76, 6133. J. Magn. Magn. Mater. 74, 341.
Jaccard, D., K. Behnia and J. Sierro, 1992a, Knopp, G., A. Loidl, K. Knorr, L. Pawlak,
Phys. Lett. A 163, 475. M. Duczmal, R. Caspary, U. Gottwick,
Jaccard, D., K. Behnia and J. Sierro, 1992b, H. Spille, E Steglich and A.R Murani, 1989,
Physica Scripta T45, 130. Z. Phys. B 77, 95.
Jarlborg, T., H.E Braun and M. Peter, 1983, Z. Kohgi, M., K. Ohoyama, T. Osakabe and
Phys. B 52, 295. M. Kasaya, 1992, J. Magn. Magn. Mater.
Kaczmarska, K., J. Pierre, A. Slebarsld and 108, 187.
R.V. Skolozdra, 1993, J. Alloys Comp. 196,
KOhler, R., C. Geibel, S. Horn, B. Strobel,
165.
Kaczorowski, D. and R. Trod, 1990, J. Phys.: S. Arnold, A. H~hr, C. K~immerer and
Condens. Matter. 2, 4185. E Steglich, 1990, Physica 165-166, 429.
Kadowaki, K. and S.B. Woods, 1986, Solid State Kondo, J., 1964, Progr. Theor. Phys. 32, 37.
Commun. 58, 507. Konno, R., 1993, Progr. Theor. Phys. 89, 51.
Kagan, Y., K.A. Kikoin and N.V. Prokof'ev, Krimmel, A., A. Severing, A. Murani, A. Grand
1992, Physica B 182, 201. and S. Horn, 1992a, Physica B 180-181, 191.
Kagayama, T., G. Oomi, K. Iki, N. Mori, Krimmel, A., P. Fischer, B. Roessli, H. Maletta,
Y. Onuki and T. Komatsubara, 1994, J. Alloys C. Geibel, C. Schank, A. Loidl and
Comp. 214, 387. E Steglich, 1992b, Z. Phys. B 86, 161.
Kalvius, G.M., A. Kratzer, K.H. MUnch, Krimmel, A., A. Loidl, P. Fischer, B. Roessli,
ER. Wagner, S. Zwirner, H. Kobayashi, A. D6nni, H. Kita, N. Sato, Y. Endoh,
T. Takabatake, G. Nakamoto, H. Fujii, T. Komatsubara, C. Geibel and E Steglich,
S.R. Kreitzmann and R. Kiefl, 1993, Physica 1993, Solid State Commun. 87, 829.
B 186-188, 412. Ku, H.C., G.P. Meissner, E Acker and D.C.
Kasaya, M., E Iga, K. Katoh, K. Takegahara and Johnston, 1980, Solid State Commun. 35, 91.
T. Kasuya, 1990, J. Magn. Magn. Mater.
Kuang, J.P., H.J. Cui, J.Y. Li, EM. Yang,
90--91, 521.
H~ Nakotte, E. Brtick and ER. de Boer, 1992,
Kasaya, M., T. Tani, H. Suzuki, K. Ohoyama
and M. Kohgi, 1991, J. Phys. Soc. Jpn 60, J. Magn. Magn. Mater. 104-107, 1475.
2542. Kuznietz, M., H. Pinto, H. Ettedgui and
Kasaya, M., K. Katoh and T. Takegahara, 1991b, M. Melamund, 1989, Phys. Rev. B 40, 7328.
Solid State Commun. 78, 797. Kuznietz, M., H. Pinto and M. Melamund,
Kawamata, S., K. Ishimoto, Y. Yamaguchi and 1990a, J. Magn. Magn. Mater. 83, 321.
T. Komatsubara, 1992, J. Magn. Magn. Kuznietz, M., H. Pinto and M. Melamund,
Mater., 1992, 51. 1990b, J. Appl. Phys. 67, 4808.
HEAVY FERMIONS AND RELATED COMPOUNDS 51

Kuznietz, M., H. Pinto, H. Ettedgui and I. Gociaska and T. Luciski, 1994, Phys. Rev.
M. Melamund, 1992a, Physica B 180-181, B 50, 9581.
55. Maeno, Y, M. Takahashi, T. Fujita, Y. Uwatoko,
Kuznietz, M., H. Pinto, H. Ettedgui and H. Fujii and T. Okamoto, 1987, Jpn. J. Appl.
M. Melamund, 1992b, Phys. Rev. B 45, Phys. 26, 545.
7282. Malik, S.K. and D.T. Adroja, 1991a, Phys. Rev.
Kyogatu, M., Y. Kitaoka, H. Nakamura, B 43, 6277.
K. Asayama, T. Takabatake, E Teshima and Malik, S.K. and D.T. Adroja, 1991b, Phys. Rev.
H. Fujii, 1990, J. Phys. Soc. Jpn 59, 1728. B 43, 6295.
Lacerda, A., R. Movshovich, M.E Hundley, Maple, B.M., J.W. Chen, Y. Dalichaouch,
RC. Canfield, D. Arms, G. Sparn, J.D. T. Kohara, C. Rossel, M.S. Torikachvili,
Thompson, Z. Fisk, R.A. Fisher, N.E. Philips M.W. McElfresh and J.D. Thompson, 1986,
and H.R. Ott, 1993, J. Appl. Phys. 73, 5415. Phys. Rev. Lett. 56, 185.
Lacerda, A., T. Graf, M.E Hundley, M.S. Mason, T.E., B.D. Gaulin, J.D. Garrett, Z. Tun,
Torikachvili, J.M. Lawrence, J.D. Thompson, W.J.L. Buyers and E.D. Isaacs, 1990, Phys.
D. Gajewski, RC. Canfield and Z. Fisk, 1995, Rev. Lett. 65, 3185.
Physica B 206-207, 358. Mason, T.E., G. Aeppli, A.P. Ramirez, K.N.
Lawrence, J., 1982, J. Appl. Phys. 53, 2117. Clausen, C. Broholm, N. Stucheli, E. Bucher
Lee, RA., T.M. Rice, J.W. Serene, L.J. Sham and and T.T.M. Palstra, 1992, Phys. Rev. Lett.
J.W. Wilkins, 1986, Comments Cond. Mat. 69, 490.
Phys. 12, 99. Mason, T.E., 1994, Private communication.
Lee, W.H. and R.N. Shelton, 1987a, Phys. Rev. Menovsky, A.A. and J.J.M. Franse, 1983, J.
B 35, 5369. Cryst. Growth 65, 265.
Lee, W.H., H.C. Ku and R.N. Shelton, 1987b, Mentink, S.A.M., G.J. Nieuwenhuys, A.A.
Phys. Rev. B 36, 5739. Menovsky and J.A. Mydosh, 1991, J. Appl.
Liang, G., N. Jisrawi and M. Croft, 1990, Phys. 69, 5484.
Physica B 163, 134. Mentink, S.AM., G.J. Nieuwenhuys, C.E. Snel,
Lin, H., L. Rebelsky, M.E Collines, J.D. Garrett A.A. Menovsky and J.A. Mydosh, 1992, J.
and W.J.L. Buyers, 1991, Phys. Rev. B 43, Magn. Magn. Mater. 104-107, 15.
13232. Mentink, S.AM., N. Bos, B.J. van Possum,
Lloret, B., B. Buffat, B. Chevalier and G.J. Nieuwenhuys, J.A. Mydosh and K.H.J.
J. Etourneau, 1987, J. Magn. Magn. Mater. Buschow, 1993, J. Appl. Phys. 73, 6625.
67, 435. Mentink, S.A.M., 1994, PhD Thesis, Leiden
Loewenhaupt, M. and K.H. Fisher, 1994, University, unpublished.
Neutron Scattering on Heavy Fermion and Mentink, S.A.M., B.J. van Rossum, G.J. Nieu-
Valence Fluctuating 4f-Systems, in press. wenhuys, J.A. Mydosh and K.H.J. Buschow,
LOhneysen, H. yon, T. Pietrus, G. Portisch, 1994a, J. Alloys Comp. 216, 131.
H.G. Schlager, A. Schrt~der, M. Sieck and Mentink, S.A.M., A. Drost, G.J. Nieuwenhuys,
T. Trappmann, 1994, Phys. Rev. Lett. 72, E. Frikkee, A.A. Menovsky and J.A. Mydosh,
3262. 1994b, Phys. Rev. Lett. 73, 1031.
Loidl, A., K. Knorr, G. Knopp, A. Krimmel, Mentink, S.A.M., G.J. Nieuwenhuys, H. Nakotte,
R. Caspary, A. B6hm, G. Sparn, C. Geibel, A.A. Menovsky, A. Drost, E. Frikkee and
E Steglich and A.R Murani, 1992, Phys. Rev. J.A. Mydosh, 1994c, to be published.
B 46, 9341. Mentink, S.A.M., G.J. Nieuwenhuys, A.A.
Luke, G.M., L.R Le, B.J. Sternlieb, Y.J. Uemura, Menovsky, J.A. Mydosh, H. Tou and
J.H. Brewer, R. Kadono, R.E Kielt, Y. Kitaoka, 1994d, Phys. Rev. B 49, 15579.
S.R. Kreitzman, T.M. Riseman, C.L. Seaman, Mentink, S.A.M., G.J. Nieuwenhuys, A.A.
Y. Dalichaouch, M.B. Maple and J.D. Garrett, Menovsky and J.A. Mydosh, 1994e, Physica
1990, Hyperfine Interact. 64, 517. 194-196, 275.
Luke, G.M., A. Keren, K. Kojima, L.R Le, Mielke, A., E.-W. Scheidt, J.J. Rieger and
B.J. Sternlieb, W.D. Wu and Y.J. Uemura, G.R. Stewart, 1993, Phys. Rev. B 48, 13985.
1994, Phys. Rev. Lett. 73, 1853. Mignot, J.-M, R. Kahn, M.-J. Besnus, J.-P.
Lunkenheimer, R, M. Kramer, R. Viana, Kappler and C. Godart, 1993, Physica B
C. Geibel, A. Loidl, W. Suski, H. Ratajczak, 186-188, 475.
52 G.J. NIEUWENHUYS

Millis, A.J., 1993, Phys. Rev. B 48, 7183. Ning, Y.B., J.D. Garrett and W.R. Datars, 1990,
Millis, A.J., 1994, Physica B 199-200, 227. Phys. Rev. B 42, 8780.
Mirambet, E, P. Gravereau, B. Chevalier, L. Trut Ning, Y.B., V.V. Gridin, C.V. Stager, W.R. Datars,
and J. Etourneau, 1993, J. Alloys Comp. 191, A. Dawson and D.H. Ryan, 1991, J. Phys.:
L1. Condens. Matter 3, 4399.
Mirambet, E, B. Chevalier, L. Fournes, P. Nishigori, S., T. Suzuki, T. Fujita, H. Tanaka,
Gravereau and J. Etourneau, 1994, J. Alloys T. Takabatake and H. Fujii, 1994, Physica B
Comp. 203, 29. 199-200, 473.
Miya, H., 1963, Progr. Theor. Phys. 29, 477. Nohora, S., H. Narnatame, A. Fijimofi and
Movshovich, P., A. Lacerda, P.C. Canfield, T. Takabatake, 1993, Physica B 186-188,
D. Arms, J.D. Thompson and Z. Fisk, 1994a, 403.
Physica B 199-200, 67. Norman, M.R. and D.D. Koelling, 1992,
Movshovich, P., A. Lacerda, P.C. Canfield, unpublished.
J.D. Thompson and Z. Fisk, 1994b, J. Appl.
Onuki, Y., Y. Shimizu and T. Komatsubara, 1984,
Phys. 76, 6121.
J. Phys. Soc. Jpn 53, 1210.
Mulder, EM., 1994, PhD Thesis, Leiden
Ott, H.R., 1987, in: Progress in Low
University, unpublished.
Murasik, A., S. Ligenza and A. Zygmunt, 1974, Temperature Physics, XI, ed. D.E Brewer
Phys. Status Solidi A: 23, K 163. (North-Holland, Amsterdam) p. 215.
Nakamura, H, Y. Kitaoka, K. Asayama, Y. Onuki Ott, H.R., 1989, Physica C 162-164, 1669.
and M. Shiga, 1994, J. Phys.: Condens. Ott, H.R., E. Felder, S. Takagi, A. Schilling,
Matter 6, 10567. N. Sato and T. Komatsubara, 1992, Philos.
Nakamura, S., T. Goto, Y. Isikawa, S. Sakatsume Mag. 65, 1349.
and M. Kasaya, 1991, J. Phys. Soc. Jpn 60, Ott, H.R., 1994, J. Low Temp. Phys. 95, 95.
2305. Ott, H.R., H. Rtidigier, Z. Fisk and J.L. Smith,
Nakotte, H., E. Brfick, ER. de Boer, A. J. 1983, Phys. Rev. Lett. 50, 1595.
Riemersma, L. Havela and V. Sechovsky, Ott, H.R., H. Rtidigier, E Delsing and Z. Fisk,
1992, Physica B 179, 269. 1984, Phys. Rev. Lett. 52, 1551.
Nakotte, H., 1994, PdD Thesis, University of Ott, H.R., H. RUdigier, E. Felder, Z. Fisk and
Amsterdam, unpublished. B. Batlogg, 1985a, Phys. Rev. Lett. 55,
Nakotte, H., R.A. Robinson, J.W. Lynn, E. Brtick 1595.
and ER. de Boer, 1993a, Phys. Rev. 47, 831. Ott, H.R., H. Rtidinger, Z. Kisk, J.O. Willis and
Nakotte, H., E. Brtick, ER. de Boer, E Svoboda, G.R. Stewart, 1985b, Solid State Commun.
N.C. Tuan, L. Havela, V. Sechovsky and 53, 235.
R.A. Robinson, 1993b, J. Appl. Phys. 73, Ott, H.R. and Z. Fisk, 1987, in: Handbook on
6551. the Physics and Chemistry of the Actinides,
Nakotte, H., L. Havela, J.E Kuang, ER. de Boer, Vol. 5, eds. A.J. Freeman and G.H. Lander
K.H.J. Buschow, J.H.V.J. Brabers, T. Kuroda, (North-Holland, Amsterdam) p. 85.
K. Sugiyama and M. Date, 1993c, Int. J. Paixao, J.A., G.H. Lander, A. Delapalme,
Mod. Phys. 7, 838.
H. Nakotte, ER. de Boer and E. Briick, 1993,
Nakotte, H., E. Brtick, K. Proke~, J.H.V.J. Bra-
Europhys. Lett. 24, 607.
bers, ER. de Boer, L. Havela, K.H.J. Buschow
Palstra, T.T.M., 1986, PhD Thesis, Leiden
and Yang Fu-ming, 1994a, J. Alloys Comp.,
University, unpublished.
in press.
Nakotte, H., K. Prokeg, ER. de Boer, E. Brtick, Palstra, T.T.M., A.A. Menovsky, J. van den Berg,
P. Svoboda, V. Sechowsky, L. Havela, A.J. Dirkmaat, P.H. Kes, G.J. Nieuwenhuys
J.M. Winand, A. Seret and J.C. Spirlet, 1994b, and J.A. Mydosh, 1985, Phys. Rev. Lett. 55,
Physica, in press. 2727.
Neifeld, R.A., M. Croft, T. Mihalisin, C.U. Segre, Palstra, T.T.M., A.A. Menovsky, J.A. Mydosh
M. Madigan, M.S. Torikachvili, M.B. Maple and G.J. Nieuwenhuys, 1986a, J. Magn.
and L.E. DeLong, 1985, Phys. Rev. B 32, Magn. Mater. 54-57, 435.
6928. Palstra, T.T.M., G.J. Nieuwenhuys, J.A. Mydosh
Nieuwenhuys, G.J., 1987, Phys. Rev. B 35, and K.H.J. Buschow, 1986b, J. Magn. Magn.
5260. Mater. 54-57, 549.
HEAVY FERMIONS AND RELATED COMPOUNDS 53

Palstra, T.T.M., G.J. Nieuwenhuys, R.F.M. Robinson, R.A., J.W. Lynn, V. Nuez, K.H.J.
Vlastuin, J. van den Berg, J.A. Mydosh and Buschow, H. Nakotte and A.C. Lawson,
K.H.J. Buschow, 1987a, J. Magn. Magn. 1993a, Phys. Rev. B 47, 5090.
Mater. 67, 331. Robinson, R.A., M. Kohgi, T. Osakabe,
Palstra, T.T.M., G.J. Nieuwenhuys, R.RM. P.C. Canfield, T. Kamiyama, T. Nakane,
Vlastuin, J.A. Mydosh and K.H.J. Buschow, Z. Fisk and J.D. Thompson, 1993b, Physica
1987b, J. Appl. Phys. 63, 4279. B 186--188, 550.
Pecharski, V.K., Yu.V. Pankevich and O.I. Bodak, Robinson, R.A., J.W. Lynn, A.C. Lawson and
1983, Sov. Phys. Crystallogr. 28, 97. H. Nakotte, 1994a, J. Appl. Phys. 75, 6589.
Peron, M.N., Y. Kergadallan, J. Rebizant, Robinson, R.A., A.C. Lawson, V. Sechovsky,
D. Meyer, S. Zwirner, L. Havela, H. Nakotte, L. Havela, Y. Kergdallan, H. Nakotte and
J.C. Spirlet, G.M. Kalvius, E. Colineau, F.R. de Boer, 1994b, J. Alloys Comp. 213,
J.L. Oddou, C. Jeandy and J.R Sanchez, 1994, 528.
J. Alloys Comp., in press. Roessler, G.M. Jr. and P.M. Tedrow, 1990,
Pleger, H.R., E. BrOck, E. Braun, E Oster, Physica 165-166, 419.
A. Freimuth, B. Politt, B. Roden and Rogl, P., 1984, in: Handbook on the
D. Wohlleben, 1987, J. Magn. Magn. Mater. Physices and Chemistry of Rare Earths,
63-64, 107. Vol. 7, eds K.A. Gschneidner and L. Eyring
Polatsek, G. and R Bonville, 1992, Z. Phys. B (North-Holland, Amsterdam) p. 1 and refences
88, 189. therein.
ProkeL K., H. Nakotte, E. Brtick, ER. de Boer, Rogl, P., B. Chevalier, M.J. Besnus and
L. Havela, V. Sechowsky and P. Svoboda, J. Etourneau, 1989, J. Magn. Magn. Mater.
1994, IEEE Trans. Magn. 30, 1214. 80, 305.
Ptasiewicz-Bak, H., J. Leciejewicz and A. Rossat-Mignod, J., L.P. Regnault, J.L. Jacoud,
Zygmunt, 1981, J. Phys. F 11, 1225. C. Vetter, P. Levy, J. Flouquet, E. Walker,
D. Jaccard and A. Amato, 1988, J. Magn.
Quezel, S., J. Rossat-Mignod, B. Chevalier,
Magn. Mater. 76-77, 376.
P. Lejay and J. Etourneau, 1984, Solid State
Rossel, C., K.N. Yang, M.B. Maple, Z. Fisk,
Commun. 49, 685.
E. Zirngiebel and J.D. Thompson, 1987, Phys.
Rainford, B.D. and D.T. Adroja, 1994, Physica Rev. B 35, 1914.
194-196, 365. Rossi, D., R. Marazza and R. Ferro, 1979, J.
Rainford, B.D., D.T. Adroja and J.M.E. Geers, Less-Common Met. 66, 17.
1994, Physica B 199-200, 556. Rossum, van B.-J., 1993, Masters Thesis, Leiden
Ramirez, A.P., B. Battlog and Z. Fisk, 1986, University, unpublished.
Phys. Rev. B 34, 1795. Sakon, T., K. Imamura, N. Takeda, N. Sato and
Ramirez, A.P., T. Seigrist, T.T.M. Palsta, T. Komatsubara, 1993, Physica B 186-188,
J.D. Garrett, E. BrOck, A.A. Menovsky and 297.
J.A. Mydosh, 1991, Phys. Rev. B 44, 5392. Sakurai, J., Y. Yamaguchi, S. Nishigori,
Rasul, J.W. and R Schlottmann, 1989, Phys. T. Suzuki and T. Fujita, 1990, J. Magn. Magn.
Rev. 39, 3065. Mater. 90--91, 422.
Rauchschwalbe, U, U. Gottwick, U. Ahlheim, Sakurai, J., K. Kegai, K. Nishimura, Y. Ishikawa
H.M. Mayer and F. Steglich, 1985, J. and K. Mori, 1993, Physica B 186-188, 583.
Less-Common Met. 111, 265. Sampathkumaran, E.V., R. Vijayaraghavan and
Rebelsky, L., M.W. McElfresh, M.S. Torikachvili, I. Das, 1991, Z. Phys. B 84, 247.
B.M. Powell and M.B. Maple, 1991, J. Appl. Sandratskii, L.M. and J. K0bler, 1994, Phys.
Phys. 69, 4810. Rev. B 50, 9258.
Reehuis, M., T. Vomhof and W. Jeitschko, 1991, Sandratskii, L.M., J. Ktibler, P. Zahn and
J. Less-Common Met. 169, 139. I. Mertig, 1994, Phys. Rev. B 50, 15834.
Riedi, P.C., J.G.M. Armitage, J.S. Lord, Santini, P. and G. Amoretti, 1994, Phys. Rev.
D.T. Adroja, D.B. Rainford and D. Fort, 1994, Lett. 73, 1027.
Physica B 199-200, 558. Sarma, B.K, S. Levy, S. Adenwalla and
Regnault, L.P., W.A.C. Erkelens, J. Rossat- B. Ketterson, 1992, in: Physical Acoustics,
Mignod, P. Lejay and J. Flouquet, 1988, Phys. Vol. 20, ed. M. Levy (Academic Press,
Rev. B 38, 4481. Boston) p. 107.
54 G.J. NIEUWENHUYS

Satoh, K., T. Fujita, Y. Maeno, Y. Uwatoko and Severing, A., E. Holland-Moritz, B. Rainford,
H. Fujii, 1990, J. Phys. Soc. Jpn 59, 692. S.R. Culverhouse and B. Frick, 1989a, Phys.
Sauls, J.A., 1994, Adv. Phys. 43, 113. Rev. B 39, 2557.
Schank, C., F. Jfihrling, L. Luo, A. Grauel, Severing, A., E. Holland-Moritz and B. Frick,
C. Wassilew, R. Borth, G. Olesch, C.D. Bredl, 1989b, Phys. Rev. B 39, 4164.
C. Geibel and F. Steglich, 1994, J. Alloys Severing, A., A.P. Murani, J.D. Thompson,
Comp., to be published. Z. Fisk and C.-K. Loong, 1990a, Phys. Rev.
Schenck, A., 1993, in: Frontiers in Solid B 41, 1739.
State Sciences, Vol. 2, eds L.C. Gupta and Severing, A, E. Gratz, B.D. Rainford and
M.S. Multani (World Scientific, Singapore) p.
K. Yishimura, 1990b, Physica B 163, 409.
269.
Schenck, A., P. Birrer, EN. Gygax, B. Hitti, Severing, A, J.D. Thompson, RC. Canfield and
E. Lippelt, M. Weber, P. BOni, P. Fisher, Z. Fisk, 1991, Phys. Rev. B 44, 6832.
H.R. Ott and Z. Fisk, 1990, Phys. Rev. Lett. Severing, A., T. Perring, J.D. Thompson,
65, 2454. EC. Canfield and Z. Fisk, 1994, Physica B
Schlottmann, P., 1989, Physics Rept. 181, 1. 199-200, 480.
Schlottmann, P., 1992, Phys. Rev. 46, 217. Sheng, Q.G., B.R. Cooper, J.M. Willis and
Schlottmann, P., 1993, J. Appl. Phys. 73, 5414. N. Kloussis, 1990, J. Appl. Phys. 67, 5194.
Schoenes, J., P. Fumagalli, H. Ruegsegger and Shigeoka, T., K. Hirota, M. Ishikawa, H. Fnjii
D. Kaczorowski, 1989, J. Magn. Magn. and T. Takabatake, 1992, Physica B
Mater. 81, 112. 186-188,469.
Schrieffer, J.R. and P.A. Wolff, 1966, Phys. Rev. Sigrist, M. and K. Ueda, 1991, Rev. Mod. Phys.
149, 847. 63, 239.
Schr6der, A., J.G. Lussier, B.D. Gaulin, Singhal, R.K., N.L. Saini, K.B. Garg, J. Kanski,
J.D. Garrett, W.J.L. Buyers, L. Rebelsky and L. Iver, P.O. Nilsson, R. Kumar and
S.M. Shapiro, 1994, Phys. Rev. Lett. 72,
L.C. Gupta, 1993, J. Phys.: Condens. Matter
136. 5, 4013.
Seaman, C.L., M.B. Maple, S. Ghamaty,
B.W. Lee, M.S. Torikachvili, J.-S. Kang, Skolozdra, R.V., Ja.F. Mikhalski, K. Kaczmarska
L.Z. Liu, J.W. Allen and D.L. Cox, 1991, and J. Pierre, 1994, J. Alloys Comp. 206,
Phys. Rev. Lett. 67, 2882. 141.
Seaman, C.L., M.B. Maple, S. Ghamaty, Steeman, R.A., E. Frikkee, R.B. Helmholdt,
B.W. Lee, M.S. Torikachvili, J.-S. Kang, A.A. Menovsky, J. van den Berg,. G.J.
L.Z. Liu, J.W. Allen and D.L. Cox, 1992, J. Nieuwenhuys and J.A. Mydosh, 1988, Solid
Alloys Comp. 181, 237. State Commun. 67, 103.
Sechovsky, V. and L. Havela, 1988, in: Steeman, R.A., E. Frikkee, S.A.M. Mentink,
Ferromagnetic Materials, Vol. 4, eds E.P. Wohl- A.A. Menovsky, G.J. Nieuwenhuys and
fath and K.H.J. Buschow (North-Holland, J.A. Mydosh, 1990a, J. Phys.: Condens.
Amsterdam) p. 309 Matter 2, 4059.
Sechovsky, V., L. Havela, P. Nozar, E. Brfick, Steeman, R.A., A.J. Dirkmaat, A.A. Menovsky,
ER. de Boer, A.A. Menovsky, K.H.J. Buschow E, Frikkee, G.J. Nieuwenhuys and J.A.
and A.V. Andreev, 1990, Physica 163, 103. Mydosh, 1990b, Physica B 163, 382.
Sechovsky, V., L. Havela, A. Purwanto, Steglich, E, J. Aarts, C.D. Bredl, W. Lieke,
A.C. Lawson, R.A. Robinson, K. Proke~,
D. Meshede, W. Franz and H. Schafer, 1979,
H. Nakotte, E. B~ck, ER. de Boer,
Phys. Rev. Lett. 43, 1982.
R Svoboda, H. Maletta and M. Winkelmann,
1994a, J. Alloys Comp. 213, 536. Steglich, E, G. Spare, R. Moog, S. Horn,
Sechovsky, V., L. Havela, R Svoboda, A. A. Grauel, M. Lang, M. Novak, A. Loidl,
Purwanto, A.C. Lawson, R.A. Robinson, A. Krirnmel, K. Knorr, A.R Murani and
K. Prokeg, H. Nakotte, E. Brfick, ER. de Boer M. Tachiki, 1990a, Physica B 163, 19.
and H. Maletta, 1994b, J. Appl. Phys. 76, Steglich, E, U. Ahlheim, C. Schank, C. Geibel,
6217. S. Horn, M. Lang, G. Sparn, A. Loidl and
Selsane, M., M. Lebail, N. Hamdaout, J.R A. Krimmel, 1990b, J. Magn. Magn. Mater.
Kappler, H. Noel, J.C. Achard and C. Godart, 84, 271.
1990, Physica B 163, 213. Stewart, G.R., 1984, Rev. Mod. Phys. 56, 755.
HEAVY FERMIONS AND RELATED COMPOUNDS 55

Stewart, G.R., Z. Fisk, J.O. Willis and J.L. Smith, Takegahara, K., H. Harima, Y. Kaneta and
1984, Phys. Rev. Lett. 52, 679. A. Yanase, 1993, J. Phys. Soc. Jpn 62, 2103.
Straub, G.K. and W.A. Harrison, 1985, Phys. Tedrow, EM. and J.H. Quateman, 1986, Phys.
Rev. B 31, 7668. Rev. B 34, 4595.
Strnat, K.J., 1988, in: Ferromagnetic Materials, Thompson, J.D., R.D. Parks and H. Borges,
Vol. 4, eds E.E Wohlfarth and K.H.J. Buschow 1986, J. Magn. Magn. Mater. 54-57, 377.
(Elsevier, Amsterdam) p. 131. Thompson, J.D. et al., 1993a, in: Transport and
Stroka, B., A. Schr6der, T. Trappmann, Thermal Properties of f-Electron Systems, eds
H. v. LOhneysen, M. Loewenhaupt and H. Fujii, T. Fujita and G. Oomi (Plenum, New
A. Severing, 1993, Z. Phys. B 90, 155. York) p. 35.
Strong, S.E and A.J. Millis, 1994, Phys. Rev. B Thompson, J.D., EC. Canfield, A. Lacerda,
50, 12611. M.E Hundley, Z. Fisk, H.R. Ott, E. Felder,
Sugiura, E., K. Sugiyama, H. Tawanaka, M. Chernikov, M.B. Maple, E Visani,
T. Takabatake, H. Fujii and M. Date, 1990, J. C.L. Seaman, M.A. Lopez de la Torre and
Magn. Magn. Mater. 90-91, 65. G. Aeppli, 1993b, Physica 186-188, 355.
Stillow, S., 1994, Private communication. Thompson, J.D., J.M. Lawrence and Z. Fisk,
S011ow, S., T.J. Gortenmulder, G.J. Nieuwenhuys, 1994a, J. Low Temp. Phys. 95, 59.
A.A. Menovsky and J.A. Mydosh, 1994, J. Thompson, J.D., Y. Uwatoko, T. GraY, M.E
Alloys Comp. 215, 223. Hundley, D. Mandrus, C. Godart, L.C. Gupta,
Suzuki, T., M. Nohara, T. Fujita, T. Takabatake EC. Canfield, A. Migliori and H.A. Borges,
and H. Fujii, 1990, Physica B 165-166, 421. 1994b, Physica B 199-200, 589.
Suzuki, T., T. Akazawa, E Nakamura, Y. Tanaka, Torikachvili, M.S., R.E Jardin, C.C. Becerra,
H. Fujisaki, S. Aono, T. Fujita, T. Takabatake C.H. Westpal, A. Paduan-Filho, V.M. Lopez
and H. Fujii, 1994, Physica B 199-200, 483. and L. Rebelsky, 1992, J. Magn. Magn.
Szytula, A., S. Siek, J. Leciejewicz, A. Zygmunt
Mater. 104--107, 69.
and Z. Ban, 1988a, J. Phys. Chem. Solids,
Trod, R., D. Kaczororowski, M. Kolende,
49, 1113.
A. Szytula, M. Bonnet, J. Rossat-Migrnod
Szytula, A., M. Slaski, B. Dunlop, Z. Sungaila
and H. No~l, 1993, Solid. State Commun.
and A. Umezawa, 1988b, J. Magn. Magn.
87, 573.
Mater. 75, 71.
Trovarelli, O., J.G. Sereni, G. Schmerber and
Takabatake, T., E Teshima, H. Fujii, S. Nishigori,
T. Suzuki, T. Fujita, Y. Yamaguchi and J.P. Kappler, 1994, Phys. Rev. B 49, 15179.
Uwatoko, Y., G. Oomi, S.K. Malik, T. Takahatake
J. Sakurai, 1990a, J. Magn. Magn. Mater.
and H. Fujii, 1994, Physica B 199-200, 572.
90-91, 474.
Takabatake, T., E Teshima, H. Fujii, S. Nishigori, Uwatoko, Y., T. Ishii, G. Oomi and S.K. Malik,
T. Suzuki, T. Fujita, Y. Yamaguchi, J. Sakurai 1995, Physica B 206-207, 199.
and D. Jaccard, 1990b, Phys. Rev. B 41, Venturini, G., B. Malaman, L. Pontonnier,
9607. M. Bacmann and D. Fruchart, 1988a, Solid
Takabatake, T., S. Miyata, H. Fujii, Y. Aoki, State Commun. 66, 597.
T. Suzuki, T. Fujita, J. Sakurai and T. Hiraoka, Venturini, G., B. Malaman, L. Pontonnier and
1990c, J. Phys. Jpn 59, 4412. D. Fruchart, 1988b, Solid State Comrnun. 67,
Takabatake, T., H. Iwasaki, G. Nakamoto, 193.
H. Fujii, H. Nakotte, ER. de Boer and Walker, M.B., W.J.L. Buyers, Z. Tun, W. Que,
V. Sechowski, 1993, Physica 183, 108. A.A. Menovsky and J.D. Garrett, 1993, Phys.
Takabatake, T., G. Nkamoto, H. Tanka, Y. Bando, Rev. Lett. 71, 2630.
H. Fujii, S. Nishigori, H. Goshima, T. Suzuki, Walter, U., Z. Fisk and E. Holland-Moritz, 1985,
T. Fujita, I. Oguro, T. Hiraoka and S.K. J. Magn. Magn. Mater. 47--48, 159.
Malik, 1994a, Physica B 199-200, 457. Wiesinger, G., E. Bauer, A. Amato, R. Feyer-
Takabatake, T., Y. Maeda, T. Konishi and herm, EN. Gygax and A. Schenck, 1994,
H. Fujii, 1994b, J. Phys. Soc. Jpn 63, 2853. Physica B 199-200, 52.
Takegahara, K., Y. Kaneta and T. Kasuya, 1990, Wolff, E.L., 1985, Principles of Electron
J. Phys. Soc. Jpn 59, 4394. Tunneling Spectroscopy, in: International
Takegahara, K. and Y. Kaneta, 1992, Progr. Series of Monographs on Physics (Oxford
Theor. Phys. Suppl. No. 108, 55. Science Publications).
56 G.J. NIEUWENHUYS

Wulff, M., J.M. Fournier, A. Delapalme, 1984, J. Low. Temp. Phys. 56, 601.
B. Gillon, V. Sechovsky, L. Havela and Yang, K.N., M.S. Torikachvili, M.B. Maple,
A.V. Andreev, 1990, Physica 163, 33l.
H.C. Ku, B.B. Pate, I. Lindau and J.W. Allen,
Yamaguchi, Y., J. Sakurai, E Teshima, H.
Kawanaka, T. Takabatake and H. Fujii, 1990, 1985, J. Magn. Magn. Mater. 47--48, 558.
J. Phys.: Condens. Matter 2, 5715. Zolnierek, Z., H. NoEl and D. Kaczorowski,
Yang, K.N., M.S. Torikachvili and M.B. Maple, 1987, J. Less-Common. Met. 128, 265.
chapter 2

MAGNETIC MATERIALS STUDIED


BY MUON SPIN ROTATION
SPECTROSCOPY

A. SCHENCK and F.N. GYGAX


Institute for Particle Physics of ETH ZBrich
CH-5232 Villigen PSI
Switzerland

Handbook of Magnetic Materials, Vol. 9


Edited by K.H.J. Buschow
01995 Elsevier Science B.V. All rights reserved

57
CONTENTS

1. Introduction ................................................................. 60
2. Muon spin rotation (/~SR) spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
2.1. Parameters + phenomenology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
2.2. Muon site and local fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.3. More on spin-lattice relaxation in the paramagnetic and ordered state . . . . . . . . . . . . . . 76
3. Review of results in elemental metals and alloys ................................... 80
3.1. Spontaneous dipole and hyperfine fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.2. Critical phenomena ...................................................... 96
3.3. Ferromagnetic 3d-element based alloys ...................................... 101
3.4. Chromium and its alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
3.5. /~-SR in Fe and Ni . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4. Review of results in intermetallic compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.1. Compounds involving transition elements .................................... 104
4.2. Intermetallic compounds containing rare earth elements . . . . . . . . . . . . . . . . . . . . . . . . . 114
4.3. Intermetallic compounds containing actinide elements .......................... 175
5. Review of results in magnetic insulators .......................................... 199
5.1. Oxides with corundum-type structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
5.2. Orthoferrites and RNiO3 perovskites ........................................ 209
5.3. Miscellaneous mostly Cu-based and layered oxides ............................ 212
5.4. MnO .................................................................. 215
5.5. Magnetic fluorides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
5.6. COC12.2H20 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
5.7. Solid oxygen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
6. Review of results in layered cuprate (high Tc) compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
6.1. La2CuO 4 and related compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
6.2. YBa2Cu30= and related compounds ........................................ 244
6,3. Bi-based (2212)-compounds ............................................... 269

58
MUON SPIN ROTATION SPECTROSCOPY 59

7. Study of magnetic order in organic compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274


7.1. (TMTSF)2X ............................................................ 274
7.2. Ni2(C2HsN2)2NO2(C104) (NENP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
7.3. p-NPNN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
List of some of the used abbreviations and symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
1. Introduction

This chapter contains a review of results on magnetic materials obtained by a single


technique, namely Muon Spin Rotation (#SR) Spectroscopy. In this respect the
present chapter is quite different from most other contributions in this series of
volumes which focus on (all) the magnetic properties of certain materials as collected
by different methods and techniques. The reader will find that the application of
#SR-spectroscopy has in many instances just confirmed what was known already
from other studies, but often has added a new flavour or twist to our knowledge
and understanding of established magnetic phenomena. Beyond that, however, when
using positive muons (#+) #SR spectroscopy has shown the potential to uncover
new and unexpected magnetic features owing to its high sensitivity to very small
magnetic fields and spatially inhomogeneous properties. Thus #SR-spectroscopy is
most powerful in the field of small moment magnetism and in all instances when
magnetic order is of a random or very short range nature and where neutron scattering
will fail. Compared to NMR-spectroscopy #SR-spectroscopy will work also in the
case of very broad lines (up to 100 MHz) and very short relaxation rates (down
to 10 -8 s). On the other hand relaxation phenomena involving Ta > 100 #s are out
of the time window accessible by #SR. In this respect #SR and NMR are rather
complementary techniques. But note that generally the implanted positive muons
(#+) are found at interstitial sites and, therefore, probe the magnetism from a different
perspective. The local probe aspect of #SR and its sensitivity to inhomogeneous
features has in particular brought the real nature of solids into light which are never
perfect in lattice structure, stoichiometry and morphology, and hence also magnetic
properties do not show up in an ideal manner. For example, in many intermetallic
compounds magnetic order is found to be established in a spatially inhomogeneous
way, leaving sometimes a fraction of the sample in the paramagnetic state even far
below the phase transition temperature. There are even examples where magnetic
order evolves so gradually in space and with temperature that the concept of a
cooperative phase transition looses its meaning and consequently no anomalies in the
specific heat are observed. It is in fact an open question whether such observations
have to be correlated only with the sample quality (sample quality certainly matters)
or whether they could also reflect more intrinsic properties. Such distinction, on the
other hand, may be quite artificial: real solid compounds are what they are and their
imperfections are an integral part of them.
To perform #SR measurements positive muons have to be implanted into the
material of interest. How innocent are the implanted #+? Do they, by their presence,

60
MUON SPIN ROTATIONSPECTROSCOPY 61

modify local properties, including magnetic ones? Note that practically only one
muon will be present at a time in the sample (average # lifetime is 2.2 #s) and that
the total number of #+ implanted during an experiment exceeds rarely 101° which
number is to be compared with the usual density of host lattice atoms. In effect
radiation damage will be of no concern whatsoever.
In metals the muons's positive charge will be screened jointly by the conduction
electrons piling up at and around the #+. The screening is usually, at normal con-
duction electron densities, accomplished within a distance of the order of the Bohr
radius which is small compared with interatomic distances in a solid. Friedel os-
cillations will cause a ripple in the conduction electron density distribution outside
of the screening cloud which will produce additional electric field gradients at the
nearest host neighbor sites (also in cubic systems). Most importantly the presence
of the #+ causes a local lattice dilatation: the nearest neighbor host atoms may
be pushed away by a few % of their rigid lattice distance from the interstitial po-
sitions, changing also locally the interatomic distances. These are all well known
effects when hydrogen is introduced into a metal in small concentrations (with #+
one probes the infinite dilution limit!). It is conceivable that the changed distances
could have an effect on the magnetic coupling of the atoms in the vicinity of the
#+, but so far there is no compelling evidence for such a possibility. In any case
#SR-data have always reflected magnetic phase transitions at temperatures in agree-
ment with bulk determinations of Tc or TN. The only clearly established effect of
a #+ induced local modification was found in PrNi5 (see section 4.2.4) where the
crystalline electric field (CEF) splitting of the ground state of the #+ nearest Pr 3+-
neighbor was significantly altered. The consequences of this observations for the
interpretation of magnetic features observed by #SR in other rare earth or actinide
based compounds are not clear yet but will require attention in all future investiga-
tions.
In insulators the #+ is observed to be present in the form of the strongly para-
magnetic muonium (#+e-) atom or is found to be bound chemically to one of the
constituents (in particular to oxygen, if present). So far muonium has only been
observed in the magnetic compound MnF2, while in all other investigated magnetic
insulators the #+ appears to be in a diamagnetic state (see section 5). No indications
for #+ induced modification in local magnetic features were ever manifest.
So the answer to the above questions is that, as far as experience teaches, the
positive muon is a fairly innocent magnetic probe, but that this has to be studied in
each case, where it could matter, with care.
Another problem in #SR studies is connected to the question of the #+ site in the
lattice after implantation. To extract quantitive information from #SR-data knowl-
edge of the site is a prerequisite, but often the site is not known. Therefore the reader
will notice that the discussion of #SR results is almost always intimately interwoven
with considerations of the muon's possible or actual site or sites.
This review attempts to include all material on #SR studies of magnetic compounds
up to the sixth international conference on #SR spectroscopy, held in June 1993 on
the island of Maui, Hawai. However, we were forced by space and time limitations
to skip one very important field of #SR-applications in magnetism, namely the very
62 A. SCHENCKand F.N. GYGAX

successful study of spin glasses. It is hoped that this very special subject will find
coverage in some future review article.

2. Muon spin rotation (pSR) spectroscopy

#SR spectroscopy is a variant of other well known hyperfine probe techniques such
as NMR, 7"/PAC and PAD, and M6ssbauer spectroscopy. Therefore in #SR spec-
troscopy one measures basically the same parameters as in the other methods. In the
following section 2.1 these parameters will be briefly recapitulated and some specia-
lities, when using the #SR technique, will be pointed out. The technique itself will
only be sketched briefly. The interested reader is referred to the many articles that
describe the technique in detail (see, e.g., Chappert 1984, Schenck 1985, Chappert
and Yaouanc 1986a, Cox 1987, Seeger and Schimmele 1992, Smilga and Belousov,
1994). Section 2.2 discusses the connection between the #+ site and local magnetic
fields and section 2.3 provides some material on #+ spin relaxation due to the tem-
poral fluctuations of the host magnetic moments. By far most of the investigations
were done by using positive muons (#+). In a very few cases also negative muons
(#-) have been applied. Positive muons are usually implanted at an interstitial site
(the same site that is usually occupied by hydrogen in metals) and probe the mag-
netism from this point of reference. In contrast negative muons are captured into
a ground state Bohr orbital of the host crystal atoms at substitutional positions. In
this respect # - S R has more in common with NMR. However, the negative charge
of the # - , close to the host nucleus, reduces the total nuclear charge seen by the
electrons to effectively ( Z - l), thereby transforming the #--atom to an impurity
atom, different in valency from the original one. In addition the captured # - has a
much reduced polarization in the lowest Bohr orbital (< 1/6 of initial polarization)
and a reduced effective free decay rate which renders # - S R much more difficult and
limited in applications.

2.1. Parameters + phenomenology


2.1.1. Brief description of the #+SR technique
#SR spectroscopy rests on the weak decay of the #+ : #+ --+ e + + ue + ~u which,
because of parity violation in the weak interaction, leads to an asymmetric distribution
of the decay positron with respect to the spin of the decaying #+:

Are+(0) oc 1 + A cos 0, (2.1)

where 0 is the angle between the e+-trajectory and the #+ spin. (Other relevant
properties of the #+ are listed in table 2.1.) Hence by measuring the positron
distribution it is possible to determine the original #+ spin direction. This, of course,
requires the observation of many decays and implies that the participating #+ all
possess initially the same spin orientation, i.e. that they are polarized. Polarized
#+-beams with polarizations up to ~ 100% are available at the so called meson
MUON SPIN ROTATION SPECTROSCOPY 63

TABLE 2.1
Some properties of the muon where me is the electron mass, mp the proton mass and /~p the
proton magnetic moment.

Property Values
Mass (m u) 206.76835(1 l)me
= 0.1126096mp
= 105.6595 MeV c -2
Charge +e, - e
Spin (I) ! h
2
Magnetic moment (/~) (in units of/~p) 3.1833455(5)
Gyromagnetic ratio (7~/2r) 13.553879 (± 0.2 ppm) kHzG -1
9 factor (g~) 2.002331848(17)
Direction o f / ~ ±l-y~lI (+ : #+, - : # - )
Lifetime (%) 2.19703(4)/~s

TABLE 2.2
List of proton accelerators with #SR facilities (1993).

Name Country Beam mode /z+-beamlines and/~SR-instrumentation


PSI Switzerland DC 2 surface (4 MeV)/~+ beamlines (1 dedicated to/~SR),
2 decay beamlines (20-50 MeV), 6 spectrometers
available, including a low temperature set-up (dilu-
tion refrigerator) and a high pressure set-up. Good for
ZF, LF, TF-measurements. Ultra slow muon beam line
under development.
ISIS(RAL) UK pulsed dedicatedsurface/~+ beamline with 3 experimental
ports, one equipped with a dilution refrigerator. Best
for ZF and LF-measurements. 1 decay beamline with
several ports (RIKEN-RAL), under construction.
JINR (Dubna) Russia pulsed 1 decay beamline, several ports 1 general purpose
or DC spectrometer(MUSPIN). LF, ZF, TF-measurements
possible. Surface muons also available in the future.
PNPI (Gatchina) Russia pulsed 1 decay beamline, ZF, TF-measurements possible.
LAMPF USA pulsed 1 surface/z + beamline, best for ZF, LF-measurements.
TRIUMF Canada DC 3 surface/z + beamlines, 2 decay beamlines. Dilution
refrigerator available, 5 T-SC-magnet, two general pur-
pose spectrometers with various cryostats.
BOOM (KEK) Japan pulsed 1 dedicated #+ surface beamline, dilution refrigerator
available, if-resonance spectrometer, superconducting
magnet for LF-measurements. Ultra-slow-muon beam-
line under development.

factories and a few other m e d i u m energy accelerator centers (see table 2.2). T h e
a s y m m e t r y parameter A in eq. (2.2) is basically given by

A = Pa, (2.2)

where P is the b e a m polarization and a an intrinsic asymmetry parameter which


is determined b y the weak interaction decay m e c h a n i s m . If all decay positrons,
64 A. SCHENCK and EN. GYGAX

irrespective of their energy 1, are detected 2 with the same efficiency, an average of
= 0.3 will result. The total asymmetry is thus quite sizable and generally much
larger than in nuclear 13-decays.
If the spin polarized #+ are stopped or implanted in a target in which they are
~bjected to magnetic interactions, their polarization/5 may become time dependent:
P(t). The evolution of/5(t) can be monitored by measuring the positron distribution
as a function of elapsed #+ life time. In fact it suffices to monitor only the positron
rate into a particular direction, say along the direction of the initial polarization
/5(0). This geometry will be assumed throughout this section. It is straightforward to
extend the discussion to other directions of observation. The positron rate dN~+ (t)/dt
as a function of elapsed #+ life time is then given by the expression (see, e.g.,
Schenck 1985).

dN~+(t) 1
d-----~ - No ~ exp(-t/%)(1 + A/5(t)./5(O)/P(O)), (2.3)

where % = 2.2 #s is the average/z + life time and the exponential factor accounts
for the decay of the #+. /5(t) •/5(0)/P(0) 2 can be identified with the normalized #+
spin auto correlation function

G'(t) - (2.4)
(s(o) 2)
This function contains all the physics involved in the magnetic interaction of the
#+ inside the target or sample. Henceforth we will define P(t) as the projection of
fi(t) onto the initial polarization/5(0), choosen to be the direction of the positron
observation, i.e. P(t) =/5(t)./5(O)/P(O) = G(t)P(O). P(t) is called the #SR signal,
sometimes it is also referred to as the asymmetry since it determines the effective
decay asymmetry in the distribution eq. (2.1).
Figure 2.1 represents in a schematic way a typical experimental arrangement for
measuring the distribution given by eq. (2.3) and to extract P(t). Since it involves
the measurement of individual #+ life times (on an event after event basis in a
continuous (dc) #+ beam; in a somewhat different fashion in a pulsed beam with
pulse width << %) it is referred to as a time differential technique. For details see
Schenck (1985). Also time integral detection schemes have been developed which
we will not discuss here. Actual #SR experiments benefit from the availability of
two types of #+-beams. The first, conventional type, is formed from pions decaying
in flight (Tr+ -+ #+ + u~) and involves relatively high #+ energies (~ 40-50 MeV)
and spin polarizations in the range (60-80)%. Muons in such a beam need first to be

1 Since the decay of the # + involves three particles (e +, Ue, r~,) the e + energy will vary continuously
between zero and a maximum energy E ~ _~ 1/2muc 2 ,-- 52 MeV. On the average the e + energy
amounts to ~ 30 MeV.
2 To detect both incoming # + and outgoing positrons plastic scintillators connected via light pipes to
photomultipliers are commonly used.
MUON SPIN ROTATION SPECTROSCOPY 65

~÷ e÷
defector ® B detector
, , , i , i , , , , , , ,

)j* PU ~ 0,1~ " POLARIZATION


be'am- = ~J*Larmorfrequency ~0.0 . ".'" ~'.=

01

target asymmetric 0 12,/3 /, S 6


decay e+ distribut. TIME[psec]
lysis
/
Sfarf _I Clock ]_Stop 3 '1 ' ' ~' ' ' ' ~ / i , i J
I I-
Hisfogramming Dispta y
memory
0 ~
0 1 2 3 4 S 6 7
TIHE[ysec]
Fig. 2.1. Schematic illustration of the/~SR-method pertaining to the transverse field technique, i.e. the
initial /z+-polarization /3(0) is perpendicular to the applied field and the /z+ will perform a Larmor
precession.

degraded by a suitable moderator (the stopping range at 50 MeV is about 15 g cm -2)


before they are brought to rest in the sample of interest, which usually has a mass
per unit area (facing the beam) smaller than the stopping range. The sample has
to have a certain minimum thickness if all #+ are to be stopped. This minimum is
related to the momentum resolution of the #+ beam and to range straggling effects
and amounts to typically several g/cm 2, which prevents the use of thin samples. The
advantage of this type of beam is that/z + will stop rather homogeneously throughout
the sample and that one can be sure, therefore, to really probe bulk properties. The
other type of beam is called a surface beam and originates from pions decaying at
rest close to the surface of the primary pion production target 3. Those #+ have
an energy of < 4.1 MeV, almost 100% spin polarization and a range in matter of
170 mg/cm 2. The main advantage of this type of beam is the possibility to use
rather thin samples (some fraction of a mm), while required lateral dimensions are
similar to those in a decay beam (,-~ (0.5-4) cm2). Occasionally it is observed that
the top layer in which the 4.1 MeV #+ are stopped is different from the bulk so that
rather misleading results are obtained.
P(t) can be monitored under three different conditions: (i) no external field is
present, P(t) evolves solely by the interaction of the #+-spin with internal fields. This
is referred to as zero field (ZF)-#SR; (ii) an external field ~r is present perpendicular
to /3(0). This is referred to as transverse field (TF)-#SR; (iii) an external field is
present parallel to/3(0). This is referred to as longitudinal field (LF)-#SR. Accessible
parameters are discussed below.

3 Pions are produced by high energy (> 180 MeV) protons impinging on a suitable target, e.g., graphite.
66 A. SCHENCK and EN. GYGAX

2.1.2. Larmor or precession frequency: wu


The Larmor frequency or the splitting frequency of the #+ (ra = +l/2)-Zeeman
states is usually measured by~ arranging a magnetic f i e l d / t to be perpendicular to
the initial spin polarization P(0) of the implanted #+ (transverse field (TF)-pSR).
The signal then observed is equivalent to the free induction decay signal in NMR.
Alternatively the splitting can be measured by magnetic resonance techniques in
which case /4 is directed parallel to /5(0). wu and H are related to each other
through the gyromagnetic ratio %, i.e.

wu=%H. (2.5)

wu is independent of the relative directions o f / ~ and/5(0). For arbitrary angles 0


between/~ and fi(0) the evolution of the #+ polarization component along/5(0) is
given by (precession on the surface of a cone with aperture 0)

P(t)/IP(0)l = cos 2 0 q- sin 2 0 cos wut. (2.6)

The field H may be extemally applied or may be intrinsically present at the #+-
stopping site. Of course there could be different fields at different sites in which
case several differently precessing components may be observed, i.e.

P(t)/IP(O)I = ~ Ai (cos 20i + sin 2 0i cos wit) (2.7)


i

with

Ai = ~ Pi(O)/P(O) = 1. (2.8)
i i

To make the splitting readily visible P(t) is often subjected to a Fourier analysis
(Brewer 1982). The measurement of wi determines the local fields Hi.

2.1.3. Signal amplitudes: Ai


Ai determines 4 the fraction of #+ or, since the #+ are implanted more or less homo-
geneously across the sample volume, the fraction of sites or of the sample volume
which is associated with a particular Hi. Because the initial /5(0) is usually well
known and because of the normalization condition, eq. (2.8), the Ai can be absolutely
determined, including the identification of a so called missing fraction which may
arise from extremely quickly depolarizing or from extremely rapidly precessing #+,
introducing a time dependence too fast to be resolved by the spectrometer. The cos 2 0
term(s) in eqs (2.6) and (2.7) may also be invisible due to their time independence.
The amplitude of the time dependent term, Ai sin 2 0i, allows to determine Oi and
therefore the direction of Hi with respect to the initial polarization/~(0). This is,

4 The Ai may also be called (partial) asymmetries as follows from eqs (2.7) and (2.8) and the fact that
P(t) determines the effective decay asymmetry.
MUON SPIN ROTATION SPECTROSCOPY 67

of course, restricted to one domain monocrystalline samples. In polycrystalline or


monocrystalline samples with randomly oriented domain magnetizations all possible
directions of Hi with respect to /5(0) may occur and one has to integrate over all
possible directions in which case
X
2 1 2
P(t)/lfi(O)l=~)--~A~ ~1 + -c ocos
it3 /) = 3 + ~ Ai cos wit. (2.9)

The amplitude of the precessing component is reduced to 2/3. The l/3-term just
reflects the fact that for a completely random orientation of/~i on the average 1/3 of
all #+ will experience a n / t i parallel to their spin and hence will not precess. More
complicated situations may be visualized in systems with complex or incommensurate
magnetic structures. In any case a careful study of signal amplitudes as a function
of sample orientation could be of great help in unravelling the structure of internal
fields.

2.1.4. Knight shifl: Ku


The Knight shift is defined as usual:

Hint-Next (47r - N ) X , (2.10)


K• -- Hext

where the second term corrects for the demagnetization and Lorentz field. N is
the demagnetization factor of the sample and X the bulk magnetic susceptibility (in
emu/cm3). For more details see section 2.2.

2.1.5. Spin lattice relaxation rate: /~i = 1/T1


It describes the rate of repopulation between the #+-Zeeman states (m = ±1/2),
if there is a deviation from thermal equilibrium, and involves spin flip transitions.
The spin flip transitions are induced by fluctuating magnetic fields with components
perpendicular to /5(0). Each spin flip transition requires an energy transfer of wh
from or to the lattice heat reservoir. The initially polarized/z + (being in one of the
two Zeeman states) will become depolarized according to

P(t) = I~(O)l exp(-Alt). (2.11)

In #SR, P(t) is measured either in zero external field (ZF-#SR): PZF(t) or in an exter-
nal field (defining the z-axis) parallel to P(0) (longitudinal field (LF)-#SR): PLF(t).
In ZF-field PzF(t) may be also affected by static field distributions (see below) and
it may not be possible to distinguish dynamic and static features. In longitudinal
fields one can decouple the #+ spin from static fields (see below) and PLF(t) will
only reflect spin lattice relaxation. However spin lattice relaxation may be affected
by the applied field (see, e.g., Abragam 1970, Slichter 1978). For more details on
spin-lattice relaxation mechanisms see section 2.3.
68 A. SCHENCK and EN. GYGAX

2.1.6. Spin-spin or transverse relaxation rate: /~2 = ] / 7 2 (homogeneous line


broadening)
It describes the loss of polarization of the ~recessing spins, usually observed by
applying a magnetic field perpendicularly to P(0) (cos 0 = 0): PTF(t). Equation (2.6)
has then to be modified to
PTF(t) = I/:3(0)le -'k2t COS totzt. (2.12)

Relaxation is a result of dephasing of the precessing #+. It does not involve any
energy transfer between the #+-spin system and the lattice. For very fast dynamics
T2 -~ Ta (see, e.g., Slichter 1978).
For arbitrary direction of H with respect to/5(0) eq. (2.6) can be written as
P(t)//5(9)1 = c°s2 0 e -*IT' + sin 20 e -t/T2 cos tout. (2.13)

2.1.7. Inhomogeneous line broadening, Gaussian relaxation rate


Inhomogeneous line broadening is a result of an inhomogeneous field distribution.
#+ at different sites will feel more or less different fields. Such an inhomogeneity
can be caused by nuclear or electronic dipole fields, which are static on the scale
of the #+ life time. Inhomogeneous line broadening can be observed in a TF-#SR
experiment. Choosing the applied field Hext to be much stronger than any fields
arising from internal sources only the distribution of the internal field components
H int along/~ext are of concern. If the spectral distribution of H int is given by F ( H int)
with f + ~ --zHintb-'(/-/int'~dH' - z = 0, eq. (2.6) (cos 0 = 0) can be written as
; -i-nt--,--z

PTF(t)/Ifi(O) I = ~ 1F 0(3
dto~ F(toz) cos (too + to~)t

= G ~ ( t ) cos coot, too = %Hext. (2.14)

The precessing spins suffer a loss of polarization (again by dephasing) which is


described by the relaxation function GTF(t). The latter is given by the Fourier
transform of the spectral distribution /;'(Hint). In favorable cases F ( H int) may be
made visible directly by Fourier transforming the measured PTF(t).
If F ( H int) is given by a Gaussian distribution: ( 1 / ~ ) e x p ( - H e z / 2 M 2 ) , GTF(t)
will be given by a Gaussian as well

( 1 2 2 ) ( 1 )
GTF(T) = exp -- -~ 7~Met = exp - - a2t2 (2.15)
2
where M2 is the second moment of the field distribution and ~r in this case is usually
referred to as the relaxation rate.
If F(H~) is given by a Lorentzian distribution GTF(t) assumes a simple exponential
behaviour as in the case of homogenous line broadening.
MUON SPIN ROTATION SPECTROSCOPY 69

2.1.8. Inhomogeneous line broadening, zero field (ZF) case


First realized with #+ and so far also restricted to #SR-applications this is a very po-
werful speciality in #SR-spectroscopy. No external field is applied but the implanted
#+ are exposed to randomly oriented internal fields which are also statistically dis-
tributed in absolute values. If each cartesian component of the internal fields is
Gaussian distributed with identical second moment A 2, P(t) (measured along/5(0))
assumes the form

1 2 (1 - ' y u2A 2t 2) exp ( - 1 2A2.2"~


P(t)/I/5(O) I -- GGKT(t)=- ~ q- ~ ~ ")'/zZ3 t ) . (2.16)

This is the famous Kubo-Toyabe function first derived by Kubo and Toyabe (1967)
and later redetected for #SR by Yamazaki (1979) and first observed by Hayano et
al. (1979). If the components are Lorentzian distributed one derives instead

1 2
3 + ~ (1 - %at) exp(-%at),
GLT(t) = ~ (2.17)

where a is the half width at half maximum of the Lorentzian distributions.

2.1.9. Decoupling from static internal fields in longitudinally applied fields (LF)
The Kubo-Toyabe expressions can be generalized to include a longitudinally (i.e.
parallel to fi(0)) applied field H. For Gaussian distributions Hayano et al. (1979)
derive

G~w(t, H) = 1 - 2 H---
~ 1 - exp i1 )
- ~ "/2AZt2 cos(%Ht) + )
H03 ao exp(-y2/2) sin(Hy/A)dy. (2.18)

Figure 2.2 displays G~KT(t,H) for various H including the zero field Kubo-Toyabe
function. The effect of H is to remove the time dependence and eventually for very
large H (i.e. Hext/A >> 1) to restore P(t) to the initial IP(0) I even at long times,
reflecting the decoupling of the #+ spin from the internal fields (the cone on which
a #+-spin precesses shrinks to zero aperture).
In case of very large A the rapid initial decay of GKT(t) may occur within the dead
time of the spectrometer and only the asymptotic GKT(Oe) may be observed. Even
in this case A can be determined by measuring the field dependence of GKT(ee, H).

2. i. i0. The effect of slow fluctuations (motional averaging) on the Kubo-Toyabe


signal
In the two previous sections static inhomogeneous field distributions were assumed
from the outset. If there is also some time dependence involved, i.e. the internal fields
70 A. SCHENCK and EN. GYGAX

1.0
_5

0.8

0.6

L~ 0.4

0.2

0.0
0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2 3.6 4.0
Time ((},uA) "1 )

Fig. 2.2. Display of the LF and ZF Gaussian Kubo-Toyabe function GGKT(t,H). The time is measured
in units of (%,A) -1 and the applied field H in units of zl. For H/A >> 1 G~T stays close to 1 reflecting
the decoupling from the internal fields.

become stochastic functions of time, the ZF Kubo-Toyabe function modifies to what


is called a dynamic Kubo-Toyabe function. In general this function cannot be written
in analytical form. Assuming a Gaussian-Markovian process Kubo and Toyabe in
their seminal 1967 paper were the first to include also stochastic fluctuations. In #SR
the dynamic Kubo-Toyabe function is usually calculated on the basis of a strong
collision model, which implies a pure Markovian process (Hayano et al. 1979). The
parameter introduced in the latter treatment is a fluctuation rate u of the local fields,
whereby each fluctuation event destroys the correlation between the internal field
distributions before and after the event completely. In order to produce an effect
the spectral distribution in the static limit has to be Gaussian or such that it has
a finite second moment. This excludes a Lorentzian distribution from the present
considerations. Figure 2.3 displays the effect of u on a Gaussian Kubo-Toyabe
signal. For very small u (i.e. u/TA << 1) only the 1/3-term is affected and decays
like

GzF(t) = G~T(t, u)lt>3/za =


1 ( 2 )ut
~ exp -- ~ (2.19)

reflecting directly the fluctuations rate u. For very fast fluctuations (i.e. u/%A >> 1)
one gets

GzF(t) = GP~T(t,u) = exp


( 22)--
27u A
/J
t (2.20)

which is of the form of eq. (2.11) and we are back in the spin lattice relaxation
regime.
MUON SPIN ROTATION SPECTROSCOPY 71

1.0

0.8

0.6

0 0.4
~\\ \ \ _ ~ v/A = 0

0.2

0.0
0 2 4 6 8 10
Time ((~t/~) -1 )

Fig. 2.3. Display of the ZF dynamic Kubo-Toyabe function OGT(t, u), where u is the fluctuation rate
of the local fields, u is expressed in units of 7 u A and the time in units of (%~A)-I.

Similarly the longitudinal field Kubo-Toyabe signal can be calculated. For very
rapid fluctuations eq. (2.11) is again recovered. For extremely rapid fluctuations we
enter again the spin lattice relaxation regime where depolarization in a longitudinal
field is only arising from spin flip transitions. See also section 2.3.

2.1.1 I. Special features in dilute magnetic systems


By dilute systems we mean on the one hand spin glasses, where the magnetic ions
are indeed only present in small concentrations but on the other hand also perfectly
ordered magnetic systems which contain a small number of randomly missing mo-
ments (magnetic holes) or which contain a few lattice defects or impurity atoms,
perturbing the magnetic structure only slightly. Implanted/z + in such systems pos-
sess a wide distribution of distances to the magnetic ions or the defect centers and
hence also a wide distribution of magnetic coupling strength with these centers. As
is well known the spectral distribution of fields probed by the #+ (or any other hy-
perfine probe) assumes under such circumstances a Lorentzian distribution (Walstedt
and Walker 1974) and the W-evolution of the #+ polarization follows eq. (2.12)
or in ZF eq. (2.17), respectively. (For refined expressions in the case of real spin
glasses see Uemura et al. 1985)
If the internal fields are allowed to fluctuate in time with a unique rate u eq. (2.11)
changes to

PLF(t) = I/3(0)1 exp ( - x/~¢) (2.21)

(McHenry et al. 1972). For very fast fluctuations GTF(t) and GGT(t) assume like-
wise this square root stretched exponential form. Note that for an ideal Lorentzian
distribution the fluctuations should have no effect on the time evolutions of the #+
polarization. However, in reality even for very dilute magnetic systems the spectral
72 A. SCHENCK and EN. GYGAX

distributions always possess a finite second moment, facilitating motional averaging


effects (see, e.g., Uemura et al. 1980).
An obvious generalization of eq. (2.21) is the stretched exponential expression

PLF(t) = IF(o)l exp ( - ( A l t ) ~) (2.22)

which sometimes describes #SR-data quite well. Its implication may not always be
clear. One cause could be a distribution of fluctuation rates u rather than an unique u.

2.2. Muon site and local fields


Like any other hyperfine field probe the #+ interacts with magnetic atoms through the
dipole-dipole interaction and more indirectly in metals through the RKKY mechanism
by means of which a non zero spin density is induced at the #+ site leading to a
contact hyperfine interaction. In insulators a non-zero spin density at the #+ may
be caused by transferred hyperfine fields involving covalency effects. In general the
effective magnetic field at the #+-site in the absence of an external field is given by

/31oc =/3¢ + B~ip, (2.23)

where/3c is the contact hyperfine field and Bdip the net dipolar field. /~dip can be
expressed as

. (ff(ei) 3(ei - ~'.)(/7(,~i) • (,~i - e.)) "~


) (2.24)

= ~ ~dip (~/z -- ~ ) ' ~(ri), (2.25)


i

where the sum runs over all magnetic atom positions ~ (lattice sum) and fi(~) is
understood to be the static component (thermal average) of the total moment at
a given site. The #+-position ~'u and the atomic sites are measured from some
convenient origin. It is clear that Bdip depends crucially on the assumed #+-position
+-~
and on the assumed orientations of the fi(Y0- The symbol Adip denotes the dipolar
tensor with tr(Adip) = 0.
In the ferromagnetic case all fi(~) will point in the same direction and are usually
of the same magnitude. The evaluation of the lattice sum is split into two parts by
separating the volume of the sample into a sphere around the ~+ (the Lorentz sphere)
and the rest. Summing over the rest yields the Lorentz field BL = (47r/3))~rs and the
+-+

demagnetization field /~dem = -- N - J~b, where -Ms is the magnetization of a single


++
domain (containing the #+) and .~rb is bulk magnetization of the whole sample. N is
the demagnetization tensor. It is hereby assumed that the Lorentz sphere is smaller
than a typical domain size. In non-magnetized ferromagnetic samples (all domain
MUON SPIN ROTATIONSPECTROSCOPY 73

magnetizations will add up to a zero total magnetization) Bde m = 0. Evaluation of the


lattice sum inside the Lorentz sphere can be done by using the Ewald method (Meier
et al. 1987) which assures a rapid convergence of the lattice sum. From symmetry
considerations it will become immediately evident that in cubic systems with the #+
at a site of cubic point symmetry (e.g., tetrahedral and octahedral interstices in an
fcc-crystal) the net dipolar field from the sources inside the Lorentz sphere will be
zero. The tetrahedral and octahedral interstices in a bcc-crystal, on the other hand,
which possess no cubic point symmetry, are associated with a non-zero net dipolar
field. In this case the two types of sites each split into up to three magnetically
inequivalent sub-sites with populations depending on the direction of the ordered
moments (or the domain magnetization). The overall cubic symmetry then requires
that the sum of net dipolar fields, weighted by the relative population of the sub-sites,
yields zero again. Hence, if the #+ should happen to be implanted at tetrahedral sites
in a ferromagnetic bcc lattice one may expect to find up to three different precession
frequencies, the sum of which weighted by the signal amplitudes yielding indeed
zero (see, e.g., the #SR-results on ferromagnetic iron, section 3.1).
In antiferromagnetically ordered systems J~dem a n d /3L are necessarily zero. For
simple antiferromagnets the lattice sum may be split into sums over ferromagnetic
sublattices. Again symmetry considerations can help to identify those interstices at
which the dipolar fields will cancel. Powerful programs have been developed which
allow to calculate the net dipolar fields or the lattice sums for any kind of anti-
ferromagnetic structure, whether single ~ or multiple q*, helical or with modulated
moments. Commensurate structures will always lead to distinct net dipolar fields
at a given type of site, but there may be many magnetically inequivalent sub-sites
leading to a corresponding number of distinct precession frequencies. Incommensu-
rate structures are more difficult to handle and usually are approximated by invoking
very large magnetic unit cells. In any case implanting #+ in such systems a more
complicated extended frequency spectrum can be expected, reflecting a more or less
inhomogeneous field distribution over the #+-sites, which will lead to relaxation by
dephasing (see section 2.1). If the local fields over the #+ sites vary sinusoidally in
one dimension P(t) will be given by the Bessel function Jo(t).
In metals the contact hyperfine field, or local spin density, respectively is induced
by the magnetic moments via the RKKY mechanism which is linear in the moments.
In ferromagnetic metals Bc is therefore proportional to the domain magnetization,
i.e.

/~c = Ac_~r~, (2.26)

where Ac is a contact hyperfine coupling constant. Although Bdip may be zero at


certain sites Bc will in general be different from zero. On the basis of the RKKY
mechanism Ac may be expressed as (Kittel 1966, Schenck 1993)

87r g] - 1
Ac - __ Jsf(~,)r/(,Ftz), (2.27)
3 gj
74 A. SCHENCKand EN. GYGAX

where gj is the Land6 factor of the electronic moment, Jsf(ft,) is an effective exchange
coupling constant which depends on the #+ site via the distance to the neighboring
electronic moments, and r/(Yu) is the so called spin density enhancement factor which
incorporates changes in the local electronic density distribution due to the presence
of the #+. r/(~'u) may be determined from Knight shift measurements in the para-
magnetic state at high temperature (only the Pauli spin paramagnetism is left).
In an antiferromagnetic structure Be may be written as
.., 8~ gj -- 1
Be(r,,) : T rl(%) ~ ~ Jsf ([g~, - ~l)fi(~), (2.28)
i
where we assumed that the spin density induced at the #+ by each electron moment
fii at position ~ can be simply superimposed. 0Vsf(l~'u- ~l) is an effective exchange
coupling constant for each moment fi(r~). This expression shows that in an anti-
ferromagnetic structure Be can become zero at certain sites of high symmetry with
r_eespect to the magnetic structure. From symmetry considerations follows that also
Bdip will vanish at such sites quite in contrast to the ferromagnetic case.
So far we assumed the #+ to be fixed at the geometrical center of an interstitial site.
In reality the #+ performs zero point vibrations around this position. As Meier (1980)
has shown this has no effect on Bdip as long as the #+-wave function is well confined
to the interstitial site volume and does not overlap with the neighboring host magnetic
moments. In contrast the contact hyperfine field/3c may be modified considerably
by both the >+ zero point vibration as well as lattice vibrations (Manninen and
Nieminen 1981, Estreicher and Meier 1982, 1984). Inclusion of such effects can
account for small deviations between the temperature dependence of/3c and _~rs or
the sublattice magnetization in antiferromagnetic systems, respectively.
In the paramagnetic phase static magnetic moments can be induced by an external
field Hext. At each magnetic atom site one has

#+ : X++
a t " /~ext, (2.29)
++
where Xat is an atomic susceptibility tensor. The moment arrangement corresponds
to a ferromagnetic order. Bdip (eq. 2.25) can then be rewritten as follows

/~dip(~'/~) = ~ ~dip,i (7~/~ -- ~i)" ~ a t " / t e x t (2.30)


i

= Adip (r/z) " ++


Xat" /~ext, (2.31)
..o.
where Adip (UIz) is again a traceless and symmetric tensor which depends on the
assumed #+-site in a given crystal structure. Its calculation is straightforward. The
total field (neglecting now BL and Bdem) at the #+ is

B. : I( xt + ETdip)l (2.32)

H xt + (&ip. B xt)/H xt + . . . . (2.33)


MUON SPIN ROTATIONSPECTROSCOPY 75

from which follows the dipolar Knight shift constant


+4 +at
Kdip= (JtIext • Adip (r/z)' Xat" [text)/H2xt • (2.34)
Similarly the contact hyperfine field is given by
/3c(r~) = Ac Xat" /r~ext, (2.35)
from which we obtain the contact hyperfine Knight shift constant
Kc = Ac (I-Iext ' ++ Hext)/Hext"
Xat" " 2 (2.36)
+-1.
This constant is isotropic as long as Xat is isotropic.
Often one has to deal with axially symmetric systems in which case the total
Knight shift is given by
1 / \
K = ~ Ac ~(XII + 2X±) + 2(Xll - x±)Pz(cos 0))

+51 Adip ((Xll -- X±) + (X± + 2XII)P2(cos O)), (2.37)

where 11, _1_refer to directions of/text parallel or perpendicular to the axis of axial
symmetry. Specifically one finds

Kit = (Ac + Adip)X[],

K± = (Ac- ~1 Adip)X±. (2.38)

If X is taken to be the molar susceptibility in emu, Ac and Adip will be given in units
of (mol/emu). If X is understood to be the atomic susceptibility measured in units of
(#B/emu) Ac and Adip will be given in units of (emu/# B) per atom. Very often Ac
and Adip are quoted in these latter units. Ac contains quite important information in
that it depends on the exchange interaction between the conduction electrons and a
local magnetic moment. The same exchange Hamiltonian is also responsible for the
Kondo mechanism. However Ac depends also on the electronic structure established
at and around the implanted #+. This feature has so far prevented, with some rare
exceptions, a detailed analysis of measured' Ac-values.
Kdip (or /~dip in the ordered state) is less affected by the presence of the #+ and
reflects more directly intrinsic properties. Some lattice relaxation around the/z + will
lead usually to a small reduction of Adip. Basically Adip is proportional to the atomic
susceptibility (or the local moment) of just the nearest neighbours which are placed
far enough away from the #+ to remain usually unaffected by its presence. Recently,
however, some evidence has been obtained that the #+ could cause a change in the
crystalline electric field splitting of rare earth atoms thereby changing their magnetic
response (see section 4.24, Feyerherm et al. 1994b). Whether this could also have
an effect on the size and orientation of ordered moments next to the #+ is not known.
In any case one has never found any indications that the phase transition temperature
measured locally by #SR is any different from its bulk value.
76 A. S C H E N C K a n d E N . GYGAX

2.3. More on spin-lattice relaxation in the paramagnetic and ordered state


#+ spin lattice relaxation or spin flip transitions are induced by fluctuations of the
local field components perpendicular to the initial spin polarization. As in the static
case the local fields may be of dipolar origin or of the contact hyperfine field typg.
The fluctuations reflect the dynamics of the electronic moments or spins at the mag-
netic atom positions. Among other hyperfine probes #SR is special in that dipolar
fields can be as strong as the contact hyperfine fields and contribute strongly to the
#+ spin lattice relaxation rate. In contrast, in NMR and M6ssbauer spectroscopy
dipolar field induced relaxation is usually negligible and not taken into account (see,
e.g., Hohenemser et al. 1989)•
Quasi-elastic neutron scattering is another method to investigate the spin dynamics.
The differential cross section dcr2/dOd03 for magnetic scattering is proportional to
the frequency and wave vector (momentum transfer) dependent Van Hove response
function S(q, co) (Lovesey 1987), which can be approximated by

v(¢)
s(¢, w)= _1kBTx(0") (2.39)
71" 032 q - / ' ( 0 " ) 2

X(0~) is the wave vector dependent susceptibility a n d / ' ( ( ) is inversely proportional


to the lifetime of excitations with wave vector ¢. The #+ spin lattice relaxation on
the other hand, can be expressed in the simplest case as

A1 --- A 2 7r ~ ,S'(~, co), (2.40)


N ¢

which in zero field or for co/l'(() << 1 is simply

kBTA 2 x(q) (2.41)


- r(¢)
q

A is proportional to the magnetic coupling strength. More complete expressions can


be found in Yushankhai (1989), Lovesey et al. (1992a, b), Yaouanc et al. (1993a)
and Keren et al. (1994b).
Three important temperature regions have to be distinguished: (i) ordered regime
(ii) critical regime just above the ordering temperature and (iii) high temperature
paramagnetic regime•
(i) In the ordered regime the dynamics arises mainly from spin wave excitations•
However, since magnon energies are usually much larger than the #+ Zeeman split-
ting energy, conservation of energy forbids a direct (single step) process in which a
single magnon is created or absorbed by a muon spin flip. Relaxation can only be
affected in lowest order by a two step or Raman process. On this basis and describ-
ing the interactions of the host magnetic moments by a Heisenberg Hamiltonian and
including a magnetic anisotropy term Yaouanc et al. (1991) have calculated in an
MUON SPIN ROTATIONSPECTROSCOPY 77

approximate way the #+-spin lattice relaxation rate in a ferromagnetic system. In


the limit of Ean (= anisotropy energy) << k B T they arrive at

-
9 (.y,g,,B)2a k T2 In (k,r) (2.42)
167r3 h:D 3 \ ~an ]"

G is a factor which depends on the #+ site and the lattice geometry (it may be
zero) and D is the spin wave stiffness constant. Note that only the dipolar coupling
of the #+ to the electronic moments leads to eq. (2.38) while the isotropic contact
hyperfine interaction does not contribute. It is also found that )~1 is dominated by
contributions from near the center of the Brillouin zone, i.e. by long wave length
magnon excitations. Another remarkable feature is the fact that only longitudinal spin
fluctuations (i.e. parallel to the easy axis) are effective in inducing/,+-relaxation -
also in zero external field - and can thus be studied by #SR under zero field conditions
while polarized neutron scattering requires a non zero applied field.
Two magnon induced relaxation in a ferromagnetic has also been considered by
Lovesey et al. (1992a) with special reference to #+ stopped in EuO.
Keren(1994) has considered a different, possibly quite efficient depolarization
mechanism for the #+ with special reference to antiferromagnetic MnFe2 by noting
that typical magnon energies could match well with the splitting of vibrational states
of the #+. In this case absorption of a single magnon could promote the #+ to an
excited vibrational level accompanied by a spin flip conserving both energy and the
total angular momentum. It is expected that the temperature dependence of the thus
induced relaxation rate will follow an Arrhenius law.
(ii) In the critical regime just above Tc the spin dynamics is crucially affected
by also the dipole-dipole interaction between the magnetic ions (Frey and Schwabl
1988, 1989b). This gives rise to the appearance of two separate terms which describe
the spin dynamics along q (longitudinal mode, L) and perpendicular to ( (transverse
mode, T). The two modes are differently effective in inducing #+ spin relaxation
depending on the relative strength of the contact hyperfine and the dipolar coupling
of the /~+ to the magnetic ions with angular momentum J and Land6 factor 9L.
For an isotropic dipolar cubic (fcc) ferromagnet Yaouanc et al. (1993a, b) derive
the following expression for zero applied field in the small q" limit (appropriate for
temperatures close to Tc)

= - dqq 2 2p 2 - - + (1 _p)2 (2.43)


v ao Ur(g) ~ '

where the weight factor p is given by

1 n.H
- - (2.44)
P=3+ 47r

H is the contact hyperfine coupling constant (in units of 9 L # B / V where v is the


volume of the primitive cell of the Bravais lattice) between the #+ and one of
78 A. SCHENCK and EN. GYGAX

the nu nearest neighbour magnetic ions and 1/3 arises from the dipolar coupling.
If nuH/47r >> 1 (dominating contact hyperfine interaction) the transverse mode is
dominantly contributing while in the other extreme of dominant dipolar interaction
the longitudinal mode is most effective.
xL'T(() and/~L,T(q) obey scaling laws in the critical temperature regime (dynamic
scaling theory, renormalization-group calculations). Actual expressions may be ob-
tained from the mode-coupling theory (Frey et al. 1988, 1989a, b). The susceptibility
can be written in the Ornstein-Zernike form like

xT(q ) _ T
q2 + ~-2 (2.45)

and

xL(q ) = r
q2 + q~ + ~ - 2 ' (2.46)

where £ is the correlation length and qD is a dipolar wave vector (Yaouanc et al.
1993a). The correlation length has a power law temperature dependence

T-Tc) -~
(2.47)
= ~o gc '

where u = 0.705 for a three dimensional Heisenberg magnet. I'L'T(() is more


difficult to evaluate and the reader is referred to Yaouanc et al. (1993a) and Lovesey
et al. (1992b). Using the scaling relation

F(¢) c<qZF(x) (2.48)

with
1
x = -- (2.49)
~q
(z is a critical exponent) and neglecting the dipolar interaction between the ionic
moments (spin conserving dynamics, z = 5/2) and between the #+ and the ionic
moments one finds (for z > 0)
~1 (3( ~3/2 --~ ¢ 2 + z - d - r t , (2.50)

where d = 3 is the dimensionality, r/~_ 0.03 for a 3D-Heisenberg system. Admitting


now also non spin-conserving dipole-dipole interactions between the ionic moments
one finds that z = 2 for the transverse mode and z = 0 for the longitudinal mode
if the dipolar interaction is dominating over the Heisenberg exchange (i.e. q << qo)-
Consequently

AT o( ~ (2.51)
MUON SPIN ROTATION SPECTROSCOPY 79

and k } is temperature independent.


In the general case AI,ZF may be written as

/kl,ZF = W [2p2/T(qo) + (1 - p)2IL(T)], (2.52)

where ~ = arctan (qD~) and the functions IT'L(~o) are universal functions which have
been evaluated by Frey and Schwabl (1988). A similar expression can be derived
for a body centered cubic (bcc) lattice (Yaouanc et al. 1993a).
Critical behaviour in an antiferromagnet has been considered by Lovesey (1992)
(see also Keren et al. 1993), in the latter paper a comparison between #SR and
neutron data has been worked out). For an isotropic antiferromagnet the #+ spin
lattice relaxation rate diverges on approaching TN like

)~1 (X {1/2 (2.53)

for both contact hyperfine and dipolar coupling. In an uniaxial antiferromagnet


critical fluctuations are ineffective in inducing #+ relaxation if the coupling to the
ionic moments is mediated by the contact mechanism. On the other hand, if the
coupling is of dipolar origin one finds

(T-TN ) -~' (2.54)

where ~z is the correlation length along the anisotropy axis. The temperature depen-
dence in eqs (2.53) and (2.54) are precisely the same predicted for the NMR line
width (l/T2). Equation (2.54) has been used to explain #SR results in MnF2 (see
chapter 5).
(iii) Spin lattice relaxation induced by paramagnetic fluctuations of local moments
at high temperatures, i.e. far above any critical regime, have not been treated to
any great extent. Assuming that the local moments fluctuate uncorrelated at high
temperatures (F(q) = 1"(0) = r c 1, x(O) e< l / T ) one finds

),1 ~x rc, (2.55)

where rc is a local correlations or relaxation time associated with the ionic mo-
ments. Two processes may be responsible: the RKKY interaction and the Korringa
mechanism. The former induces a temperature independent, the latter a linear T-
dependence of the ionic moment fluctuation rate %-z, i.e. (Hartmann et al. 1986)

7-c =
i (
c'Tc J / ( J + 1) + -£ (IsfN(eF)) , (2.56)

where e' is a constant, Tc is the ordering temperature (or the paramagnetic Curie-
Weiss temperature), Isf is the conduction electron - local spin exchange integral and
N(eF) the Fermi energy density of states. Tc is proportional to the effective RKKY
80 A. SCHENCKand F.N. GYGAX

exchange coupling constant between a local moment and all other local moments
which in turn is proportional to j2f. The RKKY part of eq. (2.52) follows also from
a treatment by Lovesey et al. (1992b).
Our discussion of the #+ spin lattice relaxation rate has been limited here to
local moment magnets, ignoring also possible effects of the crystalline electric field
(CEF) splitting of the ground state multiplet in f-electron ions. CEF-effects will be
mentioned in some of the sections in chapter 4. Spin lattice relaxation in itinerant
magnets is considered in section 4.1.

3. Review of results in elemental metals and alloys

Since in general the magnetic properties of elemental metals are very well known,
the results of #SR investigations on such systems have served primarily to develop
a better understanding of technique and role of the implanted #+ as a magnetic
probe. As a consequence the present section is organized somewhat differently than
the following sections. While in the latter ones the results are listed compound by
compound, in the present section the results on various elemental metals are collected
in different paragraphs with respect to particular properties. The most salient features
of crystal structures and magnetic properties for the magnetic elemental metals are
sketched in table 3.1. Lists of essential data obtained in # + S R experiments, together
TABLE 3.1
Schematics of the crystal structures and the magnetic orders for the magnetic 3d-transition
metals and the lanthanides. The/z+ site is also indicated; it is in parenthesis if there is
only a strong indication for that option.

Element Crystal T r a n s i t i o n Magnetic Critical /z+


or alloy structure temperature o r d e r temperature site
Fe bcc FM -+ 1044 K (O)
Co hcp -+ 690 K FM O
fcc FM ~ 1398 K
Ni fcc FM --+ 627 K (O)
Cr bcc LIAF --+ 124 K (T±)
TIAF -+ 312 K
CAF
Pd (alloys) fcc varying
Pr hex AF --+ 27 K
Nd hex AF -+ 19 K
Sm complex AF --+ 106 K
Eu bcc AF --+ 89 K
Gd hcp FM -+ 292 K 0
Tb hcp FM -+ 223 K
AF -+ 229 K
Dy orth --+85 K FM -+85K (0)
hcp AF -+ 178 K
Ho hcp FM -+20 K
AF --+ 132 K
Er hcp FM -+19K
AF -+ 85 K
M U O N SPIN ROTATION SPECTROSCOPY 81

N r~ o

-~
o #. =o

~oN
o

oo
~ ~ o • o

~ ' ~
=.~ o~

u~

¢0
O

o odo o
09

O9

~,.~ V/

~+ N F'~NN [-~
n~

+
O

9
"O
oo
o
-I- + + -l- ±

O O O O o
82 A. SCHENCK and EN. GYGAX

~ N - ~ ~ . ~ ~...~ ~'-"

~oz~z.~=&~ ~ :~z ;~oe~'°


~,~,~ ~ ~.~ ~ ~v,,
ed

~<~

d d

+ + ++

Z ~ ~u ~mm
MUON SPIN ROTATION SPECTROSCOPY 83

.. o~
+

r~ 0

t'¢3 "~

r~ o9

g
N

~2

N o o
84 A. SCHENCKand EN. GYGAX

with the basically used experimental procedures are presented in tables 3.2 (3d-
elements and their alloys) and 3.3 (lanthanides). These two tables are meant to give
in the shortest form an overview with reference lists of the experimental #+ studies
subject of the present chapter(sections 3.1 to 3.4). In section 3.5 we report briefly
the results of investigations with negative muons (#-SR) in Fe and Ni.

3.1. Spontaneous dipole and hyperfine fields


This section deals with the results obtained for the 3d-transition metals Fe, Co, Ni,
Cr and for the lanthanides in the magnetically ordered states. Ferromagnetic Gd is
presented as an introductory example in section 3.1.1. This choice is motivated by
the fact that several problems related to the method can be treated in this case and
because the #SR method has shown for Gd its best contribution to the understanding
of magnetism in elemental metals.

3.1.1. Ferromagnetic Gd: an example


The local field at the #+ in ferromagnetic Gd was first measured for two temperatures
by Gurevich et al. (1975a), but the influence of the local dipole field was not
considered in the interpretation. A thorough study was carried out in polycrystalline
Gd by Graf et al. (1977). The results, together with the studies in Co, Dy, Fe and
Ni are comprehensively reported in Denison et al. (1979).
According to section 2.2 the local field at a #+ in an unmagnetized ferromagnet
in zero field is given by

/3.(~.) =/3L + B~ip(f.) +/3c(rn). (3.1)

B~ip(/V~) is the lattice sum inside the Lorentz sphere. Gd has a hcp crystal structure.
Between the Curie temperature Tc = 292 K and 230 K the Gd spins are aligned
along the c-axis. Below 230 K the easy axis of magnetization deviates from the
c-axis. The (non-pSR) literature lists various inconsistent temperature dependences
of the angle 0 between magnetization and c-axis. Figure 3.1 shows ]B.[ versus
temperature in zero field. Clearly this behaviour is very different from Ms(T), the
domain magnetization in Gd. The interpretation (Denison et al. 1979) is summarized
in the following points.
(i) One starts with

4 7r~rs(T) ' (3.2)

where 2~rs is the domain or saturation magnetization.


(ii) #+ sites: The hcp lattice offers tetrahedral (T) and octahedral (O) interstitial sites
for the muon. A calculation of the dipole field tensor (assuming an undistorted
lattice) indicates that the fields at T sites, O sites as well as at substitutional
(S) sites are all parallel - but not parallel to the Gd moments - with of course
different magnitudes.
MUON SPIN ROTATION SPECTROSCOPY 85

(iii) The complicated behaviour of B , versus T is essentially due to Bdip, more


precisely to the variation of B~ip with 0. This is illustrated considering the
behaviour of the field vector 3~, defined as the sum of/3L and/3c, left term of
the equation

(3.3)

It is expected that/3c(T) and hence also J((T) will be approximately proportional


to ~rs(T) - note that BL, Bc and )( are all parallel to Ms. The dependence O(T)
can now be determined in the following way:
For a given #+ stop site one draws a series of X(T) curves with 0 as
parameter, using the measured values of B~(T) and the calculated B~ip(0 , r'tz)-
Such sets of curves are shown in fig. 3.2 a-c for substitutional, tetrahedral
and octahedral p,+ sites. Since only the magnitude o f / 3 , ( T ) was determined
experimentally, combinations using both B,(T) > 0 and B,(T) < 0 are retained.
The boundary condition that X(T) has to follow the 0 = 0 ° curve between 230 K
and Tc has to be considered - the points are indicated by solid circles in fig. 3.2
a-c. From the figure it is obvious that only the assumption of the octahedral site
with Bu > 0 allows X(T) to be approximately proportional to Ms (monotonous
dotted line in fig. 3.2 c).
The O(T) dependence is so determined with an accuracy exceeding that of the neu-
tron diffraction data, the resulting temperature dependence and the values obtained
by other means are presented in fig. 3.2 d. The hyperfine field Be, obtained from

I I I I I I

1.5 20

v
1.0 z g,
v

m
-t 10 tr
G_.~a LL

0.5
m
5 =

0 I r I I i hl 0
0 1 O0 200 300
Temperature (K)
Fig. 3.1. Absolute value of the measured local field B~ vs. temperature in zero external field for
polycrystaUine Gd. The right-hand scale corresponds tothe directly measured/zSR frequency. The solid
line through the data points is meant to guide the eye (Denison et al. 1979).
86 A. SCHENCK and F.N. GYGAX

1.5
1.0
0 5 l Substitutional Site~ 1~ g . o , 5
~o.o ~ 0.0
X "0.5 Bl' < 0 - i -o.5 .,o..-.,. B,, <o -I
- 70"
-1.5 0 -1.5
-2.0 - -2.0
0 1O0 200 300 0 1O0 200 300
90 ' ' 't ' ' '|
2.0 (b) 8O
(d) 1
1.5 I
7O
1.0 '90" 60
0.5 50
~v 0.0 40 °i# ?.',
X -0.5 . B,, < 0 ,"~f
30
r, '
-1.o F" 20 ?"" t'ITo
-1.5 10
-2.0 0
0 1O0 200 30( 0 1O0 200 300
Temperature (K) Temperature (K)
Fig. 3.2. In a, b and c: field 2 =/3~, - / ~ d i p = /~c + BL plotted as function of temperature for three
possible/~+ residence sites and for Bu > 0 and B~, < 0. The solid lines correspond to the indicated
fixed values of the angle 0. For T > 230 K it is known that 0 = 0, the values of X deduced from the
IB~,I measurements are displayed as solid circles for that temperature domain. Only the assumption of
the octahedral/~+ site with B~, > 0 allows X(T) to be smooth and monotonous (dotted line in (c)) over
the full temperature range. From that curve the function O(T)is determined and displayed in (d) as a solid
line; in (d) the 0 values from neutron diffraction (full circles) and torque (full triangles) measurements
are also indicated (from Denison et al. 1979).

Be = X - B L (BL = (47r/3)Ms), is finally s h o w n in fig. 3.3. Bc is o p p o s i t e ( n e g a t i v e )


to Ms. T h e i n t e r p r e t a t i o n o f B c ( T ) for G d will b e briefly d i s c u s s e d t o g e t h e r with
that for the o t h e r f e r r o m a g n e t i c e l e m e n t s i n section 3.1.3.
N e w d a t a o b t a i n e d in a s i n g l e crystal G d s a m p l e b y H a r t m a n n et al. ( 1 9 9 4 )
call for slight a d j u s t m e n t s in the c o n c l u s i o n s w h i c h h a v e j u s t b e e n p r e s e n t e d - see
s e c t i o n 3.1.2.
MUON SPIN ROTATION SPECTROSCOPY 87

I I I I I I

6 ~ , ~ i n ° r m )

4
o ~ \',

0 I I I I I II
0 1oo 200 300
Temperature (K)
Fig. 3.3. Extracted hyperfine field Bc plotted as function of temperature. The dashed curve is the
magnetization normalized to Bc(T = 0 K); Bc is opposite (negative) to the local magnetization.

The particular interpretation of the Bu(T) measurements in Gd illustrates how


new information about magnetism can be obtained from #SR. One has to stress,
however, that since certain features of the #+ in the sample material are not always
perfectly known, great importance has to be attached to the use of adequate assump-
tions, estimates of the importance of effects escaping control, and particularly to the
consistency of the over-all picture obtained after analysis.
Let's just mention two type of problems one can encounter in this context:
- As far as the dipole field calculation is concerned, it matters of course to know
the exact (interstitial) position of the #+. This achieved, the #+ zero-point
motion, #+ tunnelling, departure from the ideal site position, e.g., because of
bonding (in compound substances) or trapping at impurities may also play a
role. Moreover, the muon can deform locally (mainly expand) the lattice -
possibly modifying the local structure anisotropy.
- In case of sufficiently fast #+ motion (i.e. diffusion) a sampling of (magneti-
cally) non-equivalent sites will average the v a l u e s / ~ ( ~ i) corresponding to the
different sites i:

B.~eff= ~ a~/~u(~'.i), (3.4)


i

where a~ stands for the #+ population fraction at site i. So it is possible, as, e.g.,
in Fe, to encounter an averaging, possibly to zero, of the different dipolar field
contributions and thus to observe a single B~ff value instead of the characteristic
multiplet corresponding to the number of non-equivalent #+ sites.
88 A. SCHENCKand F.N. GYGAX

3.1.2. Local fields Bu(~'u) and dipole fields in magnetic elemental metals
Figure 3.4 a-d shows IBu[ as function of temperature for Fe, Co, Ni and Dy. As
in the case of Gd (fig. 3.1), the complexity of these dependences is essentially due
to Blip(T ). Where the dipole fields/3~i p vanish, in Ni because of its fcc structure,
or in bcc Fe because of rapid #+ diffusion among interstitial sites (see section 2.1),
Bu(T) (= BL + Be) is smooth and monotonous, whereas, e.g., in Co the drastic
effect of B~ip is obvious (Graf et al. 1976b). In hcp Co (as in hcp Gd), the ordered
magnetic moments produce a unique dipole field at each type of interstitial site.
This is, therefore, not modified by diffusion, provided that diffusion takes place only
among interstitial sites of the same type. Thus

= - & - (3.5)

B~ip(~/~) can be calculated, e.g., as described by Denison et al. (1979). From the
same ref. one see that/3 lip(T) deviates markedly from ~rs(T) if the direction of the
axis of easy magnetization varies with respect to the crystal structure as a function
of T. As for Gd, this is the case for Co, where in addition a modification from a
hcp to a fcc structure at 690 K changes the picture even more. B~ip vanishes for the
fcc structure.
Measurements on Gd single crystal samples (Hartmann et al. 1990b and 1994,
Kratzer et al. 1994a) confirm in essence the analysis presented in section 3.1.3 but
show that the spin re-orientation between 245 and 220 K is steeper than anticipated
and accompanied by a peak in the #+ depolarization rate. At a closer look a complex
precession signal, which can be separated into two frequencies, is detected in this
temperature range. Hartmann et al. (1994) conclude that the spin turning does not
occur simultaneously in two different domains of the sample. The single crystal data
imply that B~ and B~i obtained by Denison et al. (1979) cannot both be strictly
correct. Either Bc un~fergoes a change around 230 K which is directly coupled to
the spin turning, or the value of the dipolar field used in the earlier evaluation is
slightly too low.
It is noteworthy that so far all the #SR data obtained from the magnetic elements
have been explained using the dipolar field calculation under the assumption of
localized dipole moments.
The pressure dependence of B , in Gd was measured by Hartmann et al. (1990a).
The changes by pressure are much larger than those observed in the 3d-metais Fe,
Co and Ni (Lindgren et al. 1987 and Butz et al. 1987). The essential contribution to
dB,/dp comes from the turning of the axis of easy magnetization with pressure at
low temperature and from the reduction of Tc with pressure at higher temperature.
See section 3.1.4 for the effect of pressure on Bc.
Dy shows rather interesting magnetic properties - see the schematic arrangement
of the Dy spins in fig. 3.4 d. Between TN = 178 K and Tc = 85 K it is a helical
antiferromagnet. The spins lie in the basal plane of the hcp lattice, the helix axis
is parallel to the c-axis and the helix angle c~ between the ferromagnetically ordered
planes is temperature dependent. Below 85 K a discontinuous orthorhombic distor-
tion occurs, with a corresponding phase change to a ferromagnetic state. The spins
MUON SPIN ROTATION SPECTROSCOPY 89

are then along the a-axis. The striking features of the #SR measurements (Denison
et al. 1979) are: (i) B u is observed in the two magnetic phases (i.e. the first time
for an antiferromagnetic state in a metal), which can only be explained if the #+
is not diffusing through the lattice; (ii) a continuous variation of Bu occurs across
the magnetic and structural phase transition. Interestingly the calculations of B~ip(T )
show also a smooth behaviour across the phase change for #+ at the octahedral sites
and only a weak discontinuity of about 4% at this transition for #+ at the tetrahedral
sites. From the measurements neither the muon site nor the sign of Bu could be
determined. See section 3.1.3 and fig. 3.4 h for the deduced Be(T).
Several studies were made about the influence of an elastic strain on Bu in Fe
(Namkung et al. 1984, Kossler et al. 1985, Herlach et al. 1989, Fritzsche et al.
1990). The conclusion is that the observed effect stems essentially from a change
in B~ip, due to a shift in #+ population between magnetic non-equivalent interstitial
sites reduced by modified elastic energies. In this context see also the effect of
impurity induced strain in section 3.3.1.

3.1.3. Temperature dependence of the hyperfine field


For Fe, Co, Ni and Dy the dependence Bc(T) is deduced from B u at the #+, BL and
the calculated/3~i p, using eq. (3.5), and is shown in fig. 3.4 e-h (bottom section). The
hyperfine fields measured with t~- in Fe and Ni are treated separately in section 3.5.
For Fe and Ni no knowledge of the #+ site is required, as discussed previously. For
Co the octahedral interstitial site is assumed in order to obtain a smooth variation of
Be with temperature (except for the discontinuity at the 690 K structural transition),
following approximately the Co magnetization curve Ms(T). In that metal the sign
of B~ (with respect to Ms) was determined to be negative for T < 500 K, below
the spin reorientation between 500 and 600 K, and positive after completion of
the easy axis reorientation for T > 600 K. This was obtained by measuring the
precession sense of the muon (via the #SR signal phase information) in presence
of an additional external magnetic field. (Another way to obtain the sign of Bu is
to follow the measured ]B~[ as function of increasing external field values driving
the sample beyond magnetic saturation - see the case of Ni, Denison et al. (1979),
illustrated in fig. 3.5.)
In a-Fe at low temperature Schimmele et al. (1990 and 1994) have observed
an oscillating signal in a longitudinal #SR experiment on a magnetically saturated
crystal between 30 mK and 600 mK. In the experiment the high-purity single crystal
sphere was magnetically saturated in a (111) direction, say [111]. The dipolar
magnetic fields at the interstitial sites of tetragonal symmetry (like O- or T-sites)
are perpendicular to [111] (Seeger and Monachesi 1982). The fields at different
sites transform into each other through rotations by 120 ° or 240 ° around [111], their
/~(111)
absolute value is labeled ~dip . The only remaining contributions to the local field
are/~ext and/~o, both parallel or antiparallel to [111]. Therefore the local field value
is given by
{ ~,(111)\2] 1/2
Bu = (Bext -t- Bc) 2 -~- I/Zldip ) j (3.6)
90 A. SCHENCK and EN. GYGAX

(a) (b)
I I t I [ I J I I I 12
Fe 800 Co
5O
10
"I-
a ~6 40
I
% 600 I 8 !
% ~2 I
qb
OlD 3o ~ I 6 ==
~2 400
0 ,¢
2o ~ 4
1 le"q f O3

T~
!
10
200
\ 2

o i I I J I l 0 I I ~ I I I 0
2oo 400 600 800 1000 0 200 600 1000 1400
Temperature (K) Temperature (K)

(e) (f)
i i t I i 1.0 i i i ; i ' i i 6

10 0.8 ~ I :
5
o.o 4
8 0.4
I i 2
0.2 1
6
m 0.0
if-i--
.i~we~JiI T, 0

4
-0.4 Bdip h~p . ~ tcc
c ic~ci -3
I : I
2 _o.o - I ', I _,
-
Bc i ' I
-5
0.8 ~ -6
0 I I [ I I 1.0 i i i I I I r I
200 400 600 800 1000 0 200 600 1000 140
Temperature (K) Temperature(K)
Fig. 3.4. In (a), (b), (c) and (d): temperature dependence of the measured absolute value of B u for Fe,
Co, Ni and Dy respectively, for polycrystalline (o) and single crystal ([3) samples in zero field. The spin
structures in the ferro- and helical antiferromagnetic phase of Dy (d) are indicated. In (e), (f), (g) and
(h): temperature dependence of the hyperfine field Bc for Fe, Co, Ni and Dy respectively, extracted from
the IBul measurements (a~l). For Dy the two possibilities due to the unknown sign of Bu are indicated.
For Fe and Ni the magnetization Ms(T) is indicated, whereas for Co the field contributions B~ip and
BL as well as the deduced temperature dependence of Bu are displayed, together with a sketch of
the spin turning and the structural phase transition (Denison et al. 1979).
MUON SPIN ROTATION SPECTROSCOPY 91

(c) (d)

i I 14 C i I
i 1
Ni • Dy
1.5 mllhl=l~ 20 12
"Ib-,,,-
150
c T

qk
t5! 28 v ~oo
ell
u?.
,+
I m co
0.5 4 5O
orthorhombic
5
I 2 -Ferromagnelic--b*-Ant=ferromagnetic1
I
I
0 I I tl 0 0 I I t 1 0
0 200 400 600 0 5O 100 150 200
Temperature (K) Temperature (K)

(g) (h)
0.8 I I i - - T l r
24

0.6 20
-Bc =
16

vO.4

8 Ferromagnelic-
0.2
4 Bc=

0
I I I
2
200 400 600 0 5O 100 150 200
Temperature (K) Temperature (K)

Fig. 3.4. (Continued).


92 A. SCHENCK and EN. GYGAX

I I I I

g=
t
21 I
!
0 I ~s I I
0 1 2 3 4
Bext (ka)
Fig. 3.5. Local field [B~I as function of external field Bext for a single crystal Ni sphere. At the
saturation field Bs = (4rr/3)Ms the /*SR frequency starts to increase with Bext, indicating that B u
is positive, i.e. in the direction of the bulk magnetization (Denison et al. 1979).

95
AT= 30mK z~
90
O T= 70 mK
85 O T= 200 mK rh
T
Jr T = 300 mK
80 x T = 400 mK
N
32
m O T = 600 mK L~
75 R.~
v
>=
70
m A
65
60

55 I I I I I I I I I I I
0.6 0.8 1.0 1.2 1.4 1.6
Bext (T)
Fig. 3.6. /zSR frequency observed in longitudinal measurements on c~-Fe as function of the externally
applied field at different temperatures - Schimmele et al. (1994).

Figure 3.6 shows the #SR frequency ut, = Bu%/(2rr) (% = gyromagnetic ratio of
#+) obtained in the experiment for various values of Bext and temperature. The
/zSR signal is moderately relaxed, which indicates that at such low temperatures the
#+ hopping rate in Fe is slow compared to the observed #SR frequency, which is
about 60 MHz at its minimum. The measurement of uu as function of the applied
external magnetic field gives directly the dipolar magnetic field and Be, the latter
value amounts to (-1.13 + 0.02) T, in close agreement with the earlier measurements
by Denison et al. (1979) - see also table 3.4.
For Dy, where, coupled to the structural change at 85 K (from orthorhombic to
hcp), the magnetic order also changes (from ferromagnetic to a helical antiferromag-
MUON SPIN ROTATION SPECTROSCOPY 93

~ ~o

o~

+.~

I I I I
e~

I i I I
~ ~ H
.~ ~ . ~ ~
= g"Ng

. ~ ,' o

oo o o

,-~ o ~

rm

I I -t- -t- ~
-H

~=~

8
94 A. SCHENCK and EN. GYGAX

netic order), Bc can be determined along the same lines as described (Denison et
al. 1979), calculating B~i p in the appropriate way for the antiferromagnetic phase,
where of course Ms is zero. Over the full temperature range B~ip(T ) does not differ
significantly if one assumes T- or O-sites for the #+. The fact that B u is observed in
the antiferromagnetic phase of Dy can only be explained if the #+ is not diffusing
through the lattice. In the experiment it was not possible to determine the sign of
Bu, thus the two possibilities for Bc are indicated in fig. 3.4 h.
For the ferromagnetic elements a compilation of Ms, Bj, and Bc is presented
in table 3.4 for all data extrapolated to T - 0 K, basically following Denison et
al. (1979). In addition the table shows hyperfine fields B ° derived from neutron
diffraction results and values B the°r resulting from various theoretical calculations.
The fact that Bc is not just caused by the undisturbed interstitial spin density, as
measured by neutron scattering, is obvious - except for Ni. One common feature of
the Be values is the negative sign, meaning that the magnetic densities at the #+ are
opposite to the average magnetic density in the unit cell. This feature is explained by
the Daniel-Friedel model (Daniel and Friedel 1963). The temperature dependence
of both Bc(~'t,) for Fe, Co, Ni and Dy and the normalized domain magnetization Ms
for Fe and Ni are shown in fig. 3.4 e-h and for Gd in fig. 3.3.
Be is not strictly proportional to Ms. Nishida et al. (1978) defined the 'deviation'
function A(T):

Be(T) Ms(T)
- A(T) - - , (3.7)
Be(0) Ms(O)

which is shown for Fe, Co and Ni in fig. 3.7. For Fe, Co and Gd Bc is below the
normalized Ms for T # 0 K, whereas Bc is above Ms normalized in Ni. Schenck

0.10 I I I

Ni
0.05

'7,
t- 0.00

-0.05

-0.10 i i p
0.00 0.25 0.50 0.75 1.00
T/mc
Fig. 3.7. Temperature dependence of the reduced hyperfine field A ( T ) for Fe, Co and Ni:
A(T) = {Bc(T)/Bc(O)}/{Ms(T)/Ms(O)} - Nishida et al. (1978).
MUON SPIN ROTATIONSPECTROSCOPY 95

(1985) lists possible reasons for such a behaviour and reviews various theoretical
models dealing with that matter.
For rare earth elements, Campbell (1984) has pointed to a complication one may
encounter in the deduction of Bc from B , measurements. In such samples the
electric field gradient (EFG) created by the #+ charge will affect the orientation of
the neighboring lanthanide moments and so will lead to altered dipolar fields at the
#+ site. Estimates are presented for Tb, Dy, Ho, Er and Tm in the cited paper.
For Gd, which has no orbital moment contribution to the total moment, the effect is
negligible.

3.1.4. Pressure dependence of the hyperfine field


The possibility that the deviation between the temperature dependence of Bc and
Ms originates from the different manner in which thermal volume changes affect the
two quantities has been studied by measurements of the pressure dependence of Be
(Butz et al. 1980, Lindgren et al. 1987, Butz et al. 1987, Hartmann et al. 1990a,
and Kratzer et al. 1994a). Figure 3.8 show such measurements in Ni and Fe (Butz
et al. 1980). Schenck (1985) discusses some implications. The emerging picture is
not clear. Whereas for Ni half of the deviation between the temperature dependence
of Be and Ms is explained by the volume change, the effect has the opposite sign
for Fe. Other causes must be considered.

1.004

1.002

1.000 0 Ni
:=L
• Fe
t,n

v
Q. 0.998
m=

0.996

0.994

0.992
0 1 2 3 4 5 6 7
p (kbar)
Fig. 3.8. Pressure dependence of the local field B~ at 298 K for ferromagnetic Ni and Fe. B~
is normalized to unity at zero pressure (Butz et al. 1980).
96 A. SCHENCK and EN. GYGAX

3.2. Critical phenomena


In their pioneering study in Ni and Fe, Foy et al. (1973) noticed that #SR could be
used to examine critical behaviour near the magnetic critical temperature. Antifer-
romagnetic phase transitions were observed by Gurevich et al. (1976) in Dy and Ho
and then by the same group (Grebinnik et al. 1979a) for the rest of the magnetic
lanthanides. The experimental studies dealing specifically with critical phenomena
are listed in tables 3.2 and 3.3 with the indication that the #SR measurements were
performed near the critical temperatures Tc or TN or that the essential data (like the
#+ depolarization rate, F, the spin fluctuation, SF, the Knight shift, K , , etc.) were
obtained also near Tc or TN.
The first #SR studies aimed at extracting quantitative information from experi-
ments probing critical phenomena were performed by Barsov et al. (1983 and 1984)
around the N6el temperature in Er, and in parallel by Nishiyama et al. (1983 and
1984) around the Curie temperature in Ni.
In their study of Fe in the Tc region Herlach et al. (1986) stress some of the
difficulties one encounters trying to extract static and dynamical exponents using the
various #SR methods. This paper also reports K~(T) and transverse depolarization-
rate versus temperature data obtained in FeCo and FeZr alloys. The results permit
to follow the critical exponents deduced from the two temperature dependences as
function of the increasing concentration of the elements alloyed to Fe.

3.2.1. Critical behaviour of the muon Knight shift


In a first approximation we expect the measured muon Knight shift to follow the
magnetization (or susceptibility) of a paramagnetic sample. This would give a #SR
frequency shift (after correction for demagnetization field and Lorentz field) like

]¢Bext
Av~----
(T - Tc)

well above Tc, with possible deviation in the critical region.


This is effectively so in Ni, where Gygax et al. (1980) have measured Ku over a
wide temperature range above Tc. The strong T-dependence follows exactly that of
the bulk magnetic susceptibility X, also in the critical regime. Figure 3.9 a shows the
low temperature results, from which a critical exponent 3' ,-~ 1.28 is extracted, in good
agreement with the value of 3' -- 1.35±0.02 obtained for the susceptibility in the same
temperature domain. Figure 3.9 b - a Clogston-Jaccarino plot - shows that the high
temperature #+ Knight shift is also strictly proportional to X. Since the temperature
dependence of the total susceptibility is associated with the d-electrons' susceptibility
term, the slope dK~/dx can be identified with the induced contact hyperfine field at
the #+ per unpaired d-electron per atom, Ac,d (from K~,,d = Ac,d • x~t/#B , X~t is the
atomic susceptibility).
In this instance it is interesting to compare Ac,d with the induced hyperfine field per
d-electron per atom in the ferromagnetic state of Ni, which is AcV,d = Bc(f'~,)/Ms (see
Schenck 1985). The two values are -1.224(23), resp. -1.14(2) kG/(# B • atom), i.e.
MUON SPIN ROTATION SPECTROSCOPY 97

AT (K)
6 8 101215 2 0 2 5 3 0 4 0 5 0 6 0 80100
II I II I I II I II I I

-1000
/ -500 "N'l• ,AKp
AZ
= -0.223
(2~ mol
' "emu
X

y = 1 . 2 8 / X,
\
-1000
'\.
V\ / ~ = 1.35
Q.

- 1 O0 (fromz) 2 -1500
\
\,
\

(a) -2000
(b) '~",
-10 I I I -2500 I I I I
0.5 1.0 1.5 2 40 80 120 160 xlO 6
log (T-Tc) Z (emu/g)
Fig. 3.9. Muon Knight shift for paramagnetic Ni, plotted against log(T -/Pc) in the low temperature
region (a), and plotted against the bulk magnetic susceptibility with temperature as implicit parameter
in the high temperature region (b). The critical exponent deduced for K u (solid line in a) is compared
to the con'esponding value obtained from bulk susceptibility measurements (dashed line in a). The high
temperature region extends to 270 K above TC (Gygax et al. 1980).

different by about 8%. A step has also been observed by Nishiyama et al. (1983).
No explanation does exist so far - see Schenck (1985). However, the basic equality
of the induced hyperfine field in the paramagnetic phase and in the ferromagnetic
phase seems to be the rule - see, e.g., for Gd W~ickelg~rd et al. (1986). In this
same paper it is further reported that closer to T¢, say in the critical region below
1.05T¢, the muon frequency shift deviates from the simple Curie-Weiss behaviour.
Moreover, the frequency deviation observed indicates an exponent below 1.0 for
(T - T¢), whereas the susceptibility for polycrystalline Gd has an exponent around
1.24 in the same temperature range. This deviation is however trivial, it is expected
from the magnetization curves measured for Gd at the same applied field as in the
/ZSR experiment (Karlsson 1990).

3.2.2. Muon-spin relaxation or depolarization rate in 3d-elements


For Ni, Nishiyama et al. (1984) observed that above Tc the practically field indepen-
dent longitudinal/z-spin relaxation rate (A1 = l/T1) measured at low field displayed
loss of the critical behaviour when T approaches T¢ (fig. 3.10). It was ascribed to
the establishment of short range ordering around the /z+, tending to diminish the
dipolar field sum at the/z+ because of the cubic symmetry.
The temperature dependences of the #+ depolarization rate (as well as of the
muon Knight shift) have been studied in the giant moment alloys PdFe and PdNi,
98 A. SCHENCK and EN. GYGAX

Illlll t I I III1111 I I IIIIII I I 1 IIII

0.1

:=k

0.01

0.001
,,,,,,I , , ,,,,,,I , , ,,,H,I , , ,,t,

0.1 1 10
T-TC (K)
Fig. 3.10. Temperaturedependence of the/zSR damping rate for metallic Ni. The points are the data
of Nishiyamaet al. 1984. The solid line is the result of the model from Yaouancet al. (1993a), taking
the/z + dipolar interaction into account, whereas the dashed line gives the prediction when this latter
interaction is neglected.

which display ferromagnetic ordering, by Gygax et al. (1981). The measurements


were performed in the paramagnetic state for both alloys and also below Tc for
Pd0.97Fe0.03. The results are tentatively ascribed to the presence of slowly fluctuating
ferromagnetically coupled clusters.
One had to wait for Yaouanc et al. (1993a) who established a thorough scheme
to compute the paramagnetic critical zero-field muon-spin relaxation rate, at first
only for cubic ferromagnets (see section 2.3). This calculation includes the full
(long range) dipolar interaction between the muon and the lattice dipole moments
in a mode-coupling theory. It appears that the damping rate is determined by the
relative weight of the hyperfine interaction and the dipolar interaction between muon
moment and the lattice ions magnetic moments. The data from Nishiyama et al.
(1983) are well explained by this model (fig. 3.10), as well as the #+-spin relaxation
rates measured in paramagnetic Fe by Herlach et al. (1986).

3.2.3. Muon-spin relaxation in the lanthanides


Compared with 3d magnetism, the magnetism of the lanthanide 3+ ions, with a well
localized 4f shell, is characterized by much more pronounced effects of the orbital
contribution and strongly localized magnetic moments. This last fact has important
consequences for the positive muon occupying an interstitial site: (i) the hyperfine
field (contact field) is small due to the rather low spin density of the conduction elec-
trons, because of the indirect mediation of polarization from d-electrons, (ii) the local
dipolar field is high and often the dominating field since the neighboring lanthanide
moments are large and (iii) the symmetry of the dipolar coupling is low compared to
that of the hyperfine coupling, which is isotropic with respect to the crystalline axes.
The latter property means that the muon probes paramagnetic fluctuations along dif-
ferent axes with different weight. A judicious selection of the relative orientation of
MUON SPIN ROTATION SPECTROSCOPY 99

initial #+ polarization axis, crystal axes and applied magnetic field allows a selec-
tive weighting of the z-, y- and z-components of the spin correlation functions in
the investigated material (Karlsson 1990). Due to its large gyromagnetic ratio (in
comparison to nuclear hyperfine probes) the #+ can measure lanthanide relaxation
times as short a s 10 -13 s.
#SR measurements on lanthanide paramagnetism provide important information on
the collective behaviour of coupled localized spins, particularly at high temperature.
They are a valuable complement to other methods, mainly because of their sensitivity
to short-lived correlations, their general applicability - independent of the choice
of element - and their flexibility allowing to select among the different relaxation
parameters.

Gd. The #+ spin relaxation in Gd has been experimentally studied by W~ickelghrd


et al. (1986 and 1989) in the temperature region from Tc to 2Tc; additional infor-
mation and discussions are provided by Hartmann et al. (1990b), Karlsson (1990),
and Dalmas de Rrotier and Yaouanc (1994a). The major results are that magnetic
correlations extend to far above Tc, with a temperature evolution containing two
different terms, for e > 0.03 (e = ( T - T c ) / T c ) and for e < 0.03. For single crys-
tal measurements an anisotropy of the relaxation rate with respect to the crystalline
c-axis is observed, at least up to a few K above Tc (Karlsson 1990). This behaviour,
similar but weaker than in Er, is surprising for Gd which is a S-state ion. Let's
mention that anisotropy has also been observed in the susceptibility of Gd (Geldart
et al. 1989).
Dalmas de Rrotier and Yaouanc (1994a) have adapted their #SR depolarization-
rate model (Yaouanc et al. (1993a) - see section 3.2.2) to the special case of hcp Gd.
Although their theoretical results are only valid for truly zero magnetic field, they
assumed that the 10 mT transverse-field data obtained by W~ckelghrd et al. (1986)

i i ~L~II I i i i~11111 i ~ i l l ~ J I i

Ising Heisenberg
4 dipolar q isotropic

~ 1.o

0 average over single \&


crystal measurements
0.1
i t lliHI t i irll~ll i i l l r t H t i

1 10 100
T-Tc (K)
Fig. 3.11. Temperature dependence of the/~SR damping rate for Gd. The polycrystalline data are from
Wackelgftrd et al. (1986) and the single crystal data from Hartmann et al. (1990b). The full line is the
prediction of the dipolar Heisenberg model from Dalmas de Rrotier and Yaouanc (1994a); the different
temperature regimes are indicated (see text).
100 A. SCHENCK and EN. GYGAX

on a polycrystalline sample could be fitted by the theory. The result (fig. 3.11)
follows well the polycrystalline data, whereas the zero-field single-crystal data by
Hartmann et al. (1990b) are less-convincingly described. In this theory the magnetic
dipole-dipole interactions cause a crossover as the temperature is reduced in the
paramagnetic region from an isotropic Heisenberg regime to a dipolar Heisenberg
regime at about Tc + 15 K. As the temperature is reduced further, their is possibly a
second crossover temperature to Ising behaviour at roughly Tc +4 K. For a discussion
of the different features derived from #SR and susceptibility measurements as well
as the discrepancies between #SR data the reader is referred to the original work.

Er. The data obtained in partially oriented crystals by Barsov et al. (1983 and 1986)
indicated first the occurrence of an anisotropic paramagnetic relaxation rate. Later
single crystal zero-field data were presented and commented by Hartmann et al.
(1990c), Karlsson (1990) and W~ippling et al. (1993) for the temperature region
extending from TN to 2TN. In contrast to the weakly anisotropic Gd, Er shows an
extreme anisotropy of the relaxation rate which survives up to at least 2TN (fig. 3.12).
The rate ),± (#+ polarization perpendicular to the c-axis) shows the usual strong di-
vergence when approaching TN from above. This arises from the slowing down of
the 4f-electron spin dynamics due to the evolution of paramagnetic spin correlations
(Karlsson 1990). In contrast, All remains independent of temperature at the value
corresponding to the spin fluctuation in a free paramagnet, i.e. the high temperature
limit (fig. 3.12). Hence, solely the component of Er spin which undergoes mag-
netic ordering at TN develops also paramagnetic correlations and fluctuates with the
c-axis as quantization axis. (Below the N6el temperature of TN ~ 85 K first the
c-axis components of the Er magnetic moment orders in a sinusoidally modulated
incommensurate arrangement!) For the perpendicular orientation the initial signal
asymmetry drops sharply when decreasing T through TN and the signal is effectively
lost at lower temperature, whereas the parallel signal remains altogether unaffected
down to 58 K (Hartmann et al. 1990c). This confirms that magnetic order is totally
restricted to the Er spin component parallel to the e-axis and that not even short
range order develops for the perpendicular component.

6! rN' '
5-

2 - Ln I-1
I 0 ooooo¢
~ l og g 0 gi
0
0 100 200
Temperature (K)
Fig. 3.12. Muon spin depolarization rate in Er single crystal for two orientations: circles for the muon
polarization parallel to the c-axis and squares for the muon polarization perpendicular to the c-axis
(Hartmann et al. 1990c).
MUON SPIN ROTATIONSPECTROSCOPY 101

In their conclusion W~ippling et al. (1993) state that the long range nature of
the dipolar interaction between the #+ spin and the lattice spins has to be taken
into account to explain the measured relaxation rates. A correct evaluation of the
data would have to proceed along the lines proposed by Yaouanc et al. (1993a)
(section 3.2.2), but such calculations have so far not been applied to Er.

3.3. Ferromagnetic 3d-element based alloys


3.3.1. Fe, Co, Ni
Several authors have measured the variation of Bu with impurity concentration in the
elemental 3d-ferromagnets (see in table 3.2 the entries for Fe-, Co- and Ni-alloys).
The goal was to find how the deduced Bc changes with concentration in relation to
the corresponding changes in the bulk magnetization. However, it became clear that
in such experiments the #+ did not randomly sample the lattice sites because of the
attractive or repulsive interaction with the impurities. This showed up in the interplay
of trapping and diffusion of the #+ in the 'dirty' host lattice, as reported by Nishida
et al. (1979) and Kossler et al. (1979) in Fe- and Ni-alloys. Together with the studies
of Stronach et al. (1979, 1981, 1983), and of Herlach et al. (1986) in Fe- and Co-
alloys, a total of over 15 different types of dilute alloys were examined. Because
of the mentioned impurity dependent modification of the #+ behaviour these B~
dependences give only limited information on the intrinsic host magnetism. They can
however provide an interesting contribution in conjunction with the results of other
methods applied to study the magnetism in a specific alloy. Stronach et al. (1981,
1983) discuss the effect of the impurity induced strains upon Bu and Be. Herlach et
al. (1986) have studied several Fe-Co alloys as well as an amorphous FeZr alloy.
In particular the disturbing effects of sample inhomogeneities are discussed.
Solt et al. (1992) have performed first #SR measurements in neutron-irradiated
pure Fe and Fe-based copper containing alloys. This study was part of a general
investigation undertaken with various methods on the radiation generated fast pre-
cipitation from supersaturated metastable FeCu solutions. #SR has shown its great
sensitivity to both vacancy-like defects and to the Cu-depletion of the solvent ma-
trix. The average and distribution-width of Bu changes with n-irradiation fluence
as a consequence of the broad size distribution of created voids and other defects
in the pure metal and of the induced Cu-precipitates in the alloys. Annealing after
irradiation produces agglomeration of vacancies or defects, which again changes the
field distribution sampled by the muons. A modeling of the experimental results is
presented and the preliminary #SR study shows that due to the high sensitivity of
the technique samples with higher purity and less irradiation damage are required
for a better resolution of defect spectra.

3.3.2. Pd-based alloys


Pure Pd is paramagnetic, but close to a magnetic instability. Alloying it, e.g., with
Fe or Ni at atomic % concentration can strongly polarize the d holes on neighboring
Pd sites, producing 'giant moments'. These moments can show ferromagnetic order
or couple antiferromagnetically and exhibit spin-glass ordering, depending on the
alloy concentration and the temperature.
102 A. SCHENCKand EN. GYGAX

Different #SR studies were performed on Pd-based alloys (see the references in
table 3.2). The general meaning of the results is not at all clear. Again the peculiar
effects of the impurity atoms on the #+ are difficult to assess. Even in 'pure' Pd
the role of the muon is dubious. For example, Nagamine et al. (1977) interpret
their data assuming a localized #+ at low temperature (in analogy to the behaviour
of H in Pd), whereas Hartmann (1990) seems to believe that for very pure Pd the
#+ diffuses fast for temperatures at least as low as 18 K. For weakly doped Pd,
Dodds et al. (1983) assume a negligible #+ diffusion rate below 100 K (based on
the muon diffusion measured in Pd lightly doped with Gd) and interpret their data in
PdMn (2 at.%) assuming randomly distributed #+ locations, whereas the same group
consider the muon as static in PdFe0.0o35Mn0.05 at low temperature, but physically
excluded from the interstitial sites closest to the Mn sites (Gist et al. 1986).

3.4. Chromium and its alloys


The three first #SR studies in Cr (Kossler et al. 1977, Grebinnik et al. 1979b,
Weidinger et al. 1981) did not succeed in detecting spontaneous muon precession in
this antiferromagnetic metal. The measurements yielded #+ depolarization rates as
function of temperature or field and showed only a modest consistency. As Weidinger
et al. (1981) noticed, it seems that the influence of impurities on the #+ behaviour
partly biases the experimental outcomes. The question of the #+ localization was
addressed but not conclusively answered.
The Stuttgart group (Major et al. 1986, Templ et al. 1990) was able to see a #+
precession signal in a zero-field measurement in Cr below 12.5 K. The frequency of
about 84 MHz shows up with a small amplitude, implying that only a small fraction
of the muons see a distinct local field. Below 124 K Cr is in an incommensurate
longitudinal spin-density wave (SDW) state. The Stuttgart group explain their data
assuming that for T < 12.5 K the #+ stays at a tetrahedral interstitial site with
tetragonal symmetry axis perpendicular to the crystalline [100] direction (labeled T ±-
site). The other tetrahedral site and the octahedral sites are excluded and #+ hopping
has a low probability during the muon lifetime. In addition it is assumed that the
#+ resides preferentially near a field maximum of the SDW. This situation can result
either if the SDW maximum is pulled towards the #+ or if the muons thermalize
and stop at the SDW antinodes (in Cr the SDW or better the accompanying charge
density waves produce also strain waves - the effect of strain on muon trapping has
already been mentioned in sections 3.1.2 and 3.3.1).
In a further study (Grund et al. 1994) proton-irradiated Cr was used to vary the
possible pinning/trapping conditions for muons and SDW. For the low temperature
regime the data are now interpreted assuming a localization of the #+ at random
position in the host; the surviving coherent signal at long times corresponding to the
(small) fraction of the #+ stopping near the SDW maximum, as indicated by the
calculation. Therefore, the muons neither cause SDW motion nor have to undergo
long range diffusion.
In a study by Noakes et al. (1992) it is shown that the zero-field coherent #SR
frequency of 84 MHz previously observed in Cr has a dramatic impurity dependence.
MUON SPIN ROTATION SPECTROSCOPY 103

It is known that alloying Cr with small amounts of other metallic elements has
strong effects on the magnetism (Fawcett 1988). In their experiment Noakes et al.
(1992) were able to suppress, shift dramatically or split the 84 MHz #+-frequency
observed in pure Cr by alloying with, e.g., V, A1 or Mn at atomic % concentrations
and varying the temperature. So far only speculative explanations are suggested.
Another contribution dealing with the condensation of incommensurate SDW and
the onset of the AF state stems from a ZF #SR study in the Cr-like alloy Cr85Mo15
by Telling et al. (1994). As in other instances (see, e.g., section 3.3.2) the effect
of impurities on the #+ behaviour in the Cr host is possibly pathological, i.e. very
interesting for a defect therapist, but rather an annoying complication for the simple
magnetically-minded pedestrian.

3.5. #-SR in Fe and Ni


A # - stopped in a solid forms a muonic atom # - z X in which the bound # - behaves
like a heavy electron (mu = 207me). The muonic radius % = aome/(muZ) is much
less than the Bohr radius a 0, hence the pseudonucleus # - z X appears to the atomic
electrons very similar to an isolated impurity nucleus of charge Z - 1 (Yamazaki et al.
1979 and Yamazaki 1981). However, for light elements (Z < 30) the muonic atom
1s-wave function extends noticeably outside the nucleus. For example, in the case of
/z-Fe, % is about twice the nuclear radius Ro. In contrast to the previously discussed
#+SR measurements, which yield information on the hyperfine field (i.e. the electron
spin density) at the interstitial #+ position,/~-SR leads to inforrnations on the radial
distribution of the electron spin density p(r) near the (pseudo-impurity) nucleus. This
is much like NMR, M6ssbauer spectroscopy and time-differential perturbed angular
correlation, all techniques studying the hyperfine fields acting upon the nuclei of
adequately selected impurity atoms in the host. Calling B~f the average hyperfine
field acting upon the # - bound to z X and B hf the average hyperfine field acting upon
the impurity nucleus of charge Z - 1, one usually defines the hyperfine anomaly as

AX = ( ~ h f _ BhNf)/Bhf. (3.8)

Assuming that the muonic atom and the impurity atom occupy the same type of site
in the lattice and have the same electronic structure, a non-zero A value will be
due to the fact that the # - samples p(r) over a more extended region of space than
the Z - 1 nucleus. Yamazaki (1981) has discussed some aspects of the hyperfine
anomaly and their importance in the study of magnetic materials. Imazato et al.
(1984) reported the first results in a ferromagnetic material, i.e. Ni, and Keller et al.
(1987) extended the measurements to ferromagnetic iron.
In Ni a comparison of the hyperfine field for #-Ni with that for 59Co yields
nNi = -2.4(3)% to -2.8(5)% over the temperature range 23-303 K (Imazato et
al. 1984). The results indicate that the electron spin density near the Ni nucleus
decreases with increasing r more steeply than the s-state electron charge density,
a behavior which is a consequence of core polarization (see Freeman and Watson
1965, Freeman et al. 1984). The values are in good agreement with the unrestricted
Dirac-Fock calculation by Freeman et al. (1984).
104 A. SCHENCK and EN. GYGAX

25 I I I I

1.0
2O
E
0.8
"o 15
-~ 0.6 &
E
o 10 "~
"a 0.4 :~
"0

5SMn rr
5 o/, NMR 0.2
• p Fe

0 I r 1 i 0.0
0 0.2 0.4 0.6 0.8 1.0
T/T c

Fig. 3.13. Magnitude of local field in Fe as a function of reduced temperature T/TC (TC = 1043 K).
Solid triangle: data for 52MnFe (dilute limit) obtained with NMR on oriented nuclei from Hagn et al.
(1982). Open circles and open triangles: 55Mn data for (1.5 at.%) 55MnFe from Koi et al. (1964)
and Yamagata et al. (1983), respectively. The solid circles are the/~-Fe data for iron and the solid
line represents a fit using a simple mean-field model (Keller et al. 1987). For comparison the reduced
magnetization curve of pure iron is also shown (right scale).

In Fe # - S R in zero applied field was used to study the hyperfine field in the
temperature range 320-690 K (Keller et al. 1987). Bhuf departs from the magne-
tization curve of pure iron in the same way as the hyperfine field acting upon a
55Mn impurity in diluted (1.5 at.%) MnFe measured by N M R by several groups
with different N M R techniques (Koi et al. 1964, Rubinstein et al. 1966, Yamagata
et al. 1983), see fig. 3.13. This indicates that the electronic structure of # - F e is very
similar to that of a Mn impurity in Fe. AFe is found to be - 0 . 9 ( 3 ) % and temperature
independent over the temperature range investigated. The anomalous behavior of the
impurity B~f has been discussed by Jaccarino et al. (1964), Low (1966) and Shirley
et al. (1968), and can be understood on the basis of a mean-field approach. The
A value measured by Keller et al. (1987) is significantly smaller than expected if
core polarization plays a similar role in # - F e relative to Mn as in # - N i relative to
Co. The reason for the small AFe is not known. In this case, in contrast to Ni, no
unrestricted Dirac-Fock calculation are available.

4. Review of results in intermetallic compounds

4.1. Compounds involving transition elements


Only a few intermetallic compounds with no lanthanides or actinide constituents have
been studied up to now, by #SR. In this section we discuss the weak itinerant magnets
MnSi, Y9Co7 and Nbl_=Zr=Fe2 (see table 4.1). Other Y containing compounds are
included in section 4.2.
M U O N SPIN ROTATION SPECTROSCOPY 105

oo ~ oo o~

N~v

e~

© 0

--1
r~ I.~ gT., .C~

©
M
c.i
~8
cq

o~<
o~

g~ 8

0
r..)
106 A. SCHENCK and EN. GYGAX

4.1.1. MnSi
The interest in this system was motivated by the possibility to extend spin lattice
relaxation rates (Ta l) measured by 55Mn NMR (Yasuoka et al. 1978a) to much
lower temperatures and higher rates by using the #+. This extension appeared very
important since detailed predictions on the temperature dependence of T1-1 on the
basis the self consistent renormalization (SCR) theory for itinerant ferromagnetic
electrons (Moriya et al. 1973, 1974, Moriya 1977) and applied to MnSi (Makoshi
et al. 1978), could only sensibly be checked in this way. In fact MnSi is known to
be a weak helimagnet with a period of ,-~ 180 A below Tc = 29.5 K (Ishikawa et
al. 1976, Motoya et al. 1978) but since the helical structure has such a long period
it was expected that T~-1 behaves almost like the spin lattice relaxation rate in an
itinerant-electron ferromagnet (Hayano et al. 1978a).
We start with a review of the measurements in the paramagnetic phase (T > Tc).
In a first study the #+-Knight shift was measured in a transverse field experiment
with Hext = 0.29 T (Yasuoka et al. 1978b). It was found to scale well with the
magnetic susceptibility from which the coupling parameter Ac~+ could be determined
(see table 4.2 which includes also the corresponding A Mn determined by NMR).
According the SCR theory above Tc the spin lattice relaxation rate T1-1 (see also
eq. (2.41)) is predicted to be given by

Im X L(g wo)
2 ,u,,Mn 2 (4.1)
1/T1 (., Mn) = 271,(Mn) (Ac ) T Z coo

1 T
= TI(~) T - Tc (4.2)

,'~^,2 ( d,u,Mnh2 T (4.3)


C(~I/~(Mn) L~c ) T - Tc '

and hence
2 Mn 2 2
T(U)/T(Mn) =- "yfvln(Ac ) /'yu(Ac/ ,)2 . (4.4)

First results on (Ta~)-~ obtained in a longitudinal field of 700 G agreed with


the SCR-predictions appropriate for an itinerant ferromagnetic system quite well.
TABLE 4.2
Compilation of hyperfine fields and coupling parameters in MnSi
(Tc = 29.5 K).

Acu (kG/#B) B~f (T = 0) (W) A Mn (kG//~B)


T > Tc -4.8 ± 0.2 - -138(1)
T < TC -6.94 -0.273
-3.94 -0.155
MUON SPIN ROTATION SPECTROSCOPY 107

Comparison with a few NMR results above 100 K proved also the validity of the
scaling relation eq. (4.4) (Hayano et al. 1978a, 1980).
Later it was realized that ZF-#SR measurements could not only be used to de-
termine TI-t(# +) but also Tll(Mn) at the same time. This is a consequence of the
fact that in ZF the #+ will also feel the 55Mn-nuclear dipole fields. If they are static
the effect on the #+-polarization is described by the static Kubo-Toyabe relaxation
function (see section 1.8). However, since the nuclear spins relax as well under the
action of the spin fluctuations of the itinerant electrons a dynamic picture applies
and the induced/z+-depolarization has to be described by a dynamic Kubo-Toyabe
function GKT(t, u(Mn)), where u(Mn) = Ta-a(Mn) is the nuclear spin lattice relax-
ation rate. In this case the overall #+-depolarization function is then given by (two
channel relaxation)

CzF(t) = exp (-- t/T} ")) GKT(t, u(Mn)). (4.5)

Analyzing zero field data with eq. (4.5) both 1/T~u) and 1/T~ Mn) could be deter-
mined over a much wider temperature range (Matsuzaki et al. 1987, Kadono et al.
1990). The results are compiled in fig. 4.1. The solid line through the 1/T1 (Mn)-data
is a fit of the function
1 T T
- - - a - - + b . (4.6)
T1 (Mn) T - Tc (T - Tc) 2

The second term is included to account for a somewhat modified temperature depen-
dence in the vicinity of Tc following from numerical solutions of the SCR-theory

I I~ IIItl I I I I I IIII I I I i i iiii I i

101
_oc_ :,zedm_o_ ent.
10 0
"1
v

~.,._
'T'--
10-1
=2 I1~:
",it--
10 -2
~/j,, o II
"+2,<, II
,,,,,,, , ,,',',J,I ,
10-3
10-1 10 0 101 10 2
T-T c (K)
Fig. 4.1. Log-log plot of both the #+ (open symbols) and the Mn-nuclear spin (filled symbols) lattice
relaxation rates l/T1 versus temperature. The filled circles and diamonds are extracted from ZF-/zSR
measurements, the filled squares are from NMR measurements. The solid lines follow eq. (4.6), the
dashed lines represent various other theoretical models (Hayano et al. 1978a, Kadono et al. 1993).
108 A. SCHENCKand EN. GYGAX

(Moriya et al. 1973). The solid line through the 1/Ta(#)-data is obtained by just
scaling the upper solid curve with the factor (UMnAcMn/l/IzAc#)2 = 5.0.
Another nice verification of the SCR-theory follows from #SR measurements of
the temperature dependence of the electrical field gradient (EFG) at the #+ nearest
neighbour Mn-nuclei (Kadono et al. 1993). Applying the muon-quadrupolar level
crossing technique (Kreitzman et al. 1986) the EFG showed a diverging behaviour
as Tc is approached from above. This is shown in fig. 4.2 where the temperature
dependence of the quadrupolar splitting frequency UQ of the 55Mn-nucleus (I = 5/2)
is displayed. Also shown are results from NMR-measurements performed at higher
temperatures (Yasuoka et al. 1978a). The two data sets do not agree too well (and
do not scale either) in the temperature range where they overlap. The reason for this
is not known at present.
In the SCR-model the temperature dependence is explained by the coupling of
charge density to spin density fluctuations (Takahashi and Moriya 1978). In this
model the net EFG at the Mn-nucleus is given by

q/
q=qo+qel = qo+ ( 1 / x ) + d , (4.7)

where q0 is the contribution from the positive ion cores including the #+ and qel
arises from the deviation of the conduction electron charge density from a cubic
distribution. The parameter q~ and d are related to the EFG in the absence of
electron-electron interactions and of the effect of exchange enhancement. Inserting
the Curie-Weiss law for the susceptibility the temperature dependence of UQ can be

1.5 i j i t

1.25 MnSi

1
N

-1- .75

o .5
°Ooo O O
0000
.25

I I I I I
0
0 100 200 .300 400 500 600
Temperoture (K)
Fig. 4.2. Temperaturedependence of the quadrupole coupling parameter //Q in MnSi. Triangles are
from the #+ level crossing measurements, octagons stem from NMR measurements (from Kadono
et al. 1993).
MUON SPIN ROTATION SPECTROSCOPY 109

expressed as

1 3e2Q [ q~ ]
. (4.8)
UQ(T) -~ 27r 2 M ( 2 I - 1) qo + (T - T c ) / c + d

This equation can indeed be used to fit the #SR data well (Kadono et al. 1993).
Using c = 0.6 (emu/mol)K and Q = 0.55 barn one obtains q0 = 6.75(3) x 102z cm -3,
q' = 1.30(1) x 1025 c m - 3 e m u - l m o l and d = 1.68(1) x 102 mol/emu.
In summary the SCR-model can describe very well the temperature dependence
of both the 55Mn and/z+-spin lattice relaxation rate and the EFG at the Mn-nuclei
in the paramagnetic phase.
We now review results obtained below Tc (Takigawa et al. 1980, Kadono et al.
1990). ZF-measurements reveal two distinct precession frequencies (with relative
populations Pl/P2 = 0.77(9)) the temperature dependence of which is shown in
fig. 4.3 (Kadono et al. 1990, Takigawa et al. 1980). The data are compared with

0.25 [ I I I I I

0.20 -

0.15

0.10 - --. \iO

" IP

0.05
qb A . f = "6.94kg/~ B \~
I Ahf = -3.94kg/l~B

0.00 I I I I 1I
0 5 10 15 20 25 30 35
Temperature (K)
Fig. 4.3. Temperature dependence of local fields Btz in MnSi below T C. The solid lines represent
theoretical calculations of MQ(T)/MQ(O)and the dashed lines show MQ(T)/MQ(O)obtained from neutron
scattering (from Kadono et al. 1990).
110 A. SCHENCK and EN. GYGAX

the temperature dependence of the saturation magnetization MQ(T) (determined by


neutron scattering, Ishikawa et al. 1976) and a perfect scaling is observed implying
(after correcting for the Lorentz field)

Bhf(T) = A2 + MQ(T). (4.9)


The resulting coupling constants A2 are to be found in table 4.1. The weighted
average

Ac~ = pl Acu(1) + p2Acu(2) = -5.63 kOe//z B (4.10)


is 17% larger than the coupling constant in the paramagnetic state (see table 4.2). It
is suggested that this discrepancy may be due to the ordered helical structure causing
two magnetically inequivalent sites and also dipolar field contributions. Note that in
the paramagnetic state only one signal showed up in the TF-data. Possible #+-sites
in MnSi were not discussed.

400 /
¢) Ahf = -6.9kg/g B / ~)
0 Anf =-3.9kg/g g
/
300

o
o~
200

e4 =t.

100

0 5 10 15 20 25 30 35
Temperature (K)
Fig. 4.4. Temperature dependence of/z + spin lattice relaxation rates, multiplied with the respective u 2
from the oscillating components below T C in MnSi. The SCR theory predicts a linear dependence (from
Kadono et al. 1990).
MUON SPIN ROTATIONSPECTROSCOPY 111

The solid lines in fig. 4.3 represent the predictions of the SCR-theory for Me(T)
below To As can be seen the data do not follow very well the predicted Me(T ).
This is discussed further in Takigawa et al. (1980).
Information on the spin lattice relaxation rate below Tc is obtained from the
exponential decay of a third nonprecessing component in the #SR-signal. This
component, actually consisting of two contributions, arises from those #+ whose
spins happen to be parallel to the static internal fields in the randomly oriented
multi-domain sample but which spins are affected by fluctuating field components
perpendicular to their direction (see eq. 2.13). In the analysis of these non-precessing
components it is assumed that 1/T( 0 o( w2%, where wi are the frequencies of the
corresponding precessing components and % is a common correlation time arising
from the Mn-spin fluctuations. Figure 4.4 shows the temperature dependence of
(wi/21r)2/T( O. According to the SCR-theory one expects

1/T1 c( T / [ M o ( T ) ] 2 (4.11)

and hence w~i)/T1 oc T. This linear dependence is indeed seen for temperatures
T < T c / 2 but it breaks down for larger T. Takigawa et al. (1980) suggests that the
deviation of 1/T(O(T) from eq. (4.15) may be a consequence of the helical structure:
"Since it has a finite Q, although very small, the spin fluctuations in a small region
around Q may play an important role in the relaxation process, at least near Tc ". In
this case an antiferromagnetic SCR-model would be more appropriate which predicts
1/TI o( T / M e ( T ). At least qualitatively this can describe the trend of the data closer
to Tc.

4.1.2. Y9Co7
This compound shows a magnetic phase transition at Tc ~ 6 K and becomes su-
perconducting below Tc ~- 2.7 K in coexistence with the magnetic order (Sarkissian
1982). The ordered magnetic phase below 6 K is believed to be an itinerant ferro-
magnetic state (Sarkissian 1982, Huang et al. 1983a, b). In the latter reference it is
reported that application of pressure suppresses the magnetic order and increases the
superconducting transition temperature and Hc2. It is further concluded that the long
range ferromagnetic- and superconducting order parameter coexist but vary spatially.
NMR measurements (Takigawa et al. 1983b) suggest that only at the Co-sites in the
2b positions (for notation see Grover et al. 1982), forming a linear chain along the
c-axis, a sizable spin density is to be found associated with a ferromagnetic moment
of 0.2#B per atom.
ZF-pSR measurements were performed with the aim to collect further informa-
tion on the peculiar magnetic behaviour of this compound (Ansaldo et al. 1985).
Figure 4.5 displays ZF-spectra for temperatures around and below To At tempera-
tures above Tc a well developed static Kubo-Toyabe signal was visible which arises
from the 59Co-nuclear dipole fields and proves the absence of #+ diffusion up to
at least 220 K. The results shown in fig. 4.5 could only be adequately fitted by
invoking two components, one of which is the afore mentioned Kubo-Toyabe signal
112 A. SCHENCK and F.N. GYGAX

.30 ~ 6 . 1 K , Co 7

.20

O3
< .10

0 - -

0 10 20 30 40 50 60 70 80
Time (#s)
Fig. 4.5. ZF-/~SR signal in Y9Co7 at various temperatures. The solid lines represent fits with two
components. See text (from Ansaldo et al. 1985).

.30 Y9 0 ° 7
• O0 ~
Slow Signal

~>, .2o
E a) °~c__~ ©

< .10 O Fast Signal


go
o •
0 I 1 I I I L
0 1.0 2.0 3.0 4.0 5.0 6.0
Temperature (K)
4.0 ~)

~ 3.0 I

'~ b) O
~-'- 2.0 Fast Signal
.o_
slow
_~
¢ 1.0 • • Signal

0 q O OC~DO- i
0 1.0 2.0 3.0 4.0 5.0
Temperature (K)
Fig. 4.6. Temperature dependence of (a) signal amplitudes (asymmetries) and (b) relaxation rates from
a two component fit to the ZF-#SR signal in Y9Co7 (from Ansaldo et al. 1985).
MUON SPIN ROTATIONSPECTROSCOPY 113

(below Tc to be replaced by a dynamic Kubo-Toyabe function) which shrinks with


decreasing temperature while a second more rapidly relaxing signal grows by the
same proportion. The data could be equally well described by choosing for the sec-
ond 'fast' signal a heavily damped oscillating function or alternatively a dynamical
Kubo-Toyabe expression with a much more rapid overall relaxation as compared to
the first signal. Using the first approach one finds at 2 K an average field of ,,o 100 G
at the #+. The fit results from the second approach are shown in fig. 4.6. Figure 4.6a
displays impressively the shrinking of the 'paramagnetic' component and the rise of
the new 'fast' component below Tc. Qualitatively the same result is found in the
first approach. Figure 4.6b displays the fluctuation rates following from the dynamic
Kubo-Toyabe fit of the 'fast' signal. The static line width for this signal amounts to
A ~ 4.5 #s -1 corresponding to a field width of ~,, 53 G which must arise from the
ordered Co-moments. These results are quite different from the ones in MnSi and
do also not show any resemblance with what is observed in spin glasses. Similar
behaviour, however, is found in some of the compounds classified as heavy fermion
materials (see sections 4.2.6 and 4.3.4). The most remarkable feature of the present
data is the temperature dependence of the two component structure. It could imply
that the paramagnetic phase persists well into the ordered state in part of the sample
volume. It would be interesting to find out whether superconductivity is associated
with only this fraction of the total volume. TF-#SR measurements on monocrys-
talline samples should allow to check on this conjecture (see also the discussion on
CeCu2Si2, section 4.2.6). The other noteworthy result is the lack of any indication
for a truly long range magnetic order. The present data indicate a rather random or
extremely short range magnetic order. In contrast to ordinary spin glasses no slowing
down of spin-fluctuations is seen when approaching Tc from above.

4.1.3. Nbl_~ZrxFe2
NbFe2 is considered to be a weak itinerant antiferromagnet (TN = 13 K) with a
rather small ordered moment (~ 0.1#B) (Yamada et al. 1990). The antiferromag-
netic state is very sensitive to the exact stoichiometry and weak ferromagnetism is
observed in compounds with either enhanced Fe or Nb content. The mixing of dy-
namic ferromagnetic and antiferromagnetic correlations in NbFe2 may be the result
of topological frustration associated with the Kagome lattice like structure of the Fe
sites in NbFe2. The development of itinerant ferromagnetism is also observed in
compounds in which Nb is partially replaced by Zr up to a concentration of 20%.
At larger Zr-concentrations the system converts back to weak antiferromagnetism
(Yamada et al. 1984)
#SR-measurements were started with the aim to study the phase diagram of
Nbl_xZr~Fe2 in more detail. So far only preliminary results on a powdered sample
of Nb0.9Zro.lFe2 (Tc -~ 43 K) were presented (Crook et al. 1994). These results
from ZF- and LF-measurements seem to imply that already below 90 K (the temper-
ature at which the susceptibility X, on cooling, tends towards the Curie-Weiss form)
some static behaviour of the atomic moments develops in coexistence with dynami-
cal fluctuations. It is also found that the static field distribution shows a trend from
a Gaussian shape to a more Lorentzian one as the temperature is lowered down from
114 A. SCHENCKand EN. GYGAX

90 K. This could suggest the development of dilute inhomogeneities in an otherwise


regular spin arrangement, perhaps as a consequence of the topologically frustrated
lattice of Nbl_xZrzFe2. Below Tc the amplitude of the ZF-#SR signal drops rapidly
to 1/3 of its initial value, marking clearly the onset of the ferromagnetic state. From
the relaxation of the residual signal it can be concluded that also in the ordered
state some spin fluctuations must persist. No oscillating signal below Tc was seen
which could be due to the limited time resolution at the pulsed #+-source at ISIS.
These preliminary results together with magnetization data suggest that Nb0.9Zr0.1Fe2
cannot be viewed as a simple weak itinerant ferromagnet.

4.2. Intermetallic compounds containing rare earth elements


4.2.1. Rare earth aluminides: RAI2 (except R = Ce)
The rare earth aluminides with the exception of CeA12 all show a ferromagnetic
state with Curie temperatures ranging from 6 K for TmAI2 to 160 K for GdA12 (see
table 4.3). Their crystalline structure is of the cubic Laves phase type MgCu2 (C15).
The magnetism is believed to be associated with well localized 4f-moments. Whereas
it is known that in certain itinerant magnetic systems short range correlations among
the spins persist to temperatures well above the transition temperatures the situation
with respect to local moment systems is less clear. #SR-measurements were per-
formed with the specific aim to search for such correlations by studying the spin dy-
namics of the 4f-moments via #+-relaxation measurements in the paramagnetic phase
(Chappert et al. 1981, 1986b, Kalvius et al. 1984, Gradwohl et al. 1986, Hartmann et
al. 1984, 1986, Dalmas de R6otier et al. 1990b). The results on CeA12 will be treated
in section 4.2.7 (heavy fermion compounds). From TF-measurements in the non-
magnetic compound LaAI2 it was concluded that the #+ most likely resides at the so
called (2-2) site, having two R and two Al-atoms as nearest neighbours (see fig. 4.7).
ZF- and LF-measurements in DyAI2 below Tc showed a drop in the signal ampli-
tude to 1/3 of its value above Tc which is typical for a ferromagnet with randomly
oriented domain magnetizations (Gradwohl et al. 1986, Asch et al. 1986). The
remaining signal displays an exponential relaxation indicating that fluctuating field
components are still present below Tc. Detailed measurements in the other RA12
compounds below Tc are lacking.
Above Tc mostly TF-measurements were done, supplemented later by ZF- and
LF-studies (Kalvius et al. 1984, Gradwohl et al. 1986, Chappert et al. 1986b,
Dalmas de Rfotier et al. 1990b). From an analysis of the Knight shift from TF-
measurements it was established that the contact hyperfine field is vanishingly small
with the exception of GdA12, and that dipolar fields from the 4f-moments5 must be
responsible for the dynamically induced ~+-depolarization (Hartmann et al. 1986).
With respect to the extraction of the #+ spin lattice relaxation rate from the TF-
spectra a correction for powder broadening had first to be applied. Some room
temperature relaxation rates A2, AzF of the #+-polarization are collected in table 4.4.
It is seen that grosso modo the corrected )~2 and the )~zF agree reasonably well with
each other. Az,zF can be compared with predictions based on the RKKY interaction

5 Dipolarcouplingparameters for all RA12are quoted in Hartmannet al. (1986).


MUON SPIN ROTATION SPECTROSCOPY 115

r~

mm ~

~o ~
116 A. SCHENCK and EN. GYGAX

001 projection
,,~ ®

® ®

® Rare Earth
o A~iuminum
2-2 muon position
,, 3-I muon position
Fig. 4.7. Crystal structure of the cubic laves phase compounds RA12. The most probable # + location
is the so called (2-2) site.

TABLE 4.4
Compilation of contact hyperfine field constants Ac, room temperature relaxation rates A2 (from TF
measurements, extrapolated to Hexp = 0) and AZF (from ZF-measurements), calculated values Aff based
on the RKKY mechanism and assuming the/z + at the (2-2) site, calculated values Aj at 300 K including
the Korringa mechanism, and calculated total fluctuation times ~-j (ct = 1.477 × 1011 s - 1 K - 1 , eq. (2.52))
(Hartmann et al. 1986).

[Ac I >'2 (300 K) AZF(300 K) ),ff ~j (300) ~'j (300 K)


(kG//~B) (#s -1 ) (/zs -1 ) (/zs -1 ) (/zs -1 ) (10-12s)
PrA12 -~ 0 0.05 (3) 0.133 0.09 0.39
NdA12 ~- 0 0.02 (3) 0.059 0.04 0.19
GdAI2 ~_ 0.46 0.38 (6) 0.22 (6), 0.23 (1) 0.11 0.11 0.089
DyA12 -~ 0 1.05 (7) 1.4, 0.67 (8) 0.96 0.71 0.34
HoA12 "~ 0 1.16 (8) ~ 1.05 2.14 1.17 0.57
ErAlz "~ 0 0.84 (5) ~ 1.4 3.76 1.36 0.81
TmA12 "~ 0 0.43 (4) 4.42 1.02 0.98

among the 4f-moments and on the Korringa mechanism (see eq. (2.52)), assuming
the #+ to reside at the (2-2) site. The exchange interaction constant Isf, entering into
the RKKY interaction, is calculated from the Curie temperature T¢. The Korringa
mechanism involves the quantity [IsfN(EF) I which, according to L6wenhaupt et al.
1983, is of the order of 0.08(4) and includes de Gennes' factor. For PrA12 and NdA12,
A2 are rather small and nearly temperature independent. The data are consistent
with a temperature independent 4f-moment fluctuation rate of the order of several
1012 s -1 (see table 4.4) which is caused by the RKKY interaction alone. In the
heavier HoA12 AzF is much larger but shows also a relatively weak temperature
MUON SPIN ROTATION SPECTROSCOPY 117
II I I I i I I I [1111111111 I

0.20 o

-,$ 0.15 0
0
E 0
E 0.10
c,O 0
< 0.05 (30
0.00

HoAI 2
2.0
b,
1.5 0' /%oooo Ooo o
¢,.
1.0 oo o
o
0.5

rr 0.0 I Tc

III I t I I I Illlflllt~l
20 100 200
Temperature (K)
Fig. 4.8. Plot of signal amplitude (asymmetry)and relaxation rate of the ZF-/zSR signal in HoAI2 versus
the logarithm of temperature. Note the collapse of the amplitude below Tc (from Dalmas de R6otier
et al. 1990b).

dependence (see fig. 4.8). The RKKY prediction amounts to an even larger rate.
Inclusion of the Korringa mechanism brings the value of the calculated )`ZF,calc.
down to nearly the measured one at room temperature and reproduces roughly the
temperature dependence. The same seems to be true for TmA12.
The remaining compounds GdAI2, DyA12 and ErAI2 in contrast show a very pro-
nounced variation of ),2 and/or )`ZF with temperature (see fig. 4.9) displaying an
almost divergent behaviour as Tc is approached from above. Hartmann et al. 1986
take this as evidence for the evolution of short range dynamic correlations as Tc
is approached. Dalmas de Rdotier et al. 1990b argue that some of the tempera-
ture dependencies are strongly influenced by crystal field (CEF) effects. (Not, of
course, in GdAI2.) Data on a single crystal DyAI2 show an even more peculiar
behaviour. There a maximum in )`1 shows up at T ~ 100 K, clearly above Tc,
f o l l o w e d by a shoulder at ~ 125 K. Attempts were made to explain this behaviour
i n terms of #+ trapping and diffusion (Gradwohl et al. 1986). Asch et al. 1986
h a v e also found differences when investigating differently prepared polycrystalline
DyA12-samples, but all samples showed a clear temperature dependence of )`1. Fi-
nally Kaplan et al. (unpublished data) investigated another polycrystalline sample
118 A. SCHENCK and EN. GYGAX

l f I I I I I I I I llllllll~ I

I0
i
o~
v
-, 8
+ ErAI 2

6
t-
O 4
O
x
~ 2 o o Oooom
rr 0

I I I i I I I I [ I I I I I [ I I I I

20 1O0 200
Temperature (K)
Fig. 4.9. Plot of the relaxation rate of the ZF-/zSR signal in ERA12 versus the logarithm of temperature.
Note the diverging character of ~ZF when the temperature approaches TC. The signal is lost already
about 10 K above TC (from Dalmas de R6otier et al. 1990b).

which displayed a near temperature independent )~1 ,-,o 1 #s -1. From all those re-
sults it must be concluded that sample dependent features are involved and the
question whether dynamical short range correlations persist above Tc in some of
the RAlz-compounds or whether CEF-effects are responsible cannot be answered
with certainty. Systematic investigations in the LF-mode may remedy our present
ignorance.

4.2.2. Other cubic Laves phase (C15) compounds RM2 with M a 3d transition
metal
4.2.2.1. RM2 (M = Ni, Fe, Co). All compounds with M = Ni, Fe, Co, considered
here (see table 4.5), show ferromagnetic order within the local moment rare earth
sublattice and the more itinerant 3d-element sublattice. The coupling of the two
sublattices is ferromagnetic for light R elements (Ce-Sm, no such compounds have
yet been investigated by #SR) and antiferromagnetic for the heavier ones. Included
here is also the compound YFe2 in which, as well as in LuFe2, only the Fe-moments,
of course, can order.
Below Tc, in zero external field, well developed #+-precession signals were ob-
served in GdFe2, GdCo2, YFe2, LuFe2, TmFe2 and ErFe2, using polycrystalline sam-
ples (Barth et al. 1986a, Graf et al. 1981). In each case only one distinct precession
frequency was observed. The temperature dependence of these frequencies is in part
rather peculiar (see, e.g., fig. 4.10-4.13) and does not scale with the macroscopic
magnetization. The analysis is made difficult by the presence of #+ diffusion and
probably trapping at defect sites and the presence of two sublattices. #+-diffusion,
already quite rapid at the lowest temperatures (4.2 K), has to be invoked in order to
explain the occurance of just one frequency.
M U O N SPIN ROTATION S P E C T R O S C O P Y 119

v~

0 0

::i
>-,
,..ID

"ID

~
,,,el ,-' >,-,
©

0 0 0 0 0

[.. [.-.

8 O

o
tr~
~ II

==
S.~ -=

~ o~

•. ~ o ~

eq t.-q

s
120 A. SCHENCK and F.N. GYGAX

B. (T)
0.9
120
M(T)/M(O) ~ , 0.8

100
0.7
"1-

80 0.6
o
e,,,,

0.5
13"

L.I- 60
I 0.4
+
-~GdFe2--
B~xt= 0 T
40 0.3

0.2
20
0.1

0 e I = I = I i
0 200 400 600 800
Temperature(K)
Fig. 4.10. Temperature dependence of the spontaneous field/3~ from ZF-/~SR measurements in GdFe 2.
The solid lines show the temperature dependence of the saturation magnetizations (from Barth et al.
1986a).

Adopting again the (2-2) site as the site of #+ residence (this site is also found
to be occupied by hydrogen or deuterium at low concentrations (Didisheim et al.
1980)) one finds that there are 96 interstices of this type in the crystalline unit cell
of the C15-structure (see fig. 4.7) which are not all magnetically equivalent. This
leads to the appearance of different dipole fields at these sites which should have
lead to a well resolvable splitting of the #SR-signal into several distinct frequency
components. Averaging over all (2-2)-sites results, of course, in a cancelation of the
dipolar fields in accord with the overall cubic symmetry. The absence of any splitting
suggests that rapid #+-diffusion provides such an effective averaging mechanism.
The remaining static field at the #+ is then simply given by (see eq. (3.5))

471
B u = B c + -~- Ms, (4.12)
MUON SPIN ROTATION SPECTROSCOPY 121

, i i l B~ (T)

• q, - 0.8
100 • Gd 002
T - 0.7
80 - 0.6
-r 0.5
60
>" 0.4
= 40 0.3
m
EL 0.2
20 \T o= 404K
0. 1
0 I I i t
0 100 200 300 400
Temperature (K)
Fig. 4.11. Temperature dependence of the spontaneous field B~ from ZF-/~SR measurements in GdCo2.
The solid lines show the temperature dependence of the saturation magnetizations (from Barth et al.
1986a).

0.4
5 0 l l e ~-~ l ~ ' , , YFezl i _ B. (T)

40 • Bext = 0 T -
0.3
N
-1-
v
30 --
ro
0.2
t-

CT 20
.o
u.. 0.1
10

0 I I I I t d = si]5t
0 100 200 300 400 500 600
Temperature (K)
Fig. 4.12. Temperature dependence of the spontaneous field B~ from ZF-/~SR measurements in YFe2.
The solid lines show the temperature dependence of the saturation magnetizations (from Barth et al.
1986a).
122 A. SCHENCK and EN. GYGAX

' ' ' ' ' t B~ (T)


140 • • 1.0
O~ TmFe2
120 0.9
B~,~= 0 T
0.8
~" 100 ~O" O
-r • 0.7

>, 80 • 0.6
t-
0

• 0.5
o" 60
,~ • 0.4
40 - 0.3

0.2
20-
0.1
I
0 100 200 300 400 500 600
Temperature (K)
Fig. 4.13. Temperaturedependenceof the spontaneous field B~ from ZF-/zSRmeasurements in TmFe2.
The solid lines show the temperature dependence of the saturation magnetizations (from Barth et al.
1986a).

where the contact hyperfine field is the only microscopic field contribution. The
temperature dependencies of the extracted Bc are shown in fig. 4.14. The contact
hyperfine field seems to derive mainly from the M-sublattice. In YFe2 and LaFe2
this is a trivial conclusion. It is found from the study of the dependence of B~, on an
external field that/3c is antiparallel to the Fe-moment direction or the Fe-sublattice
magnetization. In GdFe2 and GdCo2, /3c is determined to be parallel to the total
magnetization. Since the total magnetization is determined by the Gd-sublattice
magnetization and in view of the antiferromagnetic coupling of the two sublattices
one arrives again at the conclusion that Be and the Fe or Co-sublattice magnetizations
are antiparallel. Finally in TmFe2 the total magnetization drops to zero at ~ 225 K
(see fig. 4.13), where the two sublattice magnetizations happen to cancel, but this
is not reflected at all in the #SR-data suggesting again that Bc is induced by the
M-sublattice alone. Extrapolating Bc(T) in fig. 4.14 to zero temperature and plotting
the Bc(0) versus the value of the ordered moment on the M-atoms one finds indeed a
reasonable linear scaling as can be seen in fig. 4.15. No scaling is observed if Bc(0)
is plotted versus the value of the ordered moment on the R-atoms. To the extend
that #+-diffusion and trapping can be ignored fig. 4.14 displays then basically the
temperature dependence of the M-sublattice magnetization in qualitative agreement
with results from neutron scattering experiments (Bargouth et al. 1971).
ZF-studies were also performed on DyNi2 and GdNi2 in the paramagnetic phase
(Chappert et al. 1986b, Dalmas de R6otier et al. 1990b) with the aim to study
MUON SPIN ROTATION SPECTROSCOPY 123

I ] ! I t I l I I

Bhf 0.4 ~
(T) 0.0
..,,.To I I
GdCo2
u u
YFe2
0.4 - ~ I°oalllD
- Oom~
0 0 i I f I t~ i i I

- ~ ErFe2
0.4 _ ~ ' w l l ~ Tc -

0.0 I p i i I~l I I
~p
0.8Fe e l l l l l l l • o e • • • TmFe2 -
/

0.4 [-- • T. -

0.0~ I I J I , ~° t I -

0"8t" • • leIO GdFe2 t


0.4n- " " " ''.~L,. 4

0 200 400 600 800


Temperature (K)
Fig. 4.14. Temperature dependence of the contact hyperfine field Be(0 K) in various RMa-compounds
(from Barth et al. 1986a).

I I I I

1.0- T = OK
"-- TmFe2 e ~
- _~.41~GdFe2
-z
m___ 0 . 5 - ~ ' ~ e 2 eC°' ErFe2
- ~ GdC°2
0.0 ~" t i I I
0 0.5 1.0 1.5 2.0 2.5
I~ of M atom (~ta)
Fig. 4.15. Correlation of/3c(0) with the M-atom magnetic moment in various RM2-compounds (from
Barth et al. 1986a).
124 A. SCHENCKand EN. GYGAX

the spin dynamics in this range. The observed ZF #+-relaxation rate in GdNi2 is
very similar to the corresponding one in GdA12. At room temperature one finds
,kzF(GdA12)/)~zF(GdNi2) = 0.23/0.42 ~_ 0.55(10) which agrees reasonably well with
a calculated ratio of 0.32 invoking only the RKKY mechanism. The temperature
dependence of ~zF(DyNi2) is also quite similar to the one in DyA12 but the room tem-
perature ratio ,kzF(DyA12)/),zF(DyNi2)= 0.67/0.37 = 1.8(4) is in severe disagreement
with a calculated value of 0.21. Various possible explanations for this disagreement
are discussed in Chappert et al. (1986b).

4.2.2.2. RMn2 (R = Y, Tb, Dy). In contrast to the compounds covered in the two
preceedings sections the cubic Laves phase compounds with M = Mn are antiferro-
magnets (see table 4.5). Of particular interest is YMn2 which displays a first order
strongly hysteretic phase transition to an antiferromagnetic, long wave length helical
state at ~ 100 K built from local moments (/~m ~ 2.7#B), while its behaviour above
TN is that characteristic for a weak itinerant electron Pauli paramagnet and is de-
scribed by the self consistent renormalization (SCR) theory of Moriya (1985). The
collaps of the local moment at TN by ,,o 30% is accompanied by a decrease of the
unit cell volume of 5%. The local moment can be destabilized by external pressure
or by replacing Mn by an element with a smaller ionic radius. On the other hand
substitution by an element with a larger ionic radius (providing 'negative' chemical
pressure) is expected to stabilize the local moment on the Mn site. Similar effects
can be induced by substitution of Y. Neutron scattering work revealed that strong
antiferromagnetic correlations persist to at least 6TN. #SR work on this system was
started with the aim to learn more about the spin dynamics and the change from
local to itinerant moment behaviour (Asch et al. 1990, Cywinski et al. 1990a, 1991,
Cywinski and Rainford 1994, Weber et al. 1994).
Pure YMn2 has been investigated by both ZF and TF-techniques. TF-results
(Hext = 22 mT) of Cywinski et al. (1991) are displayed in fig. 4.16 for both
ascending and descending temperature scans. The hysteretic nature of the phase
transition is clearly visible from the shifted drop in the signal amplitude, which
signals the onset or the destruction of the ordered phase. The loss of asymmetry
below TN reflects the appearance of a very wide internal field distribution over the
#+ sites which induces an extremly rapid #+ depolarization within the dead time of
the spectrometer. The appearance of a wide field distribution is in line with the long
wave length helical magnetic structure in the ordered state. The approach of the
transition temperature from above in both scans is accompanied by a quasi divergent
behaviour of the TF-relaxation rate )~2 = l/T2. )k2 could be affected by various
contributions and effects which were not considered in detail: powder broadening,
nuclear dipole fields, fluctuating fields from the electronic moments, #+-diffusion
etc. Weber et al. (1994) by a combination of TF, LF and ZF results find a somewhat
modified temperature dependence of the signal amplitude A as shown in fig. 4.17.
Most interesting is a drop of A starting below ~ 150 K. This drop results from the
appearance of a rapidly damped component which is not taken into account in the
analysis leading to fig. 4.18 and which signals the evolution of another magnetic
phase in part of the sample volume above TN. Interestingly this fast component is
MUON SPIN ROTATIONSPECTROSCOPY 125

0.10 I I f I I I I [ I I I I I I I I I I I I t I I I I I I

I*
0.08
"-1
04
,< 0.06

CO0~
c"
o 0.04
X
rr 0.02
(a) -
0.00 I I I IIIIIIIII111]11 IIiiiii
0 1oo 200 300
Temperature (K)
0.25
II II llll~ ..iil.I.II
~ ~warming .I6 ....
0.20

~" 0.15
E cooling
E

22
>.,
o~
< 0.10

0.05
(b)

0.00 II I IIIII IIIII II II IIIII IIII1


0 1O0 200 300
Temperature (K)
Fig. 4.16. Temperaturedependence of signal amplitude (asymmetry) and relaxation rate ),2 from TF
measurements (Hext = 2.2 mT) in YMn2. The solid lines are guides to the eye. Not shown explicitely
are the ),z-data points on cooling. Note the hysteretic behaviour (from Cywinski et al. 1991).

best fitted by a stretched-square root-exponential decay function (see eq. (2.21)) as


appropriate for a dilute spin glass above the freezing temperature. LF-measurement
show that this component is purely dynamical in origin and independent of the
applied field strength (Hext < 0.2 T). The evolution of the fast component could be
interpreted as following from the development of spin glass like clusters, starting
well above the transition into the long range ordered state. The same and even
more pronounced behaviour was also found in Tb doped samples of composition
Yo.9Tbo.lMn2 (Asch et al. 1990) and Y0.95Tb0.osMn2 (Weber et al. 1994). Here
the damping rate in the spin glass like phase could be extracted (~1 ~- 1 #s -1) and
126 A. SCHENCK and EN. GYGAX

I I I I

100

50
13.
E
<

0 1 O0 200
Temperature (K)
Fig. 4.17. Temperature dependence of the signal amplitude derived from a combination of TF, LF and
ZF measurements in YMn2. For the slight drop at 150 K see text (from Weber et al. 1994).

~ 3.014 ,5~ ~'


~ 3.013 o
> 3.012 I I
I I I I I /
Q ,..z-~3£X300.O
o °_~ ' ' - - P-- - - 0 - - 0 - 0 - 0 - 0 - . . 0 - - 0 -
4-"
2• 015
0.20
b)
1 0.20

015
,~ o.lo o.lo .~

0.05 0.05

0.0 t t t t t
0 100 200 300
Temperature (K)
Fig. 4.18. Temperature dependence of signal amplitude (asymmetry), relaxation rate ),2 and precession
frequency from TF-measurements (Hext = 22 roT) in Y(Mn097Fe0.03)2 (from Cywinski et al. 1990a).

was found to have no significant temperature dependence. Some preliminary studies


were also undertaken on Sc-doped (substituting for Y) samples (Weber et al. 1994).
Due to the limited amount of data the results will not be discussed here.
Further TF-#SR studies were performed on Fe substituted Y(Mnl_=Fex)2 (Cywin-
ski et al. 1990a). Neutron diffraction and susceptibility measurements (Cywinski
et al. 1990b) imply that as little as 2.5% substitution of Fe for Mn suffices to
suppress magnetic order altogether and to induce Pauli paramagnetism. This was
basically confirmed by #SR results on a sample with z = 0.03. However, neutron
polarization analysis measurements (Cywinski et al. 1990a) show the persistence
MUON SPIN ROTATION SPECTROSCOPY 127

TABLE 4.6
#+-hyperfine fields extrapolated to T = 0 K and
contact hyperfine coupling parameters Ac. Also
listed is the ordered moment on the M-ions
(Barth et al. 1986a).

/~hf(0) Ac /ZM/#B
(T) (T/LtB)
GdFe2 0.7 -0.43 1.62
GdCo2 0.41 -0.40 1.02
YFe2 -0.6 -0.41 1.45
TmFe2 0.8 -0.47 1.7
ErFe2 0.6 -0.35 1.7

of antiferromagnetic correlations in the paramagnetic state over a wide temperature


range. The neutron results could not distinguish between the possibilities that these
correlations are static, reflecting perhaps topological magnetic frustration of localized
Mn moments on a tetrahedral lattice, or are dynamic on the time scale of longitudinal
fluctuations of the Mn spins. The TF-#SR relaxation function is very well described
by an exponential function implying a dynamical origin and the temperature depen-
dence of the relaxation rate A2 is displayed in fig. 4.18. It is well reproduced by
the expression A2(T) = C T - x with z = 0.75, C = 1.12 # s - l ( K ) x. In contrast no
temperature dependence in the precession frequency (Knight shift) could be detected
in agreement with the picture of a Pauli paramagnet. The increase of A with de-
creasing temperature signals a slowing down of the Mn spin dynamics. It is argued
that it does not arise from a slowing down of transverse spin fluctuations but instead
from a slowing down of longitudinal spin fluctuations reflecting the localization of
the moments on the Mn(Fe) sites. In any case the #SR-data are inconsistent with
the evolution of static correlations and thus help in the interpretation of the neutron
results.
Substituting Mn by A1 instead of Fe leads to a 'negative' chemical pressure and
the expectation that it stabilizes the local moment on the Mn site. Indeed a slowing
down of spin fluctuations has been claimed in A1 substituted Y(Mnl_~AI~)2 (Motoya
et al. 1991) and for z > 0.06 the volume collapse is suppressed and a spin glass like
state is found at low temperatures. #SR-measurements on a sample with z = 0.1
clearly confirms spin glass order below Tg = 60 K (Cywinski et al. 1994b).
Powder samples of DyMn2 and TbMn2 were studied by a few TF-#SR measure-
ments only (Cywinski et al. 1992). Below TN the #SR-signal is lost completely.
Above TN Fourier transforms of the #SR-spectra show the development of a pro-
nounced asymmetric line shape as the temperature approaches TN from above. This
feature which, in part, may be due to powder broadening has so far prevented a
detailed interpretation of the results. Very preliminary ZF-measurements on TbMn2
(Weber et al. 1994) show a rather complex behaviour suggesting the coexistence of
static and dynamic correlations already at relatively high temperatures (250 K).
128 A. SCHENCKand EN. GYGAX

4.2.3. Binary compounds with NaCl and CsCl-structure


A list of compounds studied by #SR is found in table 4.7. The magnetic behaviour
of all these compounds is rather complex and is characterized by a delicate interplay
of various exchange and anisotropy mechanisms.

CsAs. W-measurements (0.02 T) in this compound display a strongly temperature


dependent exponential relaxation rate ),2 below 9 K with a sharp singularity at 7.6 K,
as shown in fig. 4.19 (Asch et al. 1988, Litterst et al. 1990). This temperature
is identified with TN. At the same temperature the onset of a spontaneous spin
precession is observed (at 4 K: u, -~ 24 MHz or B , "-~ 0.177 T). ZF-measurements
reveal additional subtle features. Here a rapidly damped component appears already
below 9 K and can be followed down to 7.6 K. At 7.3 K the spontaneous oscillating
component undergoes a drastic change in spectral appearance (see fig. 4.20). The
following scenario is suggested by these results: above 100 K the Ce-spins fluctuate
freely (paramagnetic fluctuations); below this temperature spin correlations develop
which, in part of the volume, lead to short range order below 9 K; at 7.6 K long
range order sets in with a change in magnetic structure at 7.3 K. The change in
structure at 7.3 K may reflect a transition between the single ( a n d triple q structure
which both may have been seen in CeAs (H~tlg and Furrer 1986, Buffet et al. 1984)
and which theory predicts to be nearly degenerate (Prelovsek and Rice 1987). Some
interesting correlations are found in another neutron scattering study (H~ilg et al.
1987), where magnetic Bragg peaks appear already at 8.5 K and where short range
ordered critical spin fluctuations peak at 7.3 K.
A problem is connected with the observation of a rather broad spontaneous preces-
sion signal. If, as usual for these cubic compounds (see also section 4.3.2) the #+ is
placed at the center of the cube formed by four Ce and four As atoms (see fig. 4.21)
dipolar field calculations lead for both the single ~ and the triple q"structures to zero
net dipolar fields at this site. As a way out it has been suggested that the presence

0.30
b~ 0.25 -0
v
0
0.20
t-
._o 0.15
..t-.,
tl:l
.BN
i_
0.10 O(9oO° CO
O
Q.
0.05 oO 0 0 O O
a
I I I I I I I I I I I I I I I
0.00
0 20 40 60 80 100 120 140 160
Temperature (K)
Fig. 4.19. Temperaturedependenceof transverse fieldrelaxationrate ,k2 ( n e x t = 20 mT) in CeAs (from
Litterst et al. 1990).
MUON SPIN ROTATION SPECTROSCOPY 129

Ox
O~

o~
,,~ Ox

~d

~ : ~o~=

R R ~

H
I~ ~0 O~

T~

<
t~ tr~ r~ BT~ t~,~

~ ..'~

Z Z Z ~

o
<<<
130 A. SCHENCK and EN. GYGAX

2.0

1.0 13K
Ok-
2.0

1.0 ~ ~ 8K

2.01.0~ 7.5K

2.o
o

1.0 7.4K
o
u_
0
0 4 8 12 16 20

1.0 7.OK

0
0 ~ lb 1~ 2b 2~ 3'o

1.0 4K

o
0 10 20 30 40
Frequency(MHz)
Fig. 4.20. Fourier power spectra from ZF-measurements in CeAs. Note the change of the spectral
distribution around 7.4 K (from Litterst et al. 1990).

of the #+ alters locally the spin structure which may not be unreasonable in view of
the delicate balance of the various competing exchange interactions etc.

CeSb. A total of 6 different AF phases are observed in this compound (Fischer et


al. 1978). The magnetic structure results from a temperature dependent stacking
of paramagnetic and ferromagnetic (001)-layers with the Ce 3+ spins pointing along
+(001). The observed spin modulations can be described by a single wave vector
q'0 = (27r/c)(0, 0, q) where q falls between 1/3 and 2/3 in fractions of integers and
hence indicates a commensurate behaviour of the spin modulations (see table 4.8).
MUON SPIN ROTATION SPECTROSCOPY 131

)o • (
I ® (
@

o ~:

Fig. 4.21. Crystal structure of fcc RX-compounds (X = pnictides, chalcogenides). Indicated is the most
likely/z+-site.

Placing the #+ again at the previously assumed site (see fig. 4.21) three different
magnetic environments can be identified with populations depending on the exact
value of q. At one type of site the net dipolar field vanishes ( u u = 0), at the two other
sites non zero dipolar fields are calculated corresponding to precession frequencies
of u~ = 16.5 MHz and u~ = 33 MHz. ZF-pSR measurements show the following
behaviour (Klauss et al. 1994): Below TN two signals are found, one reflecting
spin precession at 29.4(5) MHz and a second non oscillating signal displaying a
stretched square root exponential behaviour. The relative amplitudes mirror nearly
the expected populations for the sites associated with the u~, = 0 and u u = 33 MHz
sites (see fig. 4.22). The sum of the amplitudes do not add up to the full asymmetry,
seen above TN. The missing amplitude is assigned to the uu = 16.5 MHz site. Its
absence is explained in terms of a very rapid damping of the corresponding #SR-
signal arising from relatively slow fluctuations of the spins in the paramagnetic plane
adjacent to this site. Consistent with the first order nature of the transition at TN the
spontaneous frequency of 29.5 MHz is nearly constant in the whole antiferromagnetic
temperature regime. In contrast above TN a rising relaxation rate (following from
both TF and ZF measurements) with decreasing temperature proves the existence of

TABLE 4.8
List of AF phases and related quantities in CeSb (Klauss et al. 1994).

Phase Temperature q Population (in %) of sites with


range (K)
u=0 u=16.5MHz u=33MHz
I 15.9-16.1 2/3 33.3 66.7 0
II 15.3-15.9 8/13 38.5 46.1 15.4
III 13.7-15.3 4/7 42.8 28.6 28.7
IV 11.0-13.7 5/9 44.4 22.2 33.3
V 8.9-11.0 6/11 45.4 18.2 36.4
VI 2.2-8.9 1/2 50 0.0 50.0
132 A. SCHENCK and EN. GYGAX

0.05 I I I

0.04

0.03 I I
E
E
0.02

0.01

0 I I
0 5 10 15 20
Temperature (K)
Fig. 4.22. Temperature dependence of the amplitude (asymmerty) of the 29.5 MHz signal (from ZF
and 5 mT-TF measurements) in CeSb. The horizontal solid lines represent the predictions according to
table 4.8 (from Klauss et al. 1994).

spin fluctuations which slow down upon the approach of TN. This has also been
observed in diffuse critical neutron scattering (Halg et al. 1985).

DySb. This compound shows a MnO-like antiferromagnetic (AF II) structure (Felcher
et al. 1973). Dipolar field calculations predict a single frequency of 210 MHz at the
#+-site considered before. In contrast ZF-#SR measurements (Klauss et al. 1994)
reveal two signals with nearly the same amplitudes, one displaying a spontaneous
precession with uu _~ 240 MHz and zero frequency the other one. The latter signal
displays again a square root exponential relaxation with A ,-~ 10 #s -1. The relaxation
seems to be purely of dynamical origin since it could not be quenched in LF-runs
up to 0.6 T. The appearance of two signals can be understood if instead of the MnO-
like magnetic structure a CoO-type structure is considered. One then predicts two
magnetically inequivalent sites with the same statistical weight and zero frequency
at one site and ~300 MHz at the other site. Such a structure, however, seems to be
severly at odds with earlier neutron results (Felcher et al. 1973). Again the observed
spontaneous precession frequency is nearly temperature independent consistent with
the first order nature of the transition. However, the ZF relaxation rates show a
strong increase as TN is approach from both above and below.
TF-measurements above TN reveal an anomalously strong increase of the (neg-
ative) Knight shift below 20 K on approaching TN (Klauss et al. 1994). If the
CoO-type structure is adopted it can be explained by the onset of a quadrupolar
anisotropy some degrees above TN which would be in line with earlier structure
studies (Chen and Levy 1971).

CeAg, CeAgo.97Ino.03.CeAg as well as other RX (X = Cu, Ag etc.) compounds with


cubic CsC1 structure have attracted much attention due to their magnetic and struc-
tural properties which are governed by one ion magnetoelastic couplings and two
MUON SPIN ROTATION SPECTROSCOPY 133

0.20 0.20
O CeAg CeAg l_xlnx(x=O.03)
O
-...,- o b
0.15 o c
(/)
o
,$ o
E 0.10
E c9 0.2 0.10 0.2
O E >" 0 ~-
t.9 O tO 0
,< < '.=
X i X
0.05 0.1 0.05 "" 0.1
(1)
m
m •
~IID, ' • 0 I I 0
20 40 60 0 20 40 60
T (K) T (K)
Fig. 4.23. Temperaturedependence of ZF signal amplitude(asymmetry)and relaxation rate in CeAg
and CeAgo.97Ino.o3(fromHyomiet al. 1988).

ion quadrupolar interactions (Morin and Schmitt, 1990a). CeAg shows a martensitic
transition into a tetragonal phase at TQ _~ 15 K which is followed at Tc = 5 K by a
second order transition into a ferromagnetically ordered state. The structural phase
transition is driven by quadrupolar ordering (Morin, 1988). Substituting In for Ag it
is found that TQ increases strongly with In concentration (Ihrig and Lohmann, 1977).
There is some debate as to the nature of the structural instability in substituted com-
pounds CeAgl_xIn= with z > 0.01 which does not seem to arise from quadrupolar
ordering.
ZF- and TF-#SR measurements on polycrystalline samples of CeAg and
CeAg0.97In0.03 were aimed at a study of the structural phase transition (Hyomi et
al. 1988) while the magnetically ordered phase was not investigated. Previously the
#+ Knight shift was studied in the nonmagnetic series LaAga_=In= for 0 ~< z < 0.3
(Wehr et al. 1983). Figure 4.23 shows ZF relaxation rates ),ZF and the signal ampli-
tudes as a function of temperature. In CeAg0.97In0.03 Azv starts to rise below ~ 25 K,
reflecting the slowing down of the spin dynamics as the second order magnetic phase
transition is approached, but no sign of the structural phase transition is manifest in
these data. In contrast the behaviour of AZF in CeAg is totally different. Again below
about 25 K AZF starts to rise. But concomitantly the signal amplitude starts to drop
pronouncedly. This is due to the development of a very rapidly decaying component
which could not be resolved in these experiments performed at the pulsed #+ source
at KEK (Tsukuba, Japan). These results suggest that static spin correlations develop
already around and below 25 K, leading to a wide distribution of static fields at
the #+-sites quite in contrast to the situation in the In doped compound. Whether
this different behaviour is connected to the nature of the structural phase transitions
remains to be studied. Measurements with much improved time resolution would be
extremely interesting.

DyAg. This compound displays antiferromagnetic order below TN = 60 K but no


structural instability. Neutron diffraction has established that the spin order just
134 A. SCHENCK and EN. GYGAX

below 60 K is an incommensurate sinusoidally modulated structure which below


51 K transforms to a type I non-collinear triple (structure (Kaneko et al. 1987).
The presence of two transition temperatures is also indicated by bulk magnetic and
transport studies (Morin et al. 1990b, Sousa 1990). These findings are nicely cor-
roborated by ZF and LF-#SR measurements on poly and monocrystalline samples
(Asch et al. 1984, Kalvius et al. 1986, 1987, 1990). Approaching TN from above
one observes a strongly rising relaxation rate as expected for a second order phase
transition. At 300 K one finds )~ = 0.9 #s -1 which is of the same order of magnitude
as in DyAI2 (see table 4.4). At TN the relaxation function changes from exponential
to a stretched exponential without any drastic change in the overall relaxation rate.
A further change occurs at ,,~ 51 K, below which a dynamic Lorentz-Kubo-Toyabe
function describes the ZF-signal best. With further decreasing temperature a more
and more static Lorentz-Kubo-Toyabe signal develops. Figure 4.24 shows the tem-
perature dependence of the width A and of the fluctuation rate u extracted from the
dynamic Lorentz-Kubo-Toyabe fits. No oscillating signal is seen below TN. This
finds its explanation in the fact that at interstitial sites of high symmetry the sum
of all dipolar fields from the ordered moments is zero. The Lorentzian shape of
the residual field distribution at the #+ implies that the magnetic structure is nearly
perfect and that any distortions (e.g., magnetic holes) are dilute and not too close
to the #+. A surprising and unusual result is the increase of the width ,4 with in-
creasing temperature. It is interpreted as being due to an increasing de-locking of
the triple (magnetic structure and concomitantly, a progressive increase of a slight

10 , I , , , 4

DyAg
3

-1-
6

¢- 2 rr
, Width/ ,,,'
4
o

E
1
P Rate ,
J
I
~O ~

I I I L
00
10 20 30 40 50 0
Temperature (K)
Fig. 4.24. Temperature dependence of 4f-moment fluctuation rate u and coupling constant (= width of
the field distributions in the static case) A from dynamic Lorentz-Kubo-Toyabe fits to ZF-data in DyAg
(from Kalvius et al. 1990).
MUON SPIN ROTATIONSPECTROSCOPY 135

incommensurability (invisible to neutrons) as one moves toward the incommensurate


phase commencing at ~ 51 K (Kalvius et al. 1990). The stretched (power) expo-
nential relaxation behaviour between ~ 51 K and 60 K, which should be associated
with the incommensurate sinusoidally modulated structure, has not been rationalized
yet. Some #SR-measurements have also been performed on amorphous DyAg which
orders ferromagnetically at Tc = 18 K (Kalvius et al. 1986, 1987). A very rapid
relaxation is seen to develop below 100 K leading to a loss of the signal below
90 K consistent with M6sslauer measurements (Chappert et al. 1982). This signals
a massive slowing down of the spin dynamics and a wide distribution of local fields
due to random distortions of the local symmetry in the amorphous state.

4.2.4. RNi5 compounds


The RNi5 compounds crystallize in the hexagonal CaCu5 structure (see fig. 4.25) and
for heavy R (Gd,...) undergo a transition into a ferromagnetic state at temperatures
below 37 K (see table 4.9). PrNi5 is a Van Vleck paramagnet which shows hyperfine
enhanced nuclear spin ordering at 0.4 mK (Kubota et al. 1980).
ZF and TF-#SR measurements on the nonmagnetic reference compound LaNi5
indicate that the #+ is immobile up to 150 K and most likely occupies the f-site
(1/2, 0, 0) which is also found to be occupied by hydrogen in the a-phase of LaNisHx
(Dalmas de R6otier et al. 1990c). In PrNi5 the i-site, located near the f-site, was
unambiguously identified to be the #+ site from the angular dependence of the #+-
Knight shift [Feyerherm et al. (to be published), a previously claimed multiple site
occupation in PrNi5 (Kaplan et al. 1989, Hitti et al. 1990a) could not be confirmed
in subsequent studies on several new high quality samples]. As reviewed below, the
magnetic properties as sensed by the #+ are largely determined by the crystalline
electric field (CEF) split ground state multiplet of the 4f-electrons affecting both the
static and dynamic behaviour.

OU ~Ni ~AI
Fig. 4.25. Crystal structure of the hexagonal CaCu5 type. Indicatedare various possible #+-sites: b, f,
h, o, m with multiplicities of 1, 3, 4, 12, 6 per unit cell, respectively. The i-site (not shown) is found
along the c-axis above or below the f-site.
136 A. SCHENCK and EN. GYGAX

~- o

¢->

.~=

.=
o

,/ ~',~
II "~
¢-
0
©
~-~'N
(",1 t"q ~
:8"

b b o

0
0

~ m

~N
~ ~.~.~.~ ~.~.~ ~'~
.... Z Z ~
MUON SPIN ROTATIONSPECTROSCOPY 137

GdNis. Zero field measurements on a single crystalline sample with fi(0) perpen-
dicular to the c-axis at 3.3 K, i.e. below Tc, revealed two spontaneous precession
signals with Ul = 5.3 MHz and u2 = 220.9 MHz with an amplitude ratio al : a2
of ~ 3 : 2. Obviously two different #+ sites must be involved. If fi(0) is chosen
parallel to the c-axis no oscillating components are seen (Gubbens et al. 1994a).
These results imply that the spontaneous local field at the #+ site is oriented parallel
to the c-axis and suggest further that the c-axis is the easy axis. This conclusion is
inconsistent with magnetization measurements (Gignoux et al. 1976) but agrees with
M6ssbauer data (e.g., Tomala et al. 1977). The ZF-relaxation ),ZF has been mea-
sured for fi(0)H c-axis from room temperature down to 0.2 K. The results are shown
in fig. 4.26. At high temperature AzF is temperature independent as expected for
a Heisenberg magnet, neglecting the Korringa mechanism (see also section 4.2.1).
The sharp increase of AzF in the vicinity of Tc reflects the critical slowing down of
the spin dynamics. Since the temperature reading was not very accurate these data
cannot be used for the determination of critical exponents. Relaxation can also be
seen below Tc, involving the full asymmetry, which, below ~ Tc/2, varies approx-
imately like T 2 ln(T) (solid line in the insert of fig. 4.26). This behaviour points to
a two magnon scattering process (Dalmas de Rdotier and Yaouanc, 1994b). Fitting
with the appropriate function a spin wave stiffness constant of D = 4.2(2) meV ,~2 is
deduced. Note that this analysis does not require knowledge of the #+-site (Gubbens
et al. 1994a).

I'''' I''''1''''1'''' I''''1''''


. | ,i .... i .... ~ .... i .... I .... ] .

0.14 * _
5 o.12~- :
0.10 F °
N 0,08 [- "
•T" 4 -~
0.06

0.04 ~- .~

{D 0.02
3 ooo -
L_ ~ .... I .... i .... t . . . . . . . . R

• , 5 lO 15 20 25 30
.c 2 ,o , • •
Cl. •

1 j GdNis
0
! I I I I [ I i I I [ . . . . I , , , , I , , , i I . . . .

0 50 100 150 200 250


T e m p e r a t u r e (K)
Fig. 4.26. Temperature dependence of the ZF-relaxation rate in GdNi 5 for P(0)llc-axis. The insert
shows the data below Tc On an expanded scale. The solid line is a fit to a two magnon scattering
process (from Gubbens et al. 1994a).
138 A. SCHENCK and EN. GYGAX

TbNi5. This compound displays a strong magnetocrystalline anisotropy and the or-
dered moments point along the a-axis. Information on the CEF-levels of Tb 3+ can
be found in Gignoux and Rhyne (1986). ZF-#SR measurements on a single crystal
sample were performed below and up to Tc (Dalmas de R6otier et al. 1992). A
single spontaneous precession signal was observed the temperature dependence of
which is displayed in fig. 4.27. According to section 2.2

/~ 4~r
: ~ - ]~fs -}- Bdip -}- Bc (4.13)

~: gj/ZB (3~f) (4.14)

assuming that the contact hyperfine field/3c c~ (~f). As shown in section 3.1 this is
not generally true, but may be a reasonable assumption here by noting that Tc << (90.
The temperature dependence of {~f) was calculated in a mean field approximation
taking into account the CEF splitting of the Tb 3+ ground state multiplet (Gignoux

Illllllllllllll IIlllllll
80

~" 6o
4O
g
(~ 20

0
TbNi~

30
ri-

d 20 +++ -

c 10
E

IlllJltq~ll,lllltllllll
0 5 10 15 20
Temperature(K)
Fig. 4.27. Temperature dependence of the spontaneous precession frequency and associated relaxation
rate below T c in TbNi5 (from Dalmas de R6otier et al. 1992).
MUON SPIN ROTATION SPECTROSCOPY 139

J t ~ v uul~ n t i u ill] ; t

100
~,S~, CPOo

£/
N d~ o8
-i- o
10 ?
'1,...
~o~ • . •o°~5~o,, °
r-
° ~
ErNis
E
£3 o PSI
• ISIS
0.1
i i i t ~lJl I ~ i I i i ill

10 100
Temperature (K)
Fig. 4.28. Log-log plot of the ZF relaxation rate versus temperature for P(0)llc-axis and/~(0)_Lc-axis,
respectively in ErNi5. Data shown by filled (open) symbols were obtained at the ISIS facility of the
Rutherford-Appleton Lab. (at the Paul Scherrer Institute) (from Gubbens et al. 1994a).

and Rhyne, 1986). It is represented by the solid line in fig. 4.28 and reproduces the
data very well. The dashed line is a Brillouin function obtained by ignoring the CEF
interactions. Obviously CEF-effects are important and cannot be ignored. However,
choosing a different set of CEF parameters, one obtains about the same results as
represented by the solid line (Dalmas de Rdotier et al. 1992).
ZF relaxation rates measured above Tc (Dalmas de R6otier et al. 1990c) as well
as the relaxation of the precession signal below Tc have not yet been explained but,
at least for T > Tc, seem to be also significantly affected by the CEF-splitting.

DyNis. Gubbens et al. (1994a) report that their ZF-results on DyNi5 resemble basi-
cally those on TbNis.

ErNis. Like in TbNi5 a huge magneto crystalline anisotropy is found in ErNi5 which
leads in the ordered state to an easy axis parallel to the c-axis below Tc = 9.2 K.
ZF-/zSR measurements on sing~ crystals display strongly anisotropic features
(Gubbens et al. 1992, 1994a) For P(0)±c-axis no signal is found below Tc. Above
Tc the ZF signal is well described by a single exponentially damped component.
The relaxation rate Az~c rises steeply with decreasing temperature and becomes so
short below ~ 55 K that the signal is altogether lost (see fig. 4.28). For f(0)llc-axis
a single exponentially damped signal is found above about 7 K while below 7 K
two components are revealed, one of which is temperature dependent. However,
already at around 12 K, the signal amplitude drops by ~- 25% indicating that a cor-
responding fractions of #+ are relaxed extremely fast, reflecting perhaps the on-set
140 A. SCHENCK and EN. GYGAX

of static spin correlations in part of the volume already at around 12 K. The temper-
ature dependence of the relaxation rate ),lzl~ is displayed in fig. 4.28 too. It shows a
maximum at around 12 K (!) and is much reduced from ,kzX~up to ,,~ 200 K. The
observed anisotropy thus persists to temperatures T ~ 20Tc. Gubbens et al. 1992
could explain qualitatively the high temperature anisotropy in terms of a mean field
picture including the CEF splitting in a very approximate way. For P(0)_Lc-axis an
exponential temperature dependence of )~ZF at low temperatures is predicted while
)~lzl~ is only weakly temperature dependent. According to a 166Er-MSssbauer study
the spin fluctuations are very slow (quasi static) below 50 K (Gubbens 1989). This
may be in line with the up turn of ) ~ below around the same temperature. The slow
fluctuations are thought to arise from an indirect process between the two levels of
the Kramers ground state doublet. The temperature dependence of ,klZlFbelow Tc is
very well described by the expression

A(T) = a ctg ( A / kBT) + bT 7 (4.15)

with a = 0.15 MHz, A/kB /> 5 K and b = 1.1 Hz/K 7 (solid line in fig. 4.28). This
functional dependence is a clear signature for a phonon induced relaxation mechanism
which is produced by the magnetoelastic coupling of the 4f spins to lattice vibrations.
The first term corresponds to a one phonon process and the second to a two phonon
process, respectively. The fitted parameters have not been quantitatively explained
yet.

TmNis. ZF-measurements on single crystal samples above Tc (Gubbens et al. 1994a)


yield spin fluctuation rates which for both orientations/5(0)1 ]c-axis and/5(0)_Lc-axis
show identical temperature dependencies and agree very well with 169TinM6ssbauer

lo "t
=.E. t ~''
.E
"4-'~_
°
.o 1 }+C~ 4)0
;,2 bauer

o TmNi s
0.1
I .... I,,~ JL,~ JlT~,,I .... I~,~L

0 50 100 150 200 250


Temperature (K)
Fig. 4.29. Temperature dependence of the Tm-4f moment fluctuation rates in ZF normalized to cor-
responding M/~ssbauer data in TmNi5 (from Gubbens et al. 1994a).
MUON SPIN ROTATIONSPECTROSCOPY 141

results. Figure 4.29 shows the temperature dependence of the corresponding corre-
lation times "r (= reciprocal fluctuation rate). Since the #+ site and related coupling
parameters to the Tm 3+ moments are not known fig. 4.29 shows essentially ~IzlF(T)
and )~-~(T) normalized to the M6ssbauer results. The observed functional form of
the temperature dependence has not been explained yet.

YbNis. YbNi5 has a very low ordering temperature of only ~ 0.5 K (Hodges and
Bonville, unpublished results). ZF-#SR and M6ssbauer measurements on a poly-
crystalline sample were performed from 0.1 K up to room temperature (Bonville et
al. 1994). The temperature dependence of the #+ relaxation rate agrees very well
with the temperature dependence of the spin fluctuation correlation time "r extracted
from the M6ssbauer data. It is found that the fluctuation rate increases linearly with
temperature (i.e. AzF oc 1/T) pointing to a Korringa type mechanism as the driving
force. Compared to ErNi5 and TmNi5 the relaxation rates are much larger and much
less temperature dependent. It seems that CEF effects are less important in YbNi5
than in the two other compounds.

PrNis. Somewhat out of order we will also briefly review some TF-#SR results on
this singlet ground state compound which does not show electronic magnetic order.
The previous discussion of results in TbNis, DyNis, ErNis, TmNi5 and YbNi5 rested
to some extent on CEF-induced effects. It was hereby assumed that the implanted
#+ does not modify the CEF-Hamiltonian governing the nearest neighbour rare earth
ions. That this assumption may be grossly in error is demonstrated by #+ Knight
shift measurements in monocrystalline PrNi5 where the #+ site is unambigeously
identified with the interstitial i-site (Feyerherm et al. 1994b). It is found that while
the Knight shift at high temperatures (i.e. above 50 K) scales very well with the
bulk susceptibility this scaling is totally lost at lower temperatures. The result can

0.45 _~l i I I i i I I !

0.40 \ ;¢y . . . .
"5 0.35 y

0.25
T
.~ 0.20
=o_ o.15
0 0.10
m 0.05
0.00
0 20 40 60 80 1O0
Temperature (K)
Fig. 4.30. Temperature dependence of the local susceptibility in the basal plane of PrNi 5 sensed and
modified by the/z + via the Knight shift (dipolar and contact) for two different directions defined in the
insert (fromFeyerhermet al. 1994b).
142 A. SCHENCK and EN. GYGAX

PrNi 5 p e r t u r b e d by IJ+

E( 1 ) 48 K - -
5 - - 45K

r6 39 K - -
- - 37K

23 K - -

\ 23 K

7.5K

v4 o 0

Fig. 4.31. Crystal field splitting of the 3H4-ground state multiplet of Pr3+ in PrNi5 without (undisturbed)
and with (disturbed) the/~+ at the interstitial i-site (only nearest neighbour Pr3+ are considered) (from
Feyerherm et al. 1994b and private communication).

be used to extract an effective local susceptibility which in the basal plane becomes
anisotropic as shown in fig. 4.30. The solid lines represent a fit to a CEF-model
in which the CEF parameters B2°,/322 and the exchange parameter A were allowed
to vary independently. A best fit was obtained for B ° = 0.54 MeV (0.51 MeV)
B 2 = 0.08 MeV (0 MeV) and A = 2.7 mole/emu (4.73 mole/emu). The values
in parenthesis refer to bulk parameters as obtained from inelastic neutron scattering
(Amato et al. 1992a). Figure 4.31 compares the CEF level schemes with and
without the #+ induced perturbation. Of particular significance is the appreciable
lowering of the first and second excited levels. Point charge model calculations
failed to reproduce the fit results but it is clear that the #+ introduces an additional
electric field gradient and destroys the local symmetry on the nn-Pr sites. The #+
induced perturbation may be particularly important in the case of singlet ground
state ions but may be also non-negligible in other cases showing a CEF splitting. In
GdNi5, however, no #+ induced effects would be expected (see also the discussion
in section 3.2 and Campbell 1984, Schillaci et al. 1984).

4.2,5. Ternary compounds with tetragonal ThCr2Si2 structure


Table 4.10 presents a list of investigated compounds. Two more compounds with the
same structure are treated in section 4.2.6 on heavy fermion systems (see table 4.11).
ZF-#SR measurements in the nonmagnetic compounds YCo2Si2 and LaPd2P2
served to get an idea on the #+-site. Possible #+-interstitial sites in the ThCr2Si2
type structure are indicated in fig. 4.32. The observed zero field Kubo-Toyabe sig-
nals point to an occupation of site No 2 (1/2, 0, 0) or site No 4 (1/2, 1/4, 0) (Dalmas
de R6otier et al. 1990a). The question is, of course, whether in all compounds
with ThCrzSi2 structure the same site is necessarily occupied (see also discussion on
CeRuzSi2, section 4.2.6 and URu2Si2, section 4.3.4). #+ diffusion is seen to set in
above ~ 130 K (Dalmas de R6otier, 1990a).
MUON SPIN ROTATION SPECTROSCOPY 143

TABLE 4.10
List of ternary compound with tetragonal ThCr2Si2 structure investigated by #SR (see also table 4.11).

Compound Magnetic TN, TC #SR Samples References


structure (K)
YCo2 Si2 - - ZF sintered powder Dalmas de R6otier et al.
(1990a)
LaPdzP2 - - ZF sintered powder Dalmas de R6otier et al.
(1990a)
CeRhzSi2 AF:single ~',/7[[c-axis ~ 35 ZF sintered powder Dalmas de R6otier et al.
(1990a)
PrRu2Si2 FM,/Tile-axis ~ 15 ZF s.c. Yaouanc et al. (1994)
PrCo2Si2 AF, 4 metamagnetic 30 (9,17) ZF s.c. Matokawa et al. (1990),
transitions Nojiri et al. (1992)
NdRhzSi 2 AF: single ~,/Tllc-axis 57 ZE LF s.c., sintered Gubbens et al. (1994b)
powder

~ ) # 6

#3
i

.#7

© o •
Ce Cu Si
Fig. 4.32. Crystalstructureof tetragonaltenarycompoundsof the ThCr2Si2 type. Starsindicatepossible
interstitial/~+ locations.

CeRh2Si2. Z e r o field data confirm the onset o f m a g n e t i c order at ~ 35 K ( D a l m a s


de R6otier et al. 1990d), p r e v i o u s l y identified b y neutron scattering (Quezel et al.
1984a). T h e onset o f o r d e r manifests itself b y a drastically increased relaxation rate
with no loss in the signal amplitude. M e a s u r e m e n t s at 22.5 K reveal an additional
p r e c e s s i o n signal c o r r e s p o n d i n g to a spontaneous field o f ~ 11 m T at the # + position.
Since no m e a s u r e m e n t s were m a d e in the range 22.5 to 33 K no information is
available on the e v o l u t i o n o f the precession c o m p o n e n t which, as far as the limited
data can tell, is not present yet a c o u p l e o f degrees b e l o w TN. The a p p e a r a n c e o f
two signals a s s o c i a t e d with zero net field and 11 mT, suggest the o c c u p a t i o n o f two
different sites. In fact, using the m a g n e t i c structure o f Q u e z e l et al. (1984a, b), one
calculates d i p o l a r fields o f zero and 13 m T at the sites No. 2 and No. 4, respectively.
144 A. S C H E N C K and E N . G Y G A X

0o

c-
o
°i
b b b
~D
d O O O O O q q -6 d ~ q
r.~ e'~ ~ e'~ e'~

,-d ,...]
[..U PU
N
[-..,
r.s., ;.i.7 g . ,
N t'q t"q N N

8 , .4- .~- ~
t'¢3
s
t~ O tt~
t"q

t"q t"q t"q

-d ,4
P T ~D

I t'q

t.-q.

~m o= t~:r,
t'q t"q

© ,4 ,4 E
t"q
e,~
O

X
a.7o z O
< < < ~-.
i <
o

eq ~ eq

~r.)
g~r,.)
#
o ,..~ e"~ v

2
, ? eq ¢q
0
eq
< .<
0
o 6 r) 666 r,.)
"~ u6 r,.)
MUON SPIN ROTATION SPECTROSCOPY 145

c~

r,~

[-.-, ,.d [--.,

O
t"-'-

•4 .4 .4

¢-..1

.4
,.-.-k
,4 .4
eb

t',l

r..)

u-!.
o
° °

o
r..)
146 A. SCHENCK and EN. GYGAX

PrRu2Si2. In contrast to the other compounds in table 4.10 PrRu2Si2 shows ferro-
magnetic order below Tc = 15 K and has a giant magnetic anisotropy (anisotropy
field Ha = 400 T (Shigeoka et al. 1992)) which seems to be much larger than in any
other 4f-electron system reported so far. CEF effect are certainly very important but
cannot account exclusively for the magnetic properties of this compound.
ZF-/~SR measurements below Tc yield a spontaneous precession signal if fi(O)±c-
axis and only an exponentially relaxing signal if P(0)l]c-axis (Yaouanc et al. 1994).
It follows that the local field B~ at the #+ must be oriented parallel to the c-axis
(at T = 4.5 K, Bu = 0.0393(2) T). The relaxation behaviour and rates obtained
below and above Tc are partially very puzzling and have not yet found an adequate
quantitative explanation.

PrCo2Si2. This compound orders antiferromagnetically below TN = 30 K and shows


metamagnetic transitions at 1.2 T, 3.8 T, 6.7 T and at the saturation field 12.2 T if
an external field is applied parallel to the c-axis. In zero field two more transitions
in the AF-structure appear at 9 K and 17 K (Shigeoka et al. 1987, 1989a, 1989b,
Nojiri et al. 1992). ZF and LF-#SR measurements on a monocrystalline sample
with/5(0) and the applied field parallel to the c-axis were initiated with the goal to
monitor changes in the #+ spin lattice relaxation rate at the 12.2 T metamagnetic
transition, exploiting the possibility to combine pulsed magnetic fields with pulsed
#SR-spectroscopy at the pulsed muon source at KEK (Tsukuba, Japan) (Motokawa et
al. 1990, Nojiri et al. 1992). Indeed an enhanced (by a factor of ,.o 10) exponential
relaxation rate (~ 0.5 #s -1) compared to ZF-results at 4.2 K and 12 K was observed
for longitudinal fields varying from ,-~ 8 T to 13 T at effective temperatures < 12 K.
However, in view of the very large error bars, +(30-50)%, on the relaxation rates and
possible systematic uncertainties it is premature to draw some physics conclusions
from these results.
More reliable are ZF-results. From 50 K down to 4.2 K exponential relaxation
is observed without any loss of asymmetry. The temperature dependence of the
relaxation rate AzF is displayed in fig. 4.33. It shows a pronounced and rather wide

1.0
&-- 1
0.8

• 0.6 E++
c-
0
O 0.4
0 o
X

rr 0.2
oo
0.0 i t i i I
0 10 20 30 40 50
T e m p e r a t u r e (K)

Fig. 4.33. Temperature dependence of ZF relaxation rate in PrCo2Si2. The three different transitions
temperatures are indicated by the dashed vertical lines (from Nojiri et al. 1992).
MUON SPIN ROTATION SPECTROSCOPY 147

maximum at ,-~ 25 K with essentially no clear hints on the transitions at 9 K, 17 K


and 30 K. In particular no divergent type of behaviour is visible at TN. Nojiri et al.
(1992) speculate that this behaviour arises from strong spin fluctuations in the long
period AF structures between 30 K and 9 K, which can be viewed as a soliton like
motion of domain boundaries.

NdRh2Si2. Antiferromagnetic order sets in at 57 K and, differently from CeRh2Si2,


ferromagnetic (001) planes are coupled antiferromagnetically along the c-axis. ZF-
and LF-#SR measurements on monocrystalline samples yield the following results

÷,~~
I .... I ' ' ' ' 1 .... I ' ' ' ' 1 .... I .... I ' ' ' '

%-- 0.06 NdRhzSi2


-i- 0.05
0.04
0+ ÷

m 0.03
~s~ • 0 mT
o. 0.02
E
r~ 0.01
0.00 ÷~ e 200 mT

0 50 100 150 200 250 300


Temperature (K)
I . . . . I . . . . { . . . . I . . . . I . . . . I ' ' ' ' 1

l
. NdRh2Si2
ZE
zero field
v

: ~,s.
r"
.m

E 0.1
: G
a

k~°°° |o • • •
S S S l L I S : I * I t ' * ' I , * * ' I * * * L I ' * ' * I

0 50 100 150 200 250 300


Temperature (K)
Fig. 4.34. Temperature dependence of the ZF relaxation rate in NdRh2Si2. The upper part refers
to /5(0)llc-axis, the lower one to /5(0)_Lc-axis. The dashed line in the upper part below TN fol-
lows a T 2 In(T) dependence, pointing to a two magnon mechanism for the relaxation (from Gubbens
et al. 1994b).
148 A. SCHENCKand EN. GYGAX

(Gubbens et al. 1994b): ff/3(0) and Bext is applied along the c-axis an exponentially
relaxing, non oscillating signal is seen, the relaxation rate )H of which is essentially
independent of [/3extl, implying a dynamical origin of the relaxation. The data are
shown in the upper part of fig. 4.34. The temperature dependence of A1 below TN
seems to follow a T 2 ln(T)-law which points to a two magnon (Raman) process (Dal-
mas de R6otier and Yaouanc 1994b) as already indicated in the weakly anisotropic
ferromagnet GdNis (see section 4.2.4). However, no quantitative analysis of this be-
haviour has been presented so far. On the other hand if P(0) is perpendicular to the
c-axis the #SR signal is lost below TN, probably because the limited time resolution
at the pulsed muon source of ISIS prevented the observation of a fast oscillating pre-
cession component. Taken together with the nonvanishing, nonoscillating component
for/3(0)llc-axis it can be concluded that the spontaneous field at the #+ is oriented
parallel to the c-axis. The relaxation rate for fi(0)_l_c-axis above TN is displayed in
the lower part of fig. 4.34. In contrast to the other orientation a strongly divergent
behaviour is revealed which signals a slowing down of the spin dynamics as TN is
approached. Note that the rate close to TN is about two orders of magnitude larger
than for the other orientation. Combining the results from both orientations it can be
concluded that the fluctuating components are also essentially only present along the
c-axis pointing to longitudinal spin fluctuations along the same axis. Due to limited
precision in the temperature control the precise form of the critical slowing down
could not be determined.

4.2.6. Rare earth based heavy electron and related compounds


Intermetallics treated in this section possess quite different compositions, crystal
structures and magnetic properties but have in common certain low temperature
properties by which they are classified as heavy electron or heavy fermion compounds
(see, e.g., Steglich et al. 1993). Included are also some systems which show nearly
heavy fermion or Kondo system behaviour. The study of magnetic properties by
#SR spectroscopy has been reviewed in detail by Schenck (1993). Short reviews
have been presented by Amato (1994) and Heffner (1993). Table 4.11 lists all rare
earth based heavy electron and related compounds which have been studied by #SR-
spectroscopy up to date (1993). Quite unique information could be obtained which
was not available from other techniques. The most important pieces of information
are compiled in table 4.12.

CeAl2. This CeA12 moderately heavy electron compound possesses a rather complex
incommensurate and modulated antiferromagnetic structure to be described by a
multiple 0~arrangement (Shapiro et al. 1979). Interestingly the transition temperature
is found to vary for different samples and seems to cluster either around 3.4 K or
3.9 K, respectively (Gavilano et al. 1993) #SR measurements were performed on
both monocrystalline and polycrystaltine (powder) samples (Hartmann et al. 1989,
Ott et al. unpublished (1993)). TF-#SR measurements (Hext = 15 mT) on a single
crystal served primarily the purpose to determine the #+-site in this cubic Laves phase
compound which was identified as the (2-2) site (see fig. 4.7) as in the other rare
earth aluminides (see section 4.2.1) (Hartmann et al. 1989). #+-diffusion was shown
M U O N SPIN ROTATION SPECTROSCOPY 149

.,--t ..-t

0",

o
o
ii .r.
=o~ ~ ~: = <
0
o o'.,"-I.
ooo
I
~d

kO
" ~ ~ ~ oi

0
H
[e

e~

v~ v~
~-~
,I °e,i "-I,
oV/ vzS

-~~ g

~ 0 A ~.~ ~
0
0 ~
o ~o '~ ~

~'~ ~
,.~ 0 ~ II
o o [,,e o
~,~ m ~

H
~'~. ~.~.~
~o ~

o~
m "0

2
,-o

o
.,< .< r.j
o+ o
r..) ~ r~ 0
150 A. SCHENCK and EN. GYGAX

O~

,....4

0 Ob ,~

< <~

I
b~

C~
~d
m

~6
~ v~'

e-,

S ~.~ "~ - ~ ~ .
° o~ o: ~ _~ o~ ~~
•- ~ ~ ~

o
#-

o
r..,)
MUON SPIN ROTATION SPECTROSCOPY 151

I ' i L n I

0.20
ooo~O
Z"
"$ 0.15
E ° °
E
< 0.10

0.05 oe

i i r i i i i
3 4
Temperature (K)
Fig. 4.35. Temperature dependence of the paramagnetic and the magnetically ordered volume fractions
in monocrystalline CeA12 as reflected by the amplitudes (asymmetries) of the two components in the
TF-#SR signal below TN = 3.9 K (from Hartmann et al. 1989).

not to be important below ~ 80 K. The same measurements revealed the appearance


of a two component structure below 3.9 K one reflecting a paramagnetic behaviour,
the other one showing a very rapid damping rate reaching ~ 40 #s -1 at ~ 3 K.
The temperature dependence of the amplitudes of the two components is shown in
fig. 4.35. The rapidly damped component reflects the evolution of a wide static
field distribution, consistent with an incommensurate, modulated antiferromagnetic
structure. But, as fig. 4.35 implies, this order does not involve at once the total
volume of the sample, but the ordered volume grows gradually from zero at
3.9 K to ~ 80% of the total volume at ~ 3.2 K. Is this unexpected behaviour
somehow related to the observation of two different transition temperatures? Zero
field measurements on a powder sample yielded essentially the same picture (Ott et
al. 1993). The latter sample was also investigated by 27A1-NQR (Gavilano et al.
1993). No mentioning of a two component behaviour is found in this paper.
The two component structure of the #SR signal in CeA12 around TN is not unique
but can be found in several other rare earth based heavy fermion compounds (e.g.,
CeA13, CeCu2Si2, YbBiPt etc.) giving the impression that it may be a characteristic
feature of this class of materials.

CeA13. Long thought to be paramagnetic down to at least 60 m K (Murani et al.


1980) it came as a real surprise that this archetypical representative of the heavy
fermion systems revealed the presence of an ordered state below ~ 1 K in ZF-
#SR measurements (Barth et al. 1987, 1989a). Similar to CeA12 a two (or rather
three) component structure developed below 1 K, but unlike in CeA12, the additional
152 A. SCHENCK and F.N. GYGAX

0.05
0.00 ~\ + -f,,, , -...,/ - - : ~ - " ~ ' ~ - - ¢ ~ ' ~ " ~ -'--'--~"7

'$ -0.05
E
E -0.10
< -o15
-0.20 a
I I I I I I
-o.%1o 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Time (psec)
0.05
0.00
"-"~"~"~" .. - - - 2 - ~
-0.05
E
E -O.lO
69
,< -0.15
-0.20 b
-0.25 I I I I I I I
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Time (psec)
Fig. 4.36. ZF-#SR signal in CeAI3 at 0.01 K (upper part) and at 0.2 K (lower part). It can be fitted by
a sum of three components: one fast relaxing but nonoscillatingwith amplitude Af, one slowly relaxing
and nonoscillating with amplitude As and an oscillating component with essentially a node near zero
time (from Amato et al. 1994a).

component displays a precession pattern corresponding to a local field of -,~ 22 mT


rather than a wide field distribution with zero average. Figure 4.37 displays the
temperature dependence of the amplitudes of the two signals, derived from more
recent ZF-measurements (Amato et al. 1994a). To fit the data in fig. 4.36 actually a
three component structure had to be adopted consisting of one oscillating, one very
fast relaxing and one slowly relaxing component. The oscillating component showed
an unphysically large phase shift (ignored in Barth et al. 1987, 1989a) which was
later attributed to a precursor state manifest by the fast relaxing component (Amato
et al. 1994a). It now seems that the sum of a fast relaxing and of a phase shifted
oscillating components is under certain circumstances more representative of a spin
density wave phenomenon (see Le et al. 1994, and section 7.1). The amplitude in
fig. 4.37 associated with the ordered state signal represents the sum of the amplitude
of the fast relaxing and of the oscillating components extrapolated to zero time.
There is another important difference to CeA12. While the phase transition in CeA12
can also be monitored by the specific heat, which shows the characteristic cusp at TN,
no such feature is seen in CeA13. Hence it seems that the magnetic order in CeA13
evolves in a non cooperative manner. This may be also indicated by the precession
frequency which shows very little variation up to the temperature where the ordered
volume vanishes (see fig. 4.38).
MUON SPIN ROTATION SPECTROSCOPY 153

25
' 7 ' ' ' '
Amax /
//

20 /
/

15
"As
o:l

E 10
E
<

\
O-

I I I I I
0 0.1 0.2 0.3 0.4 0.5 0.6
Temperature (K)
Fig. 4.37. Temperature dependence of the zero time amplitudes (asymmetries) Af and As in CeAI3,
Af reflects the volume fraction associated with magnetic order and As the (complementary) volume
fraction remaining paramagnetic (from Amato et al. 1994a).

3.5
3.0 3-O-O
-I- 2.5
2.0
'-
i1) 1.5

1.0
LI.
0.5
0.0 ~ i I
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Temperature (K)
Fig. 4.38. Temperature dependence of the spontaneous precession frequency in CeA13. The solid line
is intended to guide the eye only (from Amato et al. 1994a).

S o m e o f the o b s e r v e d features are s a m p l e dependent. The first a p p e a r a n c e o f a


m a g n e t i c a l l y o r d e r e d v o l u m e m a y vary f r o m 2 K d o w n to 0.6 K. In Barth et al.
(1987, 1989a) it is shown that the m a g n e t i c order m a y be r a n d o m or spin glass like
a b o v e 0.8 K up to 2 K, no such feature is seen in A m a t o et al. (1984a). A t the
154 A. SCHENCKand EN. GYGAX

lowest temperature Barth et al. (1987, 1989a) find nearly no persisting paramagnetic
volume while Amato et al. report on a residual paramagnetic volume of ,,o 25% (see
fig. 4.38). NMR-measurements have provided conflicting results on the presence or
non-presence of magnetic order in CeA13 (Nakamura et al. 1988, Wong and Clark
1992). Later NMR studies by Ott et al. (1993) (priv. communication) seem to be
consistent with the #SR results. From the local field of 22 mT at the #+, considering
only dipolar fields and ignoring a possible contact hyperfine contribution, an upper
limit of ~ 0.1 I#B was estimated for the ordered moment (Schenck 1993).

CePb3. TF-#SR measurements on a polycrystalline sample at 10 K revealed a


split four-component signal with distinct Knight shifts ranging from -4200 ppm to
+1000 ppm (Uemura et al 1986a). No further results are available.

CeCus, Ce(Cul_xAIx)5, Ce(Cul_xGax)5. The system CeCu5 is not considered to be


a heavy fermion compound according to a commonly adopted classification scheme,
although its low temperature Sommerfeld constant 7 is nearly the same as for CeA12.
Replacing some of the Cu by A1 or Ga (only the 3g-sites in the hexagonal CaCu5
structure are involved) the system seems to acquire heavy fermion properties as
reflected by vastly increased 7 constants which seem to peak for z around 0.20 (Bauer
1991). Undoped CeCu5 orders antiferromagnetically below ,-~ 4 K (propagation
vector ( = (0, 0, 1/2), ~orallc-axis), with rising z a suppression of TN is observed
and for z > 0.2 long range order seems to be suppressed (Bauer 1991). The question
was whether some type of short range or spin glass order could persist though.
To answer this question ZF-#SR was applied to undoped CeCu5 and to doped
compounds with z (A1, Ga) = 0.1 and 0.2 (all polycrystalline) (Wiesinger et al.
1994a, 1994b and priv. communication). In CeCu5 a spontaneous precession signal
appeared below TN in an increasing part of the volume, however, never accounting for
more than ~ 30% of the total volume, as can be deduced from the signal amplitude
displayed in fig. 4.39 (insert). This figure shows also the temperature dependence
of the spontaneous precession frequency. The non-oscillating component shows
appreciable exponential relaxation down to the lowest temperatures (A _~ 2.2 #s -1 at
1 K). The observed behaviour seems to be incompatible with the magnetic structure
indicated above, since for that structure only one signal is predicted with Bu = 0
for the #+ at the b-site, which site is deduced from the ZF Kubo-Toyabe signal
observed above TN. The temperature dependence of the two signal amplitudes,
moreover, suggest that two different kinds of magnetic domains could coexist like
in CeA12 and CeA13.
Preliminary data on the doped compounds show that the oscillating signal, which
indicates long range order, has disappeared (or nearly disappeared in the sample with
z = 0.1) and is replaced below ~ 2 K by a wide static field distribution of Gaussian
shape, which now accounts for 75% of the total volume (Wiesinger et al. 1994b,
priv. communication). The rms width of the field distribution amounts to ~ 23 mT
which is about half the field resulting from the spontaneous precession frequency in
CeCus. The data reveal clearly that indeed some spin glass like order is present even
in compounds with z = 0.2. This raises the possibility that the enormous "),-values
MUON SPIN ROTATIONSPECTROSCOPY 155

G O

6 ° ' ' C 'e C u 5


o
~ 0 0
0
30 , , ,

~
°
~
4 ~
2O
i
~ o
oc.c io~
oc ~
,~ °° o
q)

0 i i i I ~V
0 1 2 3 4 5

Temperature (K)
Fig. 4.39. Temperature dependence of the spontaneous precession frequency in CeCu5. The insert
shows the behaviour of the amplitudes. Open circles refer to the precessing component, closed circles
to the nonprecessionbut rapidly relaxing (),ZF ~ 2.2 #s for T --+ 0 K) component(from Wiesinger et al.
1994a).

observed around z = 0.2 are of magnetic origin and arise from a high density of
excited states at small energies which are known to exist in spin glasses.

Cefu6. This is one of the few remaining heavy fermion compound which have not
shown any hints for some type of order at low temperatures. ZF-#SR measurements
performed on a single crystal down to 40 mK likewise did not reveal any indication
for static magnetic order placing an upper limit of ~ 10-3#B on any ordered moment
if such should exist (Amato et al. 1993c).

CeB6. This cubic compound shows antiferroquadrupolar ordering below TQ = 3.2 K


and antiferromagnetic order below TN = 2.3 K, described by a double 0" structure
(ql = (1/4, 1/4, 1/2), 42 = ( 1 / 4 , - 1 / 4 , 1/2)) (Effantin et al. 1985). A possible
temperature dependence was not studied. In the quadrupolar phase antiferromagnetic
order can be induced by an external field. Concerning the structure of the induced
order controversial results are obtained from UB-NMR (Takigawa et al. 1983a,
Kawakami et al. 1982) and neutron scattering (Efffintin et al. 1985; Erkelens et al.
1987).
TF-#SR measurements on monocrystalline samples above TQ proved unambigu-
ously that the #+ is located at the site (1/2, 0, 0) (or equivalently at the sites
(0, 1/2, 0) and (0, 0, 1/2)) (see fig. 4.40) (Feyerherm et al. 1994a). ZF-measu-
rements below TN displayed a very complex signal with 8 different spontaneous
precession frequencies ranging from ~ 2.3 MHz to ~ 77 MHz as can be seen from
156 A. SCHENCK and F.N. GYGAX

(~)Ce 0 B

Fig. 4.40. Crystal structure of cubic CeB6. The/z+-site is indicated.

a Fourier-transform of the ZF-#SR signal at 10 mK (fig. 4.41). The temperature


dependencies of the spontaneous frequencies are shown in fig. 4.42. Their partially
very strange behaviour could not be explained yet. It may point to a temperature
dependent reorientation of some of the Ce-moments. On the basis of the magnetic
structure suggested by Effantin et al. (1985) dipolar field calculations predict no
more than seven different fields, associated with the sites (1/2, 0, 0), (0, 1/2, 0)
and (0, 0, 1/2), out of which six appear as slightly splitted triplets (sites (1/2, 0, 0),
(0, 1/2, 0)). The calculated fields (a contact contribution is included) range from
30 mT to 90 mT (or 4.06 MHz to 12.2 MHz), far lower than the maximum observed
field of ,-~ 0.57 T (Feyerherm et al. (1994a) and to be published). In particular the
high frequency components are principally not explainable if the ordered moment
is as small as 0.28#B as claimed by Effantin et al. (1985). It must be concluded
that the AF-magnetic structure in CeB6 as well as the size of the ordered moments
are drastically different from the neutron scattering conclusions or that the pres-

100 " i

100[
i i

,~,
i i i

=,

r,,,,. = 60mK
ILl 60l T
3= 6O
13.

0 2 4 6 8 10
~ 2O

0
0 40 60 ' 80
FREQUENCY (NHz)
Fig. 4.41. Fourier power spectra of the ZF-/zSR signal in CeB6 below TN (from Feyerherm et al.
1994a).
MUON SPIN ROTATION SPECTROSCOPY 157

80 ¸
• 0 @ @0@ @ O
@ 0@0@O @O
70 @0
O
@
• &AAAA AA
• A O
50
A
40,oe ooo , o O
O OOO00 OOo0 •
A
30 0

I 20.Pure mare un II D N D N I - I unDID [] •

10 []
O
C
CD
D CeB 6
Cr
CD 8.~<~ @ spontaneous muon
m •L z ~ %, precession in
6 • o zero magnetic field

IOO OOi O O
~ O@0@0 OO@ O
4' • O
O @ O0~
m D m ~ D mmDmD [] m D l ~ D
2 ~ m mt~u N D
A A~i~

.... , , . . . . . . . . . . . *
0.0 0.5 1.0 1.5 2.0 2.5
Temperature (K)
Fig. 4.42. Temperature dependence of the eight spontaneous precession frequencies in CeB6 below TN.
Open symbols: /5(0)11(1, 1, 0), filled symbols: /5(0)11(1, O, O) (from FeYerherm et al. 1994a).

ence of the #+ alters the magnetic structure locally. However, even when admitting
the latter possibility, the appearance of more than seven frequencies is difficult to
understand.
Knight shift measurements in the quadrupolar phase are also at odds with the
suggested induced AF-structure in this regime in that no further splitting of the two
signals, seen above TQ, is observed below TQ. The observation of two signals above
TQ reflects the fact that the three sites (1/2, 0, 0), (0, 1/2, 0) and (0, 0, 1/2) which
are crystallographically equivalent, are magnetically inequivalent in the paramgnetic
state (Feyerherm et al. 1994a). Despite considerable efforts to understand CeB6 it
remains a mysterious system and is a challenge for further studies.

CeCu2Si2. This is probably the most famous and probably the most thoroughly in-
vestigated heavy fermion compound due to the fact that it was the first HF system
displaying superconductivity. It was known that some kind of AF-magnetic order de-
velops in nonsuperconducting Cu-deficient CeCuz_xSi2 or in Cel_~LaxCuzSi2, but
no magnetic order was found in superconducting (Cu-enriched) CeCuz+xSi2 (Grewe
and Steglich 1991). It therefore was a surprise when ZF-#SR measurements on a su-
perconducting CeCu2.1Si2 sample revealed the onset of some random magnetic order
below 0.8 K, i.e. at a temperature above the superconducting transition temperature
158 A. SCHENCK and EN. GYGAX

0.8 ~, v__-~ "~x! CeCu2 Si2


-

0.4 SC "\ "


~ \ o \

"./ I"° \"° \"


0 2 4 6 8
B (Tesla)
Fig. 4.43. Magnetic and superconducting phase diagram of CeCu2Si 2. The solid lines represent He2 for
HextHa-axis and llc-axis, respectively, the lower dashed curve reflects anomalies in the elastic constant
Cll, the upper dashed curve separates a not well identified ordered magnet phase from the pararnagnetic
phase (adapted from Steglich et al. 1991).

Tc ~ 0.6 K (Uemura et al. 1989). Since then NMR and other measurements have
provided indications for a magnetic phase diagram in the temperature versus field
representation as shown in fig. 4.43. For small applied fields the phase boundary
is close to Tc somewhat inconsistent with the ZF-#SR result. An important ques-
tion, never answered satisfactorily before, concernes the nature of the coexistence
of magnetic order and superconductivity: is it carried by the same electrons, is it
microscopic or macroscopic? ZF-measurements on a CeCu2.2Si2 sample (Luke et
al. 1989b, 1994) and on a CeCu2.05Si2 sample (see Amato 1994, Feyerherm et al.
1995) seem to show that magnetic order and superconductivity do not coexist on a
microscopic scale, that they are even mutually exclusive. The ZF-#SR signal starts
to develop a two component structure as the temperature is lowered below ,-~ 1.2 K.
Above this temperature relaxations is only induced by random dipolar fields from
the Cu-nuclear moments (Kubo-Toyabe signal). One of the two components be-
low 1.2 K shows a fast Gaussian decay with a strongly rising cr as the temperature
decreases (fig. 4.44). The other component show basically the same behaviour as
above 1.2 K with a somewhat enhanced relaxation rate at around To. The relaxation
rate of the fast component corresponds to a static field spread of about 140 G at
the lowest temperature. Assuming truly random order and considering all possible
and reasonable interstitial sites an upper limit of 0.25#B is estimated for the ordered
moment. The #+-site in CeCu2Si2 is not known. From the splitting of the TF-signal
above 1.2 K in a polycrystalline sample it may be concluded that two different
sites are involved (unpublished results). Most interesting is the behaviour of the
amplitudes. As fig. 4.45 shows the ordered volume fraction, as measured by the
amplitude of the fast signal, increases to about 70% at T~. Just below Tc it decreases
again, signaling a depression in the ordered volume as superconductivity sets in. Su-
perconductivity seems to be confined to the volume fraction not showing magnetic
order, but this needs to be studied in more detail. There is essentially no difference
MUON SPIN ROTATIONSPECTROSCOPY 159
4 i i I r i i i i p i i i i i

CeCu2osSi 2
121"~ ZF "
lO I
81
3 6:
4.


C
0 0.2 0.4 0.6 0.8 t.0 t.2 1.4
r (K)
Fig. 4.44. Temperature dependence of the Gaussian decay constant of the fast relaxing component
in CeCu2.05Si2 (from Amato 1994).

1.0
CeCu2.05Si2
0.8 pm

0.6

<E 0.4

0.0 i i i i i i i i i 1 i
0.0 0.2 0.4 0.6 0.8 1.0 1.2
Temperature (K)
Fig. 4.45. Temperature dependence of the amplitudes of the two components in the ZF-/zSR signal in
CeCu2.osSi2. A1 (open symbols) and A2 (filled symbols) reflect the paramagnetic and the magnetically
ordered volume fractions, respectively. The latter is identified by the fast Gaussian decay (fig. 4.44)
(from Amato 1994).

in the temperature dependence of the two amplitudes measured in CeCu2.2Si2 and


CeCu2.05Si2.
The appearance of two magnetically inequivalent domains in CeCu2Si2 resembles
to a great extend the results obtained in CeAI2, CeA13, Ce(CUl_x(Ga, A1)=)5 as shown
above. Once more one gets the impression that this behaviour is characteristic for
this class of compounds.
160 A. SCHENCK and EN. GYGAX

CeRu2Si2. Conventional measurements did not reveal any hint for magnetic order
in this compound. However, a metamagnetic like transition at Hext -~ 8 T for
T < 10 K suggests that CeRu2Si2 is close to a magnetic instability (Haen et al. 1987).
Inelastic neutron scattering show indeed (also for CeCu6 !) the evolution of intersite
antiferromagnetic correlations below 60 K (Rossat-Mignod et al. 1988b). It is argued
that quantum fluctuations prevent the thermal divergence of the correlation length
thereby preventing long range static order. This picture was changed by ZF- and LF-
#SR measurements (Amato et al. 1993a, 1994b, Amato 1994). These measurements
on two different monocrystalline samples revealed the onset of some static order at
around 2 K (see fig. 4.46). The only other signature for a phase transition is seen
in the thermoelectric power S which displays a minimum at ,-~ 2 K (see fig. 4.46).

Ill, I i i i i Illl| I I I I IIII I I I I I IIII I I 0.03

,1, ~ S.C. 2
0.02
'1/)

v
=L CeRu2S'2 _.~_~_~_@~ ......
ZF
II)
0.03 0.01
¢-

.o
0.02
rr
0.01
a

0 ml I l I II I , , i ,,,,

1.5 f,,,, l t illl l i I I I IIII i i i i i loll i i

AT_Lc
i

1.0 ~-
0
o 0 @ -
@ 0
0.5 0
o
- b
o
o %° i
0 ,,,I , , ,,i,I , , iJl,,,I ~ ' t'~'l'l I I

10 .2 10 1 10 ° 101
T (K)
Fig. 4.46. Temperaturedependence of (a) the ZF-relaxationrate and (b) the thermoelectricpower S in
CeRu2Si2 (from Amato 1994).
MUON SPIN ROTATION SPECTROSCOPY 161

This feature was previously attributed to the onset of a coherent regime in a Kondo
lattice (opening of a pseudo gap (Martin 1982)) but this interpretation may now need
revision. The width of the static internal field distribution at the #+ in the ordered
state is extremely small and of the order of 0.02 mT. From this an ordered moment of
I0-3#B is estimated (upper limit 0.7 x 10-3/ZB). This is the smallest moment ever
encountered in an ordered state. The #+-site in CeRu2Si2 is known from the angular
dependence of the Knight shift measured at higher temperatures in a transverse field
(Amato, priv. communication and to be published). It is the (1/2, 1/2, 0) position
(site No. 1 in fig. 4.32). Note that measurements in the non magnetic isostructural
compounds Y(Co)2Si2 and LaPd2P2 did not favour this site.
TF-field measurements in a very small applied field (7.4 mT) served to monitor
the #+ spin lattice relaxation which is induced by the 4f-electron spin fluctuations
(Amato, priv. communic.). The TF-relaxation rate A2 = l/T2 is mainly reflecting a
1/Tl-mechanism since static line broadening is practically absent due to the absence
of sizable nuclear moments (a small correction for those is applied) and the small
applied field. The relaxation time T2 is displayed in fig. 4.47 together with the
onsite (Fss) and inter site (/]s) line width from inelastic neutron scattering (Fss is
proportional to the single site fluctuation rate and /is to the correlated inter site
fluctuation rate) (Regnault et al. 1990). Since the fluctuations are fast we have

T2~- T1 oc F. (4.16)

As can be seen T2 scales closely with /'ss as expected. The proportionality factor
between T2 and /'ss can be used to extract a rough estimate on the size of the

3.5 ~ 150

3.0 C e R u 2 Si2

2.s , 1 - loo
"-" 2.0 ~_~L,k_.,L-i,JS..~
' ' ' ~ ' ~ I ~"
.-.s

L.- 1.5 I--~

1.0 50

-is
0.5 -

0.0 , I i I , 1 i I i 0
0 20 40 60 80 100
T (K)
Fig. 4.47. Temperaturedependenceof the/z+ TF relaxation time 7"2(open circles) and the on site (Fss)
and intersite (/]s) line width from inelasticneutron scatteringin CeRu2Si2 (from Schenck 1993).
162 A. SCHENCK and EN. GYGAX

fluctuating moment: ~ I#B. The ordered moment at low temperatures is extremely


reduced compared to this value. Since such small moments are found in other
heavy electron systems (UPt3, Ul_xThxBel3, UCus, see section 4.3.2) it seems that
the establishment of very small moment magnetic order (random or long range) is
another characteristic feature of heavy fermion systems. Note that generally almost
all ordered moments are much reduced in comparison to the Hund's rule moment of
Ce3+: # = 2.54#B. The system CeRu2Si2 shows more conventional AF-magnetic
order when other elements are placed on the Ce, Ru or Si sites. For instance
Ce(RUl_~Rhx)2Si2 has shown AF like transition for 0.05 < z < 0.3 by means of
specific heat and susceptibility measurements (Sakakibara et al. 1992). For z = 0.15
a transition temperature of TN --~ 5.5 K is estimated. ZF-#SR measurements on
polycrystalline Ce(Ru0.asRh0.15)2Si2 confirmed the onset of magnetic order below
5.5 K, which appears to be random in nature as deduced from the Gaussian character
of the relaxation behaviour below 5.5 K (Murayama et al. 1993). However, as found
in other systems only a fraction of the volume takes part in the order. From the static
field spread of N 15 m T at 4 K the ordered moment is estimated to be ~ 0.2#B.

CeTSn (T = Ni, Pd, Pt). These orthorhombic crystallographically rather complex


compounds are viewed as Kondo lattice systems with no heavy fermion properties.
However substitution of a few % Cu for Ni transforms the system CeNiSn to a HE
compound. CeNiSn is a Kondo semiconductor with as small pseudo gap ( ~ 14 K) at
the Fermi surface which is thought to arise from AF-magnetic correlations between
quasi particles. In high magnetic fields ( ~ 12 T) the semiconducting behaviour
disappears and gap formation is suppressed (Takabatake et al. 1992). Various exper-
imental results (see, e.g., NMR results of Kyogaku et al. 1992) hint at a magnetic
phase transition at low temperatures (< 0.2 K) but an actual phase transition could
not be established. In contrast CePdSn and CePtSn are Kondo metals which display
antiferromagnetic order below N 8 K, which for both compounds shows an incom-
mensurate, modulated structure according to neutron results (Kohgi et al. 1992,
Kadowaki et al. 1993).
TF-#SR measurements on monocrystalline CeNiSn reveal an unmeasurably small
relaxation rate above 2.5 K. Below this temperature the relaxation rises in a pro-
nounced manner but remains small (A2 "-~ 0.065 #s -1 at 33 mK). The observed
temperature dependence, following a power law A2 e( T -1/3, is not untypical for a
system approaching a phase transition but down to 33 mK no phase transition was
encountered (Kratzer et al. 1992). It can be concluded that CeNiSn is at all temper-
atures (i.e. down to 33 inK) a fast spin fluctuator. This may be correlated with the
other notion that CeNiSn is a valence fluctuating system, as has been deduced from
susceptibility data (Takabatake et al. 1990).
ZF measurements in CePdSn and CePtSn reveal the onset of spontaneous coherent
spin precession below 7.5 K and 8.2 K, respectively (Kalvius et al. 1994). In CePdSn
the #SR signal displays clearly two well resolved frequencies and the relaxation is
relatively slow implying relatively narrow field distributions at the #+-sites. The
latter are not known yet. This observation of narrow and distinct field distributions
MUON SPIN ROTATION SPECTROSCOPY 163

0.2
a

20G LF -J i
0.2
2.. .... ......... ,,,. t
-" . . . . "' ' " '~ ' t '"t ~ t 0.1

0.1

CePtSn 9K
CePtSn 7K ZF

0 2 4 6 8 0 I 2 3

~a~ 0.2 CePtSn 8.2K b


200G LF O. I

0.1 ~ ~ 0

lOG LF CePtSn 4.5K ZF


, i ~ I , I , I ~ -0. I I i I I t

o 0.4 o.a t.2 1.8 2 o 0.5 I

Time (#s)
Fig. 4.48. ZF-#SR signal in CePtSn at various temperatures reflecting the appearance of static random
order (b) and coherent long range order (c and d). The static order at 8.2 K is verified by the decoupling
in longitudinal fields (b) (from Kalvius et al. 1994).

appear to be inconsistent with an incommensurate modulated spin structure which is


expected to produce rather wide field distributions.
CePtSn displays a rather unexpected behaviour just above the temperature where
coherent spin precession first appears. As can be seen from fig. 4.48 the ZF-signal
changes from a weak decay above 8.5 K to a static Kubo-Toyabe behaviour below
8.5 K, reflecting the development of a random static field distribution with a width
of about 45 G (Kalvius et al. 1994), followed by the appearance of the coherent spin
precession pattern just below 8.2 K. Obviously long range order is preceeded by a
precursor spin glass like state. It could imply that the fast paramagnetic fluctuations
above 8.5 K are slowed down to rates below 1 MHz, creating a quasi static situation
with respect to the #SR time window. A further change in the coherent precession
pattern is observed below ~ 5 K (fig. 4.48d), indicating a transition to a different
magnetic structure in agreement with neutron scattering results (Kadowaki et al.
1993). Again there is a problem to reconcile the appearance of distinct and relatively
narrowly peaked frequency distributions with the claimed incommensurate structures
(also below 5 K in CePtSn).

YbBiPt, Ybo.sYo.sBiPt. YbBiPt possesses a record high Sommerfeld constant of 7 -~


8 J/(mol-Yb K 2) at low temperatures which has lead to quite some interest in this
164 A. SCHENCKand EN. GYGAX

cubic compound. Specific heat and susceptibility measurements suggest some kind
of phase transition at ~ 0.4 K (Fisk et al. 1991). Recent resistivity measurements
suggest a partial gapping of the Fermi surface at 0.4 K and the appearance of a spin
density wave (SDW). The opening up of a gap can be suppressed in fields >0.3 T
(Movshovich et al. 1994).
ZF-, TF- and LF-#SR measurements were performed on crushed powder and
crystalline samples (Amato et al. 1992b, Amato et al. 1993b, Heffner et al. 1994).
The ZF-measurements identified the transition at 0.4 K to be a magnetic phase
transition. Since no precession pattern was seen but only a wide Gaussian like
distribution of fields (~ 8 mT) with zero average it was concluded that a spin
glass like random order or some complicated incommensurate structure must be
established again in a spatially inhomogeneous way. Some information on the #+ site
was obtained from TF-measurements at higher temperature, which indicated that the
#+ is located somewhat off center at the position (1/2, 1/2, 1/2) (Feyerherm, priv.
communication). Adopting this site and the random order picture an ordered moment
of ~ 0.1#B is estimated. This is much smaller than the moment following from the
Curie behaviour of the susceptibility below 10 K (3.6#B). However, the small value
of the ordered moment is consistent with the absence of a nuclear Schottky anomaly
in the low temperature specific heat (Thompson et al. 1993).
Like in other heavy fermion systems the ZF-signal showed a two component
structure which developed already much above 0.4 K, i.e. below ~ 1.2 K. Above
0.4 K both components showed an exponential relaxation behaviour. At 0.4 K one
of the two components changed abruptly to a Gaussian behaviour with a rapidly
increasing decay rate, while the other component continued to show exponential
relaxation (see fig. 4.49) just as in the paramagnetic phase. Interestingly the Gaussian
component changed in appearance, below ~ 0.1 K, indicating a change in magnetic
structure or spin dynamics (Heffner et al. 1994) consistent with similar evidence
from susceptibility and specific heat data (Thompson et al. 1993). The Gaussian field
distribution did not change in longitudinal fields up to 2 T. This is quite significant
since it proves that the magnetic order, sensed by the #+, cannot be associated with
the SDW carried by the conduction electrons, which is destroyed in fields above
0.3 T. It therefore seems that the magnetism in YbBiPt is of twofold origin: Yb-
local moment order and a conduction electron SDW, implying that the electronic
ground state is made up of rather independent electronic subsystems. In this respect
there is a certain parallelity with the results from UPd2A13 and UCu5 (see section
4.3.2).
ZF-measurements in Yb0.sYo.sBiPt lead to rather similar results. Interestingly,
despite a 50% dilution of the Yb-moments, the phase transition or spin freezing
temperature was very little reduced. The dilution effect was clearly seen in a corre-
sponding decrease of the static field spread (Amato et al. 1993b).

Sm3Se4. Low temperature susceptibility measurements revealed a huge linear specific


heat term with 3' = 0.79 J/(K 2 mol-Sm3+) despite the fact that Sm3Se4 shows semi-
conducting behaviour with no free carriers at low temperatures (Fraas et al. 1992).
This system has been discussed within the context of "heavy electrons without free
MUON SPIN ROTATION SPECTROSCOPY 165

8 , I I I I I I I I I I

7
6
4" s o
,.~ 4 []
3
2
1
0 I I I I r I I I I I
0.6 I I I I I I I I I I

0.5 b

r " 0.4
0.3
,..=
0.2
0.1
0.0 I I I I I IIII II
I l t l I I

10080I
lOi i IoDA~A,,Ii c
o o
o
o o
6o 0°°~

40 ° 0 o o
2O 0 []
0/ I I I I I 1 1 1 1 I
0.0 0.2 0.4 0.6 0.8 1.0 1.2
T (K)
Fig. 4.49. Temperature dependence of the amplitudes and the relaxation rates from a two component
analysis of the ZF-/zSR signal in crystalline YbBiPt. The rapidly relaxing signal (Af, ¢rf, Af) changes
from an exponential behaviour (Af) to a Gaussian behaviour 0rf) at 0.5 K. The slowly relaxing signal
(As) displays an exponential behaviour with only a weakly temperature dependent ~,s (from Amato
et al. 1993b).

carriers". Sm appears in this compound in the two valence states Sm2+ and Sm 3+
with a relative occupation of 1 : 2, and no charge ordering. Relatively slow valence
fluctuations are observed above 100 K (Takagi et al. 1992, 1993). Nothing was
known about the low temperature magnetic properties of this compound.
ZF-#SR measurements on a polycrystalline sample revealed the development of
static magnetic order below 20 K by the appearance of a fast damped Gaussian
component, again in only part of the volume (see fig. 4.50). The low temperature
static field spread seen by the #+ amounts to ~ 3.5 mT, from which, making some
reasonable assumption on the #+ site, a randomly ordered moment of 0.05#B can be
estimated (Takagi et al. 1993 and unpublished results). A spin glass type structure of
randomly frozen moments (perhaps related to the statistical distribution of Sm 3+ ions
166 A. SCHENCK and EN. GYGAX

3.5 I i illllh I i i lll,,q i , Ulllll I

"7 3.0~ + + + ~ ZF
Sm3Se4
=k
v 2.5 o Gauss.
• Exp.
tr 2.0
t-
O
1.5 '~ ,~ DLF6kOe i
X
m 1.0
rr
0.5 • oOdpe oo • ~ 0 m~
0 i 0 i Illlll F I I IIIIII I I I IIllll l

100 I ' ' 'J'"l ' ' ' ''H'l ;I-'.-,~'"-U-~ "-
,
i
80 i
t

60 _o_e_o ~ . . . . . oe'"
"o
.-_=
i

~. 40 .... %.
E
,< "if,
20

0 -1 I ,,,,,,I i I iiiiiii i i iiiiill

10 100 10~ 102


T (K)
Fig. 4.50. Temperaturedependence of ZF relaxation rates (exponential, Gaussian) and amplitudes (asym-
metries) from a two component analysis of the ZF-#SR signal in Sm3Se4 below ~ 20 K. Above 20 K
only one component, decaying exponentially, is seen which is unchanged in a longitudinal field (from
Takagi et al. 1993 and unpublished results).

on the Sm sublattice) may lead to a high density of low energy magnetic excitations
- as in a conventional spin glass - and could explain the large linear specific heat
term, as discussed in the case of YbBiPt.

4.2.7. Ternary rhodium borides and chevrel phase compounds


The interest in the ternary compounds RRh4B4 (R = rare earth) stems from their
property to show superconductivity in coexistence with AF magnetic order as in
SmRh4B4 or reentrant ferromagnetism as in HoRh4B 4 and ErRh4B 4 (see, e.g., Maple
et al. 1982, Kakani and Upadhyaya 1988). The Chevrel phase systems RMo6S8
(RMo6Se8) show coexisting AF-magnetic order and superconductivity for R = Nd,
Gd to Er, Yb (Fischer 1978). The situation in EuMo6S8 and EuMo6Se6 is contro-
versial. Recent investigations claim the establishment of some complex order below
0.4 in EuMo6S8 and below 0.85 K in EuMo6S8 (Takabatake et al. 1984, Quezel
et al. 1984b). A structural phase transition from rhombohedral at high tempera-
tures to trictinic a low temperatures was detected at ~ 110 K in EuMo6S6 (Baillif
et al. 1981). At the same temperature a change from a metallic to a nonmetallic
behaviour was found in resistivity measurements (Meul et al. 1982). The system
Eu]_~SnxMo6S8 (z > 0.2) shows again superconductivity. No structural phase tran-
MUON SPIN ROTATIONSPECTROSCOPY 167

sition is seen in this pseudo ternary compound and the rhombohedral structure is
observed at all temperatures which is apparently a prerequisit for the development
of superconductivity. It has attracted particular attention because of the observa-
tion of field induced superconductivity due to the Jaccarino-Peter effect (Meul et al.
1984).

Rare earth rhodium borides. TF and ZF-#SR measurements on the nonmagnetic


homologues LaRh4B 4 (Boekema et al. 1982) and YRh4B4 (Huang et al. 1983b)
showed that the #+ is immobile up to about 200 K. Possible #+ sites were discussed
in detail by Noakes et al. 1987. Measurements on the magnetic compounds yield
the following results.

SmRh4B4, ErRh4B4. #+ relaxation observed in ZF and TF displayed a pronounced


temperature dependence (Noakes et al. 1987, 1985) from which the rare earth mo-
ment fluctuation rates could be determined. Although the Er 3+ moment is consid-
erably larger than the Sm 3+ moment the fluctuation rates are very similar in their
magnitudes and temperature dependencies (as an example see fig. 4.51). No effect
of the superconducting transition on the fluctuation rates are seen. Since the #+ re-
laxation rates became excessively large in ErRh4B4 when lowering the temperature
below 50 K the region around the onset of ferromagnetism could not be studied.
ZF-measurements below TN in SmRh4B4 did not reveal any precession signal. In-
stead the data could be best fitted with two exponentially relaxing components, one
.-(1) --~ 10 #S -1) and the other one of approxi-
displaying a very fast relaxation rate (Zzv
mately equal amplitude showing a rather slow relaxation (x(2) ~-"ZF
~ 0.17 #s -1) (Huang
et al. 1983b). It is argued that this most likely reflects an occupation of different
sites with rather different magnetic environments in the ordered state. In any case
the absence of coherent spin precession and the appearance of exponentially damped

I I I I

6
o g"J
0

-~ 4
rr
._o

¢3

II

0 I I i I
0 50 1O0 150 200 250
Temperature (K)
Fig. 4.51. Temperature dependence of the Sm 4f-moment fluctuationrate in SmRh4B4. The solid line
is a fit to a model which includes the RKKY, the Korringa mechanism and phonon induced transitions
between CEF levels (from Noakes et al. 1987).
168 A. SCHENCKand EN. GYGAX

signals points to a rather complicated antiferromagnetic structure such as has been


seen in NdRh4B 4 and TmRh4B4 (Majkrzak et al. 1982, 1983). There is no other
information on the AF structure in SmRh4B4. It also needs to be checked to what
extend the #SR-data below TN still reflect dynamic features.
The spin relaxation rates, shown in fig. 4.51, were explained in terms of phonon
induced transitions between the CEF levels of both Er 3+, Sm 3+ and the RKKY- and
the Korringa mechanism (Noakes et al. 1987). A fit of this model (solid line in
fig. 4.51) leads to values for the RKKY exchange rate P'RKKY'the Korringa constant
UKorr/T and the strength of the CEF related spin lattice relaxation which are in
rough agreement with theoretical estimates for SmRh4B6 (Kumagai et al. 1981)
(t,fi~K Y ,--o 1.2 x 10 s S- 1 , UfiKtorr/T~ 3.4 x 10 6 K - I s - l ) . For ErRh4g 4 the apparent
RKKY rate and the Korringa rate are determined to be URKKynt~-- 3 X 10Ss -1 and
U~torr/T -~ 4 x 105 K - i s -1, i.e. the RKKY-mechanism dominates the spin fluctuation
rate below ~ 50 K and the Korringa mechanism is essentially absent. This is not
understood in detail. Some excess slowing down in the spin fluctuation rate is seen
in SmRh4B4 below ,-~ 4 K when approaching TN (Noakes et al. 1987).

HoxLul_xRh4B4 (x = 1, 0.7, 0.02, 0,005). The lowest state of the J = 8 518 Ho 3+-
ground state multiplet in HoRh4B4 is a degenerate nearly pure I + 8) doublet with
no or only very small off-diagonal matrix elements of J, inhibiting or suppressing
relaxation within this doublet at low temperatures by, e.g., the Korringa mechanism.
In this respect HoRhaB4 is quite distinct from SmRh4B4 and ErRh4B 4 which both
possess degenerate CEF ground-state level which are connected by IAmjI = 1 tran-
sitions. Consistent with the ground state properties of Ho 3+ in rhodium boride a
severe slowing down of the spin fluctuations is observed in ZF- and LF-#SR mea-
surements. In the system HoxLul_xRh4B4 this is seen for all concentrations x to
the extent that even in the very dilute compounds with x = 0.02 and x = 0.005,
which do not show magnetic order, a quasi static situation develops below T "-~ 9 K
for x = 0.02 and below T _~ 6.5 K for x = 0.005 (described as an isolation of the
ground state doublet). Since the fluctuation rates are essentially independent of x
below ~ 10 K a single ion relaxation mechanism must prevail which points to the
Korringa mechanism, although quite suppressed, as the responsible one (Heffner et
al. 1985, 1984, MacLaughlin et al. 1983, Boekema et al. 1982). The onset of
magnetic order in HoRh4B4 (Tc = 6.6 K) and in HO0.TLU0.3Rh4B4 (Tc = 4.1 K) is
not reflected visibly in the ZF-#SR relaxation data. However, very similar to the
dilute compounds, quasi static behaviour develops below (9-10) K (fluctuation rates
< 1 MHz).
Above ~ 10 K the fluctuation rates change to a thermally activated behaviour
reflecting the transition from intra ground state doublet relaxation to a much faster
activated behaviour involving excited CEF levels as in SmRh4B 4 and ErRh4B 4.
The onset Of superconductivity in Hoo.TLuo.3Rh4B4 has no apparent effect on the
fluctuation dynamics.

GdRh4B4. The spin dynamics in compounds such as SmRh4B4, ErRh4B4 and


HoxLUl-xRh4B4 is obviously very much determined by the CEF-split ground state
MUON SPIN ROTATION SPECTROSCOPY 169

features. In contrast Gd 3+, which is in a pure spin s t a t e (857/2) , is unaffected by the


CEF. This is clearly reflected in ZF- and LF-#SR data which reveal an essentially
temperature independent spin fluctuation rate (~ 5 x 101° s -1, depending on the
assumed site) down to the ferromagnetic ordering temperature (_~ 5.6 K). The tem-
perature independence points to an RKKY exchange mechanism. In comparison to
S m R h 4 B 4 and E r R h 4 B 4 the rate is enhanced by two orders of magnitude (MacLaugh-
lin et al. 1983). No slowing down of the fluctuation rate close to Tc shows up when
approaching Tc from above (the critical region may be very narrow) but the #+
relaxation rate decreases steeply below Tc reflecting the freezing of fluctuations in
the ordered state (small amplitude fluctuations). The rather sharp break in the #+
relaxation rate at Tc is rounded off remarkably in a longitudinal field of 0.1 T. This
feature has not been explained. No spontaneous coherent spin precession signal is
seen below Tc. This together with the small relaxation rate at 2 K implies that the
#+ resides at a site where, in the ordered state, the internal fields must cancel rather
well.

EuMo6S7.5Seo.5, Euo.75Sno.25Mo6S7.6Seo.4.The first compound does not become super-


conducting but shows, as pointed out before, a structural phase transition at ~ 110 K.
The second compound becomes superconducting below 5.15 K. The partial substitu-
tion of Se for S was chosen in order to create a particular phase diagram in the field

'~
oO
80 a ) ~
100

60
I I I I I

O
40
>

20 ~ oo

I I I I I I I f I ~ I I I I ] I I I I I L I I I I I I I

0 50 100 150 200 250 300


Temperature (K)
100 i i i i I

80 b)
'~ 60 (~
O
~.~ 40

20

0 i i i i i i i i i i i i i i i i i 1 f i i i i i + ~ i i I

50 100 150 200 250 300


Temperature (K)
Fig. 4.52. Temperature dependence of the Eu-4f moments fluctuation rates u in (a) EuMo6S7.5Seo+5 and
in (b) Euo.75Sno.25Mo6S7.6Seo.4 extracted from ZF-relaxation rates (from Birrer 1991).
170 A. SCHENCKand EN. GYGAX

induced superconducting state which, however, is of no relevance for the present


discussion. ZF, LF and TF-measurements show consistently the presence of rather
slow spin fluctuations in the range ~ (1-300) K and some kind of magnetic order
below or around 1 K (Birrer et al. 1990, 1989b, Birrer 1991). The temperature
dependence of the Eu-4f moment fluctuation rate u is shown in fig. 4.52 for both
compounds. It was deduced from the muon spin relaxation rate AZF (= A1) by adopt-
ing a reasonable #+ site and assuming an Eu 3+ moment of # = 7.5#B (S = 7/2).
The effect of the structural phase transition is clearly manifest in the data shown
in fig. 4.52a. Interestingly u rises linearly with T in the metallic phase (also in
Eu0.75Sn0.25Mo6Sv.6Se0.4 from 10 K up to at least 130 K, fig. 4.52b), indicating
a Korringa mechanism, but is temperature independent in the non metallic phase
down to 10 K. In both compounds a change in behaviour is observed below 10 K.
While u in the Sn doped compound becomes now temperature independent at the
same value as found in the non metallic phase, u in the undoped compound starts
to rise down to ~ 0.8 K, where it saturates. However, LF measurements show that
below 0.8 K in EuMo6S0.vSe0.5 and somewhere below 3.5 K, but above 1.4 K, in
Eu0.75Sn0.25Mo6ST.6Se0.4static random fields are established signaling the onset of
some magnetic order (Birrer 1991). Whether it is spin glass like or of a complex
long range structure could not be determined. The fluctuation behavior displayed in
fig. 4.52b, in particular also the linear temperature dependence below 130 K, agrees
closely with M6ssbauer results in Eu0.25Sn0.75Mo6S8(Dunlap et al. 1979). The spin
dynamics in Eu0.75Sn0.25Mo6S7.6Se0.4is unaffected by the superconducting transition
at 5.15 K.

4.2.8. R2Fel4B
This class of rare earth intermetallic compounds has found an important application in
the construction of hard permanent magnets. Their usefulness in this respect derives
from very large magneto-crystalline anisotropies which are related to CEF effects
and strong spin-orbit interactions of the rare earth constituents (see, e.g., Buschow
1986) From a more fundamental point of view these compounds are interesting
since both the R- and the Fe-sublattice display magnetic order which sets in below a
common Curie temperature indicating a strong interplay of the localized 4f electrons
and the more itinerant 3d electrons. The sublattice order is ferromagnetic for all
compounds but the two sublattices couple ferromagnetically for the light rare earth
up to Sm and antiferromagnetically for the heavy rare earth. In each case a collinear
arrangement is found. In some compounds spin canting without destruction of the
collinearity is observed. The crystal structure is quite complicated as shown in
fig. 4.53. From studies of hydrided compounds with small hydrogen concentrations
it was deduced that hydrogen occupies the so called tetrahedral 4e sites with two
R and two Fe nearest neighbours (see fig. 4.53) (Ferreira et al. 1985). There
are four crystallographically equivalent 4e sites within the unit cell. They are also
magnetically equivalent as long as the ordered moments are parallel to the c-axis. It
is reasonable to assume that also implanted #+ will reside at the 4e site.
ZF-#SR measurements were under taken with the aim to gain additional informa-
tion on the low temperature magnetic structure of the RzFex4B system (Yaouanc et
al. 1987, Niedermayer et al. 1990).
MUON SPIN ROTATION SPECTROSCOPY 171

/,e site

[I 010~,

R2Fell.B
Fig. 4.53. Crystal structure of tetragonal R2Fel4B (R = rare earth). Indicated is a tetrahedral 4e
interstitial site which is known to be occupied by dissolved 'hydrogen (from Niedermayer et al. 1990).

In all investigated polycrystalline compounds (see table 4.14) a single spontaneous


coherent spin precession signal was seen in ZF-measurements. In the compounds
Y2Fe14B and Pr2Fel4B the spontaneous field Bu at the #+ site decreases smoothly
with temperature with no further structure consistent with no spin canting in this
two compounds. B,(T) scales roughly but not precisely with the macroscopic mag-
netization. Note that in Y2Fel4B Bu originates solely from the Fe-sublattice. The
upper part of fig. 4.54 shows the temperature dependence of z~u in Nd2Fe14B together
with the macroscopic M(T). Below 150 K, uu(T) decreases while M(T) continues
to increase. The break in ut,(T) coincides with the start of the spin canting known
from other experiments. Here the macroscopic magnetization changes continuously
from an orientation parallel to the c-axis to an orientation parallel to the [110] axis.
Interestingly uu(T) below 150 K scales with the magnetization component along the
c-axis (dashed curve in fig. 4.54). The spin canting should lead to the appearance
of two magnetically inequivalent types of 4e sites and one would expect to see two
distinct precession signals. The absence of a splitting below 150 K and the scaling of
u~(T) with the magnetization component along the c-axis points to some averaging
mechanism by which planar internal fields are averaged to zero. It is suggested that
/~+ diffusion, which must be fast in this complicated compound even at temperatures
as low as 4.2 K, is responsible for this mechanism. However, in the absence of any
strong temperature effect on the relaxation rate of the precession signal #+-diffusion
172 A. S C H E N C K and E N . G Y G A X

m m~ ~

~ ~ oo.~.~

0 O0 0 0 0 0 0 0 0 ~

r,¢? ~ M M M ~

~- ,--7. ~ ~ ~. ,~.

o
~V

~ VV
t~ V~V~ m
o0 ~0 ~0 7

•~ . ~ ~.~.

1 I

.1
0 0 0 0 0 0 0 ~

'N 0 o
~ N ~ N N O
r~

© "D

~ c5~: ~ ,~ ~
0

0 "0 0 0 0 ~ c5
L~
MUON SPIN ROTATION SPECTROSCOPY 173

0 tz~

~ o o ~ ~
g~ o
©

ca O
I
::i.

t~

::i.
m <
I

m I

m
O

o ~
O

Z~

0
174 A. SCHENCK and EN. GYGAX

I I I

180
Nd2Fe14B

_ M (30M'~-~'f0) el"~e
~
"1'-
~ 170
o
r-
I1}

" 160

I
100 200 300

400 02Fe14 B ~8

o>" 350 ,,--.


t"- O

u. 14

300
0 f I T 12
100 200 300
Fig. 4.54. Temperature dependence of the spontaneous precession frequency and of the saturation
magnetization M (solid line~ in Nd2Fe]4B (upper part) and in Ho2Fel4B (lower part). The dashed lines
show the projection of M on the c-axis (from Yaouanc et al. 1987, Niedermayer et al. 1990).

appears somewhat unlikely and it rather seems as if some aspect of the magnetic
structure in the spin canting regime is not properly understood. Very similar results
are also observed in H o 2 F e l 4 B where again uu(T) shows a break at ~ 50 K be-
low which temperature spin canting develops in this compound (see lower part of
fig. 4.54). Interestingly in this compound the macroscopic magnetization decreases
with temperature due to the AF coupling of the two sublattices, while u(T) increases
till ~ 50 K because in this case the dipolar fields from the Fe- and Ho-moments
add constructively at the 4e #+-site. Table 4.14 lists the low temperature (15 K)
B~ and the difference B~ - BL (BL is calculated from the macroscopic saturation
magnetization at low temperatures) which is given by +[/~c +/~dipt-
Figure 4.55
shows that B , - BE scales very well with the rare earth moment or its component
along the c-axis, respectively. This figure teaches that the Fe-sublattice contribution
to the microscopic local field at the #+ is rather independent of the rare earth element
present.
MUON SPIN ROTATION SPECTROSCOPY 175

I I I I 1 I I I

3.5

3.0 D Tb y
2.5
2.0
i
m= 1.5
• . . . . . . . . . . . . . . . . . . .

0.5
0 I I I I I I I I
-10 -8 -6 -4 -2 0 2 4
gnE(~B)
Fig. 4.55. Linear correlation of (Bu - BL) (low temperature limits) with the value of the rare earth
moment (from Niedermayeret al. 1990).

4.3. IntermetalIic compounds containing actinide elements


Some of the results to be discussed below have been reviewed before by Asch (1989,
1990) and Schenck (1993). Tables 4.15 and 4.16 list all the compounds studied so
far by #SR spectroscopy.

4.3.1. Cubic laves phase compounds: UAI2, UMn2, NpAl2

UAI2. Some times considered to belong to the class of heavy fermion systems
(7 = 90 mJ/mol K 2) UA12 seems to be a fast spin fluctuator (TSF ~ (25-30)K) down
to the lowest temperatures with no disposition towards magnetic order. The impor-
tance of spin fluctuations is revealed by a T 3 In T behaviour of the low temperature
specific heat. This picture is fully in line with ZF- and LF-#SR measurements on a
polycrystalline sample (Kratzer et al. 1986, Asch et al. 1987). LF-measurements in
particular reveal an extremely small relaxation rate ), ~< 0.05 #s -1 which translates
into a U-5f electron spin fuctuation rate of 1/Tc > 1013 s -1, assuming the U-5f
moment to be given by ~ 4#B, as deduced from the magnetic susceptibility above
100 K (see, e.g., Fournier et al. 1985). This rate estimate agrees well with results
from a neutron scattering measurement of the dynamic susceptibility (Loong et al.
1986).

UMn2. ZF- and LF-#SR measurements on a powder sample (Cywinski et al. 1994a)
prove UMn2 to be a fast spin fluctuator like UAI2 and no indications for long range
order involving either the 5f and for the 3d moments are seen in agreement with
neutron diffraction measurements. This seems to rule out expectations that AF order
could develop below ~ 240 K, somewhat above a structural transition in the range
176 A. SCHENCK and EN. GYGAX

_9.o -~t ,,:~-

N N N

• • . . ~ ~
.-~ ,~ ,~

o
~ ~

o0

o0

T~

~o~ ~
©

O~DZ ZZ ZZo

O9

O9

0
dn
MUON SPIN ROTATION SPECTROSCOPY 177

o.q~

b b~
d o ~ d
o9

oO ~ g~2 gT.2
"0
N N N N NN N b"
0
g
¢q
J
e- v~

0 P- P.
oo
I I I
d o ''~
v~

eq eq

t'~ I
t¢3 t¢3 ~ ~
II II

e-,
e~~ t~
P,

e,h e~ CD

N .~ o

e~
o

0 <

o Z
U
178 A. SCHENCKand EN. GYGAX

(210-230) K in analogy to U F e 2 where a rhombohedral distortion coincides with a


ferromagnetic transition at 165 K (see, e.g., Fournier et al. 1985).

NpAl2. NpA12 is a ferromagnet which orders below Tc = 56 K. Only prelimenary


ZF-#SR results are so far available (Aggarwal et al. 1990). In contrast to all other
cubic Laves phase compounds studied by ZF-#SR in the ordered state, NpA12 is the
only system in which a spontaneous spin precession signal is seen with more than
one precession frequency. Indeed if the #+ is placed at the (2-2) site, as found in
CeA12 (see fig. 4.6, section 4.2.6) several distinct local fields should be visible, but a
detailed analysis is as yet missing. TF-#SR measurements (0.02 T) show a strongly
rising relaxation rate or a slowing down of the spin dynamics, respectively as Tc
is approached from above, starting at ,-~ 75 K, which is typical for a second order
phase transition.
4.3.2. Binary compounds with NaCl structure
Like the rare earth monopnictides the corresponding U compounds show complex
magnetic structures of multiple q spin arrangements. The U-monochalcogenides are
ferromagnets (see table 4.15).

UAs. #SR-measurements in ZF, TF and LF on a mosaic of unoriented crystalline


platelets mirror very nicely the different magnetic phases and provide some additional
information (Asch et al. 1987, 1989, Kratzer et al. 1990). ZF-measurements below
62 K in the type IA, double 0"AF phase reveal spontaneous coherent spin precession
involving two frequencies and one non-oscillating component (see fig. 4.56). The
temperature dependence of the frequencies is displayed in fig. 4.57. Assuming the
#+ to occupy the same site as in CeAs (see fig. 4.21) one calculates for the given
AF structure a zero net dipolar field and by symmetry arguments also a zero contact
hyperfine field. The fact that non-zero frequencies and a splitting is seen is a result
of a tetragonal lattice distortion occuring at 62 K (c/a < 1) (also at 124 K (c/a > 1)).
This distortion is a necessary prequisite for the establishment of the type IA double
~' spin structure (Sinha et al. 1981). It changes the point symmetry at the #+
site with the effect that crystallographically equivalent sites become magnetically
inequivalent (--+ splitting of signal) and that the dipolar fields from the neighbouring
U-moments do no longer cancel (--+ non zero frequencies). In fact the distortion
explains the ratio of the signal amplitudes associated with the two oscillating and
the one nonoscillating signal. Dipolar field calculations predict a ratio of x/2 for
the two frequencies independent of the size of the distortion. The fact that at the
lowest temperature a ratio of ~ 2 is observed points to additional contact hyperfine
fields, which are also needed to account for the strange temperature dependence of
the lower frequency in fig. 4.57. A detailed understanding, however, is lacking. It
should be emphasized that neither neutron nor X-ray diffraction could detect the
tetragonal distortion, providing only an upper limit of ]c/a - 11 < 2 x 10 -11 (Sinha
et al. 1981, Knott et al. 1980). The #SR-results provide the only direct evidence
for the distortion by the change of the point symmetry but no quantitative analysis
was presented.
MUON SPIN ROTATION SPECTROSCOPY 179

0.15

0.10

0.05

-0.05
I I I I I I
-$ 1 2 3 4 5 6
E Time (gs)
E
(n
0.15

0.10

0.05

-0.05 I I
0.5 1
Time (~s)
Fig. 4.56. ZF-/zSR signal in UAs below 62 K in the type IA, double ~"AF phase. The lower part shows
an expanded early section of the signal (from Asch et al. 1987).

I I I I I I

12.0
o o O0 03 O 0 o
0
og
~, 11.5

o" 6.0
E
o8
8 5.5
o
LL o
5.0 oO o o
00
4.5 0 or I I I I I
0 10 20 30 40 50 60
Temperature (K)
Fig. 4.57. Temperature dependence of the two spontaneous precession frequencies in UAs in the type
IA, double q AF phase below 62 K (from Asch et al. 1987).

Above 62 K the coherent spin precession signal vanishes abruptly and only a
weakly damped Lorentzian Kubo-Toyabe signal is seen with no temperature depen-
dence up to TN ----- 124 K. This is consistent with the prediction that even in the
presence of a tetragonal distortion the type I single q" spin structure will not lead to
180 A. SCHENCK and EN. GYGAX

non-zero net dipolar (and contact) fields at the #+-site. The extremely small residual
field spread of N 0.8 m T at the #+-site, following from the static Lorentzian K u b o -
Toyabe behaviour implies a very perfect and static spin structure in the vicinity of
the #+ and that disturbances are dilute and more distant. Spin excitations do not
seem to play a role.
The transitions at TN = 124 K into the paramagnetic state shows up very markedly
in W-measurements. The relaxation rate drops by more than a factor of two and
the precession frequency displays a jump by more than +2500 ppm (Asch et al.
1987). The relaxation rate above 124 K is essentially temperature independent up to
,-~ 180 K, where another down jump to a very small rate is observed (see fig. 4.58).
In particular no increase of the relaxation rate (no slowing down of spin fluctuations)
is seen when approaching TN from above consistent with the notion of a first order
phase transition (Sinha et al. 1981). The change at 180 K is accompanied by
a transition from a Gaussian damping function above this temperature, reflecting
the nuclear (75As) dipolar field distribution, to a more exponential one below this
temperature (but still reflecting static features, Kratzer et al. 1990). The modified
relaxation behaviour below 180 K is ascribed to the onset of a magnetic precursor
state in which the 5f-moments no longer act as free paramagnetic spins. This is in
accord with diffuse neutron scattering results slightly ( ~ 10 K) above TN which reveal
strongly anisotropic spin fluctuations tending towards an incommensurate sinuoidally

Tt TN UAs

0.4 °o oo (a)

0.2 O

O
, , , ~ .... I .... I , , ,O, I ,Q 91 ....
0
Tt TN UP

0.8
_~o~p o o o oo o (b)
0.4 o

0 ,,,I .... I , I~'~,OI , O , ,q , , 1911 I


T N CeAs

0.2 (c)
0.1 ~b~:~
0 .... ,o, ,o,, o, ,o, ,9,,,,
0 50 100 150 200 250 300
Temperature (K)
Fig. 4.58. Temperaturedependence of TF-relaxation rates A2 in (a) UAs, (b) UP and (c) CeAs. The
first order nature of the phase transition at TN in UAs and UP is manifest by the discontinuousjump of
),2- In contrast ),2 in CeAs reflects a critical slowing down of the spin dynamics near TN and signals a
second order phase transition (from Asch 1989).
MUON SPIN ROTATION SPECTROSCOPY 181

moduled spin structure (Sinha et el. 1981). The #SR data show that the precursor
state sets in already about 50 K above TN (Asch et el. 1989).

USb. Like in UAs in the type I single q phase, ZF-measurements on USb do not
exhibit any coherent spin precession below TN (the spin structure is of type I, triple q-)
but display a Lorentzian Kubo-Toyabe behaviour with full amplitude. In contrast to
UAs the internal fields acting on the #+ are not totally static below TN and both the
static width A and the spin fluctuation rate u show a pronounced and very interesting
temperature dependence (see fig. 4.59) (Asch et el. 1990). A displays a maximum
at ~ 140 K at which temperature the fluctuation rate u starts to rise dramatically
with increasing temperature. Neutron scattering measurements signal at ~ 140 K the
collapse of spin waves, well resolved at lower temperatures, and the development
of a broad inelastic peak centered at zero frequency (Hagen et el. 1988). The #SR
results are interpreted in terms of a phase de-locking of the magnetic components in
the triple ( structure just as in DyAg (see section 4.2). Phase de-locking, induced
thermally, produces irregularities (defects) in the spin structure and hence increases
the field inhomogenity sensed by the #+. This explains the rise of A with increasing
temperature. At around 140 K a motionat narrowing effect sets in which arises from
a slow diffusive motion of the thermally induced perturbances or defects in the spin
structure and explains the step rise in u and - at least qualitatively - the decrease
of A above 140 K. Note that the disturbances/defects are still quite dilute so that a
Lorentzian Kubo-Toyabe picture is applicable.
By comparison with other compounds it is found that a dynamic Lorentz-Kubo-
Toyabe behaviour with a temperature dependent A is typical for a multiple 0"structure
which, in contrast to a single 0~structure, seems to be more easily distorted due to
de-locking of the phases between the Fourier components of the moments (Asch et
el. 1994).

UP. This compound displays rather similar properties as UAs (see table 4.15). A
subtle but important difference is that in the ordered phase, associated with the double

USb 9 6.o
i
1.3 i 5.0

~ 1.2 o/°\ - Rate---- 4.0 ~


/ 3.0
~" ,-Width / ~. ~"
~ 1.1 o /0%.u
2.0

1.0 ) ...... • ..... O "~ B "/" 1.0


n I I I I
50 100 150 200 I 250
Temperature (K) TN
Fig. 4.59. Temperature dependence of the width A and the spin fluctuation rate v from the dynamic
Kubo-Toyabe signal seen in USb (from Asch et el. 1990).
182 A. SCHENCKand F.N. GYGAX

structure, the propagation vector is given by q = (0, 0, 1) (type I) while in UAs


~ = (0, 0, 1/2) (type IA). Only in the latter case one calculates non vanishing dipolar
fields at the #+ site, if also a tetragonal distortion is present. Consequently in UP
no coherent spin precession signal is seen in ZF-#SR measurements (Aggarwal et
al. 1989). TF-measurements show abrupt changes in the relaxation rates at both
transitions temperatures (see fig. 4.58b) but no indication for critical behaviour in
line with the first order nature of the phase transitions. In contrast to UAs also no
magnetic precursor state is seen in UP above TN = 122 K (Aggarwal et al. 1989).

UN. UN orders antiferromagnetically into the type I single q structure at 53 K.


The transition is accompanied by a tetragonal distortion with c/a > 1. From the
pressure dependence of TN and #u strong evidence for itinerant magnetism was
deduced (Fournier et al. 1985). The phase transition is believed to be of second
order. ZF-#SR measurements below TN do not reveal a spin precession signal but
a static Lorentzian Kubo-Toyabe behaviour consistent with similar results in the
single qphase in UAs and UP (Mtinch et al. 1993). TF-measurements (Mtinch et al.
1993) show a drastic increase of A2 at TN but no indications for a critical slowing
down of the spin dynamics when approaching TN from above just like in UP (see
fig. 4.58b). This is unexpected in view of the assumed second order nature of the
phase transition. A2 above TN is in fact very small (~ 0.03 #s -l like in UP) and
temperature independent and implies a very fast fluctuation of the U-moments. From
the point of view of #SR there is no difference in the magnetic behaviour of UP and
UN above and below TN (except that UP possesses another phase transition at lower
temperatures).

UTe. In contrast to the U-monopnictides this monochalcogenide shows ferromagnetic


order below Tc = 104 K. ZF-#SR measurements below Tc could not detect any
coherent spin precession signal, probably because of a too rapid relaxation due to
a very wide field distribution associated with lattice irregularities (Aggarwal et al.
1989). TF-measurements (0.1 T) above Tc yield a frequency shift which scales very
well with the susceptibility. The exact form of the temperature dependence of the
TF-relaxation rate could not be extracted but its diverging trend when approaching
Tc from above is consistent with the second order nature of the ferromagnetic phase
transition.

4.3.3. Binary compounds with cubic AuCu3-structure: USn3, Uln3, U(Ino.5Sno.5)3


UIn3 shows antiferromagnetic order below TN = 88 K with a type I triple q spin
structure. USn3 on the other side is a paramgnet exhibiting strong spin fluctuations.
In the mixed system U(Sn~Inl_~)3 long range magnetic order is suppressed for
x/> 0.4 (Zhou et al. 1985). Prelimenary #SR-measurements applying the ZF-, TF-
and LF-technique yielded results which generally confirmed the present picture on
the magnetism of these compounds (Zwirner et al. 1993, Kratzer et al. 1994b, Asch
et al. 1994). The results on USn3 imply indeed a very fast spin fluctuation of the
order of 1013 s -1 or faster. The same is found in the paramagnetic state of UIn3. As
in the other compounds with type I triple-~f spin structure (USb, DyAg) the ZF-#SR
MUON SPIN ROTATION SPECTROSCOPY 183

signal in UIn3 below TN = 88 K is well described by a dynamic Lorentz-Kubo-


Toyabe function. The temperature dependence of A and u resembles closely the
results on USb displayed in fig. 4.59 and is discussed there. In the present case
the dynamic range extends only for about 10 K below TN. A possible first-order
nature of the phase transition is reflected by a temperature independent frequency
shift and relaxation rate in TF-measurements when approaching TN from above. New
information is gained on the mixed compound U(Sn0.sIn0.5)3 which is expected not
to show long range magnetic order. AF- and TF-measurements reveal, however,
the onset of spin glass order below ~ 30 K. This is evidenced by a change of the
ZF-relaxation function from a Gaussian to a Lorentzian behaviour (well described
by the spin glass relaxation function of Uemura et al. 1980) and a strong increase
of the W-relaxation rate below 50 K. Previously it was found in U(Sn0.5In0.5)3 that
the specific heat c(T)/T shows a slight up turn and the resistivity p(T) a break in
slope and a weak decrease at or below 30 K (Lin et al. 1987). These features
were not explained before but find now in the light of the #SR results a natural
explanation.

4.3.4. Tenary compounds with tetragonal ThCr2Si2 or CaBe2Si2 structure

URh2Si2. The antiferromagnetic local moment (#u = 1.95#B ) order in this compound
is identical to the one in NdRh2Si2 (see table 4.10), yet ZF- and LF-#SR measure-
ments on a polycrystalline sample display quite a different behaviour (Yaouanc et
al. 1990, Dalmas de R6otier et al. 1990d, 1994c). Below TN a one frequency
coherent spin precession signal is seen which at low temperatures implies a local
field of 0.305 T. Obviously only one type of site is occupied by the #+. The site
(1/4, 1/4, 1/4), No. 7 in fig. 4.32, can be excluded since it involves a zero net
internal field). The corresponding information is missing for NdRh2Si2 due to the
limited time resolution available at the ISIS facility. In CeRh2Si2, which has a

- URh2Si2
0.08
u_
N 0.06

t-
O
"~
0.04

0.02
"+ ÷ ÷÷t
X
Q Oo
rc 0.00
I I I I I I I I II f

10 100
Temperature (K)
Fig, 4.60. Temperature dependence of the relaxation rate ~ZF in URh2Si 2. Below TN )~ZFis really a )q
(see text) (from Dalmas de R6otier et al. 1994c).
184 A. SCHENCK and F.N. GYGAX

I ' i , I ' ~ i I l , i I ' ' ' I i i ~ I

b~ 0.06 uRh2Si2 o 30K


v=
~T Z~ 9 6 K
(ff 0.04
• 121K
t-
O 0.02
X
t~

~
cc 0.00
I k i i I , , , I , , , I ~ , I I I I I I

0 2 4 6 8 10
Magnetic field (mT)
Fig. 4.61. Dependence of the spin lattice relaxation rate -~1 on the strength of a longitudinally applied
field (/~(0)llnext) below TN in URh2Si 2 (from Dalmas de R6otier et al. 1994c).

slightly different magnetic structure, two different internal fields were seen below
TN, suggesting that the #+ is located at two different interstitial sites. Very different
is the dynamic behaviour below TN in URh2Si2 (see fig. 4.60). The longitudinal re-
laxation rate )`1 (the relaxation function is well described by an exponential decay) -
revealed by those #+ whose spins happen to be parallel to the static internal field -
is essentially temperature independent while in NdRheSi2 a quadratic temperature
dependence is seen for P(0)llc-axis (see fig. 4.34). Moreover, again in contrast to
NdRh2Si2, )`1 below TN is dependent on the strength of a longitudinally applied field
(see fig. 4.61). This feature is not understood at present. Above TN, ),1 drops quickly
with rising temperature similarly to what is seen in NdRh2Si2 for fi~(0)_Lc-axis or
in GdNi5 (see figs 4.34, 4.26), reflecting some slowing down of the spin fluctuations
as TN is approached from above. A more quantitative analysis of the data is as yet
missing.

U(Rho.35Ruo.65)2S@ While URh2Si2 is characterized as a local moment (# _~ 1.4#~)


antiferromagnet URu2Si2 is a magnetic heavy fermion superconductor exhibiting ul-
tra small moment (# _~ 0.02#B ) magnetic order below 17 K (see section 4.3.5). It
is of considerable interest to study the transition from one type of behaviour to the
other one by investigating the mixed compounds U(Rh~Rul_=)2Si2. ZF- and LF-#SR
measurements were conducted on the compound U(Rh0.35Ru0.65)2Si2 [polycrystalline
sample for which the magnetic structure is not known (Yaouanc et al. 1990, Dalmas
de R6otier et al. 1990d)]. The ZF-results are shown in fig. 4.62. A single expo-
nentially damped signal is observed. Its amplitude starts to drop smoothly at about
150 K to 1/3 of its initial value below ~ 40 K. This drop signals the development
of a very fast relaxing (or precessing) component which, since these measurements
were performed at the pulsed muon source of ISIS, could not be resolved. In any
case the low temperature value of the amplitude reveals unambigeously that the #+
are exposed to a very wide static field distribution or, alternatively, to a rather high
MUON SPIN ROTATIONSPECTROSCOPY 185
[ i i i i i i i1111111111 [ i

o o
0.15 oOOO

E
0.10
so
0.05
o Oq~O 0

0.00

- - U ( R h . 3 5 Ru.6 5 )28i2
(~ 0.08

0.06

•~ 0.04

.o 0.02 (~
,o %, O
_m
1
TN O
X

oo
or
"
0.00
I , I I ' ' I IIItqllllll[ I

20 100 200
Temperature (K)
Fig. 4.62. Temperaturedependence of the ZF-~SR signal amplitude (asymmetry) and relaxation rate
AZF in U(Rho.35Ruo.65)zSi2. Note the loss of amplitude between ~ 150 K and ~ 42 K, indicating the
transition into a magnetically ordered state (from Dalmas de Rdotier et al. 1990d).

local field leading to a rapid coherent spin precession, too rapid to be resolved.
In other words magnetic order is established in this compound somewhere below
,-~ 40 K. Whether it is long range or short range (e.g., spin glass like) cannot be
deduced from the present #SR data. The drop in the amplitude could indicate that
the development of magnet order proceeds in a spatially inhomogeneous fashion, a
not uncommon feature in heavy electron systems (see sections 4.2.6, 4.3.5). Very
interesting is also the temperature dependence of the ZF-relaxation rate AzF (see
fig. 4.62). LF-measurements prove that AZF (= At) is more or less entirely of dy-
namic origin in the whole temperature region studied. AzF peaks sharply at ~ 42 K,
similar to what is seen in polycrystalline NdRh2Si2 (Dalmas de Rtotier, 1990d) and
is a clear signature for a magnetic phase transition. Note that AzF is generally quite
small implying that above TN --~ 42 K spin fluctuations must be rather fast. The
relaxation rate below TN is independent of a longitudinally applied field, therefore
reflecting still some dynamics as in NdRh2Si2 and UPt2Si2, quite in contrast to the
field dependence seen below TN in URh2Si2.
In summary magnetic order is seen in U(Rho.35Ruo.65)2Si2 below ~ 40 K but the
dynamical behaviour in the ordered state is different from the one in the parent
compound URh2Si2.
186 A. SCHENCK and EN. GYGAX

' I ' I i I

,,-;- 0.08 UPt2Si2 7!


=1. o 0mT q
v 0.06
z~ 1 0 r o T
"~
¢-
._o
0.04
o • 200 mT

0.02

rr 0.00

I i I t I i I
20 40 60 80
Temperature (K)

Fig, 4.63. Temperature dependence of the ZF and LF relaxation rate in UPt2Si2. Note the sharp cusp
of AzF, A1 at the transition temperature of 35 K (from Dalmas de REotier et al. 1994c).

UCo2Si2. Some preliminary ZF-#SR measurements imply very fast spin dynamics
above the ordering temperature TN = 85 K. No slowing down of the dynamics is seen
on approaching TN from above. Below TN 2/3 of the signal is lost as is expected
for a polycrystalline sample, given the limited time resolution at the ISIS facility
(Dalmas de Rdotier et al. 1994c).

UPt2Si2. Again only preliminary ZF- and LF-#SR data are available (Dalmas de
R6otier et al. 1994c). The observed temperature dependence of AzF and A1 is
displayed in fig. 4.63. The small field dependence is ascribed to a decoupling
from the Pt-nuclear dipole moments. The data are quite similar to the results in
U(Rh0.36Ru0.35)2Sia and other isostructural rare earth compounds. They reflect mostly
the dynamics of the spin system and indicate a certain slowing down, which starts
already much above TN.

4.3.5. Actinide based heavy electron and related compounds


All systems investigated by #SR are listed in table 4.16. For a more detailed dis-
cussion of the #SR results see Schenck (1993). A compilation of some of the most
important #SR results is presented in table 4.17.

UPt3. The heavy fermion superconductor UPt3 is certainly one of the most fascinating
compounds among all heavy fermion systems possessing a rather intriguing low
temperature phase diagram (for a recent status see de Visser et al. 1993). First
indications for magnetic order at ~ 5 K, preceeding the superconducting transition
at Tc ~ 0.5 K were provided by ZF- and TF-#SR measurements on a polycrystalline
sample (Heffner et al. 1987, 1989a, Cooke et al. 1986). Figure 4.64 displays the
measured ZF-relaxation rate AZF which exhibits a significant increase below 5 K.
LF-measurements proved that this - sample dependent - increase is associated with
the development of small static fields of order 2 G at the #+ site or sites. Subsequent
neutron diffraction measurements confirmed the onset of an ordered state involving
MUON SPIN ROTATION SPECTROSCOPY 187

+
oo
.~[~

8
o 0
Cq
o,l

8 0
H
H
:.B ~o
A
0
+
V

?
0

g,,

d
0 ~'
.a ~o~o_~o0 ~ ~ .~'
~o ,~ ~ o ~ o ~ ~, . ~ °~ ~o ~ 0

~
'~ " o ~ ~ .~n ~.~ 0 0.~ ,.~ •~ 0 ~ ~

•~.~ .~ ~ . ~ O8

N ©
0 OI
0 II
0 0 A
-~ v
0 o
188 A. SCHENCK and EN. GYGAX

UPt 3 Gaussian relaxation


0.20

_ l+÷ ÷ + +
T¢tJ 0.16
:zL

• ZERO FIELD
¢.-
o
0.12 + +
X t I
c~

n'- 0.08
+

0.00 ~ ~
0.1 0.5 1.0 2.0 5.0 10,0 20.0
Temperature (K)
Fig. 4.64. Temperature dependence of the Gaussian ZF-relaxation rate cr in UPt 3. The dashed area
indicates the range of calculated relaxation rates due to nuclear (Pt) dipole fields for various possible
/z+ sites (from Heffner et al. 1987).

tiny moments of the order of 0.02#B (Aeppli et al. 1988). Taking the #SR-and
neutron results together it follows that the /~+ must be located at a site at which
the dipolar fields from the ordered U-moments, assuming the structure suggested by
Aeppli et al. (1988), cancel perfectly for a perfect stoichiometric and defect free
lattice (Schenck 1993). Note that no anomaly in the specific heat is seen at 5 K.
Prior to the discovery of magnetic order in UPt3 antiferromagnetic order had
already been found in the doped compounds Ua.95Th0.05Pt3 and U(Ph_=Pd~)3 (0.02 ~<
:c ~< 0.07) (Goldman et al. 1986, Frings et al. 1987). Interestingly the magnetic
structure is the same as in undoped UPt3 with TN also of the same magnitude
(TN ~ (3.5-6.5) K) but the ordered moment is much larger (,-~ 0.5#~) and the
magnetic phase transition can also be seen in transport and thermodynamic data
(Ramirez et al. 1986). A few low statistics ZF-#SR measurements on the compound
U1.95Th0.05Pt3 confirmed the onset of magnetic order at TN -- 6.5 K (Heffner et al.
1989a): below this temperature a two component precession signal appears reflecting
spontaneous local fields of 0.06 T and ~ 0.009 T. However, the total amplitude of
the two components accounts for only 20% of the implanted ~+ implying either a
severe loss of polarization due to an extremely inhomogeneous distribution of internal
fields in most of the sample volume or a zero field site for most of the implanted
#+. Since the two spontaneous fields are also associated with a large field spread
of ~ 0.007 T it seems that AF-order in the investigated polycrystalline specimen is
not so well developed, probably as a result of poor sample quality. Nevertheless the
appearance of well resolved precession signals in U1.95Th0.osPt3 and the absence of
any precession signal in pure UPt3 is somewhat of a mystery given the belief that
the AF-structure for both compounds is identical (e.g. scaling the field of 0.06 T
MUON SPIN ROTATIONSPECTROSCOPY 189

down by the moment ratio 0.02/0.5 _~ 0.04 one should have seen a local field of
24 G instead of the 2 G mentioned above).
ZF- and W-measurements on high quality mono- and polycrystalline UPt3 sam-
ples revealed another increase in relaxation rate starting at 490 mK, i.e. near the
temperature at which a second phase transition is seen in specific heat data some
60 mK below the transition into the superconducting state (see fig. 4.65) (Luke et
al. 1993a, b). Obviously the lower of the double transition around 0.5 K is asso-
ciated with a further increase in the static field spread ( ~ 0.01 mT). Judging from
TF-measurements at 0.18 T the field spread is rather isotropic. The appearance of
a split transition into the superconducting state is believed by many investigators
to be a consequence of the antiferromagnetic state below 5 K which provides a
symmetry breaking field. This couples to the superconducting order parameter and
splits the transition into an otherwise degenerate ground state. A coupling between
the magnetic and the superconducting order parameter seems also to be indicated by
neutron scattering data (Broholm 1989). Blount et al. (1990) have suggested that
the neutron data below Tc could be explained by a reorientation of the small antifer-
romagnetic moments. Luke et al. (1993b) showed that a rotation of the moments in
the basal plane by ~- 30 ° could reproduce the increase in local fields below 490 mK.
Other explanations, based on the speculation that the lower transition leads to a time
reversal invariance violating state, are discussed in Luke et al. (1993b).
The low temperature features seen by ZF-#SR in UPt3 are similar to results ob-
tained in U1L=Th~Be13, in which also a double transition is seen for 0.019 ~< z ~<
0.043. However, in the latter case no magnetic ordering seems to precede the transi-
tion into the superconducting state (see below) and the idea of a symmetry breaking

0.065 I
UPt~ ZF-/~SR
1 l l

0.060 P ,.k c

'T
0.055 I i
o
x
O

0.045 I 1
0.00 0.25 0.50 0.75 1.00 1.25
Temperoture (K)
Fig. 4.65. Temperaturedependenceof the ZF relaxationrate assuming exponentialrelaxation in a single
crystal of UPt3 with/5(0)Zc-axis. The positions of the two transition temperatures are indicated by the
arrows (from Luke et al. 1993b).
190 A. SCHENCK and EN. GYGAX

field responsible for a split transition may not be applicable. The occurence of ul-
tra small ordered moments in other nonsuperconducting HF-compounds (see, e.g.,
CeRu2Si2, UCus) moreover may suggest that the small moment magnetic order is
perhaps a phenomenon unrelated to superconductivity.

UCus. This moderately heavy electron compound orders antiferromagnetically be-


low TN = (15-16) K. The magnetic structure has been studied by neutron diffraction
and NMR spectroscopy. The data were interpreted in terms of two different struc-
tures. Murasik et al. (1974) (also Schenck et al. 1990b) explain their data in terms
o f ferromagnetic (1, 1, 1) planes and an antiferromagnetic coupling between neigh-
bouring planes (single q structure) while Nakamura et al. (1990) on the basis of
NMR data propose a quadruple q structure. Specific heat and resistivity measure-
ments reveal a second phase transition with hysteretic features at ~ 1 K (Ott et al.
1985). #SR-measurements were aimed in particular at a better characterization of
the 1 K transition. The studies started on a sample which did not show the 1 K phase
transition. ZF-measurements revealed the onset of three different signals below TN,
one associated with zero average field, but non-zero static field spread, the other two
displaying coherent spin precession corresponding to low temperature local fields of
0.146 T and 0.1 T, respectively (Barth et al. 1986b, 1988). Figure 4.66 shows the
temperature dependence of the spontaneous precession frequencies. The solid lines
in fig. 4.66 represent the temperature dependence of the ordered moment deduced
from neutron diffraction experiments (P. B6ni, priv. communication, Schenck et al.
1990b). As can be seen, only the 0.146 T signal scales with the order parameter. The
crystal structure of UCu5 is displayed in fig. 4.67. The two possible #+ interstitial
sites are indicated. Dipolar field calculations (#ord = 1.55#B) show that irrespec-
tive of whether the single-q or the quadruple-0" structure is adopted, Baip at the site
(1/2, 1/2, 1/2) is zero. For the single-0'structure one finds Bdip = 0.23 T at all equiv-
alent sites (3/4, 3/4, 3/4), while for the quadruple-q structure one finds t3di p = 0 T

I I I I I I I I
2O -1.5~
N ~9
"1- v

v
16
II)
>,,, 1.0
¢j

o" 1..-
8
0.5 E
+
o
~L t,j,
o
0 I I I I I I I i 0
0 2 4 6 8 I0 12 14 16 18 20
T(K)
Fig. 4.66. Temperature dependence of the two spontaneous precession frequencies in UCus. The solid
lines represent the temperature dependence of the ordered moment (from Schenck 1993).
MUON SPIN ROTATIONSPECTROSCOPY 191

UCu 5 • u
~,o Cu

"-'-- p+ (2)
I

" ' ' - " #+ (1)

w W

Fig. 4.67. Crystal structure of the cubic AuBe5 type. Two likely #+ interstitial sites are indicated.

at site (3/4, 3/4, 3/4) and B d i p = 0.27 T at sites (3/4, 1/4, 1/4), (1/4, 1/4, 3/4),
(1/4, 3/4, 1/4), respectively.
The occurence of two precession frequencies cannot be explained by either struc-
tures. Also the signal amplitude ratios (A(0.146 T) : A(0.1 T) : A(0 T) _~ 3 : 2 : 2)
do not correspond to the relative site populations. Since for each signal the damp-
ing rate is found to be quite small and could be explained more or less by just the
random Cu nuclear dipole fields the local fields are quite narrow and would point
to a rather perfect commensurate structure. These results together with the strange
temperature dependence of the 0.1 T signal could indicate that the magnetic struc-
ture is still more complex than a single-q or a quadruple-q structure or that the #+ is
found also at defect sites or changes by its presence the magnetic structure locally.
In view of the fact that basically the same results are also obtained in a second
high quality sample, which showed the 1-K-phase transition, and that all the lines
are rather narrow ( ~ 0.5 mT) we rather tend to believe that what is seen reflects
intrinsic properties.
The first sample investigated showed, as pointed out before, rather small damping
rates which were temperature independent from 3/4 TN down to some 10 mK. ZF-
measurements (limited to T < 2 K) on a second sample, which showed the 1 K
phase transition, produced the same temperature dependence of the spontaneous
frequencies with no break at ~ 1 K. In contrast the relaxation rates for both the
zero frequency signal and the 0.1 T and 0.145 T signals showed a dramatic increase
192 A. SCHENCK and EN. GYGAX

3.0

Z52~o ~,i (c~) 30


20
1.5

1.0 " ( ~ 10
a)
0.5

0.0 0.4
' d. 6 0.8
' '.
10 '
1.2 '
1.4 '.
16 '
1.8 2.0
,3.0
zs X2 (cfz) 3o
2,0
20

,-..,. 1.0 b) 10

Teo 0.5 Z 0:3 =


:::L
v
o.o o'.,1.0' , s'~

g ~
12 T 150

8 ~ 100

5o

2
0 | I l I # f l # I
0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
1.0

0.8 ~[~~i (A.) ~o


0.6 7.5

0.4 5

0.2 . d) 12.5

i m 1 i i i I i
0.0 0.4 0,6 0.8 1.0 1.2 1.4 1.6 15 2.0

T e m p e r a t u r e (K)
Fig. 4.68. Temperature dependence of the relaxation rates associated with the 0.145 T signal (~z, o1),
the 0.1 T signal (,X2,o'2) and the 0 T signal (,~4, o.4) in UCu5. Around 1.1 K the relaxation function
changes from Gaussian (o-) above 1.1 K to exponential (X) below. A further signal with approximately
the same frequency as the 0.145 T signal but an order of magnitude larger relaxation rate ,~s, also above
1.1 K, is attributed to disturbed magnetic domains (from Schenck et al. 1990d).
M U O N SPIN ROTATION SPECTROSCOPY 193

at ~ 1.1 K, as shown in fig. 4.68 (Schenck et al. 1990a, b). In parallel neutron
diffraction showed neither a change in the nuclear nor in the magnetic Bragg peaks
when changing the temperature across the 1 K-transition (Schenck et al. 1990b).
Obviously the antiferromagnetic structure is unchanged by the 1 K-transition, but
the local fields at the #+ sites show a pronounced broadening (NMR results point
to the opening up of a gap in the spin excitation spectrum at ,-~ 1 K, Nakamura et
al. 1990). The specific heat jump at ~ 1 K suggests that the 1 K phase transition
is associated with the heavy electrons. The appearance of obviously quasi random
static fields below 1.1 K may than be linked to the heavy electrons as well and it is
suggested that these electrons enter into some ordered state (random or in form of
a spin density wave) coexisting with the antiferromagnetic order established below
16 K and that these electrons or quasi particles must be distinct from those electrons
responsible for the 'conventional' antiferromagnetic order below TN. In other words
it is suggested that the low temperature behaviour of UCu5 is determined by the
presence of two rather independent electronic substates in the ground state ('heavy'
and 'light' mass states, involving different sections of the Fermi surface?) (Schenck
et al. 1990b). The idea that more than one kind of electrons has to be considered
has come up also for other compounds (see, e.g., Caspary et al. 1993, Feyerherm
et al. 1994c). Assuming a random picture the ordered moment of the heavy quasi
particles in UCu5 has been estimated to be ,-~ 0.01/zB. Subsequently Nakamura et al.
(1994), on the basis of NMR results, have suggested that the 1 K transition reflects
a transition from the quadruple c7structure above 1 K to the single q structure below
1 K. This interpretation is clearly inconsistent with the #SR results.

UCd11. This nonsuperconducting HF compound orders antiferromagnetically below


5 K but the magnetic structure is not known. Neutron scattering investigations place
an upper limit of ~ 1.5#B on the ordered moment (Thompson et al. 1988). In fact
the only direct information on the development of magnetic order at 5 K stems from
#SR-measurements on a polycrystalline sample (Barth et al. 1986c, Barth 1988).
ZF-, TF- and LF-measurements reveal a loss of the signal below 5 K which indicates
the onset of a very wide static field distribution exceeding several 0.01 T. UCd11
crystalize in the cubic BaHgll-structure which is rather complex and provides a
large number of magnetically inequivalent possible sites for the #+. In any case the
ordered moment must be of the order of lPB in order to produce a field spread of
several 0.01 T.
Of interest is the LF-relaxation behaviour above TN. As fig. 4.69 shows a field
independent #+ spin lattice relaxation is observed which follows a power law and
indicates a 5f-spin fluctuation rate of

l]5f O( ( T - T N ) 0"4+0"1 o( v/-T, T > TN. (4.17)


This dependence is typical for a system with Kondo resonance behaviour and is
consistent with the theory of Cox et al. (1985). A fluctuation rate of usf ~ 10 l° s -1
is estimated for T = 2TN. No such behaviour is seen in the other heavy electron
U-compounds. The fact that spin fluctuation become manifest in #SR-measurements
in UCdll is probably related to a relatively small distance between the p+ and a
nearest U-neighbour providing a strong hyperfine coupling (Schenck 1993).
194 A. SCHENCKand EN. GYGAX

I I I I I

UCdll
4 longi1'udinolfield
I000 G
-r 5 o 2000 G
::k ° 4000 G
v

"7 2
F--

0 I I I I I I
2 5 4 5 6 7 8
T/T N
Fig. 4.69. Temperature dependence of the spin lattice relaxation rate )~1 = T11 in UCdll. Note that )~1
is independent of the applied field. The solid line represents a fit of eq. (4.17) to the data (from Barth
et al. 1986c).

U2Znl7. Neutron diffraction shows the onset of a rather simple antiferromagnetic


structure in U2Znl7 below 9.7 K (Cox et al. 1986). Nearest neighbour U-moments
both in the basal plane and along the c-direction in this rhombohedral structure of the
ThEZn17 type are coupled antiferromagnetically and are oriented parallel to the basal
plane. On the basis of this structure one calculates vanishing dipolar fields at the two
high symmetry interstitial sites available to the #+. From the splitting of the TF-#SR
signal in the paramagnetic state, using a single crystal sample, and from the angular
dependence of the frequency shifts it seems that both sites are indeed occupied (Barth
et al. 1986b, 1989b, and unpublished results). However, when cooling the sample
through TN both ZF- and TF-measurements show a partial (~ 25%) loss of the signal
implying that in part of the sample volume a large static internal field spread develops
below TN. This is found in both a polycrystalline and a monocrystalline sample.
From LF decoupling measurements it is deduced that the with of the field spread
amounts to -,~ 0.1 T (Barth et al. 1986b). Interestingly the fraction of implanted #+
exposed to the large field spread below TN shows also a highly peculiar behaviour
above TN. In TF-measurements on a single crystal above TN this fraction is associated
with two (or even four) components distinct by their different frequency (Knight)
shifts and by a highly unusual angular dependence of the shifts (Schenck et al.
1992, and unpublished results). The angular dependence involves higher order (> 2)
Legendre polynomials and is at present not understood at all. It is speculated that
the mysterious behaviour of the 25% fraction is a result of competing interactions
which could also explain the development of same sort of random order in part of
the volume. The question o~f course is what makes this 25%-fraction of the sample
volume so different from the rest both in the ordered and in the paramagnetic regime.
It does not seem to be sample dependent, more-over its size renders it unlikely
MUON SPIN ROTATION SPECTROSCOPY 195

that crystal defects or the presence of foreign phases are responsible. The onset
of the random order coincides precisely with the independently determined N6el
temperature. In view of all this it seems as if the inhomogeneous magnetic features
seen by #SR in U2Znl7 are an intrinsic property.

UBel3, Ul_~Th~Be13, U1_~Th~Be13_vBv. UBe13 is one of the few remaining U-ba-


sed heavy electron systems which have not shown signs for a magnetically ordered
state. Rumors that a transition occurs at ,-~ 9 K could not be substantiated in later
studies. In particular ZF-#SR on UBe13 did not reveal any evidence for a magnetic
state with ordered moments > 10-3#B (Luke et al. 1991b, Heffner et al. 1990).
If more than 1.9% of U is replaced by Th (but less than 4.3%) a second phase
transition at Tc2 somewhat below the superconducting transition temperature Tel is
observed. The actual phase diagram is shown in fig. 4.70. It is argued that below
Tc2 a different type of superconducting phase could be entered (see, e.g., Sigrist
and Rice 1989, Sigrist and Ueda 1991). ZF-#SR measurements established that
the phase below To2 is associated with the development of static random intemal
fields of electronic origin (~ 0.18 roT) (Heffner et al. 1987, 1989a, 1989b, 1990).
This follows from an increased damping rate A of the ZF-signal (given by a static
Gaussian Kubo-Toyabe function) below Tc2. As an example see fig. 4.71, which
collects results from specific heat, ac susceptibility Xae and ZF-#SR measurements
on a U0.965Th0.035Bea3 sample. Clearly visible is the rise of ,4 at Tc2, which is
determined from the strong anomaly in the specific heat. Tel is determined by the
diamagnetic response of Xac. The rise of ,4 below Tc2 is well described by a spin 1/2
Brillouin function. Assuming a random moment order and making some educated

I I I I I I I
0.9

-~ Ul"xThxBe13

0.5

f MAGNETIC
0.0 0.I 0 I
2.0 ~
4.I 0 I I
6.0
x (%)

Fig. 4.70. Phase diagram of U1_xThxBe13 (from Heffner et al. 1990).


196 A. SCHENCK and EN. GYGAX

0.30 a) l I I I I I I I I
-!
S"
I U 0.965 T h 0.035 B e 13 -
0')
0.28 Ho = 0 Oe -
::L

13
0.26

0.24

I I I I I I 1 1 I
Y b)
,
&
-6
E
2.0
<
w
i
0 1.0 -
u_
0

o3 o I I I I I I I I I

0"" c)
U •e.~t~l k

-2 -

-4 - m

-6 --

"8 - -

-lO I I I 1• I I I l t
0 0.2 0.4 0.6 0.8 1.0

TEMPERATURE (K)

Fig. 4.71. Temperature dependence of # + ZF relaxation rate O'KT, specific heat cp and ac susceptibility
X~c in Uo.965Th0.035BeI3 (from Heffner et al. 1989b).

guess on the #+-site the ordered moment is estimated to amount to ~ 5 x 10-3/ZB


(Schenck 1993). As we have seen a similar behaviour was also observed in UPt3.
But ultra-small moment ordering is also seen in non superconducting heavy electron
systems (e.g., CeRuzSie). The question is then whether different mechanisms have
to be involved to explain these results. In the superconducting systems it has been
suggested that the phase below Tc2 (in UPt3 and UI_=Th=Bel3) could violate time
reversal invariance and would be associated with spontaneously appearing magnetic
fields (Sigrist and Ueda 1991, Heffner et al. 1990). Alternatively the small moment
MUON SPIN ROTATION SPECTROSCOPY 197

magnetism is unrelated to superconductivity and is a phenomenon not uncommon


for a wider class of heavy-electron systems.
Finally we mention that substitution of B for Be, although it has an effect on the
entropy released at Tel, does not interfere with the small moment order induced by
Th doping (Heffner et al. 1991).

URu2Si2. Also this moderately heavy fermion superconductor displays small mo-
ment magnetic order below TN = 17.5 K. The phase transition is accompanied by a
relatively huge jump in the specific heat but the ordered moment amounts to only
0.04#B as determined by neutron and X-ray diffraction measurements. (Broholm
et al. 1987, Mason et al. 1990, Isaacs et al. 1990). The same studies revealed
an antiferromagnetic structure of type I with the AF propagation vector given by
( = (0, 0, 1) and the static moment aligned along the c-axis. ZF-#SR measurements
reflect clearly the transition into the ordered state (McLaughlin et al. 1988, Luke et
al. 1990c, Knetsch et al. 1993). Figure 4.72 displays results on the ZF-relaxation
rate in a monocrystalline sample for P(0)±c-axis. At TN the relaxation rate rises
steeply within a few degrees followed by a much weaker increase as the temperature
is reduced further. This temperature dependence does not follow the temperature de-
pendence of the ordered moment as measured by neutron scattering (Broholm 1989,
Mason et al. 1990) a feature not understood at present. Adopting the proposed anti-
ferromagnetic structure one calculates dipolar fields of the order of 5 mT, which are
parallel to the c-axis. Only at the site (1/4, 1/4, 1/4) one finds Bdi p ~ 0. The internal
fields should thus be perpendicular to/5(0) and, except for the (1/4, 1/4, 1/4)-site
one expects to see a clear spin precession signal for P(0)±c-axis. However, no such
signal is observed. If the slow relaxation seen instead is interpreted as the beginning
of a cos oJr modulation on extracts B , _~ 0.2 mT. The neutron scattering data and

0.20
"'I" l""l";'l~"'l""l't"l '''

0.16 URu2Si 2
-.>
ZF, P I ~
09

0.12
TN
t-
O

0.08
¢1
n"
0.04 ¢~
¢
0.00 ,,,I,, l,,,,l,,,,l,t,,l,,,,l,,,,~P,,,
0 5 10 15 20 25 30 35 40
T (K)
Fig. 4.72. Temperature dependence of the ZF-relaxation rate ),ZF in URu2Si 2. Note the discontinuity at
TN (from Knetsch et al. 1993).
198 A. SCHENCK and F.N. GYGAX

the #SR-data, therefore, appear totally inconsistent. Several possibilities are implied
by this inconsistency:
(i) the #+ resides at the site (1/4, 1/4, 1/4) where B~ip = 0: this possibility
is questionable in view of the results in URh2Si2, which clearly rule out the site
(1/4, 1/4, 1/4) in this homologous compound (see section 4.3.4);
(ii) the #+ quenches somehow the 5f-moments at the nearest neighbour U-sites
(MacLaughlin et al. 1988); again this possibility is not supported by the #SR results
in URh2Si2 which seem to be fully determined by the known AF-structure;
(iii) magnetic structure and the magnitude of the ordered moments are not deter-
mined correctly by the neutron and X-ray scattering experiments. Also this is not a
very likely possibility. Hence the #SR-results provide an unsettled problem.
It should be noted that the small moment magnetic order is embedded in a strong
paramagnetic background as evidenced by the magnetic susceptibility and the #+
Knight shift below TN (Knetsch et al. 1993, and unpublished results).

UNi2AI3. Similar to UPt3 and URu2Si2 this heavy electron system orders first mag-
netically below ,-~ 5 K and becomes superconducting inside the antiferromagnetic
phase below ~ 1 K. ZF-#SR-measurements were the first to demonstrate that in-
deed magnetic order was established and that it was unaffected by the transition
into the superconducting phase. (Amato et al. 1992c). The same study provided
an estimate of the ordered moment of --~ 0.1#B. A typical ZF-signal is shown in
fig. 4.73. This signal was interpreted in terms of a multisite occupancy and a simple
antiferromagnetic structure with the moments aligned parallel to the c-axis of this
hexagonal system of the same type as the RNi5 compounds. Alternatively Uemura
et al. (1993) proposed that the #SR-data could also reflect an incommensurate spin

0.0

c-
.O
m

N
-10.

Q_

-20. r I i I i I r I =
0 1 2 3 4 5 6

Time (l.~sec)

Fig. 4.73. ZF-/zSR signal in polycrystalline UNi2A13 at 0.3 K (from Amato et al. 1992c).
MUON SPIN ROTATIONSPECTROSCOPY 199

density wave order. Indeed later neutron scattering measurements revealed a long
range incommensurate AF-order with a wave vector 0'= (1/2 + 3, 0, 1/2) (3 = 0.11)
and a maximum ordered moment of # ~_ 0.24#B (Lussier et al. 1994). Whether the
#SR-data and the neutron results are consistent with each other has to await a better
understanding of the #+ site in this system.

UPdzAI3. Among all heavy electron superconductors this system is special in that it
displays the highest superconducting transition temperature (~ 2 K) in coexistence
with the largest ordered moment (,-~ 0.85#B ). The magnetic structure has been
determined by neutron scattering (Krimmel et al. 1992). ZF-#SR measurements
imply that superconductivity and magnetic order coexist microscopically (Amato et
al. 1992d, Feyerherm et al. 1994c). However, no information on the magnetic
structure could be determined from #SR-measurements due to the fact that the #+
resides at the high symmetry b-site (see fig. 4.25) at which the net internal field in
the ordered state is zero. The #+-site has been determined unambigeously from the
angular dependence at the #+ Knight shift and the A1 nuclear dipole field induced
relaxation rate (Feyerherm et al. 1994c).

5. Review of results in magnetic insulators

In contrast to the #+ in a metal, a positive muon implanted into an insulator represents


a more complex system. Not only is it possible to encounter the formation of
muonium and the concomitant - more complex - #SR signal (or the corresponding
missing fraction), but other chemical effects may be of particular importance and
influence the #+ in its localization and its coupling to the environment. For detailed
descriptions of the situation see, e.g., Schenck (1985) and Patterson (1988).
A reference list and an overview of the magnetic insulator systems studied by
#SR spectroscopy is presented in table 5.1. (This table and the present chapter
do not embrace the substances related to high-temperature superconductors, which
are the subject of chapter 6.) Because the #SR studies are more complicated for
magnetic insulators, they are also more fragmentary and less systematic than the ones
on elemental metals. Most of the examined substances are oxides, and we divide
these schematically in crystals with corundum-type structures, rare-earth orthoferrites
with perovskite-type structures and rare-earth (R) perovskites RNiO3, CuO or related
compounds, and MnO. The studied non-oxide magnetic insulators are fluorides and
the COC12.2H20 compound. Finally the case of the antiferromagnetic molecular
crystal a-O2 is also presented. Disregarding perhaps Cr203 and MnO, which have
not been particularly well studied in this respect, it seems established that for all other
oxides below approximately 500 K the occurrence of oxygen-muon bonds ('muoxyle
bridges') plays an essential role in the localization of the positive muons - see, e.g.,
Boekema (1984), Boekema et al. (1985), Chan et al. (1986) and Lin et al. (1986).

5.1. Oxides with corundum-type structure


~-Fe203. In this compound a spontaneous #SR signal has been observed for the
first time in an antiferromagnet (Graf et al. 1978). Consecutively the same research
200 A. SCHENCK and EN. G Y G A X

0
~} ci
¢}

~ {~
,,_., o o ~2, .~ -~• ~ .~" .i~" . . ~ ~ N N N
V V ~ ~

N N M
0
{i}
U

r,D o

~a

.~ .~ : ~ ~ .~ ~ ::i4 .~ : ~ :~
{.}

"2".

v ~ N

S NN b~b~N ~'~'NN

~.~

II ~ II ~ II ~ II ~ II ~ II ~ II ~ II ~ II ~

~ ~.~
~.,~
c~

o.~ o~" ~E~~ =o~o~


~ c~ ~ 8~ ~° c?
MUON SPIN ROTATION SPECTROSCOPY 201

~ z~
,.., __ o~ "--" ~ V -.~

8

8
" ~ =

•~ : ~ . ~ - .~ --~ •. ~ .~ ~ . . ~ .~

"-b

N >
N

~ N
t¢3

..~

~ °
d

II ~ II ~ ~t II ~ II M II

Z
,::? rj r~

r,.) 6 = rj
202 A. S C H E N C K and EN. G Y G A X

o,
O~
~v ~

r...., r~
,-c,

~ E...-r''
ca ..-,...T, "-"

aa

V V [~
N N z
A~Av 4 4~v~ " v/ v/

© 0 0

O
"0
t.,q "~
0

0
0
MUON SPIN ROTATION SPECTROSCOPY 203

group (Rtiegg et al. 1981) noticed that below 120 K the spontaneous signal was
in fact split into three components - see fig. 5.1. Between 120 and 500 K a single
frequency signal remains and the corresponding Bu follows the magnetization curve.
At the Morin temperature, TM = 263 K, the signal shows that the local field at the
#+ changes by a factor 2.15, due to the iron spin reorientation in the host, but the
signal shows no discontinuity in the relaxation rate. The collapse of the three lines
into one at about 120 K seems to correspond to a transition of three distinct muon
states into a single one. This could be an indication for the onset of jumps between
different potential minima for the muon in the unit cell.
Potential energy calculations in c~-Fe203 (and in the non-magnetic compound c~-
A1203) favour particular sites for the #+, all located about 1 A away from their
nearest neighbour, an oxygen ion. In such an oxygen-muon bond configuration
the contact hyperfine field at the #+ may not only result from a direct overlap of
the wavefunction tails of the magnetic electron and the #+, but supertransfer of
hyperfine fields from a magnetic cation via the oxygen anion to the #+ can take
place, depending on the geometry and the wave functions of the ions. This field/3c
will add to the dipolar field to produce the local field/31oc acting on the #+:

Bloc = /~c + J~dip. (5.1)


/~dip@*/~) has to be calculated, e.g., as described by Denison et al. (1979) (notice
that for an antiferromagnetic sample J~dip = /~dip)" Thus, a comparison of the results
of hyperfine field calculations in this covalency-effect scheme, assuming the various
proposed/z + sites as starting guides, will help to find the #+ states by requiring
overall consistency (Boekema 1984). So far, however, the data allow more than one
unique interpretation for the #+ localization.

250 ' ' ' 1230' ' '

200 ~ 220 15
16

~" 150 - 1 2 1 0
I,..-,,-,.'4,.
50 100 150 200
:::L
>=- 100 m

50 TM ,,\T~

..... 3 0
0 I I i I i I I "I- -I---
0 200 400 600 800
Temperature (K)
Fig. 5.1. Temperature dependence of the zero-field /zSR frequencies and the local fields measured in
single-crystal a-Fe203. The low temperature data are expanded in the insert (Rtiegg et al. 1981).
204 A. SCHENCK and EN. GYGAX

The direction of the local field/31oc with respect to the crystalline axes was deter-
mined by applying an external field with various orientations relative to the crystal
and measuring IB, I. Since the antiferromagnetic structure is not destroyed if the
applied field is not too strong, the demagnetization field remains zero and the/z +
sees a field given by the vector equation

/~/z ~-~ /~loc 4- /~ext. (5.2)

With the correct assumption for the orientation of/31oc relative to ]~ext the absolute
value I~.1
deduced from eq. (5.2) will match the experimental value. Below the
Morin temperature it was observed that each frequency splits into two according
to

IS~l = Bext 4- IB~o~l (5.3)


for/~ext parallel to the c-axis, showing the alinement of/~lo~ along the c-axis. At
room temperature, well above TM, B]o~ is found to be perpendicular to the c-axis,
probably parallel to the crystalline a-axis.
The disappearance of the remaining spontaneous signal at 500 K and the reap-
pearance of a signal when applying an external field around 750 K is explained by
means of #+ diffusion (Rtiegg et al. 1981).
Examining the scheme for supertransfer of hyperfine fields it appears that Bo is
parallel to the 3d-moments and that in contrast to Bdip the field/3c is not changing its
value at the Morin temperature (for a discussion see Boekema 1984 and references
therein).

Cr203. In zero field two #SR frequencies are observed at low temperature, one
of them disappears around 150 K and the other is present up to TN. Both follow
approximatively the behaviour of the lattice magnetization (fig. 5.2). Part of the

~ i I- 8
1 O0 • Single crystal _
o Powder
80 - 6

g 60
4 m=
4o
2
20

0 I I I I I I 0
0 1 O0 200 300
T e m p e r a t u r e (K)
Fig. 5.2. Temperature dependence of the zero-field /~SR frequencies and the local fields measured in
Cr203. The powder-sample data scatter more than the single-crystaldata (Rilegg 1981).
MUON SPIN ROTATION SPECTROSCOPY 205
' I i I ' I i I

120

110

~.100

9o

80

70

60
i I t I i I I I i
1 2 3 4

Bext (kG)

Fig. 5.3. Field dependence of the transverse-field /zSR frequencies measured in single-crystal Cr203
at 130 K. /3ext is applied parallel to the crystalline c-axis. The solid lines are results of calculations
assuming various values for the angle a between/31oc and the basal plane. The lines correspond (from
top to bottom) to values a = +9 °, +1 ° and - 9 ° for the upper triplet, and a = +24 ° and -24 ° for the
lower doublet (Rilegg 1981).

muons observe a vanishing local field (see Rtiegg et al. 1979). The Cr203 /zSR-
signals differ sensibly from those observed in c~-Fe203: in the former system they
are weaker - with smaller asymmetries and higher relaxation rates.
Due among other things to the different spin structures of Cr203 and ~-Fe203,
the local fields at the #+ sites have different orientations. Transverse field #SR
measurements in Cr203 (see, e.g., fig. 5.3) show that/31oc is at an angle o~ with
respect to the basal plane. This angle amounts to ± 2 4 ° for the lower frequency
line, and to + 9 °, +1 ° and - 9 ° for the upper one (see the splitting of the two lines
showing up when/3ext is applied parallel to the c-axis). The azimuthal orientations
of Blo~ are also determined for the two frequencies and the various values of the
angle ~ (Rtiegg 1981). As for c~-Fe203, the question of the muon site is not solved
in a unique way. Also #+ trapping at a defect site cannot be ruled out - the lower
#SR frequency in Cr203, e.g., can best be reproduced assuming a trapped muon near
an non-magnetic A13+ ion substituting a Cr 3+ ion (Rtiegg 1981).

FeTi03. Transverse-field #SR measurements (Boekema et al. 1983) showed that


above TN free muon-like behaviour was observed, whereas below TN the muons are
experiencing local internal fields. Zero-field measurements below TN showed two
weak but observable signals (fig. 5.4), following approximatively the magnetization
curves for the Fe z+ ions. The values extrapolated to T = 0 K of the internal fields
amount to 2.0 and 3.4 T respectively. Relatively large covalent contributions are
206 A. SCHENCKand EN. GYGAX
i i i i i i i
500
.~- . . . . @
400

~" 300
, O____O_O...._. O',,,,
%" 200
• ~'.~% %%

100

I I I I I "~ I I
0 10 20 30 40 50 60 70
Temperature (K)
Fig. 5.4. The frequenciesof the zero-field#SR signal in FeTiO3 at low temperature. For comparison,
calculated magnetizationcurves for Fe2+ ions are drawn(Boekemaet al. 1983).

expected and a rough estimate, taking also dipolar contributions into account, yields
a field interval corresponding to #SR frequencies extending from 200 to 500 MHz,
in agreement with the measurements (Boekema et al. 1983). Precise calculations
were found difficult because possible effects of local #+ motion occurring at low
temperature.

Fe304. This is a ferrimagnetic oxide (TFN = 858 K) that undergoes a semimetal-to-


insulator transition at the Verwey temperature (Tv) near 121 K. In the course of a
series of #SR studies (Boekema 1980, Boekema et al. 1985 and 1986) an anomalous
change in local field and depolarization rate was observed at 247 K. The temperature
dependences of the zero field data (spontaneous frequency and relaxation rate) are
shown in fig. 5.5. The frequency follows essentially the bulk magnetization, but
a small offset is clearly visible for the interval starting at Tv and ending at about
247 K. The local field is observed to be directed along the (111) direction, which
is the easy axis of magnetization. The experiments performed with an external
field greater than the demagnetization field and applied along the (110) direction
(Boekema et al. 1986) showed that as the temperature is decreased below 247 K,
the #SR frequency line splits, indicating the onset of two local fields, i.e. two
magnetic inequivalent sites. This supports the model of a phase transition involving
the onset of a short-range order already nearly 130 K above the well known Verwey
transition. In that sense the anomaly at 247 K can be viewed as a precursor of the
semimetal-to-insulator transition.

V203. This oxide, which possesses interesting electronic properties, has been the
latest member of the group of the corundum structured sesquioxides studied by #SR.
Although V203 has the basic corundum structure in the high temperature phase where
it is a paramagnetic metal, it undergoes a combined structural, magnetic and electrical
phase transition below 155 K to a monoclinic antiferromagnetic insulating phase. At
low temperature the V 3+ moments are aligned at +71 ° with the corundum c-axis, in
alternating ferromagnetic planes normal to the monoclinic [0 1 0] or hexagonal [1 1 0]
MUON SPIN ROTATION SPECTROSCOPY 207

40 I i I I i I i

(a)
30
70
,.(
60
O "O.~ '"'"~
to 20 50
|%. 40 -1-
-'-I | ".
"q 30
10 .o,
20
"O,, "'
""lit,.. _ "... 10
0 I I I I. I 0
0 200 400 600 800 1000
Temperature (K)

70 I I i I
5.0
(b)
65
(.9
4.5 v
-1- 60 :=L
1213

55
~4.0
rv
50 I I ~ I I I
50 100 150 200 250
Temperature (K)
Fig. 5.5. Temperature dependence of the frequency and the relaxation rate of the zero-field /~SR sig-
nal observed in Fe304 single crystals (a), and detailed frequency data for the Verwey-phase-transition
temperature region (b) - from Boekema et al. (1985).

direction. In V2O 3 one can ignore the covalent contributions to the field observed
by a muon bonded to an oxygen and treat the internal field as purely dipolar to a
very good approximation (Chart et al. 1986). Zero-field #SR data were obtained
first by Uemura et al. (1984) and then by Denison et al. (1985). Below TN one
oscillating signal following the trend of the magnetization curve appears in addition
to the signal continuing to have a zero frequency (fig. 5.6). Data taken (Denison
et al. 1985) with an external field applied along the corundum c-axis show that
the oscillating signal splits into four, corresponding to orientations of the internal
field at 7 °, 66 °, 116 ° and 174 ° (±5 °) with respect to the c-axis in the absence of
an applied field (fig. 5.7). The magnitude of this internal field is about 0.11 T.
The measured relaxation rates clearly indicate the absence of diffusion, even above
208 A. SCHENCK and EN. GYGAX

20.0 i i i i

15.0 "° "o'-'@-O..o4...i,~

g
10.0

5.0

0.0 50.0 100.0 150.0 200.0


Temperature (K)
Fig. 5.6. Temperature dependence of the zero-field #SR frequencies measured in V203. Note signals
at zero frequency below the transition temperature at 134 K. This transition temperature shows a well
known hysteresis, not marked on this figure (Denison et al. 1985).

32.0 ' i i i i
/

.o_
24.0 -
S /
g 200
16.0

12.0

8.0

4.0

0.0 I I I I I
0.0 0.25 0.50 0.75 1.00 1.25 .50
Bext (kG)
Fig. 5.7. Transverse-field /zSR frequencies measured as function of Bext applied along the hexagonal
e-axis of a V203 single crystal at 65 K. The solid lines represent the calculated frequencies for the net
field I#~1, where 0 is the angle between the original/3]oc and the c-axis (Denison et al. 1985).

the transition temperature of 134 K. Let's mention at that point that the transition
temperature around 150 K shows a well k n o w n hysteresis, observed also in the # S R
signal ( U e m u r a et al. 1984).
In V203 only a subset o f the expected #+ sites in the c o r u n d u m structured oxides
is occupied ( B o e k e m a et al. 1986). This is viewed as a signature for the vanadium
pairing interaction suggested by G o o d e n o u g h (1963) as the mechanism responsible
MUON SPIN ROTATIONSPECTROSCOPY 209

for the structural, magnetic, and electrical phase transition in V203. In this picture
the #+ cannot occupy the interstitial sites of reduced volume resulting from the V 3+
ion displacement in the corundum to monoclinic structural transition. Hence, #SR
provides corroborative evidence for the mentioned phase-transition model.

5.2. Orthoferrites and RNi03 perovskites


The rare-earth (R) orthoferrites (RFeO3) and the RNiO3 compounds crystallize all
with the perovskite-type orthorhombic Pbnm (D 16) structure. The sublattice of the
Fe or Ni ions, octahedrally surrounded by oxygen ions, is nearly cubic. Though
crystallographically similar, the orthoferrites show a rich variety of magnetic prop-
erties (White 1969). The iron spins are coupled essentially antiferromagnetically,
with Ntel temperatures in the range 740-620 K, whereas the spins of the rare-earth
ions remain unordered down to typically 4 K. Due to a small spin canting, these
compounds behave as weak ferromagnets.
A systematic #SR study has been performed in six different orthoferrites (table 5.1)
and reported by Holzschuh et al. (1980, 1981), and in the comprehensive article by

I I' I

100 F ~ "
SmFeO 3
50
25 - 1"2"41-]74
0 = i
I I I I IQ---O-'O'-[ "~

25
= ,,-,._,,_., EuFe03 (F4)
• --~1~--~0~0-- 4
I I I I I I
Cor
" }-£4 " - * " " ~ " DyFe03
2 5 I }-F4
I 0 1 I I I I t

>=. 50 _ . _ ~ .~ . _ . YFeO 3 (F4)


25
0 I I I I I

50
25 HoFeO 3
Fr4
O. I I I I I

100"
75 ErFeO a
50
25 r2~4Fr4
0 I I I l I

0 100 200 300 400 500 600


Temperature (K)
Fig. 5.8. Measuredzero-field/~SRfrequenciesfor the rare-earth orthoferriteseries. The solidlines only
connect the data points. Fe-spin configurationsare indicated(Holzschuhet al. 1983).
210 A. SCHENCK and EN. GYGAX

Holzschuh et al. (1983). In fig. 5.8 the temperature dependences of the zero-field
#SR frequencies are summarized. At room temperature only one signal is observed,
whereas at low temperature two or three signals are seen - except for SmFeO3 and
DyFeO3. As in the case of a-Fe203, and possibly as in Cr203 and FeTiO3, this
suggests that different sites in the unit cell will be occupied by the muon at 0 K. At
higher temperature local #+ motion or hopping between these sites occurs.
The Fe spins of the studied orthoferrites are at the higher temperatures in the
so-called F4 configuration (notation of Bertaut 1963), i.e. basically aligned along
the crystalline -t-z-direction, disregarding small canting angles of the order of 0.5 ° .
At the lower temperatures the spin configuration is different, except for EuFeO3 and
YFeO3. There, for SmFeO3 and ErFeO3 the 1"2 configuration is encountered, with
now spins parallel to the -t-z-direction, however again with small canting angles.
The temperature domains characterized by/"2 and/"4 configurations are marked in
fig. 5.8. The discontinuities in the #SR frequencies in SmFeO3 and ErFeO3 are
obviously related to the/"2-/"4 transition. In the other compounds the disappearance
of the higher frequency components at certain temperature cannot be related to any of
the known properties of the substances. Holzschuh et al. (1983) state that the higher
frequency lines are associated with metastable #+ states, which decay with raising
temperature to a ground state configuration associated with the lower frequency line.
For this latter state a #+ site has been determined. This was done comparing the
results of refined dipolar field calculations with a thorough set of zero-field as well
as transverse-field #SR data, including the measurement of field and single-crystal
orientation dependences - see, e.g., fig. 5.9.
This most stable #+ site is located in the rare-earth-oxygen plane (z = 1/4 mirror
plane), at 1 A of the O-ion, practically at the center of the slightly distorted square
formed by the rare-earth ions. The distance to the O-ion corresponds to the bond
length in the (OH)- molecule, 271oc is only of dipolar origin, Bc and hence the
supertransfer of hyperfine fields can be neglected.

66 t I
YFeO3 I I I CI I I I t
Bext = 4kG Bext
T = 296 K

0 60 120 180 240


Angle ~ (deg.)
Fig. 5.9. Measured orientation dependence of the transverse-field/zSR frequencies for YFeO 3 with an
applied field of 0.4 T in the a-b plane. The solid lines are from a fit of a field-induced canting model.
The results are used for the/*+-site determination (Holzschuh et al. 1983).
MUON SPIN ROTATION SPECTROSCOPY 211

Lin et al. (1986) have re-evaluated the data in these compounds (Holzschuh et al.
1983) and pushed further the muon site search. They first confirm the findings of
Holzschuh et al. (1983) concerning the most stable site. They observe, however, that
the proposed site splits into two different subsites (in terms of magnetic structure),
leading to the same frequencies in SmFeO3 and ErFeO3, but to the two slightly
different frequencies in YFeO3 and HoFeO3 (the sites are now called 1 for the
ground state and 2 for the first metastable state). In addition Lin et al. (1986) find a
new site candidate, corresponding to the remaining unaccountable frequency line in
EuFeO3, YFeO3, and ErFeO3. This site ~site 3) lies symmetrically, approximatively
above and below site 2, at the usual 1 A bond distance of the O-ion also bonded
to site 2. New for site 3 is the observation that neglecting supertransfer, effects is
not nearly as good as for sites 1 and 2, situated in the mirror plane. A precise
re-evaluation including this covalent hyperfine field has not yet been performed.
Torrance et al. (1992) have performed a systematic study of the insulator-metal and
the magnetic transitions in the perovskites RNi03 (R = Pr, Nd, Sm, Eu) - fig. 5.10.

Rare Earth Ionic Radius (angstr6ms)


.00 1.10 1.20
500 i i ~ i _ i
Transitions in O Eu ~ ~ ~ \
.-. Mago:
~- 4 0 0 - h , k (Z) t@Srn " ~ ~ r ~ ( ~ / ~ _
•¢
'm [-],m (n-soatt.) ~ O1~(~)~

¢~ 3 0 0 -
{3..

INSULATOR . ~ " ~"J- ~'~


I~ Nd "
c 2OO
.o -L y / Eu METAL
-

o0
¢- r

,~ 100 - ANTIFERROMAGN. ~ -

0
0.86
INSULATOR

0.88
~
0.90
1 I
0.92
\ I
0.94
Tolerance Factor
Fig. 5.10, Insulator-metal-antiferromagnetic phase diagram for RNiO3 as a function of the tolerance
factor and (equivalently) the ionic radius of the rare earth (R). The observed insulator-metal transitions
(resistivity or differential calorimetry measurements) are indicated by large open symbols, whereas the
magnetic transitions are represented by three different solid symbols, according to the used method:
susceptibility, n-scattering or /xSR measurements (left inset). Additional ~SR measurements of the
AF-metal transition (not shown) are in excellent agreement with those from conductivity and neutron
measurements. The structure of the RNiO3 compounds is shown schematically in the right inset - see
text (from Torrance et al. 1992).
212 A. SCHENCKand EN. GYGAX

For the Eu and Sm compounds as well as for the solid solutions Sml_xNdxNiO3
the antiferromagnetic ordering-temperature data are obtained from TF-#SR measure-
ments only. For the Nd and Pr compounds and for the solid solutions Ndl_xLa~NiO3
and Prl_~La~NiO3 the transition temperatures obtained from #SR and neutron scat-
tering or conductivity measurements are in perfect agreement.
The GdFeO3 structure of the RNiO3 compounds (Demazeau et al. 1971, Lacorre
et al. 1991) is shown schematically in the inset of fig. 5.10. Regular NiO6 octa-
hedra share comers to form a three-dimensional array, with the R ions occupying
the space between these octahedra. In this structure the Ni-O-Ni bond angle, 0,
plays an important role, since the electronic bandwidth and the magnetic exchange
interaction are closely related to cos 0 (see, e.g., Sawatzky et al. 1976). This angle
is generally less than 180 ° because of the orthorhombic distortion, which is conven-
tionally discussed in terms of the tolerance factor, defined as t =_(dR-o)/X/~(dNi-O).
If the rare-earth ions were large enough to give t = 1, the rare-earth-oxygen bond
lengths (dR-o) and nickel-oxygen bond lengths (dNi-o) would be compatible with
the ideal cubic perovskite structure, i.e with 0 = 180 °. Since the rare-earth ions
are too small to satisfy this criterion, the structure becomes distorted as the NiO6
octahedra tilt and rotate in order to fill the extra space otherwise present around
the rare-earth ion. The distortion tends to be slightly reduced as the temperature is
increased.
Torrance et al. (1992) draw the general phase diagram (fig. 5.10) for the RNiO3
series as function of the tolerance factor t. It appears clearly that the transitions
observed form a coherent pattern. These transitions separate three distinct regimes:
an antiferromagnetic insulator, a paramagnetic insulator and a metal. The observed
insulator-metal transition depends strongly on R. For small R it occurs a at higher
temperature than the antiferromagnetic ordering. The observations are discussed in
the framework of a picture developed by Zaanen et al. (1985) and (1990), accord-
ing to which there are two general types of band gaps possible: the Mott-Hubbard
gap due to the Coulomb correlation energy U and the charge-transfer gap associ-
ated with an energy A. The insulator-metal transition in RNiO3 is most probably
caused by the closing of the charge-transfer gap, induced by an increase in the
electronic bandwidth either as a function of increasing temperature or ionic radius
of R.

5.3. Miscellaneous mostly Cu-based and layered oxides


Various #SR studies have been undertaken on CuO and other copper-oxide com-
pounds related to high-To superconductors. The review concerning this latter class
of materials is however mainly included in chapter 6.

CuO, BaCu02, BaY2Cu05. (See also BaCuO2 and BaY2CuO5 in section 6.2.6.) Wei-
dinger et al. (1988) have found magnetic ordering in CuO, BaCuO2 and BaYzCuO5
('green phase' of Y-Ba-Cu-O) indicated by well defined muon spin precession fre-
quencies in zero-field measurements. A single precession frequency was observed
in BaCuO2, whereas 5 frequencies were seen in both CuO and BaYeCuOs, cor-
responding to different muon stopping sites. A TN of 226 K was found in CuO,
MUON SPIN ROTATIONSPECTROSCOPY 213
Frequency (MHz)
0 10
i
2O 30
(a)
0
O_

.r-

0
I.t_

(b) 250
30
200

"r" 1-

~v20 15o E
>,.,, e
rrl
,o
,t"-

0"
100
~10
t.l_

50

,, ', ~ Ii ; ', ', ', Ii ~ ', ~ ', I ', ,,,I,


.....

(c) [

I .... I,,

...... i 6 o 200
Temperature (K)
Fig. 5.11. Zero-field #SR measurements in CuO: (a) typical Fourier-transform spectrum at 43 K;
(b) temperature dependence of spontaneous frequencies compared with S = 1/2 molecular-field model;
(c) muon depolarization rates. The five distinct lines, characterized by five corresponding point symbols
in (b) and (c), originate from different muon sites (from Niedermayer et al. 1988).

whereas ordering temperatures of only 11 K for BaCuO2 and 15 K for BaY2CuO5


were observed.
A more complete report on the magnetic properties of CuO, studied by #SR
as well as 57Fe M~3ssbauer-source spectroscopy (MS) and l°°Rh perturbed angular
214 A. SCHENCK and EN. GYGAX

correlation (PAC), is given by Niedermayer et al. (1988). For the interpretation of


the zero-field #SR results in magnetically ordered CuO, it is assumed that the fields
at the muon sites are due to a sum of dipolar fields and supertransferred hyperfine
interactions via the # + - 0 2 - bond, as, e.g., in c~-Fe203 (see section 5.1). As shown
in fig. 5.11, below 60 K four distinct signals are seen, two disappearing at 60-80 K,
and above 100 K, an additional signal is populated. The various components of the
signal sum to the full original muon polarization, indicating that the precession of all
implanted muons is observed. Clearly, antiferromagnetic order is detected by #SR
below TN ,-~ 226 K; this is also indicated by MS and PAC. The muon-depolarization
anomalies detected below TN (fig. 5.11c) are interpreted as indicating a change in
muon localization produced by diffusion from metastable traps at low temperature to
stable traps at higher temperature. The #SR fields that survive to TN are reasonably
well approximated by a S = 1/2 molecular-field model.
A high-quality polycrystalline CuO sample has also been studied with transverse-
and zero-field #SR by Duginov et al. (1994a). The data confirm most of the
zero-field features reported by Niedermayer et al. (1988). However, the highest
frequency line observed by Niedermayer et al. (1988) above 100 K is not seen
and Duginov et al. (1994a) claim that no #+ polarization is missing in the signals
they observe. In the temperature region of the incommensurate antiferromagnetic
structure detected with neutron diffraction measurements by Forsyth et al. (1988),
between 213 K and TN, no spontaneous #+ precession is seen. A peak in the
relaxation rate of the slowly relaxing component of the zero-field signal at 219 K is
supposed to be connected with the strong development of dynamic correlations near
the incommensurate-commensurate transition.

Cao.86Sro.14Cu02, Sr2Cu03, La2fu04_ v, Sr2fbt02Cl2. The infinite-layer compound


Cao.86Sr0.14CuO2 and the infinite-chain compound Sr2CuO3 have been studied by
#SR by Keren et al. (1993) and compared to the La2CuO4_y and Sr2CuO2C12 sys-
tems (section 6.1). In Cao.86Sr0.14CuO 2 spin precession is observed below 360 K
- above that temperature a rapid #+ depolarization due to the onset of #+ diffu-
sion occurs. The measured #SR frequency permits to extend nicely the sublattice-
magnetization curve observed with neutron scattering to lower temperature. Below
225 K the frequency curve splits into two lines approaching 17.9 MHz and 14.3
MHz respectively for T --+ 0, and Keren et al. (1993) discuss also the possible
corresponding #+ sites. The sublattice magnetization Ms shows a slower decay
with increasing temperature in Cao.86Sr0.14CuO2, compared with that observed in
La2CuO4_u and Sr2CuO2C12 (see fig. 6.19 in section 6), indicating that a wider
CuQ-layer separation results in more two-dimensional magnetic behavior. Calcula-
tions show that Cao.86Sr0.14CuO 2 (dcu O = 3.20 A) is more appropriately described
by a 3D model whereas Sr2CuO2C12 (dcuo = 7.76 A) is a very good example of a
quasi-2D Heisenberg system.
In SrzCuO3 ZF #SR oscillations were seen below 4.15 K, and the Ntel temperature
determined as 4.15K < TN < 6 K; the sublattice magnetization was followed down
to ,-~ 10 mK (Keren et al. 1993). Again two frequencies are present. The exchange
interaction, inferred from susceptibility measurements, is of the order of J = 103 K,
MUON SPIN ROTATIONSPECTROSCOPY 215

implying a remarkable suppression of the ordering temperature with kBTN/J ~ 0.01


in SrzCuO3. The result demonstrates a clear signature of low-dimensional magnetic
behavior in the CuO chains. For further discussion see also Uemura et al. (1994).

Y2Cu205. Duginov et al. (1994b) have performed ZF and TF #SR measurements in


Y2Cu205 ('blue phase' in the Y-Ba-Cu-O family). In ZF the temperature depen-
dence of Bu has been measured. In addition to the well established AF ordering
of the Cu 2+ moments at TN1 = 13 K the authors find the indication of a second
transition at TN2 = 7.5 K. This can be interpreted as a steady change in orientation
of the magnetic copper moments with decreasing temperature, starting at 7.5 K.

c~-Bi203. NQR and #SR measurements have been performed in c~-Bi203 by Duginov
et al. (1994c). ~-Bi203 is usually considered as diamagnetic, but a splitting of the
NQR spectral lines and the internal fields observed by #SR can be explained by the
bonds in c~-Bi203 being of partially covalent nature. According to this explication
not only 6s- and 6p-electrons but also 5d-electrons take part in the bond formation,
hence the electronic shell can get a small magnetic moment of the order of 0.1#B,
producing the observed field at a temperature of 135 K.

SrCrsGa4019 (Kagomd lattice). The layered oxide SrCrsGa4019 (frustrated Kagomr-


lattice system) has been examined by LF #+ spin relaxation technique by Keren et
al. (1994a). The results have been discussed by Uemura et al. (1994) in the
context of frustrated and/or low-dimensional spin systems. SrCraGanO19 shows an
unconventional spin-glass like behavior with very strong dynamical spin fluctuations
persisting for T/Tg --+ O.

5.4. MnO
Muon spin precession was observed in zero field in the ordered state of MnO by Ue-
mura et al. (1984). The single component #SR signal found below TN underlines the
fact that the sublattice magnetization Ms, proportional to B,, deviates quite notice-
ably from a S = 5/2 Brillouin function expected for conventional antiferromagnets.
Such an anomalous behaviour of Ms was also observed in neutron scattering (Shull
et al. 1951) and ESR (Sievers and Thinkham 1963) measurements. The #SR data,
however, provide much more accurate information, especially near TN; the internal
field of 0.68 T at T = 117 K suddenly disappears at T = 119 K. B~, extrapolated to
T = 0 K (1.14 T) is reasonably well understood assuming a #+ placed at the body
centered interstitial site of the simple cubic MnO lattice.

5.5. Magnetic fluorides


Several papers have been published on #+ in magnetic fluorides. First, De Renzi
et al. (1984a) and (1984b) intended to use the magnetically well known CoF2 and
MnF2 insulators as a test case for #SR. In their extended reports they present a strong
indication for an octahedrallike #+ location in CoF2, for which they could observe
a spontaneous /~+ frequency in the antiferromagnetic state (fig. 5.12). However,
216 A. SCHENCK and EN. GYGAX

Fig. 5.12. Summary of measured angular dependences of the #SR frequency shifts in CoF2 at 41 K
(circles and triangles) and at 13 K (squares). The crystal was rotated such that the external field of 0.3 T
scanned the a--c plane (left side of the vertical axis on the figure) or the a-b plane (right side of the
vertical axis). The solid lines are from calculationsfor the octahedral #+ site (De Renzi et al. 1984b).

in both fluorides only a limited fraction of #+ contribute to the #SR signals. This
was especially unfavorable for the MnF2 case - with no signal observed in the
antiferromagnetic phase - resulting in an uncertain #+ localization. Nevertheless, De
Renzi et al. (1984b) present several conclusions on the host magnetism, particularly
for the CoF2 sample.
The occurrence of a muonium signal in MnF2 and in site-diluted (Mn0.sZn0.5)F2
was discovered by Uemura et al. (1986b) (fig. 5.13). The formation of a F - : #+ : F -
'hydrogen'-bonded center was considered for MnF2 by Kiefl et al. (1987), in analogy
to the clearly established muon-fluorine bonds in various non-magnetic fluorides
studied by Brewer et al. (1986) (such a location for the muon is compatible with the
octahedrallike site favoured by De Renzi et al. (1984b) in CoF2); this type of state is
now generally considered for #+ in fluorides - Noakes et al. (1993) and references
therein. Thus, the #+ situation in fluorides is somewhat complicated, since the muons
can appear in 3 different electronic states: (i) as (partly screened?) bare #+, (ii) in
muonium form and/or (iii) as part of a ( F - # + F - ) - ion. This latter state especially
has a rather noticeable effect on the host: by pulling two F - together (e.g., in MnF2
the nominal distance of 1.76 A between the two fluorine ions is reduced to 1.21 A
by the intercalate #+ - Kiefl et al. (1987)) the crystal structure is quite affected in
the #+ vicinity. Also in the (F#+F) - ion the #+ interaction with the quite close
19F nuclear moments will compete with the interaction between #+ and the possibly
much larger anion moments, but also more distantly situated (Noakes et al. 1993).
To add to the complexity, it is observed that the occupation of the three #+ states is
a function of temperature (Noakes et al. 1993). Luckily the observation of a single
MUON SPIN ROTATION SPECTROSCOPY 217
A A
1.0

• O

~ 0.5
I-

• v ( 0 ) = 153 MHz
o v ( 0 ) = 1.3 GHz

0.0 I I I I I I I I
0.0 d.5 .0
v/v N
Fig. 5.13. Temperature dependence of the zero-field #SR frequencies measured in MnF2 below TN =
67 K; a low frequencymuon signal and a high frequencymuoniumsignal are observed (Uemura et al.
1986b). The low frequency signal originatesfrom a (F/~+F)- ion, as proposed by Kiefl et al. (1987).

spontaneous non-zero #+ frequency (other than a muonium frequency) in zero-field


measurements in the antiferromagnetic phase (MnF2, Uemura et al. (1986b), CoF2,
De Renzi et al. (1984b)) or of the characteristic oscillating zero-field muon spin
relaxation signal generated in (F#+F) - above the N6el temperature (MnF3, CuF2,
Noakes et al. (1993)) indicate a static behaviour of the muon in all studied fluorides,
at least certainly for T ~< TN.
So far the following informations on the magnetism of fluorides was deduced from
#SR measurements:

CoF2 (De Renzi et al. 1984b). In the zero field #SR experiment the spontaneous
field B u is proportional to the sublattice magnetization. The temperature dependence
of B~, reflects the presence of the strong magnetic anisotropy. In particular at the
lower temperature the behaviour follows the predictions of the spin-wave theory,
while on approaching TN the observed simple power-law dependence agrees with
the three-dimensional Ising model.

MnF2 and site-diluted (Mno.sZno.5)F2. The/z-spin relaxation rates 1/T1 measured in


zero field in both muonium and (F#+F) - states decrease rapidly with decreasing
temperature below TN (Uemura et al. 1986b). The mechanism of the spin relaxation
is explained above TN by the exchange fluctuations of the Mn moments, and below
TN by the Raman scattering of spin waves. With the magnetic Mn atoms of MnF2
randomly substituted by the non-magnetic Zn atoms, the diluted antiferromagnet
(Mn~Znl_~)F2 system can be used to study the effect of randomness on the spin
fluctuation and ordering. The rate 1/T1 for the diluted (Mn0.sZn0.5)F2 is significantly
218 A. SCHENCK and EN. GYGAX

I I I I

0.001 T= 20K ,L~ " I


0
T = 33K
,I .... I,. II " --~ -,i~"ll.r-WW'-l~
b 0
"6
E
E
T = 48K ~ l l i,ll
0
<
"8
0.002

0.001
C3
0 i I Illl T= 6511~1
' ~ 1 I
-0.001

-0.002
I I I I
0 0.1 0.2 0.3 0.4 0.5
Longitudinal Magnetic Field (T)
Fig. 5.14. The muonA9F level-crossing spectra in MnF2 as a function of temperature below TN =
67 K. The upper (A = 1) resonance is off scale at lower temperatures. The curves are fits of a
theoretical difference signal; the positions of the resonances scale approximativelywith the sublattice
magnetization (from Kiefl et al. 1987).

larger than the corresponding values for pure MnF2 at the same normalized temper-
ature (Uemura et al. 1986b). The difference between pure and dilute systems is
related to the large spectral weight of low-energy magnons in (Mn0.sZn0.5)F2 found
by neutron scattering (Uemura and Birgeneau 1986c). A model for #+ spin-lattice
relaxation in an antiferromagnet is presented by Keren (1994) and used successfully
by Keren et al. (1994b) to describe the 1/Ta measurements in MnF2.
The observation of muon-nuclear level-crossing resonance (LCR) in the antiferro-
magnetic state of MnF2 (Kiefl et al. 1987) demonstrates that this technique can be
used to obtain informations both on the local field at the muon and at neighboring
nuclear spins in magnetically ordered systems (fig. 5.14). The observed shift of the
local field at the 19F nuclei next to the #+ (with respect to 19F NMR data in MnF2)
, is attributed to the disturbing influence of the muon.

MnF3 (Noakes et al. 1993). Below TN no fast oscillation can be detected in the
zero-field measurements. At 10 K a large #+ spin-relaxation rate is observed for the
major fraction of the signal, indicating a distribution of local fields at the (F#F) sites
with a width of 0.1 T or more. Such apparent local magnetic disorder in a relatively
simple magnetic structure may be due to the complicated crystal structure. With
12 formula units per monoclinic unit cell, the muons in the (F#F)- ions may be
placed at a large number of slightly different positions with respect to the magnetic
structure.
MUON SPIN ROTATION SPECTROSCOPY 219

CuF2 (Noakes et aL 1993). In the antiferromagnetic phase, the relatively low sponta-
neous frequency (corresponding to a local field of 0.15 T at low temperature) and the
relatively low #+ depolarization rates of the signals indicate that the host is a rather
weak-field magnet. The large bulk susceptibility above TN (Fischer et al. 1974) is
interpreted as indicating substantial short range order. The muons as local probes
can be sensitive to local ordering, so the coherent frequency need not necessarily go
to zero at TN in this case. Unfortunately, the lowest temperature paramagnetic-state
data obtained so far in CuF2 were taken well above TN, at 100 K, and no evidence
for ordered moments was seen.

5.6. CoCl2.2H20
This compound has been comprehensively studied by proton NMR in its antiferro-
magnetic phase (Narath 1969). A unique magnetic field at the proton site in the water
molecule has been found, indicating that all proton sites are magnetically equivalent.
The field at the proton amounts to 0.42 T at the T = 0 K limit, and follows over a
wide temperature range a power law of the form

{Bp(0) -- B p ( r ) } oc T 6'5.

1.0
I

o Vl0wer
I

/
/
x Vupper
0.5

I-

'2.
0
v
:=L

0
T 4 . law
._1

-0.5

-1.0
X
I I I I
0.7 0.9 1.1 1.3
Log (T)
Fig, 5.15. Double logarithmic plot of the temperature dependence of the frequency shifts for the two
zero-field/zSR signals in antiferromagnetic COC12-2H20 - Brewer et al. (1981).
220 A. SCHENCK and EN. GYGAX

One expected to find the #+ at one of the proton sites, but the zero-field #SR
measurements (Brewer et al. 1981) showed surprising results: t w o frequencies are
found, corresponding to internal fields of 0.273 T and 0.283 T at the T = 0 K limit,
and both fields follow over the studied temperature range (from 4 to 17 K) a power
law with a quite different exponent:

{B#(0) - B~(T)} o( T 4"0.

This temperature dependence, plotted in fig. 5.15, is totally different from the be-
haviour in the other antiferromagnetic insulators, where generally Bloc follows the
sublattice magnetization. No explanation of the results is available so far.

5. 7. Solid oxygen
Storchak et al. (1994) present a ZF #+SR study of a-O2 (AF phase of solid oxygen)
in the temperature range 10-24 K. Solid oxygen is one of the most unusual molecular
crystal, as the 02 molecule possesses an electronic spin S = 1 in the ground state.
Strong direct coupling of the 02 molecules' 7r-orbitals is realized on the background
of the weak intermolecular Van der Waals interaction, closely connecting magnetic
and lattice properties. At equilibrium vapor pressure solid oxygen exists in three
crystalline structures; the low-temperature a-phase is known to be antiferromagnetic.
Informations on the/z+-solid oxygen complex is given by Storchak et al. (1992).
Below the a - f l transition temperature (T~-~ = 23.8 K) Storchak et al. (1994) ob-
serve long-lived #+ spin oscillations in ZF, manifesting the existence of an ordered
state. Figure 5.16 shows the local magnetic field at the muon as function of tem-
perature. For a comparison the normalized Brillouin curves for TN = T~-~ and
TN = 40 K (about the estimate given by Bhandari and Falicov (1973) and Slyusarev
et al. (1979)), respectively, are displayed. The abrupt drop of B~ in the vicinity of
T~-~ shows that the ordinary second order phase transition does not occur in solid

1.6 I I I I

1.2

(..9
v
0.8

0.4 'l I

0.0 I I I I
5 10 15 20 25 30
Temperature (K)
Fig. 5.16. Temperature dependence of the magnetic field at the /~+ in c~-O2; circles: experimental
points. Brillouin curves for spin S = 1, B0 = 1.27 kG, TN = Ta_;~ (dash-dotted line) and TN = 40 K
(dashed line), respectively. The solid line is a fit of a 2D-Heisenbergmodel with an anisotropyparameter
c~~ 10-2 - see text (from Storchak et al. 1994).
MUON SPIN ROTATIONSPECTROSCOPY 221

oxygen. Storchak et al. (1994) fit their data with the phenomenological model pre-
sented by Le et al. (1990a) for a 2D-Heisenberg spin-1 system with TN ----40 K and
the anisotropy parameter c~ = 0.01; a good agreement is obtained up to T = 23 K
(solid line in fig. 5.16). The behavior is similar to that found for Sr2CuOzC12 by Le
et al. (1990a) (see section 6.1).
It is known that in the/3-phase of solid oxygen a magnetic order, at least of short
range nature, is present (Stephens and Majkrzak 1986). Storchak et al. (1994),
however, do not observe the presence of a magnetic order above T~-~.

6. Review of results in layered cuprate (high To) compounds

#SR spectroscopy has contributed significantly to our knowledge of the magnetic


properties of the high temperature oxide superconductors and their magnetic parent
and related compounds. First evidence for magnetic order in the (123)- and the
(2212)-families and in NdzCuO4 (the parent compound of the electron high Tc su-
perconductors) was in fact provided by #SR (Nishida et al. 1987a, b, De Renzi et
al. 1989, Luke et al. 1989a). The potential of #SR to detect short range and random
order has been particularly of value and helped in the elucidation of phase diagrams.
Some of the work has been reviewed by Budnick et al. (1990a), Nishida (1992a, b)
and De Renzi (1992).

6.1. La2Cu04 and related compounds


Table 6.1 presents a list of the (214) cuprates and related compounds studied so far
by #SR. The possible #+ site or sites were considered in several papers (Le et al.
1990a, Hitti et al. 1990b, Torikai et al. 1993a). According to the most recent and
detailed investigations by Torikai et al. (1993a) possible #+ sites are restricted to
small areas on the (100)- or (011)- and the (ll0)-plane near to an apical oxygen
at a distance of ,-~ 1 A as shown in fig. 6.1. The site proposed by Hitti et al.
(1990b) is close to the 'a' area. The 'a' area is also close to a theoretical prediction
by Sulaiman et al. (1993). As in other oxide materials (see section 5) the #+
seems to bind to an oxygen ion forming a kind of an O-H bond (see also Boekema
1988).

La2Cu04. The appearance of antiferromagnetic order in La2CuO4+6 depends very


sensitively on the exact oxygen content. Some oxygen deficiency (~ < 0) stabilizes
the antiferromagnetic state (e.g., ~ = -0.3 --+ TN --~ 300 K), while an oxygen excess
(~ > 0) (introduction of holes into the system) suppresses the antiferromagnetic order
and eventually leads also to superconductivity with critical temperatures nearly as
high as in the Sr-doped compounds.
The presence of antiferromagnetic order in LazCuO4_6 is readily detected by #SR
by the appearance of a spontaneous precession signal below TN in a ZF-experiment.
It is found that the spontaneous field at the #+ extrapolated to zero temperature falls
within a rather narrow range of B u _~ (38.0-43.0) mT independent of TN or the
oxygen deficiency, respectively. Also the exact La/Cu ratio seems to be of no effect.
222 A. SCHENCK and EN. GYGAX

TABLE 6.1
List of La(214)-cuprates and related compounds studied by /~+SR (and #-SR). For references see
table 6.2.
Compound Crystal Mag. TN Tc /~+ SR
struct, struct. (K) (K)
La2_eCuO4_ ~ orthorhombic AF <290 - ZF
for T < 530 K 6-dep. (#-SR)
La2(CUl_~Zn~)O4 orthorhombic AF, random ~<260 - ZE ZF, LF
0 ~< :c ~< 0.10 for T < 530 K
La2(CUl_x Coz)O4+ 6 orthorhombic SG ZF
for T < 530 K
La2_zSr~CuO4_6 orthorhombic AF (x ~< 0.05) > 0 ZF, LF, TF
for T < 530 K for
0 ~ x ~< 0.15 SG (x ~ 0.13) x/>0.6 (#-SR)
Lal.875Bao.125CuO4 tetragonal AF 38 0 ZF
Lal.875Bao.075Sro,050Cu04 tetragonal AF <50 20.5 ZF, TF
La2_~_ySr~NdyCuO4 tetragonal AF < 30 ZF
La2NiO4+6 tetragonal, T~ AF, different 320 - ZF
from La2CuO4
Nd2CuO4_6 tetragonal, T~ AF (La2NiO4) ~ 245-275 - ZF
Pr2CuO4_ 6 tetragonal, T~ AF (La2NiO4) ~ 250 - ZF
Sm2CuO4_, tetragonal, T' AF (La2NiO4) ~ 250 - ZF
Nd2_~CezCuO4_ ~ tetragonal, T~ AF(x ~ 0.14) ~<250 ~< 24 ZF, TF
0 ~< x ~< 0.17 (x > 0.14)
Nd2_xSrxCuO4_6 tetragonal, T t AF(x ~< 0.10) <280 - ZF
0~<x ~<0.20
Lal.2Tbo.sCuO4 tetragonal, T* AF(La2CuO4) ~- 170 - ZF
/~ = 0.46#B <280
random
Sr2CuO2C12 tetragonal (KzNiF4) AF 260 - ZF
La2MCu206+ 6 2 layer perovskite AF >300 <45 ZF, TF
M = Ca, Sr for (M = Sr)
250
for (M = Ca)
Lal.9Y0.tCaCuO6+6 2 layer perovskite AF ~ 250 ZF

A c o m p i l a t i o n o f results is p r e s e n t e d in table 6.2. T h e temperature d e p e n d e n c e o f


Bu is d i s p l a y e d in fig. 6.2 for various 5. It can be r e p r o d u c e d b y adopting a 2D-
H e i s e n b e r g m o d e l with a small anisotropy to allow for 3D long range o r d e r (Le et
al. 1990a, b). A n i m p o r t a n t p a r a m e t e r is g i v e n b y A = gpBHA/4JMs w h e r e HA
is the a n i s o t r o p y field, J the H e i s e n b e r g e x c h a n g e p a r a m e t e r and Ms the sublattice
m a g n e t i z a t i o n in a p p r o p r i a t e units. In order to reproduce, Bu(T ) o(Ms(T), e.g., for
= 0.03, o n e finds A _~ 10 . 3 ( L e e t al. 1990a, b).
T h e n e a r l y constant B u ( 0 K ) in all the different L a 2 C u O 4 - s a m p l e s i n v e s t i g a t e d is in
m a r k e d c o n t r a s t to results f r o m neutron scattering. F i g u r e 6.3 shows the sublattice
m a g n e t i z a t i o n d e r i v e d from the m a g n e t i c B r a g g p e a k intensity o f three different
m o n o c r y s t a l l i n e s a m p l e s with different TN ( Y a m a d a et al. 1987). A s can b e seen the
sublattice m a g n e t i z a t i o n Ms for T --+ 0 K drops significantly with TN. It c o u l d be
s h o w n that the r e d u c e d Ms(0 K) does not arise f r o m only partial m a g n e t i c order in
MUON SPIN ROTATION SPECTROSCOPY 223

(c)
/ • : Cu
0:0
• La

W'p

Fig. 6.1. Crystal structure of La2CuO4. The shaded areas indicate possible/z+-locations, according to
Torikai et al. (1993a). Areas a and b are confinedto the (1 00)- and (0 1 0)-planes, respectively,and area
c is on the (I 1 0)-plane (from Torikai et al. 1993a).

the sample volume (Uemura et al. 1988). The difference in #SR- and neutron results
is explained to be caused by a loss of long range magnetic order as the number of
holes is increased and more and more frustrations are introduced into the magnetic
CuO2-planes. Since the #+ is sensitive to the nearest neighbour Cu-moments only it
will sense preliminary the short range magnetic order and that is independent of the
number of holes introduced. The neutron magnetic Bragg peaks, on the other hand,
arises only from those parts of the sample volume in which long range magnetic order
is still present. However, a certain disorder does not go unnoticed by the #+ in that
the local field distribution width ABu (0 K) increases, as reflected in an increased
decay rate of the spontaneous precession signal, when TN decreases (see table 6.2).
Sample quality is another important factor. As table 6.2 shows not all samples
investigated develop coherent (short or long range) magnetic order in the whole
volume. It seems that some samples remain in part paramagnetic down to the lowest
temperatures. Others show in part a wide field distribution with zero average. So
no precession signal is seen but only a rapidly decaying polarization PZF(t) (Hitti
et al. 1990b). The fact that in some samples in all the volume magnetic order is
established, as evidenced by the observation of a single frequency precession signal
224 A. SCHENCK and EN. GYGAX

o 0

~+ -~~
~ 0 0 O0

a.~
ii I
dd ~ ~ b G d d d d d d d d d d
~ d o o o o o o o ~ q 0 0 0 0 0

0
0
v
?1

o o

~g 0 0 0 0 0 0 ,~o
4~ ~ v~ ~ v ~ v ? II "U

,,0 ~:

o i.~
o~..
II - ~ v 0 0 0
ddddd d
?

I I I I

oo ~8 ~ ~u~ == t

~ II II II ?1 ~v~ ~ ~ tl II I1 II II
r..)
MUON SPIN ROTATION SPECTROSCOPY 225

o", o",
o'~ o',
. . . . . . - ~ ~
~ ~ " ' "~ ' " " ~ _ ~ ~ .

b b b b b b
~ ~ d d d d d d d d d d dd d d
dd dd o o

v v v
d=

0
~ 0 0 0 0 0 O0 O0 ~.0
0

illL

o
~v~-~v~ o"1
v~

~'ob

0 0 0 0 0 0 d ¢'-.
V~V~V~V~ v~'

o;o; r.,9

~ li il II I~ II II li li li II II it II il II il ~ ° V/
~ ° N~
~ V/
"0 ?-
0 r-~ 0 0 0 0 0 0 0 0 0 "~ "-~ '-"~ 0 0 0 0 V/I~ ~ ,, ,':
~,...~ ~ I ,....~ ~- ~
0 II II It II II II II I1 II II II II II II II II o =, ~ ~,._, ~, ~, A~ II
0
226 A. S C H E N C K and E N . G Y G A X

0 0 O 0
o, o-, o,~ £ # o,,

"~ "N "~ ~


,-~ o

=== N-g -~
°

d d d 6 d
O "6 "66 o o o d ~6

e-

0 ~ 0

V V VV V

¢) ',,D

~.D t"q
r-- ~ II t'N t'xl
O O O O O
~e~ ~ II
~- A
O

ON "--~v ON
r-- e-i rn ,.~ (",1 ~,"j

~
o
O
.~
¢)

V/ A M .~

aa=~ gg~
~o V/ll o .~ "~ O
•~ ~ , ~ ~ "-
~ ~s.~s o~.~

7
osdo - o~ d ; 7
o2 oo t'~ 6 , 5 ~v.qe-
u u ~6 O9

z z ££ Z o Z ~ ~Zd~ s
MUON SPIN ROTATIONSPECTROSCOPY 227

La2Cu 04.y
6l Oo. . • •
~°°V o o • °
41 ~ • o o ,~°eI,~

t°'°° o.o1 0.02


"% %
"'%Q,.
u_ 2 y - 0.03

0 t v t
0 1O0 200 3OO
Temperature(K)
Fig. 6.2. Temperaturedependenceof the spontaneous ~+-spin precession frequenciesin La2CuO4_y for
various y (from Uemura et al. 1988).

0.6 i i i

3" 0.4
m

to

0.2

0 1O0 200 300


Temperature (K)
Fig. 6.3. Temperaturedependence of the sublattice magnetization in La2CuO4_~ for various ~ as ob-
tained from neutron diffraction by Yamada et al. (1987) (figure taken from Uemura et al. 1988).

with maximum possible amplitude, rules out the occupation of more than one type
of interstitial site by the #+.
Finally Torikai et al. (1993b) have also measured the direction o f / 3 , ( 0 K) with
respect to the crystal axes (see table 6.2). The direction found together with the
suggested site implies that /3,(0 K) is not only arising from dipolar fields from
distant moments but in addition also from the contact hyperfine and a dipolar field
associated with a local (non spherical) spin density distribution in the vicinity of
the/z +.

La2(CUl_xZnx)O 4. Quite interesting results were obtained from ZF-measurements


in Zn substituted compounds Laz(Cul-=Zn=)O4 (Lichti et al. 1991b). The idea
was to study the effect of Zn substitution on the magnetic properties in a system
where the substitution takes place only at known Cu-sites. Since La2CuO4 only
228 A. SCHENCKand EN. GYGAX

300 i i i t .... I

~" La2Cul.xZnx04

200-

= 100

0 I i
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14
Zn Concentration (x)
Fig. 6.4. Magneticphase diagramof La2Cul_zznxO4. Opensymbolsare from/~SRmeasurementsand
closed symbolsare from dx/dt and Xmaxdata (fromLichtiet al. 1991b).

possesses CuO2-planes Zn substitution can only take place there, quite in contrast
to the case in YBazCu3Oz. The substitution of Zn 2+ creates a spin hole (3d-shell is
filled up in Zn) and removes a 3d charge hole from the planar network and disrupts
locally the Cu-O hybridization. It suppresses not only the magnetic order but also
superconductivity in the Sr-doped compounds. The ZF-#SR shows the presence of
an ordered phase in all samples with 0 ~< x ~< 0.10. However, just below TN the
magnetic order appeared random or of very short range only as evidenced by a fast
decaying but nonoscillating PzF(t). Only at significantly lower temperatures, T ~, was
a spin precession signal observed indicating the evolution of also 'long range' order.
The results allow to construct a phase diagram as shown in fig. 6.4. It suggests
a total suppression of magnetic order for z /> 0.11. The 'flat' part extrapolates
to an x value which is quite close to 0.41, the site percolation threshold for Zn
substituted sites. This could suggest that simple dilution controls this part of the
phase diagram. The faster decrease of TN and T p for small x may point to additional
frustration involving the coupling beyond the nearest neighbors. Quite remarkable
is also the observation that the development of first the random order and then the
'long range' order is restricted to a fraction of the volume which grows from zero
just below TN to essentially 100% at the lowest temperature. The remaining volume
stays paramagnetic. This behaviour is reminiscent of what has been observed, e.g.,
in CeA13 and also in HoBazCu307 (see later). Whether, in the present sample, it
could reflect a very inhomogeneous distribution of the substituted Zn, implying a
wide distribution of local TN or T p, or an evolution of magnetic order through a
percolative mechanism or something else remains to be seen. A comparison with
results on YBa2(Cul_~Zn=)Ox will be presented below.

La2_xSrxCu04. Doping La2CuO4 with Sr (i.e. substituting Sr2+ for La 3+) introduces
holes into the CuOz-planes just like in the case of an excess oxygen content. The
additional spin carrying holes, probably located at the oxygen sites, are believed
MUON SPIN ROTATION SPECTROSCOPY 229

to be coupled ferromagnetically to the neighboring Cu-moments and thus interfere


with the antiferromagnetic coupling of the Cu-moments, causing frustration and
eventually a suppression of the Cu-moment order. At higher concentrations the
holes will become the carriers responsible for superconductivity. It is suspected that
the mechanism leading to Cooper pair formation may be magnetic in nature. This
possibility has triggered quite some research into the question whether magnetic
order and superconductivity can coexsist on a microscopic scale or whether they
are mutually exclusive. One important model in this respect has been suggested by
Aharony et al. (1988) which predicts a spin glass type of state as a result of magnetic
frustrations.
The #SR-results on La2_xSr~CuO4_6 for x < 0.5 resemble very much the results
obtained in the pure La2CuO4_6 system when varying ~. In the ordered state the
low temperature spontaneous field at the /~+, B~ (~ 0 K), is rather independent
of x although TN drops quickly with increasing x. Also the width of the local
field, AB (~ 0 K), increases with increasing x and thus reflects nicely the disorder
introduced by the hole doping induced frustration (see table 6.2). The #SR results
on poly as well as monocrystalline samples show that long range antiferromagnetic
order is lost for a critical xc > 0.5 (since ~ is usually not determined the Xc-values
from different investigations vary somewhat) (Kitazawa et al. 1988, Weidinger et al.
1989, Grebinnik et al. 1990a, Torikai et al. 1993a). However, this does not mean
that no magnetic order is present at all. Rather a different type with spin glass like
properties emerges which can be observed up to z _~ 0.15 (Weidinger et al. 1989).
Although this is still somewhat controversial (see the comments by Harshman et al.
1989 and Heffner and Cox 1989 to Weidinger et al. 1989, and also the zero result
by Kiefl et al. 1989) there is mounting evidence that at least up to x = 0.13 random
static order can be observed.
The evidence is the following. (i) The phenomenon is observed in high quality
single crystals, (ii) the whole sample volume participates, (iii) the typical features
of #+-relaxation in a classical spin glass are seen. As an example fig. 6.5 shows
the ZF-#SR signal in a single crystal with x = 0.11 (Torikai et al. 1993c). At
20 K a Gaussian damping is observed which is induced by the random dipolar fields
from the Cu- and La-nuclei. At a slightly lower temperature (To) the Gaussian
decay is changed to an exponential one signaling a sudden influence of the Cu-3d
moments as a result of the onset of correlations and possibly a slowing down of their
dynamics. Below 8 K (= Tf) static random order is indicated by the development of
a Kubo-Toyabe signal. Very similar results have been obtained by Sternlieb et al.
(1990) who also demonstrated the static nature of the random field distribution at low
temperatures in LF-decoupling measurements. There is another distinct difference
between samples displaying long range magnetic order (x < 0.1) and samples with
random order (0.1 < x < 0.13). While in the former samples above TN the 3d-
spin dynamics is so fast as to be ineffective in inducing #+ spin lattice relaxation
(--+ 3d-spin fluctuation rate ~> 1012 s - l ) the dynamics is much slower in the latter
samples and seems to freeze out smoothly on approaching the freezing temperature.
Grebinnik et al. (1990a) find for a polycrystalline sample with x = 0.05 that PzF(t)
above the freezing temperature Tf ~ 4 K is well described by eq. (2.18), i.e.
PzF(t) = Po exp ( - (At)~). (6.1)
230 A. SCHENCK and F.N. GYGAX

0.15
20K
0.10
EE 15K
< 0.05 12.5K
8K
0 3.5K

0 1 2 3 4
Time (IJS)
Fig. 6.5. ZF-#SR signals in La1.89Sr0.11CuO4 for different temperatures. Note the appearance of a
nearly static Kubo-Toyabe signal at 3.5 K (from Torikai et al. 1993c).

Corresponding fit results are displayed in fig. 6.6. A displays a divergent type of
behaviour, and the exponent changes from 1 to 1/3 near Tf. This is a typical behaviour
for a not so dilute spin glass (see, e.g., Campbell et al. 1994) and has also been
seen in a Co substituted (214)-compound: La2Cu0.25Co0.7504. (Lappas et al., to be
published.)
Figure 6.7 presents a phase diagram for monocrystalline La2_=Sr=CuO4 proposed
by Torikai et al. (1990, 1993a) on the basis of their #SR-data. As is evident the spin
glass type of phase overlaps more or less completely with the superconducting phase.
However, long range order just disappears approximately where superconductivity
first shows up. It is also very interesting to note that Tf and To assume a maximum
at an z-value which corresponds to the concentration where the low temperature
tetragonal (LTT) phase is observed in polycrystalline samples. However, in the
present case the sample with z = 0.11 shows 100% bulk superconductivity (Torikai
et al. 1993c). The fact that both Tf and To tend to zero when z approaches 0.15 has
led Torikai et al. (1993c) to suggest that a phase boundary exists at z corresponding
to the maximum To and that the spin state above that z is qualitatively different from
conventional antiferromagnetic or spin glass order.
Finally we mention that also TF-#-SR-spectroscopy has been applied to the system
La2_=Sr=CuO4. Quite visible and distinguishable are two components arising from
# - captured by the planar and apical oxygen (Nishiyama et al. 1993, Torikai et al.
1993b). From these measurements the temperature and angular dependence of the
/z--Knight shift in the lowest muonic Bohr orbital in oxygen has been determined.
One expects that the results should correspond to equivalent data from t70-NMR
insofar as the same site is probed. This is not the case and also the assignment
MUON SPIN ROTATION SPECTROSCOPY 231

10 ..... I I I I I

ka 1.95 Sr0.05 Cu 0 4
o~
O•
o/

1.0
e-

x
t~

rr

0.1 I I I I
1.0
e0
0.5

0.2 i I I I I
4 5 6 7 8 9
Temperature (K)
Fig. 6.6. Temperature dependence of the relaxation rate )~zFand exponent/3 from a stretched exponential
fit to ZF-data in La].95Sro.05CuO4. Note that /3 approaches 1/3 as the spin freezing temperature is
approached (from Grebinnik et al. 1990a).

40 J , ' I ~ J ~ ' I ' ' ' ' J....~ ~ ~ t

• -~ _
/ \

30 - /.~-~¢'$
g /
Q~
/
20 /
Q~
t'~
E
10 ~-~~- ~ a

0 i I i i i

0 0.05 0.1 0.15 0.2


St-Concentration
Fig. 6.7. Magnetic and superconducting phase diagram of La2_=Sr=CuO4. Closed circles indicate the
onset temperature, To, for short range correlations and closed squares the temperature, Tf or TN, at which
static magnetic order is established (spin glass like for x > 0.05, and coherent long range for x < 0.5)
in monocrystalline samples. Results on Tf, TN for polycrystalline samples are indicated by the dashed
line. Superconducting transition temperatures, Te, in the single crystals are indicated by triangles (from
/~SR) and crosses (from ac - X), and in polycrytalline material by the dot-dashed curve (from Torikai
et al. 1993c).
232 A. SCHENCK and EN. GYGAX

of the two # - S R components to the two oxygen sites is not clear. The difference
between the 170-NMR (Ishida and Kitaoka 1991) and the (O#-)-SR results may
arise from the fact that (O#-) z-1 corresponds to nitrogen, thus being an impurity
ion in the lattice, or that the oxygen (or nitrogen) electronic shell modified during
the #--cascade to the lowest Bohr orbital is left in an unrelaxed paramagnetic state.

La2_:~(Sr,Ba)~Cu04. The system La2_~Ba~CuO~ displays a rather similar phase dia-


gram as La2_~Sr~CuO4, except in the vicinity of z = 0.125 where superconductivity
was found to be almost completely suppressed (Moodenbaugh et al. 1988). This
appears to be correlated with an additional structural phase transition at To ,-~ 60 K
which changes the system from the low temperature orthorhombic (LTO) phase to
a low temperature tetragonal (LTT) phase and which is only seen in the Ba doped
compounds (Axe et al. 1989). At much higher temperatures La2_~Ba~CuO4 like
La2_~Sr~CuO4 undergoes a structural transition from a high temperature tetragonal
(HTT) phase to the LTO phase. The structural phase diagram is shown in fig. 6.8.
The question that arose immediately was, of course, whether the disappearance of
superconductivity was accompanied by the appearance of magnetic order. ZF-#SR
measurements on polycrystalline samples with z in the vicinity of 0.125 revealed
indeed the presence of spontaneous #+ spin precession as can be seen from fig. 6.9
(Luke et al. 1991a). The low temperature spontaneous B~,(,,~ 0 K) amounts to
0.025 T which is considerably reduced from the maximum value of ~ 0.043 T
found in La2CuO4_& Assuming that the antiferromagnetic structure is the same as

500 I I I I I
60

400 '\ I.-


\ 40
'xe
\
300
A \ 20

I-
200 0
', 0.08 0.12 0.1(~
LTO ,\ X
\
100 \ HTT
qk
"
O\
0
I I I \ I I I
0 0.1 0.2 0.3

Fig. 6.8. Structural phase diagram of La2_~BazCuO4. The insert includes on an expanded scale the
magnetic transition temperatures TN. HTT, LTO and LTT stand for high temperature tetragonal phase,
low temperature othorhombic phase and low temperature tetragonal phase, respectively (from Kumagai
et al. 1993).
M U O N SPIN ROTATION SPECTROSCOPY 233

0.2 i t i t t

~ T=38K
q" "¢,'w" ,'.'* ~(2& ",

if)
< 0.1 '

0
0

0 I I I I I
0.0 0.5 1.0 1.5
Time (liS)
Fig. 6.9. ZF-/~SR signal in Lal.875Ba0.125CuO4 at 38 K (above TN) and 5 K (below TN). The presence
of coherent order at 5 K is clearly manifest (from Luke et al. 1991a).

in the parent compound the reduced Bu(,~ 0 K) can be used to estimate the mag-
nitude of the ordered moment to be about 0.3#B. This value is in good agreement
with estimates resulting from heat capacity measurements near z = 0.12 (Wada et al.
1990). B~,(~ 0 K) turns out to be rather independent of the exact value of z (see ta-
ble 6.2). Following the temperature dependence of B~,(T) the transition temperatures
TN could be determined. The results are also plotted in fig. 6.8. It is seen that To and
TN follow the same z-dependence (Kumagai et al. 1993). The LTO-LTr transition
was also found in the compound La2_~_vSr=NdyCuO4. Again magnetic order could
be detected by ZF-#SR in a sample of composition Lal.775Sr0.125Nd0ACuO4 with
B,(0 K) ~ 0.025 T (Kumagai et al. 1993). The results indicate that (at least long
range) magnetic order and superconductivity are mutually exclusive. The role of the
LTO-LTT transition is not yet clear (new #SR studies of the system Laz_=Sr=CuO4
have also revealed a long range ordered AF-state in the range z = 0.105 - 0 . 1 2
without a transition to the LTT phase (Watanabe et al. 1994)). It may involve a
reduction in hole concentration in the CuO2-planes, shifting the system closer to
the Lal.sSr0.aCuO4-situation. On the other hand the reduced value in the ordered
moment needs to be explained as well. The latter suggests a change in the electronic
structure within the CuOz-plane not just a dilution of the hole concentration.
To study these problems further investigations of the mixed system La2_=(Sr,
Ba)=CuO4 have been started. Neutron diffraction and TF- and ZF #SR measurements
were performed on a Lal.a79Bao.075Sr0.050CuO4 sample (Lappas et al. 1994a, Lappas
1993, and unpublished results). This sample revealed clearly the presence of a
superconducting transition at Tc - 22 K, detected by resistivity and magnetization
measurements. The neutron measurements showed a gradual increase of the LTI"
phase out of the LTO-phase below about 40 K, saturating at about 80% of the sample
volume below about 25 K. TF-#SR measurements revealed two signals: one with a
234 A. SCHENCK and EN. GYGAX

100
• Neutron
e /J, +SR
8o

60-
0

40-
LL

t
0 10 20 30 40 50 60
T (K)
Fig. 6.10. Temperature dependence of the volume fraction in Lal.875Bao.075Sro.05CuO4 showing the
LTT-phase (from neutron scattering) or magnetic order (from TF-/zSR-data) (from Lappas et al. 1994).

4.0 ' I ' I I I ' I I ' I ' I i

3.5 TN~,el
o • eoe
3.0
"i"

>=
>,
2.5

2.0
'1
o-
lD 1.5
-I
lD"
1.0
LI.

0.5

0.0 - , I , I , I , t ~ tOf91 Q 9 , -

0 5 10 15 20 25 30 35 40
Temperature (K)
Fig. 6.11. Temperature dependence of the spontaneous /z+ spin precession frequency in
La].875Bao.o75Sro.osCuO4 (from Lappas 1993).

small decay rate which also reflected the onset of superconductivity by a diamagnetic
frequency shift below Tc ~- 22 K; the other one, developing below -,~ 50 K on the
expense of the 'slow' signal, showing a much more rapid damping which corresponds
to a field spread of ~ 12.0 m T below 10 K. The amplitude of this component followed
closely the fraction of the LTT-phase and saturated likewise at a value of 80% of
the total signal amplitude. This is shown in fig. 6.10. In zero field this component
MUON SPIN ROTATION SPECTROSCOPY 235

displayed spontaneous spin precession with B , 000(~ 0 K) _~ 0.024 T. Interestingly


the onset of the spin precession signal, implying some 'long range' order, took
place only below N 27 K while the W-measurements imply an onset temperature of
50 K (compare figs 6.10 and 6.11).
In any case the #SR-measurements together with the neutron results prove that
the occurence of superconductivity is restricted to the LTO-phase, while magnetism
is associated with only the LTr phase. These findings are tentatively explained in
terms of coexsisting Sr-rich and Ba-rich microdomains. The temperature dependence
of the relative fractions remains unexplained.

La2NiO 4. Very similar in structure to La2CuO4 this system has attracted attention with
respect to the possibility that it may show superconductivity like its Cu counterpart.
Although no superconductivity was unambiguously found, magnetically it behaves
very similar to La2CuO4. In stoichiometric La2NiO4 long range 3D antiferromagnetic
order is observed below TN =320 K by neutron diffraction (Rodriguez-Carvajal et al.
1992). Its magnetic structure can be derived from the one in La2CuO4 by flipping the
spin of the center Cu atom in fig. 6.13(b). This is confirmed by ZF-#SR which shows
the presence of spontaneous spin precession below room temperature (Martinez et al.
1992). In contrast to La2CuO4_a two frequencies are identified corresponding to low
temperature local fields of B(~I)(0 K) _~ 0.26 T and/3(2)(0 K)= 0.015 T, respectively.

I I I
O ' O O qrl) • a
N
35 go
"r"
g
f- 30

In
25
I I I I I
50 150 250

I I I I I

2 b
-i-
1.5 A
¢..} ~A
t-'-
(D 1
O"

In 0.5 A A AA
AA
I - I I I I
0
0 50 150 250
Temperature (K)
Fig. 6.12. Temperature dependence of the two spontaneous precession frequencies seen in La2NiO4
(from Martinez et al. 1992).
236 A. SCHENCK and EN. GYGAX

® J

! o
) 0 • Cu
Qo~
0 0
b o 0
@
o--
Sr
0 Lo
1
Nd

0
0
°~o 0 CI
®

(o) (b) (c)


Fig. 6.13. Crystal structure of (a) Sr2CuO2CI2, (b) La2CuO 4, (c) Nd2CuO 4 (figures taken from Le
et al. 1990a).

The temperature dependence of the B(~)(T) are displayed in fig. 6.12. It is suggested
that the stronger field is associated with a #+ site in the Ni-O planes and the weaker
field with a site close to an apical oxygen (like in La2CuO4). As can be seen
B(~)(T) follows quite different temperature dependencies. While B~I)(T) shows a
shape typical for the sublattice magnetization B(~2) displays an anomalous rise below
about 75 K. This behaviour is explained to arise from a structural transition at 80 K
(orthorhombic to tetragonal when lowering the temperature) found in earlier work
(Rodriguez-Carvajal et al. 1988, 1992). The low temperature tetragonal structure
allows for the existence of a weak ferromagnetic component along the e-axis, which
appears below 80 K as demonstrated in magnetization measurements (Granados et al.
1990). The temperature dependence of B(u2) seems to track the weak ferromagnetic
component quite well (Martinez et al, 1992). It is argued that the local field at
a site close to the apical oxygen is more sensitive to the appearance of a weak
ferromagnetic component than a site within the Ni-O plane.
Long range magnetic order could not be detected by neutron diffraction in sam-
ples with a slight oxygen excess (5 > 0.05). ZF-#SR, in contrast, revealed at low
temperatures (,-~ 5 K) a spontaneous but strongly damped spin precession signal sig-
naling the presence of only extremely short range order (but not random order). Like
in La2CuO4+6 and Laz_zSrxCuO4 doping holes into the system introduces severe
frustration.

R2CuO 4 (R = Nd, Pr, Sm). Also these compounds have a crystal structure similar
to La2CuO4. In contrast to La2CuO4 they remain tetragonal at all temperatures (the
MUON SPIN ROTATION SPECTROSCOPY 237

5 i I I I

4
0 Nd2CuO4
0 O0 :-~
"i-
0 P~±8
3
v
,o o
c

o"
2 AC~A¢~A o AO A

o4
-->
LI- 1 g.ie B~_L~

I I I I
00 25 50 75 100 125
Temperature (K)
Fig. 6.14. Temperature dependence of the spontaneous/z + spin precession frequency and the direction
of/~u in Nd2CuO 4. Clearly manifest are the spin reorientation transitions at 35 K and at 75 K (from
Luke et al. 1990a).

phase is called T') and the Cu-ions are only four fold coordinated (there is no apical
oxygen) (see fig. 6.13c). They are the parent compounds of the electron doped (214)
superconductors. ZF-#SR spectroscopy was first (Luke et al. 1989a) to demonstrate
that all these compounds show Cu-moment magnetic order below typically 250 K
(see table 6.2). However, the magnetic structure appeared more complex as a func-
tion of temperature, as is best seen from the results on a Nd2CuO4_6 single crystal
sample. The ZF-spontaneous #+ precession frequency from this crystal is shown
in fig. 6.14 as a function of temperature. As can be seen/3t,(T) changes not only
its value but also its direction abruptly at certain temperatures. In accordance with
neutron diffraction results (Endoh et al. 1989) this is traced back to spin reorienta-
tion transitions at 35 K and 75 K, respectively. Above 75 K the antiferromagnetic
structure is the same as in La2NiO4, it changes between 35 K and 75 K to the mag-
netic structure established in La2CuO4_~ and reverts back to the La2NiO4 structure
below 35 K. The ordered moment amounts to generally ~ 0.5#B. The reduced value
of Bu(O K) as compared to Bu(OK) in La2CuO4 is a consequence of the different
magnetic structure. Note that B , (0 K) is of similar order of magnitude as the lower
field B(~2) in La2NiO4 (see table 6.2). A spin reorientation transition has also been
observed by/zSR in Pr2CuO4 at ~ 40 K (Akimitsu et al. 1994).

Nd2_zCe=CuO4_a. Replacing some of the trivalent Nd 3+ (or Pr 3+) by tetravalent


Ce4+ mobile electrons are introduced into the Cu-O planes. For a critical concen-
tration z ~> 0.14 superconductivity is observed with Tc,m~x~- 24 K (up to z ~- 0.18).
Another necessary condition for superconductivity seems to be a certain oxygen de-
ficiency (6 > 0.01) (Takagi et al. 1989). Thus electron doping has the same effect
as hole doping in the La2CuO4-system. The question then was whether also the
magnetic order, present in the undoped parent compounds, is affected in the same
238 A. SCHENCK and EN. GYGAX

300 i

La2.xSr,Cu04 Nd2.,CexCuO4
250

-~ 200
(11
n
AFM
E 150
c-
O
,*2--
ffJ 100
t'-
SC AFM ~C

I
0
0.2 0.1 0.( 0.1 0.2
DopantConcentration
Fig. 6.15. Magnetic and superconducting phase diagram of hole doped La2_=Sr=CuO4 and electron
doped Ndz_=CexCuO4. Solid symbols are deduced from/~SR-data (from Luke et al. 1990b).

way by electron - as by hole doping. #SR spectroscopy on Ce-doped Nd2CuO4


and Pr2CuO4 found indeed a rather abrupt suppression of Cu-magnetic order for
z > 0.14 (Luke et al. 1989a, 1990a, b, Davis et al. 1990, Akimitsu et al. 1994)
in agreement with later neutron diffraction work (Rosseinsky et al. 1989, 1991a).
These measurements served to track the ordering temperature as a function of :c. The
results are displayed in a phase diagram, combining the systems Lae_~Sr~CuO4 and
Ndz_=CexCuO4 (fig. 6.15). It is again found that magnetic order is not established
at once in the whole sample volume. A sizable paramagnetic fraction survives to
temperatures much below the ordering temperature TN. It is also found that a very
wide field distribution is sensed by the #+ - rather than a unique field, indicating
that the magnetic order is disturbed and probably not of very long range nature.
Figure 6.15 shows that TN declines much more rapidly with rising electron doping
than with rising hole doping. This is understood in the following way: electron dop-
ing will change Cu 2+ to Cu 1+, the latter being nonmagnetic like a Zn ion substitute,
diluting effectively the Cu-moment system. In contrast holes are mainly located at the
oxygen ions, changing O -2 to O-, the latter carrying spin and inducing effectively a
ferromagnetic coupling between Cu-moments and hence introducing frustration into
the antiferromagnetic order. Frustration is known to have a much stronger effect
than dilution on the suppression of magnetic order. On the other hand the maximum
of Tc appears at approximately the same level of hole or electron doping.
In contrast to the LazCuO4-system no intermediate doping regime is found where
superconductivity and magnetic order of spin glass type coexist. Rather there seems
to be a common sharp boundary separating the superconducting and nonmagnetic
phase from the magnetic and semiconducting phase.
MUON SPIN ROTATIONSPECTROSCOPY 239

Nd2_~SrxCu04_6. It is also possible to introduce holes into the T~-phase Nd2CuO4


by substituting Sr2+ for Nd 3+ (Rosseinsky et al. 1992). Due to a slight oxygen
deficiency (3 >t 0.03) holes were only created for x/> 0.1. ZF-#SR measurements
showed indeed a magnetic behaviour for x <~ 0.06 which was little changed from
the undoped parent compound. For a sample with x = 0.06, TN dropped to 240 K.
Also the spin reorientation transitions where hardly affected in this doping regime
(Rosseinsky et al. 1991b, 1992). However, for x = 0.10 (formal Cu-ionic charge:
+2.02) the #SR data revealed a drop of TN to 60 K and no signs for the spin
reorientation transitions were any longer visible. Also the decay rate of the #+ spin
precession signal was considerably faster than in the samples with smaller x, implying
a certain amount of disorder. Nevertheless over all long-range-order seemed to
prevail (at least in part of the volume) as in the other samples, as was evident from the
observation of magnetic Bragg peaks in neutron scattering (Rosseinsky et al. 1991b).
Increasing x to 0.20 (formal Cu-ionic charge: 2.04) no precession signal could be
found any more and no Cu-moment could be detected at 1.5 K by powder neutron
diffraction. Obviously long range magnetic order was now suppressed. But the ZF-
#SR signal still displayed the presence of a wide static field distribution with zero
average typical for a spin glass, which starts to develop below ~ 100 K in part of the
sample volume. Both in the x -- 0.01 and the x = 0.02 sample pronounced relaxation
was also visible in the paramagnetic regime indicating a much slowed down Cu-
moment dynamics in comparison to the hole free samples. The phenomenology of
magnetism and its dependence on hole doping in Ndz_xSr~CuO4_6 has thus very
much in common with the magnetic behaviour in the La2_xSrxCuO4_6 system.
However, no superconductivity was observed in Ndz-xSrxCuO4. Suppression of
magnetism by hole doping is thus not necessarily followed by a superconducting or
even a metallic phase.

L a 2 _ x T b x C u O 4 (x -~ 0.8). The tetragonal crystal structure of this compound is shown


in fig. 6.16. It arises from a fusion of half the T' (Nd2CuO4) and half the O/T
(La2CuO4)-unit cells, forming the so called T*-phase in which Cu is five fold coor-
dinated. Lal.zTh0.8CuO4 is a parent compound of yet another system which becomes
superconducting upon hole doping (Sawa et al. 1989). Little is known sofar about
the phase diagram of this system. ZF-pSR spectroscopy was first to reveal the
presence of magnetic order in the parent material of the T*-phase (Lappas 1993).
Subsequently the strength of the staggered Cu-moment (= 0.46(6)/zB) and the an-
tiferromagnetic structure could be determined by neutron scattering. The magnetic
structure was found to be identical to the one in La2CuO4 (Lappas and Prassides,
to be published). High resolution ZF-#SR spectroscopy displayed the onset of a
spontaneous spin precession signal below ~ 170 K, but already below ~ 220 K
evidence for some sort of random static order appeared as evidenced by a strong
relaxation. Static random order may have even survived in a small fraction (_~ 15%)
of the sample volume up to room temperature (AB, _ 2.3 mT). Below 220 K one
notices a steady increase of the magnetically ordered volume fraction which lev-
els off around 80% below 100 K. The temperature dependence of the spontaneous
240 A. SCHENCKand EN. GYGAX

Fig. 6.16. Crystal structure of T*-phase (La, Tb)2CuO4. Cu is fivefoldcoordinatedby oxygen.

B u is displayed in fig. 6.17. It can be fitted below 150 K by a simple power


law:

B.(T) = B,(O) [1 - -T-~N] ~ (6.2)

The parameters TN and /3 depend on the temperature range covered by the fit.
Two best fit curves are indicated in fig. 6.18 with/3 = 0.36(3), TN = 130(1) K and
/3 = 0.49(4), TN = 146(1) K, respectively. The persistence of magnetic order above
these temperatures could point to the appearance of a different magnetic phase in
part of the sample volume. An exponent of/3 = 0.36 is in close agreement with the
prediction (/3 = 0.367) for a 3D S=1/2 Heisenberg antiferromagnet. An exponent of
/3 = 0.5 would better correspond to a mean field model. A transition temperature
closer to 130 K is also indicated by the exponential relaxation behaviour ()~zF(T))
in the paramagnetic volume fraction. It shows a pronounced divergence at around
120 K (see fig. 6.18), a temperature at which about 50% of the sample is still in the
paramagnetic phase. This feature signals a considerable slowing down of the spin
dynamics in the paramagnetic volume at ~ 120 K. Strangely below this temperature
the dynamics appears to speed up again. The decay of the spin precession signal does
not show any anomaly at ~ 110 K and seems to arise only from the spread in static
local fields. The temperature dependence of ~ZF follows an Arrhenius behaviour
between 120 and 170 K yielding an activation energy of ~-, 0.06 eV. From the value
MUON SPIN ROTATION SPECTROSCOPY 241

350 j , , , , , , ,
"'-...'~. | ...... b = 0.36(3)
300 - -,r~.~. -- b = 0.49(4)

250 E. ~ . . "'"'"
I ''"~.,.,

~,~"" 200 i N~el


m= 1 5 0 . """,.

'°°t
50

Oil
0
i I
50
i
100
I
150
= I =
200
Temperature (K)
Fig. 6.17. Temperature dependence of the spontaneous B# in Lal.2Tb0.sCuO4. The dashed and solid
lines represent fits of eq. (6.2) to the data, see text (from Lappas 1993).

3.0 I I

2.5
if)
2.0

t-
1.5
O

1.0
CC

0.5
0 0 -O-

0.0 I I I i i
0 50 100 150 200 250 300
Temperature (K)
Fig. 6.18. Temperature dependence of the ZF-relaxation rate in the paramagnetic volume fraction of
La]2Tb0.8CuO4. The sharp cusp seems to coincide with TN, determined from B#(T) (see fig. 6.17)
(from Lappas 1993, Lappas et al. 1994b).

o f AZF at 20 K a fluctuation rate o f u ~ 109 s -1 can be estimated ( L a p p a s 1993,


L a p p a s et al. 1994b).
O n s e t o f m a g n e t i c o r d e r o f the Tb 3+ 4 f - m o m e n t s is o b s e r v e d b e l o w 20 K b y a
drastic i n c r e a s e o f the d e p o l a r i z a t i o n rate o f the spin precession signal ( L a p p a s et al.
1994b).
242 A. SCHENCK and EN, GYGAX

Sr2Cu02Cl 2. Also this compound is a layered perovskite with the tetragonal K2NiF4
structure. Compared to La2CuO4 all apical oxygen ions are replaced by CMons (see
fig. 6.13a). This leads to an increase of the unit cell along the c-axis by 20%. The
Cu-O coordination is thus four fold and in this sense Sr2CuO2CI2 resembles the
T'-phase (Nd2CuO4). Neutron scattering and magnetic susceptibility measurements
have revealed that SrzCuOzCI2 orders antiferromagnetically like La2CuO4 (Vaknin et
al. 1990). Magnetic order has been also established by ZF-#SR measurements on a
single crystal sample (Le et al. 1990a, b). A spontaneous spin precession signal was
found which disappears sharply at 260 K (see fig. 6.19). It consist of one frequency
down to 60 K and splits into two components with equal weight below 60 K. By
using different orientations of the single crystal sample with respect to fi(0) it was
determined that the spontaneous fields are directed parallel to the a - b plane. Note
that in La2CuO4 Btz is pointing out of the plane. The low temperature frequency
average (N 16 MHz) corresponds to a field of 0.12 T which is much larger than
B u (0 K) in La2CuO4. These differences are ascribed to a different #+-site which,
since there is no @ical oxygen, must be located near an oxygen in the CuO2 plane.
The splitting of Bu for T < 60 K is believed to be of intrinsic origin since there is
also a peak in the magnetization near 50 K when the external field is applied parallel
to the c-axis. The splitting may arise from a possible spin reorientation transition
not seen so far in any neutron scattering work.
The temperature dependence of B~, resembles the results obtained in La2CuO4. As
in the latter compound it has been explained on the basis of a 2D-Heisenberg model
with A -~ 10 -5 (see section 6.1). The much smaller value of A in the present case
implies a very weak anisotropy field as a consequence of the tetragonal structure (Le
et al. 1990a).

20 I I I I I

'=
61: ..... H I I c
N 15 ~Lm~ ~ Aax ~2
z
LX 0/ I i I I T"--r
Z~ O 50 100 150200250300
Zk Temperature (K) "-'
o
c- 10 I

12r o P# I l c a x i s ~LX
A
u_ 5 _ zx P # / caxis
Zk

0 I I I I I
0 50 100 150 200 250 300
Temperature (K)
Fig. 6.19. Temperature dependence of the spontaneous /z+ spin precession frequency (or frequencies)
in monocrystalline Sr2CuO2C12. A splitting develops around 50 K which correlates with a peak in the
magnetic susceptibility (see insert) (from Le et al. 1990a).
MUON SPIN ROTATION SPECTROSCOPY 243

La2MCu206+6 (M = Ca, Sr). The hole-doped double-layer system La2MCu206+6


(see, e.g., Fuertes et al. 1990) has shown superconductivity for M = Ca and for Sr
doped La2_xSr~CaCuO6+6 but not for the compound with M = Sr. Prior to #SR
measurements no magnetic order in the parent material had been identified. But
ZF-#SR spectroscopy (Ansaldo et al. 1991, 1992) proved again the presence of
magnetic order by the appearance of a spontaneous spin precession signal allowing
also to determine the transition temperature TN. In the system La2SrCu206+6 it
was found that TN decreases from more than 300 K for a sample with 6 < 0.06 to
,-, 250 K for a sample with 6 = 0.2. A TN -~ 250 K was measured in a La2CaCu206+~
with 6 -~ 0.037. Common for all samples was a two frequency precession at low
temperatures corresponding to local fields of B~, (0 K) = 0.036 T (main signal)
and B~ (0 K) = 0.13 T. The lower field is probably associated with #+ bound to an
apical oxygen while the high field is associated with #+ near a CuQ-plane. Note that
B~ (0 K) = 0.13 T is close to the corresponding value in Sr2CuO2C12 (see table 6.2).
The high frequency component becomes blurred above ~ 250 K. Very peculiar is
the temperature dependence of Bu(T) in the LaaCaCu206+6 sample (see fig. 6.20).
In contrast to the M = Sr samples both B~, drop by roughly a factor of 1.75 when
going from the lowest temperatures up to ~ 30 K (Ansaldo et al. 1991, 1992). The
authors suggest that this drop may be associated with the onset of superconductivity
due to the mobility and segregation of the excess oxygen intercalated in a Ca(La)
layer between two CuO2-1ayers. As the temperature is lowered the excess oxygen
may cluster into domains producing locally a high enough hole concentration to
facilitate superconductivity and depleting other regions from holes thereby stabilizing
the antiferromagnetic order there and restoring the moments to their full staggered
value. The presence of superconductivity below 45 K in about 15% of the sample
volume was indeed interfered from weak W-measurements (Ansaldo et al. 1992).

18 I I I I I I I I

16
14
1 12
10 Ill•
¢.-
8
6
LL
4
0 0 0
2
0
I I i I I I I I I I
0
0 100 200
Temperature (K)
Fig. 6.20. Temperaturedependence of the two spontaneous/z+ spin precession frequencies in
La2CaCu206.037. The pronounced increase of both frequenciesbelow (20-30) K is attributed to a
segregation into oxygen (hole) rich and oxygen depleteddomains (from Ansaldoet al. 1992).
244 A. SCHENCK and EN. GYGAX

6.2. YBa2Cu30x and related compounds


A list of compounds studied by #SR spectroscopy is given in table 6.3.
Possible #+-sites have been considered in a number of papers. Theoretical work
has been presented by Boekema (1988), Dawson et al. (1988), Li and Brewer
(1990), Halim et al. (1990), Lichti et al. (1990a, b) and experimental work has
been discussed by Dawson et al. (1990, 1991a), Nishida et al. (1990a), Brewer et
al. (1990) and Weber et al. (1990). There is a general consensus that, like in the
L a 2 C u O 4 system, the #+ will be bound to an oxygen, forming a kind of an O - H bond
with a bond length of ~ 1 ,~. More specifically, in the fully oxygenated (123)-system
(x = 7) the #+ is preferentially found near a chain oxygen at a position (0.14, 044,
0.071) (Weber et al. 1990, see fig. 6.21). The same site is also taken by hydrogen
in hydrogenated YBa2Cu307 (Weidinger et al. 1994). With less probability the #+
is also found near an oxygen in the CuO2-plane. This position is not very precisely
known. In the oxygen depleted (x = 6) system all the chain oxygen is absent and
the #+ is now found near an apical oxygen.

6.2.1. YBa2Cu30~
The very first indication for magnetic order in an unspecified member of the super-
conducting YBaCuO-family was provided by Nishida et al. (1987a), using ZF-#SR.
Subsequently the same group (Nishida et al. 1987b) was first to prove the existence
of long range magnetic order in the tetragonal phase of YBa2Cu3Ox, the parent

b/ 1', r. sa / ___k

Fig. 6.21. Lower part of the crystallographic unit cell of YBa2Cu3Oz with possible/z + sites indicated
by olden circles and filled circles. The cluster of points to the right (center is at (0.14, 0.44, 0.071) about
1.0 A away from the chain oxygen 0(4)) and equivalent positions is found for x ----7, the more spread
out distribution of sites close to the apical oxygen O(1) (distance is again of the order of 1.0 ,~) and
equivalent positions for x -- 6 (from Weber et al. 1990).
MUON SPIN ROTATION SPECTROSCOPY 245

[..
F

d c~
,,d
V

v~ o

v~

,,,.4,

I v~ ~ ~ v~

e- c4 ~:
+
::::1. ~t II II II

0 =L&:&

<~
0

o
~ tl u u
q~ tt~
N c5
# ~v II "~ o--

?
~ A ~ V

t~
0
v~v/~v/~v/v~V/V ~v/ ~ o~
v/Zv/ ~v/av/vav/~/~.v/ ~¢~
246 A. SCHENCKand EN. GYGAX

500 I i I I I I f I I I

YBa2Cu3Ox

400 gSR [2] quenched


[2] annealed

n.s. [41
[51
300 I

g
AFM ~ ~ []
!
200

~ell- j~
annealed~ I Ortho-I
100 I
'l
Ortho.ii r ~ )O00"-

i~ i i i i

6.0 6.5 7.0


X
Fig. 6.22. Magnetic and superconductingphase diagram of YBa2Cu30=. The magnetic part stems
from/zSR and neutron scattering measurements: (1) Nishida et al. (1988b), (2) Brewer et al. (1988),
(3) Tranquada et al. (1988), (4) Rossat-Mignodet al. (1990), (5) Rebelsky et al. (1989) (from Nishida
1992b).

material for the high temperature superconductor YBa2Cu307. Long range antifer-
romagnetic order was confirmed shortly after by neutron diffraction measurements
on an oriented powder sample (Tranquada et al. 1988). Since then #SR experiments
on the YBazCu30= system investigated the dependence of magnetic order on the
oxygen content and in particular tried to verify a possible coexistence of magnetic
order and superconductivity in a certain range of z like in the La2CuO4-system (see
table 6.3). Results on the transition temperatures following from these experiments
are summarized in fig. 6.22, including data from neutron scattering (Rossat-Mignod
et al. 1990, Rebelsky et al. 1989, Tranquada et al. 1988) and superconducting
transition temperatures (Nishida 1992a, b), As in the LazCuO4 system TN decreases
with the hole concentration (oxygen content), most rapidly in the range z = 6.2-6.4,
depending on the sample preparation method: samples prepared by an annealing
procedure show a drop of TN at smaller z-values while in samples quenched from
high temperatures the drop of TN is shifted to slightly larger z-values (Brewer et
MUON SPIN ROTATION SPECTROSCOPY 247

al. 1988). Weidinger et al. (1990) even found TN _~ 60 K in a quenched sample


with z = 0.5. The difference in behavior is ascribed to differences in oxygen or-
dering in the Cu-O chain planes for nonstoichiometric oxygen concentrations. By
very slow and careful annealing the orthorhombic phase can be stabilized down to
z = 6.2. For quenched samples the tetragonal phase is observed which converts to
the orthorhombic phase at higher x-values (for a detailed discussion and references
see Nishida 1992b). Independent of preparation technique TN is rather constant up
to z = 6.2. It is argued that the addition of oxygen into the chain position up to
z = 6.2 does not introduce holes into the CuO2-planes and thus cannot disturb the
intra plane antiferromagnetic exchange coupling (Rossat-Mignod et al. 1990).
Figure 6.22 shows that magnetic order (not necessarily of long range nature) and
superconductivity seem to coexist around and below z = 6.5. Actually all available
data show that magnetism and superconductivity exist in the same sample, if Tc is
below approximately 50 K (see Weidinger et al. 1990). Again the question arises on
what scale this coexistence takes place. A macroscopic separation into a magnetic
and a superconducting phase can be ruled out since TN and Tc vary in a continuous
way with the oxygen content. Although the fraction of #+ sensing magnetism
declines with rising z or Tc, respectively, the corresponding volume fraction is not

100

90

80

70

o~ 60 0.57
t-
O

50
.2

\7
'$
t-

t~
40
3;
0.67
30
\
YBa2Cu306+x \
20
\
\
10 \
\
I I } I I
0
10 20 30 40 50 60
Tcrnidp°int (K)

Fig. 6.23. Fraction of magnetically ordered volume in YBa2Cu306+= as a function of z. The fraction
is obtained from the amplitude of the corresponding/~SR-signal (from Weidinger et al. 1990).
248 A. SCHENCKand EN. GYGAX

320 I I I I

300
c I P. Site#1

280

L9
260
.._=
o~
C
0 240

1300
i I
,ev ~
I I
8 MHz
[-
0
._J 1260

1220

1180 ii i l i~~
0 50 1O0 150 200
Temperature (K)
Fig. 6.24. Temperature dependence of the two spontaneous /z+ precession frequencies in an oriented
powder sample of YBa2Cu306.0 (from Brewer et al. 1989).

associated with a constant TN, as one may suspect if macroscopic magnetic domains
would be present. The #SR data also show that the spontaneous static internal
fields at the #+ decrease very rapidly when the superconducting phase is entered
(Weidinger et al. 1990, Brewer et al. 1989). Also this is inconsistent with a
macroscopic phase separation picture. However, one cannot exclude the coexistence
of microdomains of orthorhombic and tetragonal structure which are so small that
they will influence each other, explaining the continuous variation of TN and To. The
presence of a spontaneous #+ spin precession signal or of at least coherent short
range magnetic order places a lower limit on the tetragonal domain size of ~ 50 .~.
This is still small enough to not inhibit superconductivity~ due to a proximity effect
vis-a-vis a superconducting coherence length of (20-30) A. Of course the amplitude
of the spin precession signal must be reduced (as is observed, see fig. 6.23) and
#+ in the orthorhombic domains will still feel some random fields, spilling over
from the tetragonal domains, causing depolarization (as is also observed). It has
been also suggested that an electronic phase separation could produce hole rich and
hole depleted microdomains (Emery et al. 1990). The general picture provided by
#SR on the phase diagram of YBa2Cu30~ fits well with other results, notably from
neutron scattering experiments.
We now discuss some more specific #SR results. ZF-#SR in the ordered state
displayed the presence of clearly two frequencies (Brewer et al. 1989, Budnick et
al. 1990a, Barsov et al. 1994). The dominant frequency (B, (0 K) ~_ 0.03 T) is seen
MUON SPIN ROTATION SPECTROSCOPY 249

18 ~illl • YBa2Cu30x
17 il

16
+ a

~ Q x II~ •

g • 6.08
~OAcL ° ~'
+
LL x 8.16 4~
2
0 6.31 (~
• 6.32

I I I I I I I I I I l I I

40 80 120 160 200 240 280

• 6.08 /i
4 × 6.16 /~jt J

-- 2 b

n- 1

I t I I I I I I I I 1 I I I I
0 40 80 120 160 200 240 280
Temperature(K)
Fig. 6.25. Temperature dependence of (a) spontaneous /~+ precession frequencies and (b) associated
relaxation rates in samples with various oxygen concentrations (from Barsov et al. 1994).

for :c up to 6.5 (depending on the preparation procedure). A second high frequency


component ( B , (0 K) = 0.13 T II c-axis) is only seen in samples with x < 6.3.
Occasionally a third frequency is seen corresponding to a Bu (0 K) -- (0.008-
0.015) T (Nishida et al. 1990a). The three frequencies correspond to the three sites
discussed above and are consistent with the antiferromagnetic structure determined
by neutron diffraction measurements (Tranquada et al. 1988). The dominating
B , (0 K) is rather insensitive to the oxygen concentration for z < 6.2-6.32 (Kiefl et
al. 1989, Weidinger et al. 1990, Barsov et al. 1994). In contrast and analogous to
the La2CuO4-case neutron scattering showed a pronounced decrease of the magnetic
Bragg peak intensity when z was changed from 0.15 to 0.38 (Rossat-Mignod et
al. 1990). The temperature dependence at B~ = 0.03 T and B~ = 0.13 T in an
oriented z ~_ 6.0 powder is displayed in fig. 6.24 (Brewer et al. 1989). Very peculiar
is the rise of Bu below ~ 80 K (T/TN ~- 0.2) which is more pronounced for the
dominant 0.03 T-signal which arises from ~+ located near the apical oxygen. This
250 A. SCHENCK and EN. GYGAX

anomaly is also found in sample with z -- 6.08-6.32 (Nishida et al. 1988b, Barsov
et al. 1994). Interestingly a similar anomaly at T/TN ~- 0.2 is also found in the
(2212)-family (see below). At 10 K another anomaly shows up in that Bu(T ) drops
sharply for the 0.13 T-signal and slightly for the 0.03 T-signal. This drop has not
been seen in the data of Nishida et al. 1988b (z _~ 6.2) and of Barsov et al. 1994
(z = 6.08-6.32, see fig. 6.25a). Its origin in the z = 6.0 oriented powder sample
is not explained. The 80 K anomaly shows also up in the staggered magnetization
determined from the intensity of the magnetic Bragg peaks in neutron scattering for
samples with z > 6.2 but apparently not in samples with z < 6.20 (Rossat-Mignod et
al. 1990). In contrast to the #SR results the magnetization decreases below ~ 80 K.
Interestingly also the spread ABe(T) of the local field B,(T) increases below 80 K
(Barsov et al. 1994, see fig. 6.25b). Taken together it seems that long range order
gets disturbed below 80 K (at least in samples with z > 6.08) with the effect that the
Bragg peak intensity decreases and the field spread at the #+ site is increased. That
also Bu(T) increases more rapidly below 80 K may be a subtle result of forming the
vector sum of dipolar fields at the #+ position (the position near an apical oxygen
may be particularly sensitive to slight changes in the moment order since it is situated
almost half way between the CuO2-planes). The appearance of a certain disorder
below 80 K may be associated with a freezing of holes at the oxygen sites.
Finally it has to be emphasized strongly that from the perspective of the implanted
#+ the onset of magnetic order never involves the total sample volume at once.
Depending on the sample preparation magnetic order in most of the sample volume
may only be obtained at temperatures significantly below TN. This may be explained
in terms of a distribution of transition temperatures due to an inhomogeneous distri-
bution of the chain oxygen ions, but this explanation cannot very well be applied to
samples with z ~< 6.2, where TN varies only little.
Exploratory TF-#-SR-spectroscopy on YBa2Cu30~ (z = 6,6.5,7) has been re-
ported by Nishida et al. 1991, Nishida 1993).

6.2.2. Yl_vPrvBa2Cu30~ (x = 6,7)


Replacing Y in Y B a 2 C u 3 0 7 by other rare earth atoms the superconducting properties
of the (123)-O7 system are hardly effected. A notable exception is a replacement by
Pr which - in contrast to Y or other rare earth substitutes - is believed to have an
ionic charge of ~ 4 + instead of 3 + and hence donates electrons 6. Therefore one may
expect that Pr-substitution has a similar effect on the hole concentration in the CuO2-
planes as removing oxygen. Indeed PrBa2Cu307 does not show superconductivity
but antiferromagnetic order of the Cu-moments in the CuO2-planes instead (Felner
et al. 1989). A remarkable difference to YBa2Cu306 is the fact that PrBazCu307
retains the orthorhombic structure. ZF-#SR measurements on polycrystalline sam-
ples served to determine the phase diagram of Yl_vPrvBa2Cu307 (Cooke et al.
1990a-d). Long range antiferromagnetic order, again in evidence via a spontaneous

6 This view is challenged by X-ray absorption (XANES) results which show that Pr in the (123)-
compounds is in the trivalent state (Pr 3+) (U. Staub, priv. communication). The appearance of Cu 2+-
moment order for y > 0.5 and the destruction of superconductivity must, nevertheless, imply that the
holes in the CuO2-planes are removed and instead are probably located close to the pr3+-ions.
MUON SPIN ROTATIONSPECTROSCOPY 251

spin precession signal, was observed for 0.5 < y ~< 1. The temperature dependence
of the spontaneous frequency u~ is displayed in fig. 6.26 for various y. The abrupt
drop in u, below 20 K for y = 1 and below 6.5 K for y = 0.8 is probably arising from
the onset of magnetic order of the pr4+-moments. This phase transition at TN2 has
also been seen in susceptibility, specific heat and neutron scattering measurements. It
seems, however, that pr4+-moment ordering alone is not sufficient to explain the drop
of ur, satisfactorily. Rather it appears necessary to invoke in addition a small (0.02#B)
ordered moment on the Cu-chain sites (Dawson et al. 1991b). Such a small moment
may have escaped detection in NMR/NQR measurements (Ltitgemeier 1989).
The spontaneous field at the #+, extrapolating the values above 20 K to zero
temperature, B~ (0 K) = 0.016 T, is clearly smaller than the corresponding value
in YBa2Cu306 (see table 6.4). The difference arises from the different sites that
are taken by the #+ in the 06- and the O7-compounds and is not the result of
different values of the ordered Cu-moments. The value of 0.016 T is close to what
is found in hydrogenated YBa2Cu307 (see below) and occasionally in YBa2Cu306+6
(5 < 0.5) and confirms a #+-site assignment close to a chain oxygen. As before in
YBa2Cu306+~, B~(0 K) seems to be rather independent of the Pr-content y or the
hole concentration, respectively.
At y = 0.5 no long range magnetic order could be detected. Instead a fast relaxing
signal was seen which could be well described by a function typical for spin glasses
with nearly static features (Cooke et al. 1990d). From the temperature dependence
of B~ for y > 0.5 and of the relaxation rate for y = 0.5 the N6el temperatures or a
spin freezing temperature TN1 could be determined which are displayed in fig. 6.27
as a function of y. Included are also superconducting transition temperatures Tc
for compounds with y < 0.6 as well as the additional magnetic transition tempera-
tures TN2 at low temperatures. The resulting phase diagram is very similar to the
YBa2Cu306+6-phase diagram including a narrow regime of coexisting superconduc-
tivity and spin glass order. This feature has to be commented in the same way as
for YBa2Cu306+6.

i i iiii I ~1 i i i iiii I i i

......~..... ___.....
2.0

g 1.5

°>" 1.0

IJ_
•~ o.s

2 5 10 20 50 100 200 300


Temperature (K)
Fig. 6.26. Temperaturedependenceof spontaneousprecession frequenciesin various
(Yl-=Prx)Ba2Cu3OT-samples (fromCooke et al. 1990d).
252 A. SCHENCK and EN. GYGAX

d OX ~ OX O'x

0
8
.'2. ~b
z N m m

o
o
~oEE".'~ ~~ . ~> - ~ ~ -~~ ~ ~~ ,'~~
b b"Sb -~l~,-.==
K8 K8 o
r~

0 O

~o v 8 A~ V/ V/ II ~ ii
o
a=
o o o ~ , , II

A~ A~

A~ A~
0
O

ca

d
O

g 8
O
c~
0
I 6
,eq
::l
0 oc~ d c5 c~ci o o o

e-, d c~ c~oo
0 "0 c~
N .... V/ v~ www~
0
o ~ v~ v~ V/V/V/~ c~ ~.
0 0
r..) r...)
M U O N SPIN R O T A T I O N S P E C T R O S C O P Y 253

. . . . . ~.-~ Ox ~-~ ~'0~ ~ ~-~" O~

• ~ ~ ~ ~ "~."~ ..-~ . ~ .-~ .--~ . .

0 ~o o ~ ~ "-~ "8 o o o

c~
0

II II
V/o

t¢'5

z A~
~ ~ ~
~ ~ °~ "
e~ e~ F~

cq
m

o.

o.

o
d
"~ v~
254 A. S C H E N C K and EN. G Y G A X

oo
6" o~

=..~

aeN.o~ N .~.~.~.~

o ,?_. o ,~ o ~ X ~ o

u u o
b o o b o b b
?~ _~
"~"6 o ~ o o o o o

e~
0
~. A~ ~g;,.

~s v/
~ U~ ~ ° . ~ o , o
~,, • . .~.~ ~

~: ~ II ~ ~,.o o~
"0

1¢3 1¢3

A A A~
~e ~e

o~. ~ ,~

o d ~. "~.~

, ,. ~' ~
v/ ~ , ,-,0~
II V/ ~:~ II II II II II v~
6
MUON SPIN ROTATION SPECTROSCOPY 255

"~ -~ "~ .~

•~ , ~ --~ ~ .

0 0 0

0 0 0 0

~ ~ o ~ ~.~

~ . o o o

II
V

d d d d
II II II II
256 A. SCHENCK and EN. GYGAX

5 I I I 1 I I I I I

3OO
(Yl.xPrx)Ba2Cua07 , ,t
p
&,
//
m 200
TNI,"
It /
E
~. 100<
- ~"-.,~' T.2
o ' ' ' a-im~J--e-~--~--
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
X (Pr concentration)
Fig. 6.27. Magnetic and superconducting phase diagram of (YI_mPr=)Ba2Cu307. TN1 indicates the
onset of Cu(II)-ordering, TN2 possibly the onset of Pr-ordering and perhaps some ordering of the chain
Cu-moments as well TN1 and TN2 follow from/~SR-measurements (from Cooke et al. 1990 a~l).

In PrBa2Cu306 two spontaneous frequencies showed up, one corresponding to


the frequency seen in PrBa2Cu307 the other one being close to (in fact 16% larger
than) the value seen in YBa2Cu306+6. This points to the presence of an appreciable
amount of chain oxygen ions so that #+ can be found both near chain oxygen and
apical oxygen sites, respectively.

6.2.3. HzYBa2Cu307_6
Doping YBa2Cu307 with hydrogen has the same effect as replacing y3+ by Pr 4+
ions, namely it donates electrons to the CuO2-planes and reduces the hole con-
centration. This conjecture was first fully proved by ZF-#SR-measurements which
showed the presence of long range antiferromagnetic order in HzYBa2CuaO7 for
z > 0.5 (GliJckler et al. 1989, 1990a, 1990b, Niedermayer et al. 1989), following a
first indication for some sort of local magnetic order by NMR (Goren et al. 1989).
Persistence of superconductivity up to hydrogen concentrations of z = 0.2 was ob-
served before by Reilly et al. (1987). The ZF-#SR results revealed two spontaneous
precession signals corresponding to B (1) (0 K) = 0.015 T and _/~R (1) (0 K) -- 0.03 T.
The temperature dependence of R (i) for a sample with z = 0.9 (6 = 0) is repro-
duced in fig. 6.28. The higher frequency signal appeared only weakly while in
oxygen deficient samples it was the dominant component. This fact indicates that
the low frequency signal is associated with #+ bound to the chain oxygen and the
high frequency signal with #+ bound to the apical oxygen site. The near perfect
agreement of the frequency values with the corresponding values in YBa2Cu306
and Y l _ y P r y C u 3 0 7 (see table 6.4) proves that the antiferromagnetic structure of the
ordered Cu-moments in the CuO2-planes as well as the moment value is not affected
to any significant degree by H- or Pr-doping.
The ZF-#SR measurements show that, depending on the level of hydrogen doping,
only a fraction of the sample volume contributes to the spontaneous precession
MUON SPIN ROTATION SPECTROSCOPY 257

5 i i i i i i i

7H09
~. 4 - ' -
-r-

3
O
t.--
2
{:T

u_ 1

0 i
0 50 100 150 200 250 300 350 400
Temperature (K)
Fig. 6.28. Temperature dependence of two spontaneous/~+ precession frequencies in YBa2Cu307Ho. 9.
The solid lines represent spin 1/2-Brillouin functions (from Niedermayer et al. 1989).

30

25

o~ 20
g

0.5 1.0 1.5 2.0 2.5


Hydrogen concentration
Fig. 6.29. Fraction of the magnetically ordered volume in YBa2Cu3OTHz as a function of z. The
fraction is determined from the amplitude of the spontaneous spin precession signal (from Niedermayer
et al. 1989).

signals. As fig. 6.29 reveals this fraction is zero below z = 0.5 and saturates at about
20% above z ~_ 1.5. The data show that the remaining volume fraction is associated
with a nonoscillating, slowly and exponentially decaying signal, reflecting random
internal fields of a few 0.1 roT. Niedermayer et al. (1989) speculate that this signal
could be associated with a third site with very small residual fields in the ordered
state. More likely may be an explanation in terms of a non random distribution of
the dissolved hydrogen atoms (e.g., a phase separation into a hydride phase and a
largely hydrogen free (z < 0.13) phase; Reilly et al. 1987) leaving a large part of
258 A. SCHENCKand EN. GYGAX

the sample volume in a non magnetic but also non superconducting state since no
superconductivity is indicated for z > 0.8, as follows from TF-#SR-measurements
(Gltickler et al. 1990a, b). Consistent with the picture of hydrogen as an electron
donor, compensating holes in the CuOz-planes, is the observation that magnetic order
becomes manifest already at smaller z-values in oxygen deficient compounds (see
table 6.4) (Gltickler et al. 1989).

6.2.4. YBa2(Cul_vTy)3Ox,T = Zn, Fe


The replacement of Cu by trivalent ions, such as Fe 3+, mainly involves the Cu chain
sites while divalent Zn 2+ substitutes preferentially for Cu in the CuO2-planes. Both
types of substitutions reduce Tc and eventually suppress superconductivity altogether.
In this respect the 'nonmagnetic' Zn is much more effective than even the magnetic
substituents Co, Ni and Fe. Zn substitution does not alter the orthorhombic structure,
the oxygen stoichiometry and the hole concentration while in the case of Fe substi-
tution the hole concentration is reduced and the orthorhombic structure changes to
the tetragonal one for y/> 0.04. This structural change can be avoided by employing
a special thermal treatment (Katsuyama et al. 1990). The interest in studying the
effects of substitutions on the Cu sites is motivated by the expectation to learn more
about the possible role of magnetic interactions in the pairing mechanism. While the
effects of substitutions on the superconducting properties have been studied in some
detail much less is known about the magnetic features induced by the substitutions.
#SR-spectroscopy was applied to gain more information in this respect.
Zn-substituted compounds were studied by Garcia-Munoz et al. (1991) and Men-
dels et al. (1994) with respect to magnetic properties. Superconducting properties
were studied by #SR by Albanese et al. (1992).
Zero field #SR and neutron polarization analysis were applied in a search for
localized moments in a YBa2 (Cu0.96Zn0.04)306.95-sample. No evidence for static or
slowly fluctuating local atomic spins were detected. However, Tc was reduced to
,-~ 47 K. More extended ZF and TF-#SR measurements by Mendels et al. (1994) on
samples with y = 0.04 and varying oxygen deficiencies allowed to draw a complete
phase diagram which is shown in fig. 6.30. The interesting part of it is found in the
range 6.4 < z ~< 6.67 where a disordered magnetic structure is clearly revealed by
the ZF-data which extends right up to the metallic threshold but does not seem to
overlap with the superconducting phase. Note the difference to pure YBazCu3Oz.
Obviously Zn-doping has a destructive effect on long range magnetic order but
stabilizes a random magnetic order to a higher level of hole doping. The latter is
ascribed to the effect of Zn-doping on the hole dynamics which is slowed down as
the missing Cu ions limit the spatial extension of the wave function of the holes.
Comparing Zn-doped YBazCu307 with Zn-doped La2CuO4 it appears as if Zn-doping
interferes differently with the magnetic order in the two systems. One reason may be
that, although to a lesser extend, Zn replaces also Cu in the Cu-O chains. Another
remarkable result of Mendels et al. (1994) is the observation of a Curie behaviour of
the transverse field relaxation rate in the paramagnetic regime for 6.43 ~< z ~< 6.88.
This points strongly to the presence of local moments also in the paramagnetic phase.
MUON SPIN ROTATIONSPECTROSCOPY 259
i.-- i-.--' ' I ' i i I i , i I i l i I ' ' '

q~ll,~-
- =~=i ....= ....~
\ 4'/o Zn
o

- \ \ ---- pure
._. 100 AF ~ i//~" .... ~'/

lo / so
/k TN,~SR L
0 Tc,~ac ~..'~. DM J .
1 , , , i , , ,,"-~
~ - , ", -, \, ', ,~V , i , ,

6 6.2 6.4 6.6 6.8 7


Oxygen concentration, x
Fig. 6.30. Magnetic and superconducting phase diagram of YBa2(Cu0.96Zno.04)30=. Magnetic critical
temperature TN and Tf are determined from NMR- and/zSR-data. For comparison the phase diagram
of pure YBa2Cu30= is also schematically shown (from Mendels et al. 1994).

These moments must be created due to the presence of the Zn-substituents. Again
this may be a consequence of a modified hole dynamics.
Fe-substituted compounds were studied by Matsui et al. (1990), Okuma et al.
(1990), Kossler et al. (1990) and Baggio-Saitovitch et al. (1990). Previously
MSssbauer spectroscopy (for a list of references see Okuma et al. (1990) and Matsui
et al. (1990)) and neutron time of flight measurements (Mirabeau et al. 1989) had
already provided evidence for a freezing of the Fe-moments, the latter experiment
suggesting a type of spin glass order in an y = 0.06 sample below 18 K in coexistence
with superconductivity. In contrast MSssbauer data obtained by Baggio-Saitovitch
et al. (1990) seem to indicate some non random but probably very short range
order in a semiconducting sample with g = 0.05. All #SR-data, mostly taken in
the zero field configuration, are consistent with a completely random order of the
Fe-moments: PZF(t) did not show any oscillation in samples with y >/0.04 but was
well described by a stretched exponential function (see eq. (2.22)) above the spin
freezing temperature Tf and by a modified spin glass Kubo-Toyabe function below
Tf (Okuma et al. 1990). The spin glass order below Tf was not completely static
but showed a fluctuation rate of the Fe-moments of uFe ~-- 2 x 105 s -1 at 4 K in a
tetragonal sample with y = 0.08 (Okuma et al. 1990). The ZF-#SR data allowed to
determine the freezing temperature Tf with some precision. Figure 6.31 displays a
phase diagram of YBa2(Cu~_uFeu)30= which includes M6ssbauer, susceptibility and
/zSR data, and shows nicely the over all consistency on the spin freezing temperatures
(Matsui et al. 1990) (see also table 6.4). As mentioned above, by a proper heat
treatment YBa2(CUl_uFey)307 can be prepared to remain in the orthorhombic phase
for y > 0.04. Such an 'ortho' sample with y = 0.08 was also investigated by Okuma
et al. (1990). It was found from TF-/zSR measurement that about 1/3 (increasing to
about ~ 45% at ~ 25 K and decreasing again to ~ 20% at ,,~ 4 K) of the volume
260 A. SCHENCKand EN. GYGAX

t t t2
100 ~ YBa2(CUl_yFey)30x
0 Tc (Meissnereffect
" ~ _ N(CW) x Tg susceptibility
x 4 . . • ~+sR
Z~ Nasuet al
g 50 - S U~BTamakietal
F-- O %

o , o ,
0 o.o5 0.10 0.15
Y (Fe)
Fig. 6.31. Magnetic and superconducting phase diagram of YBa2(Cul_yFe#)30= (from Matsui et al.
1990).

remained superconducting with Tc = 90 K and no signs of magnetic order, while


the remaining volume fraction displayed random magnetic order below ~ 33 K.
These observations are explained in terms of the formation of Fe-rich domains,
which enter into a spin glass phase at low temperatures, and almost Fe-free domains
which become superconducting below 90 K by forming a percolative network. The
observed complementary temperature dependencies of the two volume fractions are
not explained.

6.2.5. Lanthanide substituted (123)-compounds


As mentioned in section 6.2.2 the full replacement of Y by lanthanide ions, with the
exception of Ce, Pr, Tb and Pm, does not affect superconductivity in the (123)-system
with Tc staying close to 90 K in fully oxydized samples. At low temperatures antifer-
romagnetic ordering of the lanthanide 4f-moments is observed without any adverse
effect on the superconductivity, although the lanthanide moments are sandwiched
by the superconducting CuO2-planes. Apparently the superconducting electrons are
very narrowly confined to the CuO2-planes, as also evidenced by the extremely short
coherence length along the crystallographic c-axis (~c -~ 3 A), and do not overlap
with the local 4f-electrons. The absence of conduction electrons at the rare earth
sites is also implied by M6ssbauer data (Alp et al. 1987, Smit et al. 1987). In view
of these features it is not so clear by what mechanism the observed 3D magnetic
order is driven. It is argued that dipole-dipole interactions may be responsible (Fel-
steiner 1989). Nevertheless, it is found that the amount of oxygen deficiency can
have a pronounced effect on the magnetic structure of the 4f-moments. For exam-
ple, in ErBa2Cu30= the antiferromagnetic structure changes from 2D for x < 6.5 to
3D for z > 6.5 (Maletta et al. 1990a). Also #SR-measurements show that the 4f-
spin dynamics is strongly affected by the amount of oxygen deficiency (see below).
These results indicate that at least the coupling between the rare earth-planes is of
the super exchange type involving the chain oxygen ions and then probably also
MUON SPIN ROTATION SPECTROSCOPY 261

oxygen ions in the CuO2-planes without inducing any pair breaking. The system
HoBa2Cu30~ is special in that Ho 3+ possesses a non magnetic singlet ground state
level and the observed magnetic order requires the formation of moments out of
this singlet ground state by either overcritical exchange or nuclear hyperfine effects
(Roessli et al. 1993).

GdBa2Cu30:~. Zero field/zSR measurements on two GdBa2Cu307_a samples with


Tc = 60 and 90 K revealed two spontaneous spin precession signals below TN =
2.3 K, confirming the onset of antiferromagnetic order of the Gd-4f moments (Golnik
et al. 1987). The temperature dependence of these frequency is shown in fig. 6.32.
The higher frequency in only marginally visible in the 90 K-sample, but is well
developed in the 60 K-sample. Both frequencies can be consistently accounted for
by considering the antiferromagnetic structure of the Gd-sublattice and assigning the
#+ to the apical oxygen site (high frequency component) and to the chain oxygen
site (low frequency component).
Very interesting is the observation of/z+-relaxation (see fig. 6.33) above TN in the
60 K-sample as well as in a z = 6.2 sample, indicating some slowing down of the
Gd-4f spin dynamics but not in the 90 K sample (Budnick et al. 1990a, Niedermayer
et al. 1993). This implies that the spin dynamics in the 90 K sample is much faster
than in the 60 K-sample. This phenomenon is particularly intriguing because the
N6el temperature seems to be rather independent of the oxygen deficiency. It will
be interesting to study these features in more detail.
A possible interaction between the Cu(II)-sublattice order and the Gd-sublattice
order was investigated by ZF-fSR in a GdBa2Cu306.3 sample (Niedermayer et al.

Gd BCl2 Cu 30T_v
7
6

C"

~3

0 1 2 3
Temperature (K)
Fig. 6.32. Temperature dependence of the two spontaneous precession frequencies in GdBa2Cu3OT_~.
The circles are from a Tc = 60 K sample and the squares from a Tc = 90 K sample. Lines are guides
for the eye (from Golnik et al. 1987).
262 A. SCHENCK and EN. GYGAX

15
I

7- t/\, • G Bo :u3O ,
_ 10

0
CI2
.....2
o0_5

10 100
Temperature (K)
Fig. 6.33. Temperature dependence of the ZF-relaxation rate in three different GdBa2Cu306_ = samples
above TN of the Gd-sublattice (z = I, Te -~ 90 K; z = 0.7, Tc = 60 K; ~c= 0.2, not superconducting)
(from Budnick et al. 1990a).

10
9 .- • GdBa2Cu306+x(tetragonal)
~'7>"~-r" ~568, . . o [] YBa2Cu306+x(tetragonal)

--~ 4 ............ "~r'[]-----~D--_r-~


o 3
u. 2
0 I I i i I l I I I I I i I
o. o ddd ~ ~ oloo
T N (Gd) T N (Cu(II))

Temperature (K)
Fig. 6.34. Temperature dependence of the single spontaneous /z+ precession frequency observed in
GdBa2Cu306.3 . For comparison a few data points from a YBa2Cu306+y-sarnple are also shown (from
Niedermayer et al. 1993).

1993). A single spin precession signal was found below ~ 300 K, the temperature
dependence of the frequency u~, is displayed in fig. 6.34. Analysis of these data
show that below the Gd-sublattice ordering temperature T~ d = 2.3 K the measured
frequency can be understood by simply considering the vector sum of dipolar fields
originating from the ordered Cu(II)-sublattice and the ordered Gd-sublattice, employ-
ing for the magnetic structures the information gained from neutron scattering work
MUON SPIN ROTATIONSPECTROSCOPY 263

(McK. Paul et al. 1988, Mook et al. 1988, Rossat-Mignod et al. 1988a). Hence it
seems that below TN Ga the two magnetic sublattices are not affected by the presence
of the other one. At high temperatures, i.e. above --, 150 K, u,(T) follows closely
the behaviour seen in YBa2Cu306.3 (see fig. 6.34) indicating that also in this tem-
perature range the magnetic structure of the Cu(II)-sublattice is not modified due the
presence of Gd-moments. Quite a different behaviour is observed in the intermediate
range TN Gd < T < 150 K. Clearly the Cu(II)-magnetic sublattice is now drastically
affected by the Gd-4f moments. Whether this is a result of the development of
static correlations among the 4f-spins, induced by and perhaps also modifying the
Cu(II)-magnetic sublattice, or of the slowed down Gd-4f-spin dynamics in oxygen
deficient material, destabilizing to a certain degree also the Cu(II)-moment ordering,
is not known. It should be noted that again TN Gd and TCu(II) are not affected to any
noticeable degree.

DyBa2Cu307_6. Attempts to measure a Dy-4f moment induced #+-Knight shift


above Tc produced a zero result within the achieved precision (Schenck et al. 1990a).
No studies at low temperatures have been performed so far.

HoBa2Cu3Ox. This system has been repeatedly investigated by #SR-spectroscopy.


Generally it was observed in ZF-#SR measurements that the /z+ relaxation rate
increases drastically below 10 K and saturates below ~ 2 K (Nishida et al. 1988a,
Kuno et al. 1988, Birrer et al. 1989a, Grebinnik et al. 1990c). The low temperature
relaxation rate in samples with z ~ 7, if interpreted as dephasing, would correspond
to an internal field spread of (5-13) mT, pointing to a random freezing of the Ho-
4f moments already around 2 K, while specific heat and neutron diffraction data
indicate a transition to long range antiferromagnetic order around (140-190) mK
(Dunlap et al. 1987, Fischer et al. 1988, Roessli et al. 1993). More detailed ZF-
/~SR measurements by Birrer et al. (1989a), extending down to 40 mK, revealed
also some perhaps restricted, long range order starting already below 5 K in a small
fraction Of the polycrystalline sample volume. Their data can be analyzed in terms
of a two component structure of the ZF-#SR-signal, which develops below ,-~ 50 K.
In the range 50 K to 5 K random order appears in a fraction of the sample volume
growing from ~ 0% at 50 K to ~ 30% at 5 K, the remaining volume staying
paramagnetic. Below 5 K the random order transforms to some coherent, but short
range order, as evidenced by the emergence of an oscillating but heavily damped
signal. The temperature dependence of the corresponding frequency u, is displayed
in fig. 6.35. In parallel, the ordered volume fraction grows at the expense of the
paramagnetic fraction and reaches essentially 100% around 2 K. The oscillating
signal shows a pronounced and rather temperature independent Gaussian relaxation,
implying a static field spread of 8.2 roT, which explains why the oscillating behaviour
has been overlooked in the earlier work. A clear oscillation is only directly evident at
the lowest temperatures (see insert in fig. 6.35). As fig. 6.35 shows ut, changes rather
abruptly at ~ 100 inK. This change was attributed to a spin reorientation transition
from fiHo being parallel to the e-axis below 100 mK (also predicted theoretically
by Misra et al. 1992) to being parallel to the e-axis above 100 inK. The ordered
264 A. SCHENCK and EN. G Y G A X

- 0
2.5
2.5 - 0 0
:~ 2.0

~, 1.5
2.0
e-. ~)oo 0
N =~ 1.o
"i-
0
v
o
>~ 1.5 cOoo Oo% u_ 0.5

I I I
!
I~
c- 2 3 4 5
Temperature (K)
o" 1.0 0.1|,e 0.039K
0 • oo c
LL 0
! - 0 . 1 ~
0.5 ~ -0.2
-0"~.0 0.2 0.4 0.6 0.8 1.0 12. 1.4 1.6 1.8 2.0
I I IIIII ~rne i(~)I I I l l t J I I I @11111 I I I I III
0.0
10-1 10 0 101 10 2
Temperature (K)
Fig. 6.35. Temperature dependence of the spontaneous # + precession frequency in HoBa2Cu307_~.
The insert in the upper right corner shows the same on a linear temperature scale. The insert in the
lower left corner shows the ZF-/~SR signal at 39 mK (from Birrer et al. 1989a).

moment was estimated to be #Ho ~- 2"6#B" The most remarkable feature of the #SR-
results is the absence of a clear cut cooperative phase transition which is inconsistent
with both the specific heat and neutron results. In view of this one is tempted
to associate the anomaly in uu at 100 mK with the true phase transition to long
range antiferromagnetic order (implying that the temperature calibration was in error)
and to interpret the results above "100 inK" as an extremely short range precursor
phenomenon. More recent neutron diffraction measurements revealed a clear onset
of long range antiferromagnetic order at (190 4- 10) mK with a propagation vector
q = (0, 1/2, 1/2) and #Ho = 2.8#B aligned parallel to the c-axis (Roessli et al.
1993). The long range order is limited to a correlation length of ~ 30 A along the
c-axis due to the occurence of stacking faults along this direction. It was further
established that the nuclear hyperfine interaction is the driving force in the long
range magnetic order of the Ho3+-moments which evolves out of a singlet level of
the Ho3+-518 ground state multiplet. The magnetic structure proposed by Roessli
et al. (1993) is in marked disagreement with the #SR-results. On the basis of this
structure one calculates a net dipolar field of 0.64 mT at the #+-site while the #SR-
measurements reveal a field of 18.5 roT. It is impossible to explain the difference in
terms of an additional contact hyperfine field in view of the zero Knight shift result
in the paramagnetic phase (see below). The unique and indisputable information
provided by the #SR results is the random freezing of the Ho-moments already far
above the Nrel temperature. With respect to this several interestingquestions can be
asked: what causes frustration; which mechanism is responsible for the coupling of
MUON SPIN ROTATIONSPECTROSCOPY 265

the moments (dipole-dipole, super exchange); which mechanism is responsible for


the formation of a local moment out of the singlet ground state level also at higher
temperatures (nuclear hyperfine, overcritical exchange, population of higher excited
levels); etc? Answers to these questions are not yet available. Also, the oxygen
content seems again to be of importance since the freezing of the Ho-moments is
less visible in a HoBaeCu306.2 sarnple (Kuno et al. 1988). Dilution of Ho by Y has
also a negative influence on the freezing process (Grebinnik et al. 1990c).
TF-#SR measurements on a HoBa2Cu307_~ (To = 90 K)-sample produced no
visible Ho-induced #+-Knight shift (Weber 1991). Earlier claims (Schenck et al.
1990a,c, Maletta et al. 1990b) to have seen a very sizable Knight shift in HoBa2Cu307
are incorrect and resulted from neglecting effects due to powder broadening.

ErBa2Cu30~. ZF-#SR investigations in a series of ErBa2Cu3Ou-compounds with


different z (see table 6.5) yielded in all cases a spontaneous spin precession signal
below the respective Ntel temperatures (Maletta et al. 1990c). Of course, for
z /> 6.6, only the Er-sublattice enters into a (3D) antiferromagnetic phase below
TN = 0.6 K (Maletta et al. 1990a). The corresponding #SR-signal displayed a very
small single frequency v~, the temperature dependence of which is shown in fig.
6.36. Interestingly the precession pattern appeared to persist up to temperatures of at
least 1.2 K. At the transition temperature TN a significant increase of v~ is observed,
which is more pronounced in the z = 6.6 sample. The shift of TN and of v~ (~ 0 K)
with the oxygen content z is well in line with neutron diffraction experiments on
the same samples (Maletta et al. 1990a). In fact using the magnetic structure and
#Zr = 4"4#B (1[b'axis) derived from the neutron work one predicts a field of ~ 0.6 mT
at the/z + position near a chain oxygen in excellent agreement with the experiment.
The persistence of a precession signal above TN is also well in line with neutron
results which reveal short range magnetic correlations up to ,,o 1 K. Unfortunately
the #SR-measurements did not extend beyond ,-~ 1.2 K. (But see also Grebinnik et
al. 1990b, c.) In the whole covered temperature range the damping of the precession
signal is very small (and temperature independent below TN) (compare with the case
in HoBa2Cu3Ox !) but this has not to be taken as evidence for a very perfect long
range magnetic structure. It merely reflects the fact that the net dipolar field at the
#+ site is close to zero (this is a consequence of fiZr being aligned along the b-axis).
The difference in the relaxation behaviour (see fig. 6.36) between the z = 7.0 and
z -- 6.6 sample below 1 K is puzzling and not explained yet.
For samples with x ~< 6.4 the onset of magnetic order of the Cu(II) sublattice
is seen again by the appearance of spontaneous spin precession in the ZF-#SR
signal. No such signal was seen in a compound with z = 6.53 down to 3 K.
The temperature dependence of the single frequency v~,, seen in compounds with
z -- 6.11, 6.34 and 6.40, is displayed in fig. 6.37. In particular the data from the
x -- 6.34 sample indicate that, if it were not for the presence of the Er-moments,
v~ would approach ~ 4 MHz as in YBa2Cu306. A sample with z = 6.20 was
also investigated below 1 K. No evidence for magnetic order of the Er moments
was found. The anomalous increase of u~, below 10 K, particularly in the z =
6.34 sample, is somewhat reminiscent of a similar behaviour in GdBa2Cu306+~
266 A. S C H E N C K and EN. G Y G A X

,--Z 0 0 ~,

0
i"
t",l

+
o

• = o~
o
I~ ~ ~ ~ ~'~

8 ~ = ~ o ~ _ = ~ ~ ~
0 ~ o ~'~ "~ ~

0
~ o . ~ ~

=~ ,~ ~,~,~

0 0
o
I v~ I

0c q
9
o 0
0

I 0 I

O~
II
c~ cS

8
0
~.~

0
0
m~ t'N o~

irJ
"0
oo C
R
r~ 0

0
MUON SPIN ROTATION SPECTROSCOPY 267

v~vvv~

,.,.,

~o

~,~ ~ ~
o ~ o o

<~
on ~ V on
A

oc C
o

..... ~

o r',!,,.q.

o,..o
268 A. SCHENCK and F.N. GYGAX

ErBa2Cu30 x
0.09
-1-
o,.,o
0.06

g~a 0.03
¢._
U_

"7
0.00 I I I I I I I I I I I I

::L

0.30
1:3
I._.

0.20
.$...

0.10
C~
I I I I f f I | [ I ! I
0.00
0.5 1.0
Temperature (K)
Fig. 6.36. Temperature dependence of the spontaneous /z+ precession frequencies and the associated
relaxation rates in two superconducting ErBa2Cu30= samples (z = 6.6, 7.0) (from Maletta et al. 1990c).

(Niedermayer et al. 1993) and reflects clearly some effect of the Er-moments on the
Cu(II)-magnetic ordering. Neutron diffraction measurements on the same samples
showed the onset of magnetic correlations among the Er-moments in the a-b plane
below 1 K which essentially saturate below 230 mK without the formation of a 3D
long range antiferromagnetic structure (Maletta et al. 1990a). Since the Er-moments
do not display a 3D-antiferromagnetic order the increase of uu below ,-~ 10 K cannot
be understood in the same way as in GdBa2Cu306+6. It rather seems that the Cu(II)
magnetic sublattice is somehow modified in an as yet undetermined fashion. It has to
be emphasized that the Cu(II)-long range magnetic order is again not seen in the full
sample volume. Particularly in the z = 6.4 sample quite a fraction of it is associated
with a very rapidly but nonoscillating signal. This fraction decreases with increasing
temperature and disappears at around 300 K. It seems to reflect spin glass behaviour
in part of the sample. Whether this behaviour reflects a nonhomogeneous oxygen
distribution or is induced by the Er-moments remains to be seen. It is not known
whether this sample could become superconducting.
ZF-#SR measurements on a ErBa2Cu306.2 sample by Lichti et al. (1990b) revealed
two frequencies in the spin precession signal appearing below ,-~ 300 K. In addition
MUON SPIN ROTATION SPECTROSCOPY 269

o 6.11
,.~ _ ~.#...... o ~ _ _ • 6.34
-r e"..........-,,, ~ ~ ,', 6.40
" - "

e'O-~ 3 "'......~ ~
~ _ °.
• 2

i
I

0 I I I I I I1[ t I I I I I I I I

101 10 2

T e m p e r a t u r e (K)
Fig. 6.37. Temperature dependence of the spontaneous/z + precession frequencies in ErBa2Cu30= with
x = 6.11 (tetragonal), 6.34 (orthorhombic) and 6.40, The latter sample shows a weak superconducting
response below 32 K (from Maletta et al. 1990c).

to the familiar 0.036 T-component a second frequency corresponding to a field of


0.055 T was seen. The latter may be associated with a site near the CuOz-plane and
could correspond to the 0.13 T signal seen in YBazCu306+~ (5 < 0.3) by Brewer et
al. (1989) (see table 6.4). The same authors performed also TF-#SR measurements.
From an anomaly in the W-relaxation rate, seen at 65 K, they conclude that a second
magnetic phase transition may be present (Lichti et al. 1990b).
A search for a Er induced #+ Knight shift in samples with z = 6.53 and z = 7
(T > To) produced again a zero result (Schenck et al. 1990a). Earlier reports to
the contrary (Schenck et al. 1990c, Maletta et al. 1990b) are again unsubstantiated.
Knight shift results below Tc are discussed by Lichti et al. (1991a).

6.2.6. Other compounds


Long range antiferromagnetic order was detected for the first time by ZF-#SR in the
compounds BaCuO2 and BaYzCuO5 (green phase of YBaCuO). From these data the
transition temperatures could be determined (see table 6.3) (Weidinger et al. 1988).

6.3. Bi-based (2212)-compounds


For an overview of investigated compounds (all have an orthorhombic crystal struc-
ture) see table 6.6. In fact magnetic order in insulating or semiconducting members
of the (2212)-family7 was first demonstrated in #SR experiments almost simultane-
7 Substituting trivalent y3+ for divalent Ca2+ or Sr2+ the carrier (hole) concentration is reduced
(Tamegai et al. 1988). BizSr2YCu208 and BizSrCaYCu208 are insulators. Holes are primarily intro-
duced by some oxygen excess 5 > 0.
270 A. S C H E N C K and EN. GYGAX

oo ~o oo

"G

0
:t r~ ~o
P~ p~ p~
O

r~

+ ~ "~

0 ~ 0

r~

O (-q

:~ d d d d c5 c5
e

t'N
~ A A ~ A A~ ~1

b.., [..-,
N N t"q t'q t'q
.1

© t..)

r..)
)

e.i eq
M U O N SPIN ROTATION SPECTROSCOPY 271

© o
;~ ee3
o ~ ~

rm o
o ~.

~n 0

O ¢xl e q

¢¢3

II II

O
o

A~

r.~ [.r.. O
N N

u
[q
©
o

0
272 A. SCHENCK and EN. GYGAX

ously by Yang et al. (1989), Nishida et al. (1988c) and De Renzi et al. (1989).
ZF-measurements at ~ 5 K in compounds Bi2Sr2YI_~Ca~Cu20~8 with z ~< 0.2
revealed three oscillating components (see table 6.6) in addition to a nonoscillat-
ing one. The temperature dependencies of the three frequencies or local fields in a
Bi2Sr2YCu20~8 sample are displayed in fig. 6.38 together with the relaxation rate
of the low field (0.4 MHz) component (Nishida et al. 1990b). As can be seen, the
0.027-T-component disappears at 80 K. At the same temperature ~r drops abruptly
by a factor of ,.o 3. These features are not understood at present.
In Bi2Sr2YCu20~8 a fourth high frequency component was seen with, however,
an extremely fast relaxation rate (,.o 10 7 s - l ) .
The appearance of several components may not be a surprise in these more com-
plicated compounds and points to the existence of several #+-sites. Possible #+-sites
have not yet been considered in detail but the pair of signals centered at 0.03 T may
indicate sites near an apical oxygen similar as in YBa2Cu306. The high frequency
signal may correspond to the 0.12 T signal in YBa2Cu306+6, i.e. a site close to a
CuO2-plane may be occupied. The low frequency signal may be associated with a

Bi2Sr2YCu20Y
0.4 i i

bo
0.3 - ~mo
:b o (Pa • •
0.2

o.1

Im-,nlliIRm-, mm m mmm
0 l t m[
1.0 j i i

b
0.8 ~m

0.6 F WI i

0.4

¢,... .'¢ ,
0 I I I
0 1 O0 200 300 400
T (K)
Fig. 6.38. Temperature dependence of (a) the 3 spontaneous local fields at the /~+ observed in
BizSr2YCu20~ 8. The 0.027 T local field disappears above ~ 80 K. Below the same temperature
the relaxation rate (b) associated with the lowest local field signal increases abruptly (from Nishida
et al. 1990b).
MUON SPIN ROTATION SPECTROSCOPY 273

site near an oxygen in the Bi-O plane, which has no counterpart in the (123)-system.
In any case the close correspondence of some of the frequencies with observed fre-
quencies in the (123)- and the LaCuO-systems indicates that the Cu-moment order
must be quite similar in all the cuprate layered compounds. In particular the magni-
tudes of the Cu-moments must be quite comparable. The ZF-#SR data obtained in
Bi2Sr2Y0.6Cao.4Cu208, Bi2Sr2.5Y0.5Cu20~8 and Bi2SrCaYCu2Os+6 also show the
presence of magnetic order but it is now of a random (spin glass like) structure.
This phase is followed, when further decreasing the Y-content, by the appearance of
superconductivity (see fig. 6.39).
Like in the (123)- and the LaCuO-families the development of magnetic order
starts first in only a tiny fraction of the sample volume and even at very low temper-
atures may not involve all of the sample volume. Figure 6.40 presents an example
for how the paramagnetic fraction of the sample volume decreases with decreas-
ing temperature in compounds BizSr2Yt_xCa~Cu208 with z = 0, 0.1, 0.2 and 0.3.
These data are obtained from TF-#SR-measurements which allow to identify the

400 I

Bi2 Sr2Y1 -x CaxCu2Oy

3O0
+
2OO
b-

100
FIND
[]
AF SG SC
D

0 i I I I I I I I I
0.0 0.5 1.0
X
Fig. 6.39. Magnetic and superconducting phase diagram of Bi2Sr2_xY~Cu208. The magnetic part is
deduced from/~SR-data showing AF-order for x<0.3 and spin glass like order for 0.5 > z > 0.3. The
arrow for z = 0 indicates that TN is above the instrumental T limit (from Nishida et al. 1990b).
274 A. SCHENCK and EN. GYGAX

Bi2 Sr2Yl_xCax Cu20,


10 i i

gO.8 ~'x:o.3 x:o-2 -

."-~
~-0.6
_u ~ ,~ I f
~X:0.1
~ 0.4 • o -

a_ 0.2

0(3 t I
0 100 200 300 400
T (K)
Fig. 6.40. Fraction of paramagnetic volume in various Bi2Sr2Yl_~CaxCu20~t samples as a function of
temperature. The fraction is determined from TF-/~SR measurements (from Nishida et al. 1990b).

fraction of #+ in a paramagnetic environment. It can be seen that, e.g., in a sample


with z = 0.2 magnetic order sets in at around 300 K but at ~ 250 K more than half
of the volume is still paramagnetic and this fraction does not drop below 30% at
the lowest temperatures. Even for a more or less stoichiometric sample with z -- 0
magnetic order develops in a spatially inhomogeneous fashion. The latter result,
assuming that the oxygen content is close to 8 per unit formula, could suggest that
the inhomogeneous magnetic ordering is an intrinsic property of these layered sys-
tems. #SR-spectroscopy should be in a position to explore this property in a more
systematic way.
Interpreting the onset temperatures of magnetic order as N6el or freezing temper-
atures a phase diagram can be drawn as shown in fig. 6.39 (Nishida et al. 1990b).
By comparison with fig. 6.22 it is seen that all layered cuprate families behave in
the same way. In the insulating or semiconducting phase the CuO2-1ayers show
antiferromagnetic order. With the introduction of holes into the CuOz-planes mag-
netic order becomes frustrated and eventually a spin glass phase is entered. On
further increasing the hole concentration magnetism is lost altogether and a super-
conducting phase evolves. Whether a small region exists where spin glass order and
superconductivity coexist in a truly microscopic manner is still a disputed matter.

7. Study of magnetic order in organic compounds

All compounds studied are listed in table 7.1.

7.1. (TMTSF)2X
The family of the tetramethyl-tetraselena-fulvalene (TMTSF)2-X compounds displays
many intriguing properties such as superconductivity, anion ordering, quantum Hall
effect, spin density wave (SDW) ordering and field induced spin density waves
M U O N SPIN ROTATION SPECTROSCOPY 275

eee ~ e

+
d

¢q ¢q ¢q ,,~

~ a a

c~
c~
~c;~ z

r..)
276 A. SCHENCK and EN. GYGAX

owing to their highly anisotropic electronic structure (low dimensionality) (J6rome


and Schulz 1982, Ishiguro and Yamaji 1990). The spin density wave properties have
been studied by a variety of techniques but many detailed features, e.g., elementary
excitations, sublattice magnetization, SDW wave vector and the order of the transition
are not too well determined. The usefulness of #SR-spectroscopy in addressing some
of these problems rests again in the possibility to perform the experiments in zero
applied field.

(TMTSF)zPF6. Antiferromagnetic SDW order is observed in this compound at am-


bient pressure below TN = 122 K. The anion PF6 is located in a centrosymmetric
position. ZF and LF-#SR measurements were aimed at a determination of the exact
SDW structure (commensurate versus incommensurate) and the order of the phase
transition (Le et al. 1991, Le et al. 1994). Previous NMR measurements (Takahashi

0.14 ~N~"
~ (TMTSF)2PF6 (a)
0.14 ~ H =0
b 0.14: ~ ~ . ~,xt

0.12 % "~ " ~ ' ~ I 12.25K

<b
o.o8 ' ,? ?' '
%
E
oo,
0

0.00 I I I
0 2 4 6
Time (/zs)

(b)

13_
2.0
E
.<
1.0
n--

0.0 I I /~
0.0 0.2 0.4 0.6 0.8
Frequency (MHz)
Fig. 7.1. (a) ZF-/~SR signals in (TMTSF)2PF 6 for various temperatures (b) Fourier transform of the
signal taken at 3.7 K (from Le et al. 1994).
MUON SPIN ROTATION SPECTROSCOPY 277

et al. 1986, Delrieu et al. 1986a, b) had detected the SDW-state, but the applied
field was rather high and the system was in the spin-flop state. Figure 7.1a displays
some typical ZF-#SR spectra around the transition temperature. As can be seen
below 12.2 K a precessional pattern develops but the actual frequency spectrum, as
revealed by Fourier transforming the signal, is quite broad, showing a shoulder on
the low frequency side and a peak at a frequency of ~ 0.55 MHz, corresponding
to an internal field of 0.004 T (see fig. 7.1b). This distribution appears incompat-
ible with a commensurate structure (provided only one type of site is occupied by
the #+) but can be well reproduced by a distribution of dipolar fields arising from
an incommensurate sinusoidally modulated intermolecular spin density wave with
wave vector ~) = (0.5, 0.24, -0.06) and maximum amplitude of 0.1#B/molecule
(Takahashi et al. 1986). The shape of the frequency spectrum does not depend very
much on the assumed #+ site and also not on the exact value of Q as long as the
modulation is incommensurate.
In this picture the peak frequency uu,p (see fig. 7.1b) provides a good measure
of the SDW amplitude M or the order parameter. The temperature dependence
of Uu,p is displayed in fig. 7.2 together with the corresponding results on the two
other (TMTSF)2-X compounds with X = NO3 and X = C104. The data show little
temperature dependence below TN/2 and quite an abrupt change at TN. The latter
is taken as evidence for a first order phase transition. It may be ascribed to a cross
over from a 1D to a 3D behaviour whereby the already existing, albeit fluctuating
order parameter in the 1D-regime is frozen out at a finite value at the transition
temperature.
The temperature dependence of Uu,p below ~ TN/2 is, although weak, much
stronger than one would expect from a mean field treatment of the SDW transition

0.7

(TMTSF)2-X
0.6 Hext=O

0.5
"T-

>,
0.4 ', ~"4 p;6"~
NO 3
0
c 0.3 Ci04 4 "

E)-
a) 0.2
k_
t,

0.1
t
o.o , I
0 5 I0 15
Temperature (K)
Fig. 7.2. Temperature dependencies of the spontaneous iz+ precession frequencies in (TMTSR)2-X with
X = C104, NO3, PF6. The solid lines are guides to the eye (from Le et al. 1994).
278 A. SCHENCKand EN. GYGAX

1.02 I i i i

1.00
0o0
0.98
0
0.96

I.-- 0.94

0.92

0.90 Heisenberg (NN)

0.88 1 I I I

0 2 4 6 8 10
Temperature (K)
Fig. 7.3. Temperature dependence of uj,(T)/u~(O) = M(T)/M(O) in (TMTSF)2PF6 below TN/2 and
various model predictions. The data are very well described by a 2D-spin wave excitationmodel (from
Le et al. 1994).

(Yamaji 1982). The opening up of an energy gap A should largely suppress sin-
gle particle excitations and consequently the magnetization or Uu,p should be largely
temperature independent below ,,~ TN/2. It is therefore argued (Le et al. 1994) that
the variation of U~,p or M at low temperature is determined by collective spin wave
excitations. Indeed the temperature dependence of tJ~,p(T)/u#,p(0) = M(T)/M(O)
can be very well accounted for by such an approach (solid line in fig. 7.3) leading to
an average spin wave stiffness constant of 13 = 200 K. This value is much larger (by
a factor 3-5) than what is expected from a local moment nearest neighbour Heisen-
berg model. However, the large value can be understood, at least qualitatively, on
the basis of an itinerant electron model (Le et al. 1994).

(TMTSF)2N03, (TMTSF)2CI04. In these compounds the anions NO3 or C104 are


located at non-centrosymmetric positions with the consequence that an anion ordering
transition is found at temperature TAo (= 41 K in the NO3-compound). The ground
state of these systems depends on the degree of anion order-disorder produced by
the cooling rate through TAO. In the NO3-compound a SDW state below 12 K can
be obtained by slow cooling, while in the C104-compound a SDW state is achieved
below ~ 4 K by very rapid cooling (quenching).
ZF-#SR measurements revealed spectra very similar to the ones found in the PF6-
compound indicating an incommensurate spin density wave state in each case (Le et
al. 1994).
The temperature dependence of the peak frequency Uu,p for both compounds is
included in fig. 7.2. As can be seen the low temperature u~,v (or the spin density
wave amplitude M) is essentially the same in all three compounds despite the fact
MUON SPIN ROTATION SPECTROSCOPY 279

that rather different transition temperatures are observed. This behaviour is taken
as support for a nesting model for the SDW condensation (Takahashi et al. 1989).
According to this model the SDW gap ,4 and the order parameter M are mainly
determined by the on site Coulomb energy U and the on-chain coupling t, which are
independent of the anion X. The decrease of the transition temperature, on the other
hand is a result of the loss of perfect nesting as the dimensionality of the systems
increases (i.e. becoming more 2D), e.g., as a function of pressure or type of anion,
but this has no effect on M as long as the SDW ground state is established.

7.2. Ni2(C2H8N2)2N02(CI04) (NENP)


NENP is a nearly ideal realization of a spin-1 linear-chain Heisenberg antiferromag-
net. For such a system Haldane (1983) has predicted the existence of a quantum
energy gap between the nonmagnetic singlet ground state and the lowest triplet ex-
cited state. No such gap occurs for half integer spin systems. Admitting also some
inter chain coupling above a certain critical value long range magnetic order may
be expected (Affleck 1989). However, the inter chain coupling in the NENP system
appears too weak to allow for the development of long range magnetic order. In
fact magnetization measurements place an upper limit of 1.2 K on the transition
temperature to an ordered magnetic phase (Renard et al. 1987). Nevertheless the
low temperature properties of NENP remain an important issue.
ZF- and LF-#SR measurements on a polycrystalline specimen reveal indeed in-
teresting new information which indicates that NENP could be very close to static
magnetic order (Sternlieb et al. 1992). Applying a small (0.01 T) longitudinal field
in order to decouple the #+ spin from nuclear dipolar fields the LF-relaxation follows
well a stretched exponential (exp(-iv/~-/at)) which is indicative for a broad fluctuating
local field distribution. Figure 7.4 displays the temperature dependence of ),1. As
can be seen ),1 rises steeply between 3 K and 2 K followed by a saturation down
to 20 mK. The fast rise of ),1 below 3 K points to a slowing down of the spin
dynamics. However, no static limit is reached. This follows from a persisting field

1.4 I I I I I I

o
,~ 1,2 p o
0 NENP LF=IOOG _
~7 1.o
o.8

~- 0.6 0
O

0.4
O
rr 0.2 O
O
0.0 I ~ t I I
0 2 3 4 5 6 7
Temperature (K)
Fig. 7.4. Temperature dependence of A1 (measured in an external field of 0.01 T) in NENR Note the
rapid rise below 3 K (from Sternlieb et al. 1992).
280 A. SCHENCK and EN. GYGAX

dependence, even at 20 mK, which is roughly described by ALF O( 1/co~ and implies
very small fluctuation rates of the order of 40 MHz at 20 mK. The local fields are
of the order of Bloc ~ 90 G. The latter value is too large to be caused by loose
chain ends or impurities but, on the other hand, too small to be representative of all
Ni-moments. Therefore, if A1 and Bloc are reflecting a truly intrinsic behaviour, it
must be concluded that only a limited portion of the Ni-spin fluctuation-spectrum is
affected by a slowing down of the spin dynamics (Sternlieb et al. 1992).

7.3. p-NPNN
p-NPNN (p-nitrophenyl nitronyl nitroxide, C13H16N304) is a chemical radical which
has been identified to show ferromagnetic order (Awaga et al. 1989, Takahashi et al.
1991). Although various techniques have been used in the study of this system the
most fundamental parameter, namely the spontaneous magnetization has remained
uninvestigated, but has later been measured in a ZF-#SR experiment on t-phase
monocrystalline material (Le et al. 1993). The onset of magnetic order below
0.7 K is clearly reflected in the appearance of a long lived spontaneous coherent
spin precession signal with only one frequency uu implying a unique #+ site in
this complex system. The temperature dependence of uu, which is proportional to
the spontaneous magnetization M is displayed in fig. 7.5. The Curie temperature
following from this plot is 0.67 K and indicates by comparison with other samples
a slight sample dependence. The magnetization has been previously measured at
0.44 K with the result M = 0.5#B/molecule (Tamura et al. 1991). Rescaling
this value with the present u(T) one gets M(0 K)= 0.6#B/molecule. Expected is
M = l#B/molecule. The difference is ascribed to systenaatic uncertainties in the
magnetization measurements. The #SR data indicate that the internal spontaneous
field is nearly parallel to the crystalline b-axis.

2.5 I I I I I I I

H,~=0
2.0-
"1-

> ", 4.5

a~ 1.0

u_ 0.5

0.0 I i w I J I T i
0 100 200 300 400 500 600 700 800
Temperature (mK)
Fig. 7.5. Temperature dependence of the spontaneous /~+ precession frequency in p-NPNN. The solid
line represents a fit of eq. (7.1) to the data (from Le et al. 1993).
MUON SPIN ROTATION SPECTROSCOPY 281

The temperature dependence of uu c( M is very well fitted (solid line in fig. 7.5)
by the expression

M ( T ) oc (1 - (T/Tc)'~) ~ (7.1)
with c~ = 1.86 and/3 = 0.32. The low temperature behaviour can be expressed as
(M(0) - M ( T ) ) c< T '~ which is close to the magnon induced temperature dependence
in a 3D system with ~ = 1.5. Near Tc one has M ( T ) ~x (Te - T ) ~ which, with
/3 = 0.32, is in excellent agreement with the value /3 = 1/3 expected for a 3D
Heisenberg ferromagnet. The behaviour of/3-phase p-NPNN is thus consistent with
that of a 3D Heisenberg system. This is in contrast to susceptibility and magnetization
measurements in the ?-phase of p-NPNN which could be well explained in terms of
a quasi 1D-ferromagnet (Takahashi et al. 1991).

Acknowledgements

We are indebted to the many colleagues who have sent us re- and preprints, who
responded quickly to questions and even provided us with high quality figures. Many
thanks go to Dr. Alex Amato who read carefully the manuscript and was patient
enough to provide advice and help whenever needed. Finally we have to thank
Mrs. M. Sekolec and Mrs. R. Bachli for preparing skilfully the typed manuscript and
Mrs. I. Kusar of PSI for drawing or redrawing a major fraction of the figures.

List of some of the used abbreviations and symbols

ABBREVIATIONS
(#)LCR (muon) level crossing resonance
#SR muon spin rotation, relaxation and resonance
AF antiferromagnetic
anneal annealing
bcc body centered cubic
CAF commensurate spin density wave
cor corundum
fcc face centered cubic
FI ferrimagnetic
FM ferromagnetic
hcp hexagonal close-packed
hex hexagonal
ins insulator
LCR level crossing resonance
LF-pSR longitudinal field #SR
LIAF longitudinal incommensurate spin density wave
met metal
mon monoclinic
282 A. SCHENCK and EN. GYGAX

Mu muonium (#+e-)
n-irrad neutron irradiated
O, oct octahedral
OPC oriented polycrystals
orth orthorhombic
P paramagnetic
PC polycrystalline, polycrystal
per perovskite
p-irrad proton irradiated
rut rutile
SC, s.c. single crystal
scu simple cubic
SDW spin density wave
sem semimetal
SF spin fluctuations
SG spin glass
T tetrahedral
T± tetrahedral with tetragonal axis perpendicular to [100] direction
TF-#SR transverse field #SR
TIAF transverse incommensurate spin density wave state
USC unoriented single crystals
WA weak antiferromagnetic
WF weak ferromagnetic
ZF-#SR zero field #SR
SYMBOLS
F,A muon depolarization or dephasing rate, general
/~ #+ depolarization - several components
% muon gyromagnetic ratio (see table 2.1)
z~ 2 second moment of one cartesian field component in ZF
AB u rms width of local field distribution
strain
0 orientation or angle between field and c-axis
~1 = 1/TI spin lattice relaxation rate
spin lattice relaxation rate in ZF
/~2 = 1/T2 transverse relaxation rate
AZF ZF-relaxation rate from exponential decay
magnetic moment
/./ fluctuation rate
uu - u x level crossing resonance frequency muon - nucleon X
u u (= w u / 2 7 r ) muon Larmor or precession frequency
o Gaussian depolarization rate (cr2 = M2)
~-c (= 1/.) correlation time
orientation or angle between field projection on the basal a-b
plane and the a-axis
MUON SPIN ROTATIONSPECTROSCOPY 283

X susceptibility tensor (usually in emu/mol)


Xat atomic susceptibility tensor (emu/atom)
A amplitude or asymmetry of #SR signal
Ac contact hyperfine coupling constant
Adip dipole coupling tensor
Ai amplitude or asymmetry of i-th component of #SR signal
AB u rms width of local field distribution
B~ total local magnetic field at the muon
field at the muon - several components or sites
Bc (contact) hyperfine field at the muon
Bdem demagnetization field
Bdip dipolar field
Bext applied magnetic flux density
BL Lorentz field
BMu field at the #+ in the Mu-state
a~T Kubo-Toyabe relaxation function for Gaussian field
distribution
% Kubo-Toyabe relaxation function for Lorentzian field
distribution
GTF depolarization function in transverse field
Hext applied magnetic field
K. muon Knight shift
M2 second moment of field distribution (usually in MHz2), related
to AB~
4-4
N demagnetization tensor (factor)
p, press pressure
P(t) time dependence of muon spin polarization, projected on
P(t = 0)
PLF(t) P(t) measured in longitudinal field (/~ext]lfi(0))
PTF(t) P(t) measured in transverse field (/lext±/3(0))
PZF(t) P(t) measured in zero field
T temperature
T1 longitudinal spin-lattice relaxation time
T2 transverse (spin-spin) relaxation time
Tc Curie temperature
TFN ferrimagnetic ordering temperature
TM Morin temperature
TN N6el temperature
Tv Verwey temperature
G superconducting transition temperature
Tf freezing temperature in a spin glass-like system
284 A. SCHENCK and EN. GYGAX

References

Abragam, A., 1970, The Principles of Nuclear Zirnmermann, 1993a, Physica B 186-188,
Magnetism (Clarendon Press, Oxford). 276.
Aeppli, G., E. Bucher, C. Broholm, J.K. Kjems, Amato, A., RC. Canfield, R. Feyerherm, Z. Fisk
J. Baumann and J. Hufnagl, 1988, Phys. Rev. Z, EN. Gygax, R.H. Heffner, E.A. Knetsch,
Lett. 60, 615. D.E. MacLaughlin, H.R. Ott, A. Schenck,
Aeppli, G., D. Bishop, C. Broholm, E. Bucher, J.D. Thompson and U. Zimmermann, 1993b,
K. Siemensmeyer, M. Steiner and N. Stiissi, Physica B 186-188, 615.
1989, Phys. Rev. Lett. 63, 676. Amato, A., R. Feyerherm., EN. Gygax, D.
Affleck, I., 1989, Phys. Rev. Lett. 62, 474. Jaccard, A. Schenck, J. Sierro, E. Walker and
Aggarwal, K., L. Asch, F.N. Gygax, O. U. Zimmermann, 1993c, Physica B 186-188,
Hartmann, G.M. Kalvius, A. Kratzer, EJ. 273.
Litterst, K. Mattenberger, A. Schenck and O. Amato, A., R. Feyerherm, F.N. Gygax, A.
Vogt, 1989, Hyperfine Interact. 51, 935. Schenck and D. Jaccard, 1994a, Hyperfine
Aggarwal, K., L. Asch, S. Fredo, J. Gal, EN. interact. 85, 369.
Gygax, B. Hitti, G.M. Kalvius, A. Kratzer, Amato, A., R. Feyerherm, EN. Gygax, A.
EJ. Litterst, K.H. Mtinch and A. Schenck, Schenck, J. Flouquet and P. Lejay, 1994b,
1990, Hyperfine Interact. 64, 401. Phys. Rev. B 50, 619.
Aharony, A., R.J. Birgenean, A. Coniglio, M.A. Ansaldo, E.J., D.R. Noakes, J.K. Brewer, R.
Kastner and H.E. Stanley, 1988, Phys. Rev. Keitel, D.R. Harschman, M. Senba, C.Y.
Lett. 60, 1330. Huang and B.V.B. Sarkissian, 1985, Solid
Akimitsu, J., J. Amano, M. Yoshinari, M. State Commun. 55, 193.
Kokubum, T. Iwasaki, S. Okuma, N. Nishida, Ansaldo, E.J., J.H. Brewer, T.M. Riseman, J.E.
K. Nishiyama and N. Mori, 1994, Hyperfine Schirber, E.L. Venturini, B. Morisin, D.S.
Interact. 85, 187. Ginley and B. Sternlieb, 1989, Phys. Rev. B
Albanese, L., C. Bucci, P. Laretta, R. De Renzi, 40, 2555.
G. Guidi, R. Teteschi, E Licci., C. Paris, G. Ansaldo, E.J., Ch. Niedermayer, H. Gltickler,
Calestani, M.G. Francesconi, S.EJ. Cox and C.E. Stronach, T.M. Riseman, R.S. Cary,
C.A. Scott, 1992, J. Magn. Magn. Mater. D.R. Noakes, X. Obradors, A. Fuertes, J.M.
104-107, 509. Navarro, P. Gomez, N. Casan, B. Martinez,
Alp, E.E., L. Soderholm, G.K. Shenoy, D.G. F. Perez, J. Rodriguez-Carvajal and K. Chow,
Hinks, D.E. Capone II, K, Zhang and B.D. 1991, Physica 185-189, 1213.
Dunlap, 1987, Phys. Rev. B 36, 8910. Ansaldo, E.J., Ch. Niedermayer, H. Gltickler,
Amato, A., 1994, Physica B 199+200, 91. C.E. Stronach, T.M. Riseman, D.R. Noakes,
Amato, A., W. Btihrer, A. Grayevsky, EN. X. Obradors, A. Fuertes, J.M. Navarro, P.
Gygax, A. Furrer, N. Kaplan and A. Schenck, Gomez, N. Casan, B. Martinez, E Perez, J.
1992a, Solid State Commun. 82, 767. Rodriguez-Carvajal, R.S. Cary and K. Chow,
Amato, A., P.C. Canfield, R. Feyerherm, Z. Fisk, 1992, Phys. Rev. B 46, 3084.
EN. Gygax, R.H. Heffner, D.E. MacLaughlin, Asch, L., 1989, Physica B 161, 299.
H.R. Ott, A. Schenck and J.D. Thompson, Asch, L., 1990, Hyperfine Interact. 64, 351.
1992b, Phys. Rev. B 46, 3151 Asch, L., B. Boucher, J. Chappert, O. Hartmann,
Amato, A., C. Geibel, F.N. Gygax, R.H. Heffner, G.M. Kalvius, E. Karlsson, J. Litterst, K.
E. Knetsch, D.E. MacLaughlin, C. Schank, A. Nagamine, K. Nishiyama, R. Wappling, T.
Schenck, F. Steglich and M. Weber, 1992c, Yamazaki and A. Yaouanc, 1984, Physica B
Z. Physik B 86, 159. 130, 469.
Amato, A., R. Feyerherm, EN. Gygax, Asch, L., G.M. Kalvius, A. Kratzer, EJ. Litterst,
A. Schenck, M. Weber, R. Caspary, P. U. Potzel, R. Keitel, D. Noakes, E. Ansaldo,
Hellmann, C. Schank, C. Geibel, E Steglich, O. Hartmann, E. Karlsson, R. W~ippling, J.
D.E. MacLaughlin, E.A. Knetsch and R.H. Chappert and Y. Yaouanc, 1986, Hyperfine
Heffner, 1992d, Europhys. Lett. 19, 127. Interact. 31, 337.
Amato, A., C. Baines, R. Feyerherm, J. Flouquet, Asch, L., S. Barth, F.N. Gygax, G.M. Kalvius,
EN. Gygax, P. Lejay, A. Schenck and U. A. Kratzer, EJ. Litterst, K. Mattenberger, W.
MUON SPIN ROTATION SPECTROSCOPY 285

Potzel, A. Schenck, J.C. Spirlet and O. Vogt, Nikol'skii, A.V. Pirogov, A.N. Ponomarev,
1987, J. Magn. Magn. Mater. 63+64, 169. V.I. Selivanov, G.V. Shcherbakov, S.N.
Asch, L., EJ. Litterst, A. Kratzer, K. Aggarwal, Shilov, V.A. Suetin and V.A. Zhukov, 1986,
W. Potzel, G.M. Kalvius, EN. Gygax, B. Hyperfine Interact. 31, 339.
Hitti, A. Schenck, S. Barth, O. Vogt and Barsov, S.G., A.L. Getalov, V.P. Koptev,
K. Mattenberger, 1988, J. Phys. (Paris) 49, L.A. Kuzmin, S.M. Mikirtychyants, G.V.
C8-495. Shcherbakov, A.H. Vasilijev, V.K. Fedotov,
Asch, L., G.M. Kalvius, A. Kratzer, EJ. Litterst, V.I. Kulakov, R.K. Nicolaev, N.S. Sidorov,
A. Schenck, B. Hitti, EN. Gygax, C.A. Y.M. Mukovskii, A.S. Nigmatulin and S.E.
Scott, K. Mattenberger and O. Vogt, 1989, Strunin, 1990, Hyperfine Interact. 63, 161.
Europhys. Lett. 10, 673. Barsov, S.G., A.L. Getalov, V.P. Koptev,
Asch, L., G.M. Kalvius, EJ. Litterst, M. Weber, L.A. Kuzmin, S.M. Mikirtychyants and G.V.
K.H. MUnch, A. Kratzer, K. Aggarwal, A. Shcherbakov, 1994, Hyperfine Interact. 86,
Schenck, EN. Gygax, R. Ballou, J. Deportes 543.
and J. Chappert, 1990, Hyperfine Interact. 62, Barth, S., 1988, Dissertation ETH Nr. 8563
381. (ETH, ZUrich), unpublished.
Asch, L., G.M. Kalvius, A. Kratzer and EJ. Barth, S., E. Albert, G. Heiduk, A. Mt~slang,
Litterst, 1994, Hyperfine Interact. 85, 193. A. Weidinger, E. Recknagel and K.H.J.
Awaga, K. and Y. Maruyama., 1989, Chem. Buschow, 1986a, Phys. Rev. B 33, 430.
Phys. Lett. 158, 556; J. Chem. Phys. 91, Barth, S., H.R. Ott, F.N. Gygax, A. Schenck,
2743. T.M. Rice and Z. Fisk, 1986b, Hyperfine
Axe, J.D., A.H. Moudden, D. Hohlwein, D.E. Interact. 31, 397.
Cox, K.M. Mohanty, A.R. Moodenbaugh and Barth, S., H.R. Ott, E Hulliger, EN. Gygax,
Youwen Xu, 1989, Phys. Rev. Lett. 62, A. Schenck and T.M. Rice, 1986c, Hyperfine
2751. Interact. 31, 403.
Baggio-Saitovitch, E., EJ. Litterst, K. Nagamine, Barth, S., H.R. Ott, F.N. Gygax, B. Hitti, E.
K. Nishiyama and E. Torikai, 1990, Hyperfine
Lippelt, A. Schenck, G. Baines, B. van den
Interact. 63, 259 Brandt, T. Konter and S. Mango, 1987, Phys.
Baillif, R., A. Dunand, J. Mtiller and K. Yvon,
Rev. Lett. 59, 2991.
1981, Phys. Rev. Lett. 47, 672.
Barth, S., H.R. Ott, F.N. Gygax, B. Hitti, E.
Bargouth, M.O. and G. Will, 1971, J. Phys. C
Lippelt and A. Schenck, 1988, J. Magn.
1, 675.
Magn. Mater. 76+77, 455.
Barsov, S.G., A.L. Getalov, G. Grebinnik,
Barth, S., H.R. Ott, EN. Gygax, B. Hitti, E.
I.I. Gordeev, I.I. Gurevich, V.A. Zhukov,
Lippelt, A. Schenck and C. Baines, 1989a,
A.I. Klimov, S.R Kruglov, L.A. Kuz'min,
A.B. Lazarev, S.M. Mikirtychyants, N.I. Phys. Rev. B 39, 11695
Moreva, B.A. Morozova, B.A. Nikol'skii, Barth, S., H.R. Ott, F.N. Gygax, B. Hitti, E.
A.V. Pirogov, V.I. Selivanov, V.A. Suetin, Lippelt, A. Schenck and Z. Fisk, 1989b,
S.V. Fomichev and G.V. Shcherbakov, 1983, Hyperfine Interact. 50, 711.
Zh. Eksp. Teor. Fiz. 84, 1896; [1983, Sov. Bauer, E., 1991, Adv. Phys. 40, 417.
Phys. - JETP 57, 1105]. Bertaut, E.E, 1963, in: Magnetism, Vol. III,
Barsov, S.G., S.V. Fomichev, A.L. Getalov, G. eds G.T. Rado and H. Suhl (Academic Press,
Grebinnik, I.I. Gordeev, I.I. Gurevich, A.I. New York).
Klimov, S.R Kruglov, L.A. Kuzmin, A.B. Bhandari, R. and L.M. Falicov, 1973, J. Phys. C
Lazarev, S.M. Mikirtychyants, N.I. Moreva, 6, 479.
B.A. Morozova, B.A. Nikol'skii, A.V. Bitter, R, 1991, Dissertation ETH No. 9636
Pirogov, A.N. Ponomarev, V.I. Selivanov, (ETH, Ztirich), unpublished.
G.V. Shcherbakov, V.A. Suetin and V.A. Birrer, P., EN. Gygax, B. Hitti, E. Lippelt, A.
Zhukov, 1984, Hyperfine Interact. 17-19, Schenck, M. Weber, S. Barth, E Hulliger and
485. H.R. Ott, 1989a, Phys. Rev. B 39, 11449.
Barsov, S.G., A.L. Getalov, G. Grebinnik, Birrer, P., F.N. Gygax, B. Hitti, E. Lippelt, A.
I.I. Gurevich, B.F. Kirillov, A.I. Klimov, Schenck, D. Cattani, J. Cors, M. Decroux,
S.R Kruglov, L.A. Kuzmin, A.B. Lazarev, ¢1. Fischer and S. Barth, 1989b, Hyperfine
S.M. Mikirtychyants, N.I. Moreva, B.A. Interact. 50, 503.
286 A. SCHENCK and EN. GYGAX

Birrer, R, EN. Gygax, B. Hitti, E. Lippelt, A. Brewer, J.H., J. Carolan, R.F. Kiefl, S.R.
Schenck, M. Weber, D. Cattani, J. Cors, M. Kreitzman, G.M. Luke, T.M. Riseman, P.
Decroux and ~. Fischer, 1990, Hyperfine Schleger, B.J. Sternlieb, Y.J. Uemura, D.L1.
Interact. 61, 1125. Williams and B.X. Yang, 1989, Physica C
Blount, E.I., C.M. Varma and G. Aeppli, 1990, 162-164, 157.
Phys. Rev. Lett. 64, 3074. Brewer, J.H., R.E Kiefl, J.E Carolan, P.
Boardman, C.J., R. Cywinski, S.H. Kilcoyne and Dosanjh, W.N. Hardy, S.R. Kreitzmann, Q.
Li, T.M. Riseman, P. Schleger, H. Zhou, E.J.
C.A. Scott, 1994, Hyperfine Interact. 86, 525.
Ansaldo, D.R. Noakes, L.P. Le, G.M. Luke,
Boekema, C., 1980, Philos. Mag. B 42, 409. Y.J. Uemura, K. Hepburn-Wiley and C.E.
Boekema, C., 1984, Hyperfine Interact. 17--19, Stronach, 1990, Hyperfine Interact. 63, 177.
305. Broholm, C., 1989, Dissertation, Ris~-M-2731
Boekema, C., 1988, in: The Time Domain in (Ris~ Nat. Laboratory).
Surface and Structural Dynamics, eds G.J. Broholm, C., J.K. Kjems, W.J.L. Buyers, P.
Long and F. Grandjean (Kluwer Academic Matthews, T.T.M. Palstra, A.A. Menovsky
Publ., Dordrecht). and J.A. Mydosh, 1987, Phys. Rev. Lett. 58,
Boekema, C., R.H. Heffner, R.L. Hutson, M. 1467.
Leon, M.E. Schillaci, J.L. Smith, S.A. Dodds Burgouth, M.O. and G. Will, 1971, J. Phys. C
and D.E. MacLaughlin, 1982, J. Appl. Phys. 1, 675.
53, 2625. Burlet, P., S. Quezel, J. Rossat-Mignod and O.
Boekema, C., A.B. Denison and K. Rtiegg, 1983, Vogt, 1984, Symposium on Neutronscattering,
J. Magn. Magn. Mater. 36, 111. (Hahn-Meitner Institute, HMI-B411).
Boekema, C., R.L. Lichti, V.A.M. Brabers, A.B. Budnick, J.I., A. Golnik, Ch. Niedermayer, E.
Denison, D.W. Cooke, R.H. Heffner, R.L. Recknagel, M. Rossmanith, A. Weidinger, B.
Hutson, M. Leon and M.E. Schillaci, 1985, Chamberland, M. Filipkowski and D.P. Yang,
1987, Phys. Lett. A 124, 103.
Phys. Rev. B 31, 1233.
Budnick, J.I., B. Chamberland, D.P. Yang,
Boekema, C., R.L. Lichti, A.B. Denison, V.A.M.
Ch. Niedermayer, A. Golnik, E. Recknagel,
Brabers, D.W. Cooke, R.H. Heffner, R.L. M. Rossmanith and A. Weidinger, 1988,
Hutson and M.E. Schillaci, 1986, Hyperfine Europhys. Lett. 5, 651.
Interact. 31, 487. Budnick, J.I., M.E. Filipowski, Z. Tan,
Bonville, R, J.A. Hodges, R Imbert, G. le Bras, B. Chamberland, Ch. Niedermayer, A.
R Pari, R Dalmas de R6otier, A. Yaouanc and Weidinger, A. Golnik, R. Simon, M.
RC.M. Gubbens, 1994, Preprint. Rauer, H. Gliickler, E. Recknagel and C.
Brewer, J.H., 1982, in: Fourier, Hadamard Baines, 1990a, in: Int. Seminar on
and Hilbert Transformations in Chemistry, ed. High Temperature Superconductivity, eds
A.G. Marshall (Plenum, New York). V.L. Aksenov, N.N. Bogolubov and N.M.
Brewer, J.H., E. Koster, A. Schenck, H. Schilling Plakida (World Scientific, Singapore) p. 172.
and D.L1. Williams, 1981, Hyperfine Interact. Budnick, J.I., M. Filipkowski, Z. Tan, Ch.
8, 619. Niedermayer, H. G1Uckler, A. Weidinger and
Brewer, J.H., S.R. Kreitzman, D.R. Noakes, E.J. E. Recknagel, 1990b, Hyperfine Interact. 61,
Ansaldo, D.R. Harshman and R. Keitel, 1986, 1017.
Phys. Rev. B 33, 7813. Buschow, K.H.L, 1986, Mater. Sci. Repts 1, 1.
Brewer, J.H., E.J. Ansaldo, J.E Carolan, A.C.D. Butz, T., J. Chappert, J.F. Dufresne, O.
Hartmann, E. Karlsson, B. Lindgren, L.O.
Chaklader, W.N. Hardy, D.R. Harshman,
Norlin, P. Podini and A. Yaouanc, 1980, Phys.
M.E. Heyden, M. Ishikawa, N. Kaplan, Lett. A 75, 321.
R. Keitel, J. Kempton, R.E Kiefl, WJ. Butz, T., L. Asch, G.M. Kalvius, B. Lindgren,
Kossler, S.R. Kreitzman, A. Kulpa, Y. Kuno, O. Hartmann, E. Karlsson, R. W~ipplingand
G.M. Luke, H. Miyatake, K. Nagamine, A. Yaouanc, 1987, Hyperfine Interact. 35,
Y. Nakazawa, N. Nishida, K. Nishiyama, 761.
S. Okuma, T.M. Riseman, G. Roehmer, R Campbell, I.A., 1984, J. Phys. Lett. (Paris) 45,
Schleger, D. Shimada, C.E. Stronach, T. L27.
Takabatake, Y.J. Uemura, Y. Watanabe, D.L1. Campbell, I.A., A. Amato, F.N. Gygax, D.
Williams, T. Yamazaki and B. Yang, 1988, Herlach, A. Schenck, R. Cywinski and S.H.
Phys. Rev. Lett. 60, 1073. Kilcoyne, 1994, Phys. Rev. Lett. 72, 1291.
MUON SPIN ROTATION SPECTROSCOPY 287

Caspary, R., R Hellmann, M. Keller, G. Sparn, Cywinski, R., S.H. Kilcoyne, S.EJ. Cox, C.A.
C. Wassilew, R. KOhler,C. Geibel, C. Schank, Scott and O. Scharpf, 1990a, Hyperfine
E Steglich and N.E. Phillips, 1993, Phys. Interact. 64, 427.
Rev. Lett. 71, 2146. Cywinski, R., S.H. Kilcoyne, C. Ritter and
Chan, K.C.B., R.L. Lichti, C. Boekema, A.B. J. Rodriguez, 1990b (see Cywinski et al.
Denison, D.W. Cooke and M.E. Schillaci, 1990a).
1986, Hyperfine Interact. 31, 481. Cywinski, R., S.H. Kilcoyne and C.A. Scott,
Chappert, J., 1984, in: Muons and Pions in 1991, J. Phys.: Condens. Matter 3, 6473.
Materials Research, eds J. Chappert and R.I. Cywinski, R. and B.D. Rainford, 1992, J. Magn.
Grynszpan (Elsevier, Amsterdam), p. 35. Magn. Mater. 104-107, 1424.
Chappert, J., G.M. Kalvius, L. Asch, A. Yaouanc, Cywinski, R., S.H. Kilcoyne, T. Holubar and G.
O. Hartmann, L.O. Norlin and E. Karlsson, Hilscher, 1994a, Hyperfine Interact. 85, 221.
1981, Hyperfine Interact. 9, 595. Cywinski, R. and B.D. Rainford, 1994,
Chappert, J., L. Asch, M. Bog6, G.M. Kalvius Hyperfine Interact. 85, 215.
and B. Boucher, 1982, J. Magn. Magn. Dalmas de R6otier, R, B. Licciardi, A. Yaouanc,
Mater. 28, 124. B. Chevalier, R Chaudou~t, R. Madar and J.R
Chappert, J. and A. Yaouanc, 1986a, in: Topics Sanchez, 1990a, Hyperfine Interact. 59, 317.
in Current Physics, ed. U. Gonser, Vol. Dalmas de R6otier, R, J.R Sanchez, E Vulliet,
40, Microscopic Methods in Metals (Springer, A. Yaouanc, S. Harris, O. Hartmann, E.
Heidelberg) p. 297. Karlsson, R. W~ippling, B. Barbara and Ph.
Chappert, J., A. Yaouanc, O. Hartmann, E. L'H6ritier, 1990b, Hyperfine Interact. 64,
Karlsson, E. W~ickelgard, R. Wappling, L. 383.
Asch and G.M. Kalvius, 1986b, Hyperfine
Dalmas de R6otier, R, J.R Sanchez, A. Yaouanc,
Interact. 31, 331.
S.W. Harris, O. Hartmann, E. Karlsson,
Chen, H.H. and P.M. Levy, 1971, Phys. Rev.
R. Wappling, D. Gignoux, B. Gorges, D.
Lett. 27, 1383, •385.
Schmitt, Ph. L'H6ritier, A. Weidinger and
Cooke, D.W., R.H. Heffner, R.L. Hutson,
RC.M. Gubbens, 1990c, Hyperfine Interact.
M.E. Schillaci, J.L. Smith, J.O. Willis, D.E.
64, 389.
MacLaughlin, C. Boekema, R.L. Lichti, A.B.
Dalmas de R6otier, R, J.R Sanchez, A. Yaouanc,
Denison and J. Oostens, 1986, Hyperfine
Interact. 31, 425. B. Chevalier, R Chaudou~t and R. Madar,
Cooke, D.W., R.L. Hutson, R.S. Kwok, M. 1990d, Hyperfine Interact. 64, 457.
Maez, H. Rempp, M.E. Schillaci, J.L. Smith, Dalmas de R6otier, R, A. Yaouanc, RC.M.
J.O. Willis, R.L. Lichti, K.C.B. Chan, C. Gubbens, D. Gignoux, B. Gorges, D. Schmitt,
Boekema, S.P. Weathersby and J. Oostens, O. Hartmann, R. W~ppling and A. Weidinger,
1989, Phys. Rev. B 39, 2748. 1992, J. Magn. Magn. Mater. 104-107,
Cooke, D.W., M.S. Jahan, R.S. Kwok, R.L. 1267.
Lichti, T.R. Adams, C. Boekema, W.K. Dalmas de R6otier, R and A. Yaouanc, 1994a,
Dawson, A. Kebede, J. Schwegler, J.E. Crow Phys. Rev. Lett. 72, 290.
and T. Mihalisin, 1990a, Hyperfine Interact. Dalmas de R6otier, R and A. Yaouanc, 1994b,
63, 213; 1990b, Physica B 163, 675; 1990c, Hyperfine Interact. 85, 233.
J. Appl. Phys. 67, 5061. Dalmas de R6otier, R, A. Yaouanc, RC.M.
Cooke, D.W., R.S. Kwok, R.L. Lichti, T.R. Gubbens and B. Chevalier, 1994c, Hyperfine
Adams, C. Boekema, W.K. Dawson, A. Interact. 85, 227.
Kebede, J. Schwegler, J.E. Crow and T. Daniel, E. and J. Friedel, 1963, J. Phys. Chem.
Mihalisin, 1990d0 Phys. Rev. B 41, 4801. Solids 24, 1601.
Cox, D.L., N.E. Bichers and J.W. Wilkins, 1985, Davis, M.R., C.E. Stronach C.E., R.S. Cary,
J. Appl. Phys. 57, 3166. WJ. Kossler, X.H. Yu, H.E. Schone, W.J.
Cox, D.E., G. Shirane, S.M. Shapiro, G. Aeppli, Lankford, A.R. Moodenbaugh, J. Oostens and
Z. Fisk, J.L. Smith, J.K. Kjems and H.R. Ott, E Boolchand, 1990, Hyperfine Interact. 63,
1986, Phys. Rev. B 33, 3614. 305.
Cox, S.EJ., 1987, J. Phys. C 20, 3187. Dawson, W.K., K. Tibbs, S.P. Weathersby, C.
Crook, M.R. and R. Cywinski, 1994, Hyperfine Boekema and K.-C.B. Chan, 1988, J. Appl.
Interact. 85, 203. Phys. 64, 5809.
288 A. SCHENCK and EN. GYGAX

Dawson, W.K., C.H. Halim, S.E Weathersby, Duginov, V.N., G. Grebinnik, R. Horyn, B.E
J.A. Flint, J.C. Lam, T.J. Hoffman, C. Kirillov, J. Klamut, I.A., K.I. Krivosheev,
Boekema, K.-C.B. Chan, R.L. Lichti, D.W. K.I., T.N. Mamedov, V.G. Olshevsky,
Cooke and M.S. Jahan, 1990, Hyperfine A.V. Pirogov, V.Yu. Pomjakushin, A.N.
Interact. 63, 219. Ponomarev, A.J. Zaleski and V.A. Zhukov,
Dawson, W.K., C. Boekema, R.L. Lichti and 1994b, Hyperfine Interact. 85, 311.
D.W. Cooke, 1991a, Physica C 185-189, Duginov, V.N., G. Grebinnik, T.N. Mamedov,
1221. V.G. Olshevsky, V.Yu. Pomjakushin, V.A.
Dawson, K.W., J.C. Lam, C. Boekema, D.W. Zhukov, B.F. Kirillov, B.A. Nikolsky, V.G.
Cooke and J.E. Crow, 1991b, J. Appl. Phys. Orlov, A.V. Pirogov, A.N. Ponomarev,
69, 5385. V.A. Suetin and E.A. Kravchenko, 1994c,
Hyperfine Interact. 85, 197.
Delrieu, J.M., M. Roger, Z. Toffano, E. Wope
Dunlap, B.D., K.G. Shenoy, EY. Fradin, C.D.
Mbougue, P. Fanvel, R. Saint James and K.
Burnet and C.W. Kimball, 1979, J. Magn.
Beechgaard, 1986a, Physica B 143, 412.
Magn. Mater. 13, 319.
Delrieu, J.M., M. Roger, Z. Toffano, A. Dunlap, B.D., M. Slaski, D.G. Hinks, L.
Moradpour, K. Beechgaard, 1986b, J. Phys. Soderholm, M. Beno, K. Zhang, C. Segre,
(Paris) 47, 839. G.W. Grabfree, W.K. Kwok, S.K. Malik, I.K.
Demazeau, G., A. Marbeuf, M. Pouchard and P. Schuller, J.D. Sorgensen and J. Sungaila,
Hagenmuller, 1971, J. Solid State Chem. 3, 1987, J. Magn. Magn. Mater. 68, L 139.
582. Effantin, J.M., J. Rossat-Mignod, P. Burlet, H.
Denison, A.B., H. Graf, W. KiJndig and P.E Bartholin, S. Kunii and T. Kasuya, 1985, J.
Meier, 1979, Helvetica Physica Acta 52, 460. Magn. Magn. Mater. 47+48, 145.
Denison, A.B., C. Boekema, R.L. Lichti, K.C.B. Emery, V.J., S.A. Kievelson and G.H. Lin, 1990,
Chan, D.W. Cooke, R.H. Heffner, R.L. Phys Rev. Lett. 64, 475.
Hutson, M. Leon and M.E. Schillaci, 1985, J. Endoh, Y., M. Matsuda, K. Yamada, K. Kakurai,
Appl. Phys. 57, 3743. Y. Hidaka, G. Shirane and R.J. Birgneau,
De Renzi, R., 1992, J. Magn. Magn. Mater. 1989, Phys. Rev. B 40, 7023.
104-107, 461. Erkelens, W.A.C., L.P. Regnault, P. Burlet, J.
De Renzi, R., G. Guidi, C. Bucci, P. Podini, Rossat-Mignod, S. Kunii and T. Kasuya,
R. Tedeschi and S.EJ. Cox, 1984a, Hyperfine 1987, J. Magn. Magn. Mater. 63+64, 61.
Interact. 17-19, 479. Estreicher, S. and P.F. Meier, 1982, Phys. Rev.
De Renzi, R., G. Guidi, P. Podini, R. Tedeschi, B 25, 297.
C. Bucci and S.EJ. Cox, 1984b, Phys. Rev. Estreicher, S. and P.F. Meier, 1984, Hyperfine
B 30, 186 and 197. Interact. 17-19, 327.
Fawcett, E., 1988, Rev. Modern Phys. 60, 209.
De Rcnzi, R., G. Guidi, P. Carretta, G. Calestani
Felcher, G.P., %O. Brum, R.J. Gambino and M.
and S.EJ. Cox, 1989, Phys. Lctt. A 135,
Kuznietz, 1973, Phys. Rev. B 8, 260.
132.
Felner, I., U. Yaron, I. Novik, E.R. Bauminger,
de Visser, A., N.H. van Dijk, K. Bakker and Y. Wolfus, E.R. Yacoby, G. Hilscher and N.
J.M.M. Franse, 1993, Physica B 186-188, Pillmayr, 1989, Phys. Rev. B 40, 6739.
212. Felsteiner, J., 1989, Phys. Rev. B 39, 7248
Didisheim, EE, K. Yvon, P. Fischer and D. Ferreira, L.P., R. Guillen, P. Vulliet, A. Yaouanc,
Shaltiel, 1980, J. Less-Common Met. 73, D. Fruchart, P. Wolfers, P. L'Heritier and R.
355. Fruchart, 1985, J. Magn. Magn. Mater 53,
Dodds, S.A., G.A. Gist, D.E. MacLaughlin, 145.
R.H. Heffner, M. Leon, M.E. Schillaci, G.J. Feyerherm, R., A. Amato, F.N. Gygax, A.
Nieuwenhuys and J.A. Mydosh, 1983, Phys. Schenck, Y. (3nuki and N. Sato, 1994a,
Rev. B 28, 6209. Physica B 194-196, 357.
Duginov, V.N., G. Grebinnik, K.I. Gritsaj, Feyerherm, R., A. Amato, EN. Gygax, A.
T.N. Mamedov, V.G. Olshevsky, V.Yu. Schenck, U. Zimmermann, A. Grayevski and
Pomjakushin, V.A. Zhukov, B.F. Kirillov, N. Kaplan, 1994b, Hyperfine Interact. 85,
I.A. Krivosheev, A.V. Pirogov and A.N. 329.
Ponomarev, 1994a, Hyperfine Interact. 85, Feyerherm, R., A. Amato, C. Geibel, EN.
317. Gygax, T. Kamtsubara, N. Sato, A. Schenck
MUON SPIN ROTATION SPECTROSCOPY 289

and E Steglich, 1994c, Physica B 199+200, P6rez, J. Fontauberta, C. Miravitlles, J.


103, and Phys. Rev. Lett. 73, 1849. Rodriguez-Carvajal and B. Martinez, 1990,
Feyerherm, R., A. Amato, C. Geibel, F.N. Physica C 170, 153.
Gygax, P. Hellmann, R.H. Heffner, D.E. Garcia-Mt~noz, J.L., X. Obradors, S.H. Kilcoyne
MacLaughlin, R. MUller-Reisener, G.I. Nieu- and R. Cywinski, 1991, Physica C 195-189,
wenhuys, A. Schenck and E Steglich, 1995, 1085.
Physica B 206+207, 596. Gavilano, J.L., J. Hunziker, O. Hudak, T. Sleator,
Fischer, 0 , 1978, Appl. Phys. 16, 1. F. Hulliger and H.R. Ott, 1993, Phys, Rev. B
Fischer, P., W. H~ilg, D. Schwarzenbach and H. 47, 3438.
Gamsjfiger, 1974, J. Phys. Chem. Solids 35, Geldart, D.J.W., P. Hargraves, N.M. Fujiki and
1683. R.A. Dunlap, 1989, Phys. Rev. Lett. 62,
Fischer, P., B. Lebech, G. Meier, B.D. Rainford 2728.
and O. Vogt, 1978, J. Phys. C 11, 345. Gignoux, D., D. Givord and A. Del Moral, 1976.
Fischer, P., K. Kakurai, M. Steiner, K.N. Solid State Commun. 19, 891.
Clausen, B. Lebech, E Hulliger, H.R. Ott, P. Gignoux, D. and J.J. Rhyne, 1986, J. Magn.
Brtisch and P. Untern~ihrer, 1988, Physica C Magn. Mater. 54-57, 1179.
152, 145. Gist, G.A., S.A. Dodds, D.E. MacLaughlin, D.W.
Fisk, Z., P.C. Canfield, W.P. Beyermann, J.D. Cooke, R.H. Heffner, R.L. Hutson, M.E.
Thompson, M.E Hundley, H.R. Ott, E. Schillaci, C. Boekema, J.A. Mydosh and G.J.
Felder, M.B. Maple, M.A. Lopez de la Torre, Nieuwenhuys; 1986, Phys. Rev. B 34, 1683.
P. Visani and C.L. Seaman, 1991, Phys. Rev. Gltickler, H., A. Weidinger, A. Golnik, Ch.
Lett. 67, 3310. Niedermayer, M. Rauer, R. Simon, E.
Forsyth, J.B., P.J. Brown and B.M. Wanklyn, Recknagel, J.I. Budnick, W. Paulus and R.
1988, J. Phys. C: Solid State Phys. 21, 2917. SchOllhorn, 1989, Physica C 162-164 149.
Fournier, J.M. and R. Troc, 1985, in: Handbook Glticker, H., Ch. Niedermayer, G. Novitzke,
on the Physics and Chemistry of the E. Recknagel, A. Weidinger and J.I. Budnick,
Actinides, Vol. II, eds A.J. Freeman and G.H. 1990a, J. Less-Common Met. 164+165, 1016.
Lander (North-Holland, Amsterdam) p. 29. Glticker, H, Ch. Niedermayer, G. Novitzke,
Foy, M.L.G., N. Heiman, W.J. Kossler and C.E. A. Golnik, R. Simon, E. Recknagel, A.
Stronach, 1973, Phys. Rev. Lett. 30, 1064. Weidinger, J. Erxmeyer and J.I. Budnick,
Fraas, K., U. Ahlheim, P.H.P. Reinders, C. 1990b, Hyperfine Interact. 63, 155.
Schank, R. Caspary, E Steglich, A. Ochiai, T. Goldman, A.I., G. Shirane, G. Aeppli, B. Batlogg
Suzuki and T. Kasuya, 1992, J. Magn. Magn. and E. Bucher, 1986, Phys. Rev. B 34, 6564.
Mater. 108, 220. Golnik, A., Ch. Niedermayer, E. Recknagel, M.
Freeman, A.J. and R.E. Watson, 1965, in: Rossmanith, A. Weidinger, J.I. Budnick, B.
Magnetism, Vol. IIA, eds G.T. Rado and H. Chamberland, M. Filipkowski, Y. Zhang, D.P.
Suhl (Academic Press, New York) chapt. 4. Yang, L.L. Lynds, F.A. Olter and C. Bains,
Freeman, A.J., J.V. Mallow, J.P. Desclaux and 1987, Phys. Lett. A 125, 71
M. Weinert, 1984, Hyperfine Interact. 17-19, Goodenough, J.B., 1963, Magnetism and the
865. Chemical Bond (Wiley, New York).
Frey, E. and E Schwabl, 1988, Z. Physik B 71, Goren, D., C. Korn, V. Voterra, M. Schaefer,
355. H. Riesemeier, E. R6ssler, H. Stenschke,
Frey, E., F. Schwabl and S. Thoma, 1989a, Phys. H.M. Vieth and K. Ltiders, 1989, Solid State
Rev. B 40, 7199. Commun. 70, 279.
Frey, E. and E Schwabl, 1989b, Hyperfine Gradwohl, B.A., A. Grayevski, N. Kaplan, EN.
Interact. 50, 767. Gygax, A. Schenck and A.J. van der Wal,
Frings, EH., B. Renker and G. Vettier, 1987, J. 1986, Hyperfine Interact. 31, 319.
Magn. Magn. Mater. 63+64, 202. Graf, H., W. K0ndig, B.D. Patterson, W.
Fritzsehe, A., M. Hampele, D. Herlach, K. Reichart, P. Roggwiller, M. Camani, EN.
Maier, J. Major, L. Schimmele, A. Seeger, Gygax, W. Rtiegg, A. Schenck and H.
W. Staiger, W. Tempi and C. Baines, 1990, Schilling, 1976a, Helvetica Physica Acta 49,
Hyperfine Interact. 64, 691. 730.
Fuertes, A., X. Obradors, J.M. Navarro, Graf, H., W. Ktindig, B.D. Patterson, W.
P. Gomez-Romero, N. Casafi-Pastor, E Reichart, P. Roggwiller, M. Camani, EN. Gy-
290 A. SCHENCK and EN. GYGAX

gax, W. Ri~egg, A. Schenck, H. Schilling and Gubbens, RC.M., A.M. van der Kraan and K.H.J.
P.E Meier, 1976b, Phys. Rev. Lett. 37, 1644. Buschow, 1989, Phys Rev. B 39, 12548.
Graf, H., W. Hofmann, W. K0ndig, P.E Meier, Gubbens, P.C.M., P. Dalmas de R6otier, J.P.
B.D. Patterson and W. Reichart, 1977, Solid Sanchez, A. Yaouanc, C.E. Snel, R. Verhoef,
State Commun. 23, 653. F. Kayzel, J. Song-Quan and J.J.M. Franse,
Graf, H., W. Hofmann, W. Ktindig, P.E Meier, 1992, J. Magn. Magn. Mater. 104-107,
B.D. Patterson, W. Reichart and A. Rodriguez, 1269.
1978, Hyperfine Interact. 4, 452. Gubbens, P.C.M., A.A. Moolenaar, P. Dalmas
Graf, H., G. Balzer, T. M/Sslang, E. Recknagel, de R6otier, A. Yaouanc, F. Kayzel, J.J.M.
A. Weidinger and K.H.J. Buschow, 1981, Franse, K. Prokes, C.E. Snel, P. Bonville,
Hyperfine Interact. 8, 605. J.A. Hodges, P. Imbert and E Pari, 1994a,
Granados, X., et al., 1990, L. Less-Common Hyperfine Interact. 85, 239.
Met. 164+165, 853. Gubbens, P.C.M., P. Dalmas de R6otier, A.
Yaouanc, A.A. Menovsky and C.E. Snel,
Grebinnik, V.G., I.I. Gurevich, V.A. Zhukov,
1994b, Hyperfine Interact. 85, 245.
V.A. Nikol'skii, V.I. Selivanov and V.A.
Gurevich, I.I., A.I. Klimov, V.N. Maiorov, E.A.
Suetin, 1979a, Zh. Eksp. Teor. Fiz. 76, Meleshko, I.A. Muratova, B.A. Nikol'skii,
2178; [1979, Soy. Phys. - JETP 49, 1100]. V.S. Roganov, V.I. Selivanov and V.A. Suetin,
Grebinnik, V.G., I.I. Gurevich, A.I. Klimov, V.N. 1974, Zh. Eksp. Teor. Fiz. 66, 374; [1974,
Majorov, A.E Manych, E.V. Melnikov, B.A. Sov. Phys. - JETP 39, 178].
Nikol'skii, A.V. Pirogov, A.N. Ponomarev, Gurevich, I.I., A.I. Klimov, V.N. Maiorov, E.A.
V.I. Selivanov, V.A. Suetin and V.A. Zhukov, Meleshko, B.A. Nikol'skii, A.V. Pirogov, V.S.
1979b, Hyperfine Interact. 6, 103. Roganov, V.I. Selivanov and V.A. Suetin,
Grebinnik, V.G., V.N. Duginov, V.A. Zhukov, S. 1975a, ZhETF Pis. Red. 21, 16. [1975,
Kapusta, A.B. Lazarev, V.G. Olshevsky, V.Yu. JETP Lett. 21, 7].
Pomjakushin, S.N. Shilov, I.I. Gurevich, B.E Gurevich, I.I., A.I. Klimov, V.N. Maiorov, E.A.
Kirrillov, B.A. Nikolsky, A.V. Pirogov, A.N. Meleshko, B.A. Nikol'skii, V.I. Selivanov and
Ponomarev, V.A. Suetin, I.E Borovinskaya, V.A. Suetin, 1975b, Zh. Eksp. Teor. Fiz. 69,
M.D. Nersesyan, A.G. Peresada, Yu.E Eltzev, 439; [1976, Sov. Phys. - JETP 42, 222].
V.R. Karasik and O.E. Omelyanovski, 1990a, Gurevich, I.I., A.I. Klimov, V.N. Maiorov, E.A.
Hyperfine Interact. 61, 1085. Meleshko, B.A. Nikol'skii, A.V. Pirogov, V.S.
Grebinnik, V.G., V.N. Duginov, V.A. Zhuhov, Roganov, V.I. Selivanov and V.A. Suetin,
S. Kapusta, A.B. Lazarev, V.G. Olshevsky, 1975c, Zh. Eksp. Teor. Fiz. 69, 1453; [1976,
V.Yu. Pomjakushin, S.N. Shilov, D.T. Sov. Phys. - JETP 42, 741].
Bezhitadze, I.I. Gurevich, B.E Kirillov, E.E Gurevich, I.I., A.I. Klimov, V.N. Maiorov, B.A.
Krasnoperov, B.A. Nikolsky, A.V. Pirogov, Nikol'skii, V.S. Roganov, V.I. Selivanov and
A.M. Ponomarev, V.A. Suetin and G.E V.A. Suetin, 1976, ZhETF Pis. Red. 23, 345;
Tavadze, 1990b, Hyperfine Interact. 61, 1089. [1976, JETP Lett. 23, 310].
Grebinnik, V.G., V.N. Duginov, V.A. Zhukov, S. Gygax, EN., W. Rtiegg, A. Schenck, H.
Kapusta, A.B. Lazarev, V.G. Olshevsky, V.Yu. Schilling, W. Studer and R. Schulze, 1980, J.
Pomjakushin, S.N. Shilov, I.I. Gurevich, B.E Magn. Magn. Mater. 15-18, 1191.
Gygax, EN., A. Hintermann, W. RUegg, A.
Kirrillov, E.E Krasnoperov, B.A. Nikolsky,
A.V. Pirogov, A.N. Ponomarev, V.A. Suetin Schenck, W. Studer, H. Schnidt, B. Scheerer,
W. Oestreich, E Goetz and G. Czjzek, 1981,
and A.I. Morozov, 1990c, Hyperfine Interact.
Hyperfine Interact. 8, 491.
63, 241
Haen, P., J. Flouquet, E Lapierre, P. Lejay and
Grewe, N. and E Steglich, 1991, in: Handbook G. Remenyi, 1987, J. Low Temp. Phys. 67,
on the Physics and Chemistry of Rare Earths, 391.
Vol. 14, eds K.A. Gschneider and L. Eyring Hagen, M., W.G. Stirling and G.H. Lander, 1988,
(Elsevier, Amsterdam). Phys. Rev. B 37, 1846.
Grover, A.K., B.R. Coles, B.V.B. Sarkissian and Hagn, E., E. Zech and G. Eska, 1982, J. Phys.
H.E.N. Stone, 1982, J. Less-Common Met. F: Met. Phys. 12, 1475.
86, 29. Haldane, F.D.M., 1983, Phys. Lett. A 93, 464.
Grund, T., M. Hampele, J. Major, L. Schimmele H~ilg, B., A. Furrer and O. Vogt, 1985, Phys.
and A. Seeger, 1994, Preprint. Rev. Lett. 54, 1388.
MUON SPIN ROTATION SPECTROSCOPY 291

H~tlg, B. and A. Furrer, 1986, Phys. Rev. B 34, Hayano, R.S., Y.J. Uemura, J. Imazato, N.
6258. Nishida, K. Nagamine, T. Yamazaki, Y.
H~ilg, B., A. Furrer and J.K. Kjems, 1987, Phys. Ishikawa and H. Yasuoka, 1980, J. Phys. Soc.
Rev. Lett. 59, t034. Jpn 49, 1773.
Halim, C.H., W.K. Dawson, W.A. Baldwin and Heffner, R.H., 1993, Invited talk at Actinides-93,
C. Boekema, 1990, Physica B 163, 453. Int. Conf., Sep. 20-24, Santa Fe, NM.
Harshman, D.R., G. Aeppli, G.P. Espinosa, A.S. Heffner, R.H., D.W. Cooke, R.L. Hutson, M.
Cooper, J.P. Remeika, E.J. Ansaldo, T.M. Leon, M.E. Schillaci, J.L. Smith, A. Yaouanc,
Riseman, D.L1. William, D.R. Noakes, B. S.A. Dodds, L.C. Gupta, D.E. MacLaughlin
Ellman and T.E Rosenbaum, 1988, Phys. and C. Boekema, 1984, J. Appl. Phys. 55,
Rev. B 38, 852. 2007.
Harshman, D.R., G. Aeppli, B. Batlogg, G.P. Heffner, R.H., W.D. Cooke, R.L. Hutson, M.E.
Espinosa, R.J. Cava, A.S. Cooper, L.W. Rupp, Schillaci, J.L. Smith, P.M. Richards, D.E.
E.J. Ansaldo, and D.L1. Williams, 1989, MacLaughlin, S.A. Dodds and J. Oostens,
Phys. Rev. Lett. 63, 1187. 1985, J. Appl. Phys. 57, 3107.
Hartmann, O., 1990, Hyperfine Interact. 64, 641. Heffner, R.H., D.W. Cooke, D.E. MacLaughlin,
Hartmann, O., E. Karlsson, R. W~tppling, J. 1987, in: Theoretical and Experimental
Chappert, A. Yaouanc, L. Asch and G.M. Aspects of Valence Fluctuations and Heavy
Kalvius, 1984, Hyperfine Interact. 17-19, Fermions, eds L.C. Gupta and S.K. Malik
491. (Plenum, New York) p. 319.
Hartmann, O., E. Karlsson, R. Wfippling, J. Heffner, R.H., D.W. Cooke, A.L. Giorgi, R.L.
Chappert, A. Yaouanc, L. Asch and G.M. Hutson, M.E. Schillaci, H.D. Rempp, J.L.
Kalvius, 1986, J. Phys. F: Met. Phys. 16, Smith, J.O. Willis, D.E. MacLaughlin, C.
1593. Boekema, R.L. Lichti, J. Oostens and A.B.
Hartmann, O, R. Wappling, A. Yaouanc, P. Denison, 1989a, Phys. Rev. B 39, 11345.
Dalmas de R6otier, B. Barbara, K. Aggarwal, Heffner, R.H., J.O. Willis, J.L. Smith, P. Birrer,
L. Asch, A. Kratzer, G.M. Kalvius, EJ. C. Baines, EN. Gygax, B. Hitti, E. Lippelt,
Litterst, EN. Gygax, B. Hitti, E. Lippelt and H.R. Ott, A. Schenck and D.E. MacLaughlin,
A. Schenck, 1989, Hyperfine Interact. 51,
1989b, Phys Rev. B 40, 806.
955. Heffner, R.H., J.L. Smith, J.O. Willis, P. Birrer,
Hartmann, O., B. Lindgren, R. Wappling, T.
C. Baines, EN. Gygax, B. Hitti, E. Lippelt,
Butz and G.M. Kalvius, 1990a, Hyperfine
H.R. Ott, A. Schenck, E.A. Knetsch, J.A.
Interact. 64, 367.
Mydosh and D.E. MacLaughlin, 1990, Phys.
Hartmann, O., R. Wfippling, E. Karlsson, G.M.
Rev. Lett. 65, 2816.
Kalvius, L. Asch, EJ. Litterst, K. Aggarwal,
K.H. Munch, KN. Gygax and A. Schenck, Heffner, R.H., W.P. Beyermann, M.E Hundley,
J.D. Thompson, J.L. Smith, Z. Fisk, K.
1990b, Hyperfine Interact. 64, 369.
Hartmann, O., S.W. Harris, R. W~ippling,G.M. Bedell, P. Birrer, C. Baines, EN. Gygax, B.
Kalvius, L. Asch, P. Dalmas de R6otier and Hitti, E. Lippelt, H.R. Ott, A. Schenck and
A. Yaouanc, 1990c, Hyperfine Interact. 64, D.E. MacLaughlin, 1991, J. Appl. Phys. 69,
381. 5481.
Hartmann, O., E. Karlsson, R. W~ippling, L. Heffner, R.H., A. Amato, P.C. Canfield, R.
Asch, S. Henneberger, G.M. Kalvius, A. Feyerherm, Z. Fisk, EN. Gygax, D.E.
Kratzer, H.H. Klauss, J. Litterst and M.A.C. MacLaughlin, A. Schenck, J.D. Thompson
de Melo, 1994, Hyperfine Interact. 85, 251. and H.R. Ott, 1994, Physica B 199+200, 113.
Hayano, R.S., Y.J. Uemura, J. Imazato, N. Heffner, R.H. and D.L. Cox, 1989, Phys. Rev.
Nishida, T. Yamazaki, H. Yasuoka and Y. Lett. 63, 2538.
Ishikawa, 1978a, Phys. Rev. Lett. 41. 1743. Herlach, D., K. Ft~rderer, M. F~nle and L.
Hayano, R.S., Y.J. Uemura, J. Imazato, N. Schimmele, 1986, Hyperfine Interact. 31,
Nishida, K. Nagamine, T. Yamazaki and H. 287.
Yasuoka, 1978b, Phys. Rev. Lett. 41, 421. Herlach, D., V. Claus, K. Fiarderer, J. Major,
Hayano, R.S., Y.J. Uemura, J. Imazato, N. A. Seeger, L. Schimmele, M. Schmolz,
Nishida, T. Yamazald and R. Kubo, 1979, W. Staiger and W. Tempi, 1989, Z. Phys.
Phys. Rev. B 20, 850. Chem. N.E 164, 1041.
292 A. SCHENCK and EN. GYGAX

Hitti, B., R Birrer, A. Grayevski, EN. Gygax, J6rome, D. and H.J. Schulz, 1982, Adv. Phys.
N. Kaplan, E. Lippelt, A. Schenck and M. 31, 299.
Weber, 1990a, Hyperfine Interact. 59, 377. Kadono, R., T. Matszaki, T. Yamazaki, S.R.
Hitti, B., R Birrer, K. Fischer, EN. Gygax, Kreitzman and J.H. Brewer, 1990, Phys Rev.
E. Lippelt, H. Maletta, A. Schenck and M. B 42, 6515.
Weber, 1990b, Hyperfine Interact. 63, 287. Kadono, R., J.H. Brewer, K. Chow, S.R.
Hofmann, W., W. Ktindig, RE Meier, B.D. Kreitzman, Ch. Niedermayer, T.M. Riseman,
Patterson, K. Rtiegg, O. Echt, H. Graf, E. J.W. Schneider and T. Yamazaki, 1993, Phys.
Recknagel, A. Weidinger and T. Wichert, Rev. B 48, 16803.
1978, Phys. Lett. A 65, 343. Kadowaki, H., T. Ekino, H. Iwasaki, T.
Hohenemser, C. and N. Rosov, Kleinhammes Takabatake, H. Fujii and J. Sakura, 1993, J.
1989, Hyperfine Interact. 49, 267. Phys. Soc. Jpn 62, 4426.
Holzschuh, E., C. Boekema, A.B. Denison, W. Kakani, S.L. and U.N. Upadhyaya, 1988, J. Low
Ktindig, P.E Meier and K. R0egg, 1980, Temp. Phys. 70, 4.
Hyperfine Interact. 8, 77. Kalvius, G.M, K. Nishiyama, K. Nagamine,
Holzschuh, E., C. Boekema, W. K[indig, K. T. Yamazaki, J. Chappert, O. Hartmann, E.
Riaegg and B.D. Patterson, 1981, Hyperline Karlsson, R. W~ippling, A. Yaouanc and L.
Interact. 8, 615. Asch, 1984, Hyperfine Interact. 17-19, 497.
Holzschuh, E., A.B. Denison, W. Ktindig, RE Kalvius, G.M., L. Asch, D.R. Noakes, R.
Meier and B.D. Patterson, 1983, Phys. Rev. Keitel, J.H. Brewer, E.J. Ansaldo, EJ. Litterst,
B 27, 5294. B. Boucher, J. Chappert, A. Yaouanc, T.
Huang, C.Y., C.E. Olsen, W.W. Fuller, J.H. Yamazaki, K. Nagamine, K. Nishiyama, O.
Huang and S.A. Wolf, 1983a, Solid State Hartmann and R. W~ippling, 1986, Hyperfine
Commun. 45, 795. Interact. 31, 303; 1987, J. Magn. Magn.
Huang, C.Y., E.J. Ansaldo, J.H. Brewer, D.R.
Mater. 70, 285.
Harshman, K.M. Crowe, S.S. Rosenblum,
Kalvius, G.M., L. Asch, EJ. Litterst, D. Noakes,
C.W. Clawson, Z. Fisk, S. Lambert,
J.H. Brewer, E.J. Ansaldo, J. Chappert, P.
M.S. Torikachvili and M.B. Maple, 1983b,
Morin, K. Nishiyama, O. Hartmann and R.
Hyperfine Interact. 17-19, 509.
Wappling, 1990, Hyperfine Interact. 64, 439.
Hyomi, K., H. Amitsuka, T. Nishioka, S.
Kalvius, G.M., A. Kratzer, K.-H. Mtinch,
Marayama, Y. Miyako, K. Nishiyama, K.
T. Takabatake, G. Nakamoto, H. Fujii,
Nagamine, T. Yamazaki and P. Morin, 1988,
R. W~ippling, H.-H. Klauss, R. Kiefi,
J. Magn. Magn. Mater. 76+77, 462.
Imazato, J., K. Nagamine, T. Yamazaki, B.D. S. Kreitzmann and D.R. Noakes, 1994,
Patterson, E. Holzschuh and R.F. Kiefl, 1984, Hyperfine Interact. 85, 411.
Hyperfine Interact. 17-19, 857, and Phys. Kanamori, J., H.R. Yoshida and K. Terakura,
Rev. Lett. 53, 1849. 1981, Hyperfine Interact. 8, 573.
Ihrig, H. and W. Lohmann, 1977, J. Phys. F 7, Kaneko, T., H. Yoshida, M. Ohashi and S. Abe,
1957. 1987, J. Magn. Magn. Mater. 70, 277.
Isaacs, E.D., D.B. MacWhan, R.N. Kleiman, D.J. Kaplan, N., A. Grayevski, P. Birrer, F.N. Gygax,
Bishop, G.E. Ice, P. Zschack, B.D. Gaulin, B. Hitti, E. Lippelt and A. Schenck, 1989,
T.E. Mason, J.D. Garrett and W.J.L. Buyers, Phys. Rev. Lett. 62, 2732.
1990, Phys. Rev. Lett. 65, 3185. Karlsson, E.B., 1990, Hyperfine Interact. 64,
Ishida, K. and Y. Kitaoka, 1991, J. Phys. Soc. 331.
Jpn 50, 3516. Katayama, H., K. Terakura and J. Kanamori,
Ishiguro, T. and K. Yamaji, 1990, in: Organic 1979, Solid State Commun. 29, 431.
Superconductors (Springer, Berlin). Katsuyama, S., Y. Ueda and K. Kosuge 1990,
Ishikawa, Y., K. Tajima, D. Bloch and M. Roth, Physica C 165, 404.
1976, Solid State Commun. 19, 525. Kawakami, M., K. Mizuno, S. Kunii, T. Kasuya,
Jaccarino, V., L.R. Walker and G.K. Wertheim, H. Enokiya and K. Kume, 1982, J. Magn.
1964, Phys. Rev. Lett. 13, 752. Magn. Mater. 30, 201.
Jena, P., 1974, Solid State Commun. 15, 1509. Keller, H., R.E Kiefl, Hp. Baumeler,
Jepsen, O., R.M. Nieminen and J. Madsen, 1980, W. Ktindig, B.D. Patterson, J.Imazato, K.
Solid State Commun. 39, 575. Nishiyama, K. Nagamine, T. Yamazaki and
MUON SPIN ROTATION SPECTROSCOPY 293

R.I. Grynszpan, 1987, Phys. Rev. B 35, Lankford, B.D. Patterson, W. Ktindig and R.I.
2008. Grynszpan, 1985, Phys. Rev. B 32, 293.
Keller, J., 1979, Hyperfine Interact. 6, 15. Kossler, W.J., X.H. Yu, A. Greer, H.E. Schone,
Keren, A., 1994, Hyperfine Interact. 85, 281. C.E. Stronach, M. Davis, R.S. Cary, W.E
Keren, A., L.P. Le, G.M. Luke, B.J. Sternlieb, Lankford, A. Moodenbaugh and J. Oostens,
W.D. Wu, Y.J. Uemura, S. Tajima and S. 1990, Hyperfine Interact. 63, 253.
Uchida, 1993, Phys. Rev. B 48, 12926. Kratzer, A., EJ. Litterst, EN. Gygax, L. Asch, A.
Keren, A., L.P. Le, G.M. Luke, W.D. Wu, Y.J. Schenck, G.M. Kalvius, S. Barth, W. Potzel
Uemura, Y. Ajiro, T. Asano, H. Kuriyama, and J.C. Spirlet, 1986, Hyperfine Interact. 31,
M. Mekata and H. Kikuchi, 1994a, Hyperfine 309.
Interact. 85, 181. Kratzer, A., G.M. Kalvius, L. Asch, EJ. Litterst,
Keren, A., L.P. Le, G.M. Luke, W.D. Wu and EN. Gygax, B. Hitti, A. Schenck, C.A. Scott,
Y.J. Uemura, 1994b, Hyperfine Interact. 85, K. Mattenberger and O. Vogt, 1990, Hyperfine
363. Interact. 64, 449.
Kiefl, R.E, G.M. Luke, S.R. Kreitzman, M. Kratzer, A., G.M. Kalvius, T. Takabatake, G.
Celio, R. Keitel, J.H. Brewer, D.R. Noakes, Nakamato, H. Fujii and S.R. Kreitzmann,
Y.J. Uemura, AM. Portis and V. Jaccarino, 1992, Europhys. Lett. 19, 649.
1987, Phys. Rev. B 35, 2079. Kratzer, A., K. Mutzbauer, S. Henneberger,
Kiefl, R.E, J.H. Brewer, J. Carolan, P. Dosanjh, G.M. Kalvius, O. Hartmann, R. W~ippling,
W.N. Hardy, R. Kadono, J.R. Kempton, R. H.H. Klauss, M.A.C. de Melo, J. Litterst and
Krahn, P. Schleger, B.X. Yang, H. Zhan,
Th. Stammler, 1994a, Hyperfine Interact. 87,
G.M. Luke, B. Sternlieb, Y.J. Uemura, W.J. 1055.
Kossler, X.H. Yu, E.J. Ansaldo, H. Takagi, S.
Kratzer, A., H.-H. Klauss, S. Zwirner, G.M.
Uchida and C.L. Seaman, 1989, Phys. Rev.
Kalvius and J.C. Spirlet, 1994b, Hyperfine
Lett. 63, 2136.
Interact. 85, 431.
Kitazawa, H., K. Katsumata, E. Torikai and K.
Kreitzman, S.R., J.H. Brewer, D.R. Harshman,
Nagamine, 1988, Solid State Commun. 67,
R. Keitel, D.L1. Williams, K.M. Crowe and
1191.
J. Ansaldo, 1986, Phys. Rev. Lett. 56, 181
Kittel, C., 1966, Quantum Theory of Solids
(Wiley, New York). Krimmel, A., P. Fischer, B. Roessli, H. Maletta,
Klauss, H-H., M.A.C. de Melo, EJ. Litterst, C. Geibel, C. Schank, A. Grauel, A. Loidl
L. Asch, A. Kratzer, S. Henneberger, G.M. and E Steglich, 1992, Z. Phys. B 86, 159.
Kalvius, K. Mattenberger and E Hulliger, Kubo, R. and T. Toyabe, 1967, in Magnetic
1994, Hyperfine Interact. 85, 293. Resonance and Relaxation, ed. R. Blinc
Knetsch, E., A. Menovsky, G.J. Nieuwen- (North-Holland, Amsterdam)
huys, J.A. Mydosh, A. Amato, R. Feyerherm, Kubota, M., H.R. Folle, Ch. Buchal, R.M.
EN. Gygax, A. Schenck, R.H. Heffner and Mueller and E Pobell, 1980, Phys. Rev. Lett.
D.E. MacLaughlin, 1993, Physica B 186-188, 22, 1812.
300. Kumagai, K., I. Watanabe, K. Kawano, H.
Knott, W.H., G.H. Lander, M.H. Mueller and O. Mutoba, K. Nishiyama, K. Nagamine, N.
Vogt, 1980, Phys. Rev. B 21, 4159. Wada, M. Okaji and K. Nara, 1991, Physica
Koghi, M, K. Ohojama, T. Osakabe and M. 195-189, 913.
Kasaya, 1992, J. Magn. Magn. Mater. 108, Kumagai, K., Y. Inoue, Y. Kohori and K.
187. Asayama, 1981, in Ternary Superconductors,
Koi, Y., A. Tsujimura and T. Hihara, 1964, J. eds G.K. Shenoy, B.D. Dunlap and EY.
Phys. Soc. Jpn 19, 1493. Fradin (Elsevier, New York) p. 185.
Kossler, W.J., A.T. Fiory, D.E. Murnick, C.E. Kumagai, K., K. Kawano, K. Fujiwara, I.
Stronach and W.F. Lankford, 1977, Hyperfine Watanabe, K. Nishiyama and K. Nagamine
Interact. 3, 287. 1993, Hyperfine Interact. 79, 929.
Kossler, W.J., A.T. Fiory, W.F. Lankford, K.G. Kuno, Y, N. Nishida, H. Miyatake, S. Okuma,
Lynn, R.P. Minnich and C.E. Stronach, 1979, Y. Watanabe, T. Yamazaki, M. Ishikawa, T.
Hyperfine Interact. 6, 93. Takabatake, Y. Nakazawa, J.H. Brewer, S.R.
Kossler, W.J., M. Namkung, B. Hitti, Y. Li, J. Kreitzman and T.M. Riseman, 1988, Phys.
Kempton, C.E. Stronach, L.R.Jr. Goode, W.F. Rev. B 38, 9276.
294 A. SCHENCK and EN. GYGAX

Kyogaku, M., Y. Kitaoka, K. Asayama, T. G.M. Kalvius, 1987, J. Magn. Magn. Mater.
Takabatake and H. Fujii, 1992, J. Phys. Soc. 67, 130.
Jpn 61, 43. Litterst, EJ., L. Asch, G.M. Kalvius, A.
Lacorre, P., J.B. Torrance, J. Pannetier, A.I. Kratzer, A. Schenck, EN. Gygax, B. Hitti, K.
Nazzal, P.W. Wang and T.C. Huang, 1991, J. Mattenberger and O. Vogt, 1990, Hyperfine
Solid State Chem. 91, 225. Interact. 64, 443.
Lappas, A., 1993, Dissertation (Univ. of Sussex). Ltiwenhaupt, M., B. Frick, U. Walter and E.
Lappas, A., J. Osborne, K. Prassides, A. Amato, Holland-Moritz, 1983, J. Magn. Magn.
R. Feyerherm, EN. Gygax and A. Schenck, Mater. 31-34, 187.
1994a, Physica B 194, 353. Loong, C.K., M. Loewenhaupt and M.L. Vrtis,
Lappas, A., K. Prassides, A. Amato, R. 1986, Physica B 136, 413.
Lovesey, S.W., 1987, Theory of Neutron
Feyerherm, F.N. Gygax and A. Schenck,
Scattering from Condensed Matter (Oxford
1994b, Hyperfine Interact. 86, 555.
University Press, Oxford).
Le, L.P., G.M. Luke, B.J. Sternlieb, Y.J. Uemura,
Lovesey, S.W., 1992, Hyperfine Interact. 72,
J.H. Brewer, T.M. Riseman, D.C. Jor~ston and 389.
L.L. Miller, 1990a, Phys. Rev. B 42, 2182. Lovesey, S.W., E.B. Karlsson and K.N.
Le, L.P., G.M. Luke, B.J. Sternlieb, Y.J. Uemura, Trohidou, 1992a, J. Phys. C 4, 2043.
J.H. Brewer, T.M. Riseman, D.C. Johnston, Lovesey, S.W., K.N. Trohidou and E.B.
L.L. Miller, Y. Hidaka and H. Murakami, Karlsson, 1992b, J. Phys. C 4, 2061.
1990b, Hyperfine Interact. 63, 279. Low, G.G., 1966, Phys. Lett. 21, 497.
Le, L.R, G.M. Luke, B.J. Sternlieb, W.D. Wu, Luke, G.M., B.J. Sternlieb, Y.J. Uemura,
Y.J. Uemura, J.H. Brewer, T.M. Riseman, J.H. Brewer, R. Kadono, R.E Kiefl, S.R.
R.V. Upasani, L.Y. Chiang and P.M. Chaikin, Kreitzman, T.M. Riseman, J. Gopalakrishnan,
1991, Europhys. Lett. 15, 547. A.W. Sleight, M.A. Sabramanian, S. Uchida,
Le, L.P., A. Keren, G.M. Luke, W.D. Wu, Y.J. H. Takagi and Y. Tokura, 1989a, Nature 338,
Uemura, M. Tamura, M. Ishikawa and M. 49.
Kinoshita, 1993, Chem. Phys. Lett. 206, Luke, G.M. et al., 1989b, Bull. Am. Phys. Soc.
405. 34, 976.
Le, L.P., A. Keren, G.M. Luke, B.J. Sternlieb, Luke, G.M., L.E Le, B.J. Sternlieb, Y.J. Uemura,
W.D. Wu, Y.J. Uemura, J.H. Brewer, T.M. J.H. Brewer, R. Kadono, R.F. Kiefl, S.R.
Riseman, R.V. Upasani, L.Y. Chiang, W. Kreitzman, T.M. Riseman, C.E. Stronach,
Kang, P.M. Chaikin, T. Csiba and G. Grtiner, M.R. Davis, S. Uchida, H. Takagi, Y. Tokura,
1994, Phys. Rev. B 48, 7284. Y. Hidaka, T. Murakama, J. Gopalakrishnan,
Li, Q. and J.H. Brewer, 1990, Hyperfine Interact. A.W. Sleight, M.A. Subramanian, E.A. Early,
63, 169. J.T. Markert, M.B. Maple and C.L. Seaman,
Lichti, R.L., T.R. Adams and T.L. Gibson, 1990a, 1990a, Phys. Rev. B 42, 7981.
Hyperfine Interact. 63, 199. Luke, G,M., L.P. Le, B.J. Sternlieb, Y.J. Uemura,
Lichti, R.L., K.-C.B. Chan, T.R. Adams, C. J.H. Brewer, R. Kadono, R.F. Kiefl, S.R.
Boekema, W.K. Dawson, J.A. Flint, D.W. Kreitzman, T.M. Riseman, C.E. Stronach, M.
Davis, S. Uchida, H. Takagi, Y. Tokura,
Cooke, R.S. Kwok and J.O. Willis, 1990b, J.
Y. Hidaka, T. Murakami, E.A. Early, J.T.
Appl. Phys. 67, 5055.
Markert, M.B. Maple and C.L. Seaman,
Lichti, R.L., D.W. Cooke and C. B.oekema,
1990b, Hyperfine Interact. 63, 311.
1991a, Phys. Rev. B 93, 1154.
Luke, G.M., L.P. Le, B.J. Sternlieb, Y.J. Uemura,
Lichti, R.L., C. Boekema, J.C. Lam, D.W. J.H. Brewer, R. Kadono, R.E Kiefl, S.R.
Cooke, S.EJ. Cox, S.T. Ting and J.E. Crow, Kreitzman, T.M. Riseman, C.L. Seaman, Y.
1991b, Physica C 180, 358. Dalichaouch, M.B. Maple and J.D. Garret,
Lin, T.K., R.L. Lichti, C. Boekema and A.B. 1990c, Hyperfine Interact. 64, 517.
Denison, 1986, Hyperfine Interact. 31, 475. Luke, G.M., L.E Le, B.J. Sternlieb, W.D. Wu,
Lin, C.L., L.W. Zhou, C.S. Jee, A. Wallash and Y.J. Uemura, J.H. Brewer, T.M. Riseman, S.
J.E. Crow, 1987, J. Less-Common Met. 133, Ishibashi and S. Uchida, 1991a, Physica C
67. 185--189, 1175.
Lindgren, B., O. Hartmann, E. Karlsson, R. Luke, G.M., L.P. Le, B.J. Sternlieb, W.D.
Wappling, A. Yaouanc, T. Butz, L. Asch and Wu, Y.J. Uemura, J.H. Brewer, R. Kadono,
MUON SPIN ROTATION SPECTROSCOPY 295

R.E Kiefl, S.R. Kreitzman, T.M. Riseman, Vol 2, eds M.B. Maple and O. Fischer
Y. Dalichaouch, B.W. Lee, M.B. Maple, C.L. (Springer, New York) p. 99.
Seaman, P.E. Armstrong, R.W. Ellis, Z. Fisk Martin, R.M., 1982, Phys. Rev. Lett. 48, 362.
and J.L. Smith, 1991b, Phys. Lett. A 157, Martinez, B., X. Obradors, E.J. Ansaldo, Ch.
173. Niedermayer, D.R. Noakes, M.J. Sayagu6s,
Luke, G.M., A. Keren, L.P. Le, W.D. Wu, Y.J. M. Vallet and J. Gonz~lles-Calbet, 1992, J.
Uemura, D.A. Bonn, L. Taillefer and J.D. Magn. Magn. Mater. 164-107, 941.
Garratt, 1993a, Physica B 186--188, 264. Mason, T.E., B.D. Gaulin, J.D. Garrett, Z. Tun,
Luke, G.M., A. Keren, L.P. Le, W.D. Wu, Y.J. W.J.L. Buyers and E.D. Isaacs, 1990, Phys.
Uemura, D.A. Bonn, L. Taillefer and J.D. Rev. Lett. 65, 3189.
Garrett, 1993b, Phys. Rev. Lett. 71, 1466. Matsui, M., I. Ishikawa, H. Matsuoka, M.
Luke, G.M., A. Keren, L.P. Le, Y.J. Uemura, Doyama, K. Nishiyama, K. Nagamine and M.
W.D. Wu, D. Bonn, L. Taillefer, J.D. Garrett Mekata, 1990, J. Nuclear Mater. 170, 211.
and Y. (3nuN, 1994, Hyperfine Interact. 85, Matsuzaki, T., K. Nishiyama, K. Nagamine, T.
397. Yamazaki, M. Senba, J.M. Bailey and J.H.
Lussier, J.G., A. Schrtider, B.D. Gaulin, J.D. Brewer, 1987, Phys. Lett. A 123, 91.
Garret, W.J.L. Buyers, L. Rebelsky and S.M. McHenry, M.R., B.G. Silbernagel and J.H.
Shapiro, 1994, Physica B 199+200, 137. Wernick, 1972, Phys. Rev. B 5, 2958.
Liitgemeier, H., 1989, Physica C 162-169, 1367. Meier, P.F., 1980, in: Exotic Atoms'79, eds
MacLaughlin, D.E., S.A. Dodds, C. Boekema, K.M. Crowe, J. Duclos, G. Fiorentini and G.
R.H. Heffner, R.L. Hutson, M. Leon, M.E. Torelli (Plenum, New York) p. 355.
Schillaci and J.L. Smith, 1983, J. Magn. Meier, RF., W. Ktindig, B.D. Patterson and K.
Magn. Mater. 31-34, 497. Rtiegg, 1987, Hyperfine Interact. 5, 311.
MacLaughlin, D.E., D.W. Cooke, R.H. Heffner, Mekata, M., S. Okamoto, S. Onoe, Y. Ajiro,
R.L. Hutson, M.W. McElfresh, M.E. Schilla- H. Kikuchi, T. Inami, E. Torikai and K.
chi, H.D. Rempp, J.L. Smith, J.O. Willis, E. Nagamine, 1990, Hyperfine Interact. 59, 415.
Zirngiebl, C. Boekema, R.L. Lichti and J. Mekata, M., S. Onoe, H. Kuriyama, B.J.
Oostens, 1988, Phys. Rev. B 37, 3153. Sternlieb, Y.J. Uemura and K. Nagamine,
Majkrzak, C.E, D.E. Cox, G. Shirane, H.A. 1992, J. Magn. Magn. Mater. 104-107, 825.
Mendels, E, J.H. Brewer, H. Alloul, E.J.
Mook, H.C. Hamaker, H.B. Mackay, Z. Fisk
and M.B. Maple, 1982, Phys. Rev. B 26, Ansaldo, D.R. Noakes, Ch. Niedermayer,
G. Collin, J.E Marucco, C.E. Stronach, G.D.
245.
Morris, T.L. Duty and S. Johnston, 1994,
Majkrzak, C.E, S.K. Satija, G. Shirane, H.C.
Phys. Rev. B 49, 10035; Hyperfine Interact.
Harnaker, Z. Fisk and M.B. Maple, 1983,
86, 577.
Phys. Rev. B 27, 2889.
Meul, H.W., M. Decroux, R. Odermatt, R. Noer
Major, J., J. Mundy, M. Schmolz, A. Seeger, K.P.
and O. Fischer, 1982, Phys. Rev. B 26,
D~Sring, K. Fiirderer, M. Gladisch, D. Herlach 6431.
and G. Majer, 1986, Hyperfine Interact. 31,
Meul, H.W., G. Rossel, M. Decreoux, O.
259. Fischer, G. Reminyi and A. Briggs, 1984,
Makoshi, K. and T. Moriya, 1978, J. Phys. Soc. Phys. Rev. Lett. 53, 497.
Jpn 44, 80. Mirabeau, I., M. Hennion, G. Goddens, T.E.
Maletta, H., E. P/3rschke, T. Chattopadhyay and Phillips and K. Moorjani, 1989, Europhys.
P.J. Brown, 1990a, Physica C 166, 9. Lett. 9, 181.
Maletta, H., M. Weber, P. Birrer, EN. Gygax, Misra, S.K. and J. Felsteiner, 1992, Phys. Rev.
B. Hitti, E. Lippelt and A. Schenck, 1990b, B 46, 11033.
Hyperfine Interact. 61, 1121. Moodenbaugh, A.R., Xu Youwen, M. Suenaga,
Maletta, H., M. Weber, P. Bitter, EN. Gygax, T.J. Folkerts and R.N. Shelton, 1988, Phys.
B. Hitti, E. Lippelt and A. Schenck, 1990c, Rev. 38, 4596.
Hyperfine Interact. 63, 235. Mook, H.A., D. McK. Paul, B.C. Sales, L.A.
Manninen, M. and R.M. Nieminen, 1981, J. Boartner and L. Cussen, 1988, Phys. Rev. B
Phys. F: Met. Phys. 11, 1213. 38, 12008.
Maple, M.B., C. Hamaker and L.D. Woolf, 1982, Morin, P., 1988, J. Magn. Magn. Mater. 71,
in: Superconductivity in Ternary Compounds, 151.
296 A. SCHENCK and EN. GYGAX

Morin, P. and D. Schmitt, 1990a, in: Ferromag- A. Kleinhammes, N. Rosov, J. Saylor, R.


netic Materials, Vol. 5, eds K.H.J. Bu- Schuhmann, L. Takacs, A. Teh, G. Zhang,
schow and E.E Wohlfahrt (North-Holland, C. Hohenemser and J.I. Budnick, 1988, Phys.
Amsterdam) p. 1. Rev. B 38, 2836.
Morin, P. et al., 1990b, cited in Kalvius et al. Niedermayer, Ch., H. Gltickler, R. Simon,
(1990). A. Golnik, M. Rauer, E. Recknagel, A.
Moriya, T., 1977, Physica B 86-88, 356. Weidinger, J.I. Budnick, W. Paulus and R.
Moriya, T., 1985, Spin Fluctuations in Itinerant SchOllhorn, 1989, Phys. Rev. B 40, 11386.
Electron Magnetism (Springer, Berlin). Niedermayer, Ch., A. Golnik, E. Recknagel, A.
Moriya, T. and A. Kawabata, 1973, J. Phys. Soc. Weidinger, A.J. Yaouanc, Ph. L'Heritier, D.
Jpn 34, 639. Fruchart, J.I. Budnick and K.H.J. Buschow,
Moriya, T. and K. Ueda, 1974, Solid State 1990, Hyperfine Interact. 64, 405.
Commun. 15, 169. Niedermayer, Ch., H. Gltickler, A. Golnik,
Motokawa, M., H. Nojiri, M. Uchi, S. Watamura, U. Binninger, M. Rauer, E. Recknagel, J.I.
K. Nishiyama and K. Nagamine, 1990, Budnick and A. Weidinger, 1993, Phys. Rev.
Hyperfine Interact. 65, 1089. B 47, 3427.
Motoya, K., H. Yasuoka, Y. Nakamura and J.H. Nishida, N., 1992a, Mechanisms of Supercon-
Wernick, 1978, J. Phys. Soc. Jpn 44, 833. ductivity, JJAP Series 7, 178.
Motoya, K., S.M. Shapiro, L. Rebelski and M.S. Nishida, N., 1992b, in: Perspectives of Meson
Torikachivili, 1991, Phys. Rev. B 44, 183. Science, eds T. Yamazaki, K. Nakai and K.
Movshovich, R., A. Lacerda, EC. Canfield, D. Nagamine (Elsevier, Amsterdam).
Arms, J.D. Thompson and Z. Fisk, 1994, Nishida, N., 1993, Hyperfine Interact. 79, 823.
Physica B 199+200, 67. Nishida, N., R.S. Hayano, K. Nagarnine, T.
Murani, A.P., K. Knorr, K.H.J. Buschow, A. Yamazaki, J.H. Brewer, D.M. Garner, D.G.
Benoit and J. Flouquet, 1980, Solid State Fleming, T. Takeuchi and Y. Ishikawa, 1977,
Commun. 36, 523. Solid State Commun. 22, 235.
Murasik, A., S. Ligenza and A. Zygmunt, 1974, Nishida, N., K. Nagamine, R.S. Hayano, T.
Phys. Status Solidi A: 23, K163. Yamazaki, D.G. Fleming, R.A. Duncan, J.H.
Murayama, S., C. Sekine, H. Takano, K. Brewer, A. Ahktar and H. Yasuoka, 1978, J.
Hoshi, K. Nishiyama and K. Nagamine, 1993, Phys. Soc. Jpn 44, 1131.
Physica B 186-188, 500. Nishida, N., K. Nagamine, R.S. Hayano,
Mtinch, K.H., A. Kratzer, G.M. Kalvius, L. Y.J. Uemura, J. Imazato, T. Yamazaki, H.
Asch, EJ. Litterst and K. Richter, 1993, Miyajima, S. Chikazumi, D.G. Fleming and
Hyperfine Interact. 78, 435. J.H. Brewer, 1979, Hyperfine Interact. 6, 87.
Nagarnine, K., S. Nagamiya, O. Hashimoto, N. Nishida, N., H. Myatake, D. Shimada, S. Hikami,
Nishida and T. Yamazaki 1976, Hyperfine E. Torikai, K. Nishiyama and K. Nagamine,
Interact. 1, 517. 1987a, Jap. J. Appl. Phys. 26, L799.
Nagamine, K., N. Nishida, S. Nagamiya, O. Nishida, N., H. Miyatake, D. Shimada, S.
Hashimoto and T. Yamazaki, 1977, Phys. Okuma, M. Ishikawa, T. Takabatake, Y.
Rev. Lett. 38, 99. Nakazawa, Y. Kuno, R. Keitel, J.H. Brewer,
Nakamura, H., Y. Kitaoka, K. Asayama and J. T.M. Riseman, D.L1. Williams, Y. Watanabe,
Flouquet, 1988, J. Phys. Soc. Jpn 57, 2644. T. Yamazaki, K. Nishiyama, K. Nagamine, J.
Nakamura, H., Y. Kitaoka, M. Inoue, K. Ansaldo and E. Torikai, 1987b, Jpn. J. Appl.
Asayama and Y. (3nuki, 1990, J. Magn. Phys. 26, L 1856.
Magn. Mater. 90+91, 459. Nishida, N., H. Miyatake, S. Okuma, Y. Kuni,
Nakamura, H., Y. Kitaoka, K. Asayama, Y. Y. Watanabe, T. Yamazaki, S. Hikami, M.
Onuki and M. Shiga 1994, J. Phys.: Condens. Ishikawa, T. Takabatake, Y. Nakazawa, S.R.
Matter, in press. Kreitzman, J.H. Brewer and C.-Y. Huang,
Namkung, M., W.J. Kossler, C.E. Stronach, 1988a, Jpn. J. Appl. Phys. 27, L94.
R.I. Grynszpan and B.D. Patterson, 1984, Nishida, N., H. Miyatake, D. Shimada, S.
Hyperfine Interact. 17-19, 321. Okuma, M. Ishikawa, T. Takabatake, Y.
Narath, A., 1969, Phys. Rev. 136A, 766. Nakazawa, Y. Kuno, R. Keitel, J.H. Brewer,
Niedermayer, Ch., A. Golnik, E. Recknagel, T.M. Riseman, D.L1. Williams, Y. Watanabe,
M. Rossmanith, A. Weidinger, X.S. Chang, T. Yamazaki, K. Nishiyama, K. Nagamine,
MUON SPIN ROTATION SPECTROSCOPY 297

E. Ansaldo and E. Torikai, 1988b, J. Phys. Paul, D.McK., H.A. Mook, A.W. Hewat, B.C.
Soc. Jpn 57, 599. Sales, L.A. Boatner, J.R. Thompson and M.
Nishida, N., H. Miyatake, S. Okuma, T. Tamegai, Mostoller, 1988, Phys. Rev. B 37, 2341.
Y. Iye, R. Yoshizaki, K. Nishiyama and K. Prelovsek, P. and T.M. Rice, 1987, Phys. Rev.
Nagamine, 1988c, Physica C 156, 625. Lett. 59, 1248.
Nishida, N. and H. Miyatake, 1990a, Hyperfine Qu6zel, S., J. Rossat-Mignod, B. Chevalier, P.
Interact. 63, 183 Lejay and J. Etourneau, 1984a, Solid State
Nishida, N., S. Okuma, H. Miyatake, T. Tamegai, Commun. 49, 685.
Y. lye, R. Yoshizaki, K. Nishiyama, K. QuEzel, S., P. Burlet, E. Roudant, J. Rossat-
Nagamine, R. Kadono and J.H. Brewer, Mignod, A. Benoit, J. Flouquet, O. Pefia, R.
1990b, Physica C 168, 23. Horyn, R. Chevrel and M. Sergent, 1984b,
Nishida, N., S. Okuma, S. Shiratake, Y. Ueda, Ann. Chim. Fr. 9, 1057.
A. Hayashi, S. Kambe, H. Yasuoka, K. Rainford, B.D., R. Cywinski, S.H. Kilcoyne and
C.A. Scott, 1994, Hyperfine Interact. 85, 323.
Nishiyama, K. Nagamine and T. Yamazaki,
Ramirez, A.P., B. Batlogg, E. Bucher and A.S.
1991, Physica C 185-189, 1091.
Cooper 1986, Phys. Rev. Lett. 57, 1072.
Nishiyama, K., K. Nagamine, T. Natsui, S. Rath, J., M. Manninen, P. Jena and C. Wang,
Nakajima, K. Ishida, Y. Kuno, J. Imazato, H. 1979, Solid State Commun. 31, 1003.
Nakayama, T. Yamazaki and E. Yagi, 1983, Rebelsky, L., J.M. Tranquadal G. Shirane, Y.
J. Magn. Magn. Mater. 31-34, 695. Nakazawa and M. Iskikawa, 1989, Physica C
Nishiyama, K., E. Yagi, T. Ishida, T. Matsuzaki, 160, 197.
K. Nagamine and T. Yamazaki, 1984, Regnault, L.P., J.L. Jacaud, J.M. Mignod, J.
Hyperfine Interact. 17-19, 473. Rossat-Mignod, C. Vettier, P. Lejay and J.
Nishiyama, K., K. Kojima, K. Nagamine, Flouquet, 1990, Physica 163, 606.
E. Torikai, I. Tanaka, H. Kojima and H. Reilly, J.J., M. Suenaga, J.R. Johnson, P.
Kitazawa, 1993, Hyperfine Interact. 79, 873. Thompson and A.R. Moodenbaugh, 1987,
Noakes, D.R., E.J. Ansaldo, J.H. Brewer, D.R. Phys. Rev. B 36, 5694.
Harshman, C.Y. Huang, M.S. Torikachvili, Renard, J.P., M. Verdaguer, L.P. Regnault,
S. Lambert and M.B. Maple, 1985, J. Appl. W.A.C. Erkelens, J. Rossat-Mignod and W.G.
Phys. 57, 3197. Stirling, 1987, Europhys. Lett. 3, 945.
Noakes, D.R., J.H. Brewer, D.R. Harshman, E.J. Risemann, T.M., J.H. Brewer, E.J. Ansaldo,
Ansaldo and C.Y. Huang, 1987, Phys. Rev. P.M. Grant, M.E. Lopez-Morales and B.M.
B 35, 6597. Sternlieb, 1990, Hyperfine Interact. 63, 249.
Noakes, D.R., E. Fawcett, E.J. Ansaldo, C. Rodriguez-Carvajal, J., J.L. Martfnez, J. Pannetier
Niedermayer and C.E. Stronaeh, 1992, J. and R. Saez-Puche, 1988, Phys. Rev. B 38,
Magn. Magn. Mater. 114, 176. 7148.
Noakes, D.R., E.J. Ansaldo and G.M. Luke, Rodriguez-Carvajal, J., M.T. Fern~dez-Diaz and
1993, J. Appl. Phys. 73, 5666. J.L. Martinez, 1992, J. Phys.: Condens.
Nojiri, H., M. Uchi, S. Watamura, M. Motokawa, Matter 3, 3215.
Roessli, B., P. Fischer, U. Staub, M. Zolliker and
H. Kawai, Y. Endoh and T. Shigeoka, 1991,
A. Furrer, 1993, Europhys. Lett. 23, 511.
J. Phys. Soc. Jpn 60, 2380.
Rossat-Mignod, J., P. Burlet, M.J. Jurgens, L.P.
Nojiri, H., M. Motokawa, K. Nishiyama, K. Regnault, J.Y. Henry, C. Ayache, L. Forro, C.
Nagamine and T. Shigeoka, 1992, J. Magn. Vettier, H. Noel, M. Potel, M. Gougeon and
Magn. Mater. 104--107, 1311. J.C. Lecet, 1988a, J. Phys. (Paris) 49, 2119.
Okuma, S., H. Miyatake, N. Nishida, Y. Ueda, Rossat-Mignod, J., LP. Regnault, J.L. Jacoud,
S. Katsuyama, K. Kosuge and J.H. Brewer, G. Vettier, P. Lejay and J. Flouquet, 1988b, J.
1990, Hyperfine Interact. 63, 265. Magn. Magn. Mater. 76+77, 376.
Ott, H.R., H. Rudigier, E. Felder, Z. Fisk and B. Rossat-Mignod, J., L.P. Regnault, M.J. Jurgens,
Batlogg, 1985, Phys. Rev. Lett. 55, 1595. C. Vettier, P. Burlet, J.Y. Henry and G.
Ott, H.R., H. Rudigier, E. Felder, Z. Fisk and Lapertot, 1990, Physica C 162-164~ 1269.
J.D. Thompson, 1987, Phys. Rev. B 35, Rosseinsky, M.J., K. Prassides and P. Day, 1989,
1452. J. Chem. Soc., Chem. Commun. 22, 1734.
Ott, H.R. et al., 1993, Private communication. Rosseinsky, M.J., K. Prassides and P. Day,
Patterson, B.D., 1988, Rev. Mod. Phys. 60, 69. 1991a, Inorganic Chem. 30, 2681.
298 A. SCHENCK and EN. GYGAX

Rosseinsky, M.J., K. Prassides and C.A. Scott, RM. Richards, D.E. MacLaughlin and C.
1991b, Inorganic Chem. 30, 3367. Boekema, 1984, Hyperfine Interact. 17-19,
Rosseinsky, M.J., K. Prassides and C.A. Scott, 351.
1992, J. Magn. Magn. Mater. 104-107, 599. Schimmele, L., A. Seeger, W. Staiger, W.
Rubinstein, M., G.J. Strauss and J. Dweck, 1966, Templ, C. Baines, A. Fritzsche, M. Hampele,
Phys. Rev. Lett. 17, 1001. D. Herlach, K. Maier and J. Major, 1990,
Riiegg, K., 1981, Doctoral Thesis (University of Hyperfine Interact. 64, 671.
Zurich). Schimmele, L., A. Seeger, T. Stammler, W.
Rtiegg, K., C. Boekema, W. Hofmann, W. Tempi, C. Baines, T. Grund, M. Hampele, D.
Kt~ndig and RE Meier, 1979, Hyperfine Herlach, M. Iwanowski, K. Maier, J. Major
Interact. 6, 99. and T. Pfiz, 1994, Hyperfine Interact. 85,
RUegg, K., C. Boekema, W. Kandig, RE Meier 337.
and B.D. Patterson, 1981, Hyperfine Interact. Seeger, A. and E Monachesi, 1982, Philos. Mag.
8, 547. B 46, 283.
Sakakibara, T., H. Sekine, H. Amitsuka and Y. Seeger, A. and L. Schimmele, 1992, in:
Miyako, 1992, J. Magn. Magn. Mater. 108, Perspectives of Meson Science, eds T.
193. Yamazaki, K. Nakai and K. Nagamine
Sarkissian, B.V.B., 1982, J. Appl. Phys. 53, (North-Holland, Amsterdam) p. 293.
8070. Shapiro, S.M., E. Gurewitz, R.D. Parks and L.C.
Sawa, H., S. Suzuki, M. Watanabe, J. Akimitsu, Kupferberg, 1979, Phys. Rev. Lett. 43, 1748.
H. Matsubara, H. Watabe, S. Uchida, K. Shigeoka, T., N. Iwata, H. Fujii, T. Okamoto and
Kokusho, I-I. Asano, F. Izumi and E. Y. Hashimoto, 1987, J. Magn. Magn. Mater.
Takayama-Muromachi, 1989, Nature 337, 70, 239
347. Shigeoka, T., H. Fujii, K. Yonenobu, K.
Sawatzky, G.A., W. Geertsma and C. Haas,
Sugiyama and M. Date, 1989a, J. Phys. Soc.
1976, J. Magn. Magn. Mater. 3, 37.
Jpn 58, 394.
Schenck, A., 1981, Helvetica Physica Acta 54,
Shigeoka, T., N. Iwata, Y. Hashimoto, Y. Andoh
471.
and H. Fujii, 1989b, Physica B 156-157, 741.
Schenck, A., 1985, Muon Spin Rotation
Spectroscopy (Adam Hilger, Bristol) Shigeoka, T., N. Iwata and H. Fujii, 1992, J.
Schenck, A., 1993, in: Frontiers in Solid State Magn. Magn. Mater. 104--107, 1229.
Sciences, Vol. 2, eds L.C. Gupta and M.S. Shirley, D.A., S.S. Rosenblum and E. Matthias,
Multani (World Scientific, Singapore) p. 269. 1968, Phys. Rev. 170, 363.
Schenck, A., P. Bitter, EN. Gygax, B. Hitti, E. Shull, C.G., W.A. Strauser and E.O. Wollan,
Lippelt, H. Maletta and M. Weber, 1990a, 1951, Phys. Rev. 83, 333.
Hyperfine Interact. 63, 227 Sievers III, A.J. and M. Thinkham, 1963, Phys.
Schenck, A., P. Bitter, Z. Fisk, EN. Gygax, B. Rev. 129, 1566.
Hitti, E. Lippelt, H.R. Ott and M. Weber, Sinha, S.K., G.H. Lander, S.M. Shapiro and O.
1990b, Hyperfine Interact. 64, 511. Vogt, 1981, Phys. Rev. B 23, 4556.
Schenck, A., P. Birrer, F.N. Gygax, B. Hitti, Sigrist, M. and T.M. Rice, 1989, Phys. Rev. B
E. Lippelt, H. Maletta, H.R. Ott and M. 39, 2200.
Weber, 1990c, in: International Seminar Sigrist, M. and K. Ueda, 1991, Rev. Mod. Phys.
on High Temperature Superconductivity, eds 63, 239.
V.L. Aksenov, N.N. Bogolubov and N.M. Slichter, C.P., 1978, Principles of Magnetic
Plakida (World Scientific, Singapore) p. 203. Resonance (Springer, Berlin).
Schenck, A., R Birrer, EN. Gygax, B. Hitti, E. Slyusarev, V.A., Yu.A. Freiman and R.R
Lippelt, M. Weber, P. BOni, P. Fischer, H.R. Yankelevich, 1979, Pis'ma Zh. Eksp. Teor.
Ott and Z. Fisk, 1990d, Phys. Rev. Lett. 65, Fiz. 30, 292; [JETP Lett. 30, 270].
2454. Smilga, V.P. and Yu.A. Belousov, 1994, The
Schenck, A., A. Amato, R Birrer, EN. Gygax, Muon Method in Science (Nova Science,
B. Hitti, E. Lippelt, S. Barth, H.R. Ott and Z. Commack).
Fisk, 1992, J. Magn. Magn. Mater. 108, 97. Smit, H.H.A., M.W. Dirken, R.C. Thiel and L.J.
Schillaci, M.E., R.H. Heffner, R.L. Hutson, M. de Jongh, 1987, Solid State Commun. 64,
Leon, D.W. Cooke, A. Yaouanc, S.A. Dodds, 695.
MUON SPIN ROTATION SPECTROSCOPY 299

Solt, G., W.B. Waeber, E Frisius and U. Properties of Hydrogen in Metals, eds P. Jena
Zimmermann, 1992, Mater. S c i . Forum and C.B. Satterthwaite (Plenum, New York).
97-99, 273. Sulaiman, S.B., N. Sahoo, S. Srinivas, E
Sousa, J.B., 1990, cited in Kalvius et al. (1990). Hagelberg, T.P. Das, E. Torikai and K.
Steglich, E, U. Ahlheim, A. B6hm, C.D. Bredl, Nagamine, 1993, Hyperfine Interact. 79, 901.
R. Caspary, C. Geibel, A. Grauel, R. Heffrich, Takabatake, T., R.M. Mc Callum, M. Kubota
R. KOhler, M. Lang, A. Mehner, R. Modler, and E Pobell, 1984, J. Low Temp. Phys. 55,
C. Schank, C. Wassilew, G. Weber, W.
111.
Assmus, N. Sato and T. Komatsubara, 1991,
Physica C 185-189, 379. Takabatake, T., F. Teshimea, H. Fujii, S.
Steglich, E, U. Ahlheim, C.D. Bredl, G. Geibel, Nishigori, T. Suzuki, T. Fujita, Y. Yamaguchi,
M. Lang, A. Loidl and G. Sparn, 1993, in: J. Sakurai and D. Jaccard, 1990, Phys. Rev.
Frotiers in Solid State Sciences, eds L.C. B 41, 9607.
Gupta and M.S. Multani (World Scientific, Takabatake, T., M. Nagasawa, H. Fujii, G. Kido,
Singapore) Vol 1. M. Nohara, S. Nishigori, T. Suzuki, T. Fujita,
Stephens, P.W. and C.E Majkrzak, 1986, Phys. R. Helfrich, U. Ahlheim, K. Fraas, C. Geibl
Rev. 33, 1. and E Steglich, 1992, Phys. Rev. B 45,
Sternlieb, B.J., G.M. Luke, Y.J. Uemura, J.H. 5740.
Brewer, R. Kadono, J.R. Kempton, R.E Takagi, H., S. Uchida and Y. Tokura, 1989, Phys.
Kiefl, S.R. Kreitzman, T.M. Riseman, D.L1. Rev. Lett. 62, 1197.
Williams, J. Gopalakrishnan, A.W. Sleight, Takagi, S., H. Suzuki, A. Ochiai and T. Suzuki,
A.R. Strzelecki and M.A. Subramanian, 1989, 1992, J. Magn. Magn. Mater. 116, 77.
Phys. Rev. B 40, 11320. Takagi, S., H. Suzuki, A. Ochiai, T. Suzuki, A.
Sternlieb, B.J., G.M. Luke, Y.J. Uemura, T.M. Amato, R. Feyerherm, F.N. Gygax and A.
Riseman, J.H. Brewer, P.M. Gehring, K. Schenck, 1993, Physica B 186-188, 422.
Yamada, Y. Hidaka, T. Murakami, T.R.
Takahashi, M., P. Turek, Y. Nakazawa, M.
Thurston and R.J. Birgeneau, 1990, Phys.
Tamura, K. Nozawa, D. Shiomi, M. Ishikawa
Rev. B 41, 8866.
Sternlieb, B.J., L.P. Le, G.M. Luke, W.D. Wu, and M. Kinoshita, 1991, Phys. Rev. Lett. 67,
Y.J. Uemura, T.M. Riseman, J.H. Brewer, Y. 746.
Ajiro and M. Mekata, 1992, J. Magn. Magn. Takahashi, T., Y. Maniwa, H. Kawamura and G.
Mater. 104-107, 801. Saito, 1986, J. Phys. Soc. Jpn 55, 1364;
Storchak, V., B.E Kirillov, A.V. Pirogov, V.N. Physica B 143, 417.
Duginov, G. Grebinnik, A.B. Lazarev, V.G. Takahashi, T., H. Kawamura, T. Ohyama, Y.
Ol'shevsky, V.Yu. Pomjakushin, S.N. Shilov Maniwa, K. Murata and G. Saito, 1989, J.
and V.A. Zhukov, 1992, Phys. Lett. A 166, Phys. Soc. Jpn 58, 703.
429. Takahashi, Y. and T. Moriya, 1978, J. Phys. Soc.
Storchak, V., B.E Kirillov, A.V. Pirogov, V.N. Jpn 44, 850.
Duginov, G. Grebinnik, V.G. Ol'shevsky, Takigawa, M., H. Yasuoka, Y.J. Uemura, R.S.
A.B. Lazarev, V.Yu. Pomjakushin, S.N. Hayano, T. Yamazaki and Y. Ishikawa, 1980,
Shilov and V.A. Zhukov, 1994, Hyperfine J. Phys. Soc. Jpn 49, 1760.
Interact. 85, 345.
Takigawa, M., H. Yasuoka, T. Tanaka and Y.
Stronach, C.E., W.J. Kossler, J. Lindemuth, K.G.
Ishizawa, 1983a, J. Phys. Soc. Jpn 52, 728.
Petzinger, A.T. Fiory, R.E Minnich, W.E
Lankford, J.J. Singh and K.G. Lynn, 1979, Takigawa, M., H. Yasuoka, Y. Yamaguchi and S.
Phys. Rev. B 20, 2315. Ogawa, 1983b, J. Phys. Soc. Jpn 52, 3318.
Stronach, C.E., K.R. Squire, A.S. Arrott, B.D. Takita, K., H. Akinaga, K. Masuda, H. Asano,
Patterson, B. Heinrich, W.E Lankford, A.T. Y. Takeda, M. Takuno, K. Nishiyama and K.
Fiory, W.J. Kossler and J.J. Singh, 1981, J. Nagamine, 1989, in: Proc. Tsukuba Seminar
Magn. Magn. Mater. 25, 187. High Tc Superconductivity, Tsukuba, p. 11.
Stronach, C.E., K.R. Squire, A.S. Arrott, B. Tamegai, T., A. Watanabe, K. Koga, I. Oguro
Heinrich, W.E Lankford, W.J. Kossler and and Y. Iye, 1988, Jpn. J. Appl. Phys. 27,
J.J. Singh, 1983, in: Electronic Structure and L1074.
300 A. SCHENCK and EN. GYGAX

Tamura, M., Y. Nakazawa, D. Shiomi, K. Uemura, Y.J., T. Yamazaki, Y. Kitaoka, M.


Nozawa, Y. Hosokoshi, M. Ishikawa, M. Takigawa and H. Yasuoka, 1984, Hyperfine
Takahashi and M. Kinoshita, 1991, Chem. Interact. 17-19, 339.
Phys. Lett. 186, 401. Uemura, Y.J., T. Yamazaki, D.R. Harshman, M.
Telling, M.T.F., S.H. Kilcoyne and R. Cywinski, Senba and E. Ansaldo, 1985, Phys. Rev. B
1994, Hyperfine Interact. 85, 209. 31, 546.
Templ, W., M. Hampele, D. Herlach, J. Major, Uemura, Y.J., W.J. Kossler, B. Hitti, J.R.
J. Mundy, A. Seeger and W. Staiger, 1990, Kempton, H.E. Schone, X.H. Yu, C.E.
Hyperfine Interact. 64, 679. Stronach, W.F. Lankford, D.R. Noakes, R.
Thompson, J.D., A.C. Lawson, M.W. McElfresh, Keitel, M. Senba, J.H. Brewer, E.J. Ansaldo,
A.R Sattelberger and Z. Fisk, 1988, J. Magn. Y. Onuki, T. Komatsubara, G. Aeppli, E.
Magn. Mater. 76+77, 437. Bucher and J.E. Crow, 1986a, Hyperfine
Thompson, J.D., P.C. Canfield, A. Lacerda, Interact. 31, 413.
M.F. Hundley, Z. Fisk, H.R. Ott, E. Felder, Uemura, Y.J., R. Keitel, M. Senba, R.E
M. Chernikov, M.B. Maple, R Visani, C.L. Kiett, S.R. Kreitzman, D.R. Noakes, J.H.
Seaman, M.A. Lopez d e l l a Torre and G. Brewer, D.R. Harshman, EJ. Ansaldo, K.M.
Aeppli, 1993, Physica B 186-188, 355. Crowe, A.M. Portis and V. Jaccarino, 1986b,
Tomala, K., G. Czjzek, J. Fink and H. Schmidt, Hyperfine Interact. 31, 313.
1977, Solid State Commun. 24, 857. Uemura, Y.J. and R.J. Birgeneau, 1986c, Phys.
Torikai, E., I. Tanaka, H. Kojima, H. Kitazawa Rev. Lett. 57, 1947.
and K. Nagamine, 1990, Hyperfine Interact. Uemura, Y.J., W.J. Kossler, X.H. Yu, J.R.
63, 271. Kempton, H.E. Schone, D. Opie, C.E.
Torikai, E., K. Nagamine, H. Kitazawa, I. Stronach, D.C. Johnston, M.S. Alvarez and
D.P. Goshorn, 1987, Phys. Rev. Lett. 59,
Tanaka, H. Kojima, S.B. Sulaiman, S.
1045.
Srinivas and T.P. Das, 1993a, Hyperfine
Uemura, Y.J., W.J. Kossler, J.R. Kempton, X.H.
Interact. 79, 921.
Yu, H.E. Schone, D. Opie, C.E. Stronach,
Torikai, E., H. Ishihara, K. Nagamine, H.
J.H. Brewer, R.E Kiefl, S.R. Kreitzman, G.M.
Kitazawa, I. Tanaka, H. Kojima, S.B.
Luke, T. Riseman, D.L1. Williams, EJ.
Suleiman, S. Srinivas and T.P. Das, 1993b,
Ansaldo, Y. Endoh, E. Kudo, K. Yamada,
Hyperfine Interact. 79, 915.
D.C. Johnston, M. Alvarez, D.R Goshorn, Y.
Torikai, E., K. Nagamine, H. Kitazawa, I. Tanaka
Hidaka, M. Oda, Y. Enomoto, M. Suzuki and
and H. Kojima 1993c, Hyperfine Interact. 79, T. Murakami, 1988, Physica 153-155, 769.
909. Uemura, Y.J., W.J. Kossler, X.H. Yu, H.E.
Torikai, E., K. Nishiyama, K. Kojima, K. Schone, J.R. Kempton, C.E. Stronach, S.
Nagamine, H. Kitazawa, I. Tanaka and H. Barth, F.N. Gygax, B. Hitti, A. Schenck,
Kojima, 1993d, Hyperfine Interact. 79, 879. C. Baines, W.F. Lankford, Y. Onuki and T.
Torrance, J.B., R Lacorre, A.I. Nazzal, E.J. Kamatsubara, 1989, Phys. Rev. B 39, 4726.
Ansaldo and Ch. Niedermayer, 1992, Phys. Uemura, Y.J. and G.M. Luke 1993, Physica B
Rev. B 45, 8209. 186--188, 223.
Tranquada, J.M., D.E. Cox, W. Kaunmann, Uemura, Y.J., A. Keren, L.P. Le, G.M. Luke,
H. Moudden, G. Shirane, M. Suenaga, B.J. Sternlieb and W.D. Wu, 1994, Hyperfine
P. Zolliker, D. Vaknin, S.K. Sinha, M.S. Interact. 85, 133.
Alvarez, A.J. Jacobson and D.C. Johnson, Vaknin, D., S.K. Sinha, C. Stassis, L.L. Miller
1988, Phys. Rev. Lett. 60, 156. and D.C. Johnston, 1990, Phys. Rev. B 41,
Uemura, Y.J., T. Yamazaki, R.S. Hayano, R. 1926.
Nakai and C.Y. Huang, 1980, Phys. Rev. Wackelg&d, E., O. Hartmann, E. Karlsson, R.
Lett. 45, 583. W~ippling,L. Asch, G.M. Kalvius, J. Chappert
Uemura, Y.J., N. Nishida, J. Imazato, R.S. and A. Yaouanc, 1986, Hyperfine Interact.
Hayano, M. Takigawa and T. Yamazaki, 31, 325.
1981a, Hyperfine Interact. 8, 771. Wfickelg~d, E., O. Hartmann, E. Karlsson, R.
Uemura, Y.J., J. Imazato, N. Nishida, R.S. Wappling, L. Asch, G.M. Kalvius, J. Chappert
Hayano, M. Takigawa and T. Yamazaki, and A. Yaouanc, 1989, Hyperfine Interact.
1981b, Hyperfine Interact. 8, 725. 50, 781.
MUON SPIN ROTATION SPECTROSCOPY 301

Wada, N., S. Ohsawa, Y. Nakamura and K. Wiesinger, G., E. Bauer, Th. H~iufler,A. Amato,
Kumagai, 1990, Physica B 165+166, 1345. R. Feyerherm, EN. Gygax and A. Schenck,
Walker, M.B., W.J.L. Buyers, Z. Tun, W. Que, 1994b, Preprint (SCES'94, Amsterdam), to
A.A. Menovsky and J.D. Garrett, 1993, Phys. be published in Physica B.
Rev. Lett. 71, 2630. Wong, W.H. and W.G. Clark, 1992, J. Magn.
Walstedt, R.E. and L.R. Walker, 1974, Phys. Magn. Mater. 108, 175.
Rev. 89, 4857. Wu, W.D., A. Keren, G.M. Luke, Y.J. Uemura,
W~ippling, R., O. Hartmann, S. Harris, E. C.L. Seaman, Y. Dalichaouch and M.B.
Karlsson, G.M. Kalvius, L. Asch, A. Kratzer, Maple, 1993, Physica B 186-188, 344.
P. Dalmas de R6otier and A. Yaouanc, 1993, Yamada, K, E. Kudo, Y. Endoh, Y. Hidaka,
J. Magn. Magn. Mater. 119, 123. M. Oda, M. Suzuki and T. Murakawi, 1987,
Watanabe, I., K. Nishiyama, K. Nagamine, K. Solid State Commun. 64, 753.
Kawano and K. Kumagai, 1994, Hyperfine Yamada, Y., Y. Kitaoka, K. Asayama and A.
Sakata, 1984, J. Phys. Soc. Jpn 53, 3198;
Interact. 86, 603.
3634.
Weber, M., 1991, Dissertation ETHZ Ztirich,
Yamada, Y., H. Nakamura, Y. Kitaoka, K.
No. 9681 (unpublished).
Asayama, K. Koga, A. Sakata and T.
Weber, M., P. Birrer, EN. Gygax, B. Hitti, E. Murakami, 1990, J. Phys. Soc. Jpn 59, 2976.
Lippelt, H. Maletta and A. Schenck, 1990, Yamagata, H. and M. Matsumura, 1983, J. Magn.
Hyperfine Interact. 63, 207 Magn. Mater. 31-34, 65.
Weber, M., L. Asch, A. Kratzer, G.M. Kalvius, Yamazaki, T., 1979, Hypertine Interact. 6, 115.
K.-H. MUnch, R. Ballou, J. Deportes, Yamazaki, T., 1981, Hyperfine Interact. 8, 463.
R. Wappling, EJ. Litterst, H.-H. Klauss, Yamazaki, T., R.S. Hayano, Y. Kuno, J. Imazato,
Ch. Niedermayer and J. Chappert, 1994, K. Nagamine, S.E. Kohn and C.Y. Huang,
Hyperfine Interact. 85, 265. 1979, Phys. Rev. Lett. 42, 1241.
Wehr, H., K. Knoll, EN. Gygax, A. Hintermann, Yang, B.X., R.F. Kiefl, J.H. Brewer, J.E Carolan,
A. Schenck and W. Studer, 1983, J. Phys. F: W.N. Hardy, R. Kadono, J.R. Kempton,
Met. Phys. 13, 885. S.R. Kreitzman, G.M. Luke, T.M. Riseman,
Weidinger, A., G. Balzer, H. Graf, T. MOslang, D.L1. Williams, Y.J, Uemura, B. Sternlieb,
E. Recknagel, T. Wichert, J. Bigot and R.I. M.A. Subramanian, A.R. Strzelecki, J.
Grynszpan, 1981, Hyperfine Interact. 8, 543. Gopalakrishnan and A.W. Sleight, 1989, Phys.
Weidinger, A., J.I. Budnick, B. Chamberland, A. Rev. B 39, 847.
Golnik, Ch. Niedermayer, E. Recknagel, M. Yaouanc, A.J., J.I. Budnick, E. Albert, M.
Rossmanith and D.P. Yang, 1988, Physica C Hamrna, A. Weidinger, R. Fruchart, Ph.
153-155, 168. L'Heritier, D. Fruchart and P. Wolfers, 1987,
Weidinger, A., Ch. Niedermayer, A. Golnik, J. Magn. Magn. Mater. 67, L286.
R. Simon, E. Recknagel, J.I. Budnick, B. Yaouanc, A., P. Dalmas de R6otier, B. Chevalier
Chamberland and C. Baines, 1989, Phys. and Ph.L. L'H6ritier, 1990, J. Magn. Magn.
Rev. Lett. 62, 102; see also comments by Mater. 90+91, 575.
Yaouanc, A. and P. Dalmas de R6otier, 1991, J.
Harshman D.R. et al., 1989, Phys. Rev. Lett.
Phys. C 3, 6•95.
63, 1187 and by Heffner R.H. and D.L. Cox,
Yaouanc, A., P. Dalmas de R6otier and E. Frey,
1989, Phys. Rev. Lett. 63, 2538 and replies
1993a Phys. Rev. B 47, 796.
by Weidinger et al., ibid. Yaouanc, A., P. Dalmas de R6otier and E. Frey,
Weidinger, A., Cb. Niedermayer, A. Gltlckler, 1993b, Europhys. Lett. 21, 93.
G. Novitzke, E. Recknagel, H. Eikenbusch, Yaouanc, A., P. Dalmas de R6otier, P.C.M.
W. Paulus, R. Sch0llhorn and J.I. Budnick, Gubbens, A.A. Moolenaar, A.A. Menovsky
1990, Hyperfine Interact. 63, 147. and C.E. Snel, 1994, Hyperfine Interact. 85,
Weidinger, A., J. Erxmeyer, H. Gltickler, Ch. 351.
Niedermayer, O. Laforsch, J. Gross and M. Yasuoka, H., V. Jaccarino, R.C. Sherwood and
Mehring, 1994, Hyperfine Interact. 86, 609. J.H. Wernick, 1978a, J. Phys. Soc. Jpn 44,
White, R.L., 1969, J. Appl. Phys. 40, 1061. 842.
Wiesinger, G., E. Bauer, A. Amato, R. Yasuoka, H., R.S. Hayano, N. Nishida, K.
Feyerberm, EN. Gygax and A. Schenck, Nagamine and T. Yamazaki, 1978b, Solid
1994a, Physica B 199+200, 52. State Commun. 26, 745.
302 A. SCHENCK and EN. GYGAX

Yushankhai, V.Yu., 1989, Hyperfine Interact. 50, Zhou, L.W., C.L. Lin, J.E. Crow, S. Bloom,
775 R.P. Guertim, S. Foner and G. Stewert, 1985,
Zaanen, J., G.A. Sawatzky and J.W. Allen, 1985, Physica B 135, 99.
Phys. Rev. Lett. 55, 418. Zwirner, S., J.C. Spirlet, K.H. Miinch, A.
Zaanen, J. and G.A. Sawatzky, 1990, J. Solid Kratzer, L. Asch and G.M. Kalvius, 1993,
State Chem. 88, 8. Physica B 186--188, 798.
chapter 3

INTERSTITIALLY MODIFIED
INTERMETALLICS OF RARE EARTH
AND 3D ELEMENTS

HIRONOBU FUJII
Faculty of Integrated Arts and Sciences
Hiroshima University
Higashi-Hiroshima 739
Japan

and

HONG SUN
Materials Science Research Department
Research and Development Center
Sumitomo Metal Industries, Amagasaki 660
Japan

Handbook of Magnetic Materials, Vol. 9


Edited by K. H.J. Buschow
01995 Elsevier Science B.V. All rights reserved

303
CONTENTS

1. Introduction ................................................................. 305


2. Formation of the interstitially modified intermetallic compounds . . . . . . . . . . . . . . . . . . . . . . . 307
2.1. Arc-melting method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
2.2. Melt-spinning method ..................................................... 309
2.3. Gas-phase interstitial modification method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
2.4. Solid-solid (or liquid) reaction and plasma nitriding methods ..................... 314
2.5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
3. Interstitial compounds of the 2:17-type structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
3.1. Crystallographic structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
3.2. Curie temperature and exchange interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 328
3.3. Magnetic anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
3.4. M6ssbauer and NMR studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
3.5. Substitution effect ........................................................ 349
4. Interstitial compounds of the l:12-type structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
4.1. Location of N atoms in the tetragonal structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 354
4.2. Structural and intrinsic magnetic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 356
4.3. Substitution studies ....................................................... 364
4.4. Interstitial modification study on compounds with various structures . . . . . . . . . . . . . . . . 364
5. Electronic band structure calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
5.1. Calculations of the 2:17-type interstitial compounds R2Fel7Z3 ( Z = N or C) . . . . . . . . . . 367
5.2. Calculations of the l:12-type interstitial compounds RFe12_~TzZ u ( Z = N or C) . . . . . . 375
5.3. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380
6. Applications ................................................................. 382
6.1. Improvement of the thermal stability ......................................... 382
6.2. Development of permanent magnets ......................................... 383
7. Prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 396
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397

304
1. Introduction

One of the methods for fabricating new magnetic materials is to interstitially in-
troduce non-metallic atoms like H, B, C or N with small atomic-radius into host
metals or compounds. As interstitially modified compounds, hydrides have been
well known since 1960s and the magnetism has been covered in chapter 6 of volume
VI in this series (Wiesinger and Hilscher 1991). Generally speaking, no significant
change in the magnetism was observed upon hydrogenation. On the other hand, the
possibility of dramatic improvement of magnetic properties was pointed out in the
nitride FeI6N2 twenty years ago, which had been accidently found in the Fe-N thin
films formed by evaporating Fe in a Nz-gas atmosphere (Kim and Takahashi 1972).
This compound is regarded to be one of the interstitially modified compounds. The
deduced Fe moment was reported to be 2.9#B, named 'giant magnetic moment'.
Metallic iron crystallizes in the body centered cubic structure below 912°C and
has an atomic moment of 2.2#u at 0 K, which is smaller than 3#B/Fe-atom expected
for a localized d-state in an Fe atom with seven 3d-electrons. This moment reduction
is due to the hybridization of the 3d-3d electron states or the overlap between the
3d-electron wave functions. If the hybridization or overlap is reduced or removed
by lattice expansion due to interstitial modification, we can expect an increase of the
moment per Fe atom. Since a confirmation of the 'giant moment' (~3.0/zB/Fe-atom)
was made by Sugita et al. (1991) in a single crystalline Fex6N2 thin firm, much
attention has been paid to developing Fel6N2 in bulk form, which is technically
promising material for the applications as magnetic medium in high performance
metal recording tapes or as soft magnetic materials. However, data of the magnetic
properties of Fel6N2 phase widely scatters, and there are still controversies as to
whether the giant moment exists in the Fel6N2 compound or not.
Under these circumstances, it is very interesting to apply the interstitial modifica-
tion technique to rare earth (R) transition metal (TM) intermetallic compounds for
searching new promising magnetic materials. Before introducing the magnetism on
interstitially modified R-TM intermetallic compounds which have been discovered in
recent years, we will briefly trace the historical background of rare earth permanent
magnets.
Until now, the intermetallics composed of rare earth and 3d elements have been
mainly developed as high-performance permanent magnets. In the 1960s, hexago-
nal SmCo5 compound with the CaCu5-type structure appeared on the stage as the

305
306 H. FUJII and H. SUN

first rare earth high-performance magnet. The compound has quite favorable mag-
netic properties as permanent magnets, especially, (1) the extremely large uniaxial-
magnetocrystalline anisotropy (#0Ha "~ 28 T) originating from the single-ion aniso-
tropy of the Sm sublattice, (2) the relatively large saturation magnetization due to
ferromagnetic coupling between Sm and Co moments (Ms '-~ 1.14 T) and (3) the high
Curie temperature (To = 1000 K) (Strnat 1967). The development of the liquid-phase
sintering technique made fully dense and stable SmCo5 magnet possible (Das 1969,
Benz and Martin 1970). In alloys of Sm with Co, the saturation magnetization of
Sm2Co17 (Ms N 1.25 T) is larger than that of SmCo5 and the Curie temperature
(Tc = 1193 K) is also higher. Although the uniaxial anisotropy field of Sm2Co17
is only 7 T which is smaller than that of SmCos, Sm2Co17 has the possibility of
leading to better permanent magnets than SmCo5 magnet. As a result of much ef-
fort, a nice combination between the high saturation magnetization of Sm2Co17 and
the high magnetic hardness of SmCo5 was realized by controlling the kinetics of
the precipitation reaction in a system with the approximate composition SmCo7.4, in
which some of the Co was replaced by Fe and small amounts of Cu and Zr (Ojima
et al. 1977, Yoneyama et al. 1978). The highest energy product which has been
achieved is more than 240 kJ/m 3 (Mishra et al. 1981).
As the main components Sm and Co are particularly expensive, it is desirable to
use iron-based compounds in place of cobalt-based compounds. Unfortunately, the
rare earth iron compounds with the CaCus-type structure do not exist. The R2Fe~7-
type compounds which are isomorphous to R2Co17 exist, but the Curie temperature
is too low to be used as permanent magnets and the magnetic anisotropy at room
temperature is not uniaxial which is not suitable for permanent magnets.
In the 1980s, scientific research on rare earth iron compounds containing a small
amount of non-metallic elements had been done quite intensively. As a result of con-
tinuing efforts, a novel type of permanent magnet material was discovered in 1984,
which was based on the ternary compound Nd2Fel4B with the tetragonal structure
(Sagawa et al. 1984, Croat et al. 1984). The magnetic characteristics of Nd2Fe14B
are Ms = 1.60 T, #0Ha = 7 T at room temperature and Tc = 588 K. The achieved
energy product was 360 kJ/m 3, which significantly exceeded all previous values. Be-
cause of its lower cost and superior properties, the Nd-Fe-B magnets have rapidly
replaced the Sm-Co magnets and the spectrum of applications has continuously ex-
panded. The main problem of the Nd-Fe-B magnets is the poor temperature stability
due to the relatively low Tc which leads to a practical temperature limit of 150°C.
Also their corrosion resistance is weak. Thus it is still necessary looking for new
magnetic materials with better thermal and magnetic properties. Here, it should be
noted that B is not an interstitial atom in the tetragonal structure and NdzFel4B does
not belong to the interstitially modified compounds.
In 1990, the interstitially modified compound Sm2Fea7N3 was discovered by ap-
plying the gas-phase interstitial modification technique to Sm2Fel7 (Coey and Sun
1990). This interstitial nitride was prepared by heating SmzFea7 at 450-500°C un-
der a nitrogen or ammonia gas atmosphere. The crystal lattice of Sm2Fe17 expands
more than 6% to accommodate three nitrogen atoms at the interstitial sites. The
Curie temperature Tc increases dramatically from 398 K to 752 K. The saturation
INTERSTITIALLYMODIFIED INTERMETALLICS 307

magnetization of Sm2Fel7N3 (Ms = 1.54 T) is comparable to that of NdzFel4B and


the uniaxial magnetic anisotropy (#0Ha = 26 T) is three times as strong as that
of NdzFel4B. By applying a similar method, the nitrides of Nd(Fel_zTz)lzN with
T = Ti, V and Mo, which crystallize in the tetragonal ThMnlz-type structure, were
also successfully formed (Yang Y.C. et al. 1991, Anagnostou et al. 1991a, Wang
and Hadjipanayis 1991). The Curie temperature and uniaxial anisotropy field of
these Nd-containing 1:12 nitrides are comparable with those of NdzFel4B, but the
magnetizations are somewhat lower.
Similarly to the R-TM intermetallics with the 2:17-type and l:12-type structures,
many other structure type compounds can also be interstitially modified, and the
structural and magnetic properties can be drastically altered. Hydrocarbon gas can
be substituted for nitrogen to make interstitial carbides. Therefore interstitial mod-
ification has opened a wide field for scientific research besides the generation of
great technical interest in the application of SmzFe17 interstitial nitride or carbide as
permanent magnets.
Since the first scientific publication of the discovery of the 2:17-type nitrides in
early 1990 (Coey and Sun 1990), world wide efforts have been made and are still
being devoted to the study of interstitial modifications of various kinds of R-TM
intermetallic compounds. The study covers not only every aspect of the structural and
intrinsic magnetic properties, but also the development of the promising interstitial
compounds into hard magnets. In addition, much attention has been paid to the
study of the gas-phase interstitial modification (GIM) process itself. In this chapter,
we will review the studies of the interstitially modified intermetallic compounds
in the following order: the study of the GIM process; the effects of interstitial
modifications on the structural and magnetic properties of the 2:17-type compounds,
l:12-type compounds and also compounds of other structure types; the electronic
band structure calculations of the interstitial compounds and finally the research
aimed at applications of the interstitial compounds as permanent magnets. The
emphasis is on the 2:17-type and 1:12-type interstitial nitrides and carbides.

2. Formation of the interstitially modified intermetallic compounds

Hydrogen, nitrogen and carbon are the well known elements that have been success-
fully used for interstitial modifications. The interstitial hydrides have been known
for a fairly long time (see the review of Wiesinger and Hilscher 1991). In the
field of magnetism, besides the improvement of magnetic properties (mainly the
Curie temperature) by hydrogenation, the HDDR (Hydrogenation-Decomposition-
Desorption-Recombination) process has been used as a novel technique for the pro-
duction of permanent magnets. Another field where the hydrides are of great interest
is their applications as hydrogen storage and hydrogen purification materials, and
furthermore as hydrogen batteries. The interstitial carbides and nitrides were dis-
covered and developed in recent years. C and N were found to have much stronger
effects than H on the magnetic properties. While the hydrides and nitrides can only
be produced by the Gas-phase Interstitial Modification methods (GIM), the carbides
308 H. FUJII and H. SUN

can also be made by arc-melting and melt-spinning in addition to the GIM process.
In this section, various kinds of processing methods for producing the interstitially
modified materials will be described. The emphasis is on the GIM method.

2.1. Arc-melting method


Detailed investigations on the conditions under which the tetragonal R2Fel4C phase
is formed have shown that the corresponding ternary systems contain another new
phase with the composition RzFea7Cy (Liu N.C. et al. 1987, Gueramian et al.
1987, de Mooij and Buschow 1988), which is the interstitial modified 2:17 phase.
As shown in fig. 2.1, the solid line defines temperature regions below which the
tetragonal R2Fe14C phase is stable and above which the hexagonal or rhombohedral
R 2 F e l 7 C v phase is stable.
Because of the high melting point of carbon, Fe3C is made first by arc-melting.
Then it is mixed together with appropriate amount of R and Fe and melted again to
form the as-melted ingot. High temperature and long time (above 1000°C for a few
weeks) annealing is necessary for forming the R2Fe17C u phase. In order to avoid the
phase transition into R2Fel4C during cooling (fig. 2.1), samples have to be quenched
into water from the annealing temperature. The carbon content y is continuously,
variable, but the maximum amount is less than 1.6 for heavy rare earths and less
than 1.0 for light rare earths. Both the maximum contents are much less than 3,
which is the available interstitial sites in the hexagonal or rhombohedral structures
(Sun et al. 1990a, Coene et al. 1990, Zhong et al. 1990).
The substitution of Ga for Fe can stabilize the 2:17 structure with higher in-
terstitial carbon content. Carbon concentration as high as 2.5 was achieved in
R2Fe17_zGazCy(z = 2 and 3, 0 ~< y ~< 2.5) by arc-melting (Shen et al. 1993,

I I I I I I ! I I F

12oo

R2Fe17Cy~ ~
v
t--- lOOO
R2Fe14 C

800
I I I I I I I I [ I
LaCe Pr NdPmSmEuGd.TbDyHoEr
Fig. 2.1. Transformation temperature Tt of the R-Fe-C systems. The R2FelTCv phase is stable above
the solid line and the RzFel4C phase is stable below the solid line. The broken line indicates the
temperature range in which the reaction rate is too low for formation of the R2Fel4C phase from the
non-equilibrium phases of the as-cast melt. Replotted from de Mooij and Buschow (1988).
INTERSTITIALLYMODIFIEDINTERMETALLICS 309
I I

#_
Sm2Fe1702.0

r"

Sm2Fe14Ga3C2.0 o 1~

[ 1
30 40 50 60
20 ( degree )
Fig. 2.2. X-raydiffractionpatterns of Sm2FelTC2.0and Sm2FeI4Ga3C2.0. The compoundswereprepared
by arc-meltingand annealedin vacuumat 1270 K for 12 hours (after Shen et al. 1993).

1993b). Shen et al. also reported similar stabilization effects by the substitutions of
Si or A1 for Fe. Annealing at temperatures higher than 1000°C is still necessary, but
the time period requested is reduced to less than 24 hours. The effect of Ga can be
clearly seen in the X-ray diffraction patterns of Sm2FelTCz0 and SmzFe14Ga3C2.0 in
fig. 2.2. While the later compound is of nearly pure 2:17 phase with the ThzZn17
rhombohedral structure, the former one is a mixture of a-Fe and probably the 2:17
phase.

2.2. Melt-spinning method


Interstitial compounds R2Fe17Cv with y = 0-2.8 can be prepared by arc-melting
and subsequent melt-spinning (Shen et al. 1992, Kong et al. 1992, 1993a, b,
Cat et al. 1992a, b, 1993a, b). The alloys with the corresponding stoichiometric
compositions were first arc-melted, and then by using a suitable quenching rate (10-
20 m/s for Y and Tb), almost pure 2:17 phases were obtained. The maximum carbon
content y increases as the rare earth atomic radius decreases (or as the atomic number
increases). Only the preparation of carbides containing heavy rare earths has been
reported.
Common features of the arc-melted and melt-spun interstitial carbides are their
high temperature stability (contrary to the metastability of the gas-phase intersti-
tially modified interstitial materials) and the structure transition from the hexagonal
Th2Ni17-type structure to the rhombohedral Th2ZnlT-type structure with the increas-
ing carbon content for heavy rare earth compounds. For interstitial compounds made
by the GIM process, no such structure transformation has been reported. These fea-
tures will be discussed later.
310 H. FUJII and H. SUN

2.3. Gas-phase interstitial modification method


2.3.1. Gas-phase interstitial modification process
It has been known for long that rare earth-iron intermetallics can absorb prodigious
quantifies of hydrogen, and that their magnetic properties are altered (Wallace 1978,
Buschow et al. 1982). This idea was extended to nitrogen and carbon absorptions
by using gasses containing N or C (Coey and Sun 1990, Coey et al. 1991a, b, Sun
et al. 1990b and 1992). The gas absorption characteristics were studied by the latter
authors using a thermopiezic analyzer (TPA), which involved the measurement of the
gas pressure variation as a function of temperature. The basic principle is that when
a sample which absorbs or desorbs gas is heated in a closed volume at a constant
rate, the pressure-temperature curve p(T) and its derivative dp(T)/dT provide a
fingerprint of the material, and allow the amount of gas-envolving or gas-absorbing
phase to be determined. The TPA traces of Y2Fel7 in H2, N2, NH3 and Chill0 are
shown in fig. 2.3. Y2Fel7 acts as a catalyst and it induces breaking of the strong
N -= N bond even at temperature as low as 400°C. When making carbides using this
method, normally a hydrocarbon gas is used. At the same time of the formation of
the carbides, hydrogen gas also forms, which has to be pumped out at temperatures
around 500°C in order to avoid the formation of hydrides during cooling to room
temperature. As can be seen from the TPA trace of Y2Fe]7 in H2 shown in fig. 2.3,
the H2 gas pressure at about 500°C is roughly the same as at room temperature,
which indicates that nearly no hydrogen remains in the compound. Therefore, the
hydrogen content in the carbides is negligible.
The gas-phase interstitial modification process can be carried out using fixed
or flow gas atmospheres (nitrogen or ammonia for nitrogenation, and hydrocarbon
gasses for carbonation) in conventional furnaces. Sometimes a certain fraction of
hydrogen gas is added in order to activate nitrogenation or carbonation by hydrogena-
tion which occurs at lower temperatures (fig. 2.3). Liu J.P. et al. (1991) interpreted

' ' J ' ' I ' ' I ' ' I i , , i , , , i , , , i , , , i |


130

, 120 H2
E

e
11o
100 \
o. 80

70 ,,,I,r,;

140

r /
F / -"°
a.
,oor; T; I , , , i , , , I ~ , J , , I , r , I , , i 1--100
200 400 600 800 200 400 600 800
Temperature ('C) Temperature ('c)

Fig. 2.3. Thermopiezicanalysisof Y2Fe17 heatedin H2, NH3, N2 and C4H10at 10 K/min.
INTERSTITIALLYMODIFIEDINTERMETALLICS 311

the nitrogenation reaction process using the same model as for the ternary hydrides
by considering the reaction enthalpy (AHf) of the following two transformations,

2R2Fe17 + 3N2 --+ 2R2Fe17N3, (2.1)

and

R2Fe17 + N2 -+ 2RN + 17Fe. (2.2)

The reaction enthalpy AH~ of the interstitial nitride formation process is sufficiently
negative, which makes the formation of the 2:17 nitrides possible, once the tempera-
ture has been raised high enough to overcome the activation energy for the absorption
process. Although the enthalpy change AH~ of the decomposition process is more
negative, the activation energy is also much higher. Thus the nitride phase can be
formed in a limited temperature range as a metastable phase. If the nitrogenation
process is performed at too high temperatures or for too long times, degradation
takes place and the harmful impurity phases appear in the nitrides. By carefully
controlling the nitrogenation condition, the decomposition process can be depressed,
but it can not be completely eliminated.
Many studies have been devoted to the preparation of high quality nitrides. High
pressure nitrogenation, pre-hydrogenation treatment etc. have been found to be help-
ful. Fujii et al. (1992a, b) have claimed that one of the reasons for the decomposition
of the nitride phase into SmN and ~-Fe is that the temperature at the surface of the
Sm2Fe17 particles increases as a result of the heat formation during the reaction
process and might exceed the decomposition temperature. Thus, the relative bet-
ter thermal conductivity of the high pressure (up to 100 atm.) N2 gas could help
to reduce the particle surface temperature and depress the decomposition process.
Fukuno et al. (1991, 1992a) reported that hydrogen treatment before nitrogenation
significantly increased the gas-solid reaction area by inducing many cracks in the
Sm2Fe17 particles. This hydrogenation process thus promoted nitrogenation at a
lower temperature and the formation of impurity phases could be minimized. Lower
temperature and longer nitrogenation time is preferred in order to obtain high quality
nitrides.

2.3.2. Reaction mechanism


The nitrogenation process has been observed and studied by many different exper-
imental techniques, which include the Kerr microscopy observations by Mukai and
Fujimoto (1992), in situ neutron powder diffraction studies by Isnard et al. (1992a),
micrograph studies applying metallography and EPMA techniques by Colucci et al.
(1992, 1993a, b) and Fujii et al. (1994b). The nitrogenation process is understood
to proceed preferentially through extended defects such as phase and grain bound-
aries, dislocations and dislocation arrays. Nitrogen diffuses into the particles along
one such path and then bulk diffusion occurs perpendicular to this path inside the
particle. The bulk diffusion coefficient is very small compared with the short circuit
diffusion through defects.
312 H. FUJII and H. SUN

The reaction between the gas phase and the rare earth intermetallics was studied
by Skomski and Coey using the lattice gas model (Coey et al. 1992, Skomski
and Coey 1993a). The net reaction energy U0 consists of three parts; the energies
which are necessary to dissociate the gas molecules, the energy for expanding the
lattice and the energy gain due to the gas-lattice interaction. For the Sm2Fel7N v
system, it was estimated that U0 = - 5 7 + 5 kJ/mol, and the diffusion parameters
Do = 1.02 mmZ/s, Ea = 133 kJ/mol. Diffusion constant at a certain temperature T
can be calculated by the expression DT = Do exp(-Ea/kT). The reaction kinetics
of N2 absorption by Sm2Fel7 were experimentally investigated using volumetric and
gravimetric measurements by Uchida H.-H. et al. (1993) and they found that Ea was
in the range 100-163 kJ/mol, in agreement with Skomski and Coey (1993a).
Skomski and Coey (1993a) also investigated the stress and strain in the inho-
mogeneously nitrogenated grains in the linear elastic approximation. The nitrogen
diffusion progresses from the grain surface into the center. At the initial stage of
nitrogenation or in the case of not fully nitrided materials, the stress and strain lead
to an expanded nitrogen-free particle center with an increased Curie temperature but
negative anisotropy constant/(1. This soft center could act as a nucleation center
for domain walls and destroy coercivity.
Fujii et al. (1994b) studied the nitrogen absorption process in SmzFel7 particles at
733 K under various Nz-gas pressures. The observation of EPMA-line profiles of the
N-element in the grains led these authors to claim that nitrogenation process consists
of the following two mechanisms: The nitrogen diffusion process and the phase
transformation from the nitrogen poor Sm2Fel7Nx to the fully nitrided Sm2Fel7N3.
Both mechanisms progress concurrently under Nz-gas pressures above 0.1 MPa.
However, the diffusion process becomes dominant under low N2-gas pressures below
0.05 MPa and at this pressure a homogeneous phase with any value of z is stabilized,
suggesting that the SmzFeI7-N solid solution exists above 733 K.

2.3.3. Nitrides with intermediate nitrogenation content


It has been an open question whether the metastable nitride is a simple gas-solid
solution with continuous range of intermediate nitrogen contents or whether it is a
two-phase mixture of nitrogen-poor and nitrogen-rich phases. The attractive inter-
atomic long-range interaction energy arising from the lattice deformation around the
interstitial site is important for answering this question (Coey et al. 1992). Below a
critical temperature Tcri, the attractive interaction dominates and the interstitial atoms
form macroscopic clusters, and consequently a two-phase mixture is formed. If the
nitrogenation is carried out above Tcri, then a solid solution phase should be formed.
For Sm2FeavNv, Tcn was estimated to be about room temperature. As the nitrogena-
tion is usually performed at temperatures much higher than room temperature, the
nitrides should be present as a gas-solid solution phase with a continuous range of
intermediate nitrogen contents, according to the estimation by Coey et al. (1992).
However, not all the experimental results support this conclusion.
Katter et al. (1992a) reported that the Sm2Fel7 nitrides existed over the whole
concentration range of 0 ~< y ~< 3. SmzFel7Nu with intermediate nitrogen contents
(0 ~< y ~< 2.94) was successfully prepared and the nitrogen content dependence of the
INTERSTITIALLYMODIFIEDINTERMETALLICS 313

unit cell volume, the Curie temperature, the saturation polarization, the anisotropy
field and the thermal stability was examined. All these properties showed a continu-
ous increase with increasing nitrogen content y. Christodoulou and Takeshita (1993e)
also reported that the Sm2Fe17 nitride is not a line compound but exhibits a large sol-
ubility of nitrogen. However, it was difficult to form homogeneous interstitial com-
pounds with y between 0 and 3, which requires either partial nitrogenation followed
by long time heat treatment in argon atmosphere or nitrogenation in a predetermined
amount of nitrogen. Mukai and Fujimoto (1992) observed the domain patterns of
SmzFelyNy and concluded that nitrides with an intermediate nitrogen content exist.
They found that the magnetic domain patterns were sensitive to the nitrogen content
and the domain width became narrower as the nitrogen content decreased. The pos-
sibility of the existence of an intermediate nitrides was also suggested by Uchida H.
et al. (1992) by measuring volumetrically the pressure-composition isotherm of the
SmzFelv-N system at 823 K. No plateau region was observed in the P - C - T curve at
nitrogenation concentrations lower than 3, which was taken as an indication of the
existence of solid solutions between Sm2Fe17 and N.
On the other hand, there are many publications in which it is reported that nitrides
with intermediate nitrogen content do not exist. By X-ray diffraction experiments,
Coey et al. (1991a) examined the reaction products after heating YzFe17 in the TPA
to temperatures ranging from 300 to 850°C. The results showed that at temperatures
lower than the temperature at which fully nitrided sample could be formed (600°C),
the products were always mixtures of the pure 2:17 phase and the fully nitrided
phase. Isnard et al. (1992a) reached the same conclusion by in situ neutron powder
diffraction study of the nitrogenation process of NdzFel7. They found that even
at the beginning of the reaction, the nitride was a highly charged interstitial phase,
no progressive filling of the interstitial site was observed and only the ratio of the
nitride phase increased as nitrogenation progressed. Similar results were obtained in
a metallography study of the nitrogen diffusion patterns in Nd2Fe17 by Colucci et
al. (1993b). No formation of a nitrogen solid solution phase was observed, and the
fully nitrided phase precipitated directly from the phase free of nitrogen.

2.3.4. Formation of carbonitrides and other multiple interstitial compounds


of H, C and N
Carbonitrides were formed by subsequent nitrogenation of the arc-melted carbides
(Coey and Sun 1990, Nagata and Fujii 1991, Kou et al. 1991a, b, Yang Y.C.
et al. 1992a). This is a combined arc-melting and GIM process. Altounian et al.
(1993) and Chen et al. (1993a, b) made the carbonitrides by the GIM process
only. The carbonitriding process can be done either using mixtures of N2 gas and
hydrocarbons or sequentially one followed by another. A mixed gas of nitrogen and
methane proved to be successful for obtaining carbonitrides of R-Fe intermetallics
with various nitrogen to carbon ratios. The reaction was written as follows,
1
yN2 + zCH4 + R2Fe17 ~ R2Fe17CxN v + 2xH2. (2.3)

Similarly to the preparation of carbides by the GIM process, H2 should be pumped out
to avoid the formation of the undesired hydrides. For the sequential process, it was
314 H. FUJII and H. SUN

found that nitrogenation should be done first, followed by the carbiding reaction.
If carbonation was followed by nitrogenation, then only carbides were produced
because the carbon layer formed on the surface of the particles prevented nitrogen
atoms from entering into the particles.
Christodoulou and Takeshita (1993f) reported the preparation of SmzFelv-carboni-
tildes, carbonhydrides, nitrohydrides and carbonitrohydrides. The carbide was made
first, by direct melting or gas phase carbonation, then it was nitrogenated and finally
hydrogenated. The nitrohydride can be produced directly by nitrohydrogenation in
a mixture of N2 + H2 or N}-I3 gasses. The SmzFely-carbonitrohydrides can also be
prepared by gas phase reaction of Sm2Fel7 powders with a mixture of hydrocarbon,
N2 and H2 gasses.

2.4. Solid-solid (or liquid) reaction and plasma nitriding methods


A new route for the synthesis of metalloids interstitials is solid-solid (or liquid)
reactions. This has been tried for synthesizing carbides using heavy hydrocarbon
organic compounds, such as benzene (C6H6) and toluene (C7H8) etc. (Fruchart et
al. 1994). Interstitial modifications of Y2Fel7 by reaction with solid boron, silicon
and sulfur were tested by Skomski et al. (1993), but were not very successful.
Compared to the gas-phase interstitial modification, the solid-solid reaction has the
following advantages; (1) it can be conducted at relative low temperatures and thus
the decomposition process could be depressed, and (2) liquid and solid materials are
easier for handling. In order to avoid the formation of hydrides, the reaction should
be conducted at temperatures higher than 400°C. For this purpose, Fruchart et al.
(1994) suggested to use hydrocarbon (and hydrocarbonitrogen) materials of heavier
formula weight, which have both a higher thermal stability and a lower H/C(N) ratio.
Plasma nitriding of SmzFe~7 was done by Machida et al. (1993). It was performed
by glow-discharge between the electrodes under a differential pumping condition at
2 Torr of a N2-H2 mixed gas with a molar ratio of 1:2. The advantage is that the
reaction temperature was only 423 K, which is much lower than the conventional
thermal gas-solid or solid-solid reaction.

2.5. Summary
Sm2Fe17nitride is a most interesting material for permanent magnet applications. For
making the nitride, GIM is the most suitable way. High quality (fully nitrided and
free from impurity phases) nitrided material is essential for making the final magnet
products with useful magnetic properties. The nitrogenation process is a relative
complicated process which is not only affected by the nitrogenation conditions, such
as the temperature, reaction period, gas pressure and composition of gasses, but is
also affected by the initial Sm2Fel7 particle size and surface condition. The basic
features and diffusion patterns of the GIM process have been fairly well understood.
However, further studies are necessary for improvement and better controlling of the
GIM process.
INTERSTITIALLYMODIFIEDINTERMETALLICS 315

3. Interstitial compounds of the 2:17-type structure

Among binary rare earth iron intermetallics, compounds with the 2:17-type structure
are the most iron rich ones and they are particularly stable. The R2Fel7 compounds
exist across the whole lanthanide series, from Ce to Lu, except for La itself. The
magnetic properties of the series have been studied in great detail (Buschow 1977,
Wallace 1985). The magnetization is fairly high because of the high iron content.
However, none of these compounds exhibits an easy-axis anisotropy at room tem-
perature, and only Tm2Felv does so in the liquid nitrogen temperature range. Fur-
thermore, the Curie temperatures are surprisingly low (240-480 K) for compounds
containing so much iron. Therefore the R2Fe17 compounds have been disregarded
as potential permanent magnet materials.
It has been found that elements with small atomic radius can occupy interstitial
sites of the R2Fel7 host lattice. These interstitial atoms cause displacements of the
metal atoms from their regular sites, and the resulting crystal lattice distortions give
rise to various kinds of changes in physical properties which are interesting from
both fundamental and applied points of view. Interstitial modifications by H, C and N
have been studied intensively and proved to be successful for improving the magnetic
properties of R2Fe17. In this section, various studies of the structural and magnetic
properties of interstitial R2Fe17 carbides and nitrides will be summarized. Studies of
the R2Fe17-hydrides have been already reviewed by Wiesinger and Hilscher (1991).
For this reason we will omit them here.

3.1. Crystallographic structures


The RzFe17 compounds crystallize in the rhombohedral ThzZn17-type structure (space
group R3m) for rare earths lighter than Gd and in the hexagonal ThzNily-type struc-
ture (space group P6/mmc) for rare earths heavier than Tb. The compounds with
R = Gd, Tb and Y can exist in both forms depending on the high temperature
annealing condition. Hexagonal form of Ce2Fel7 has also been reported to coexist
with the rhombohedral one (Buschow and van Wieringen 1970). Both of the struc-
tures are derived from the CaCus-type structure by the ordered substitution with a
(dumbbell) pair of Fe atoms for each third rare earth atom in the basal plane (Florio
et al. 1956, Zarechnyuk and Kripyakevich 1962). When these substituted layers are
stacked in the sequence abcabc, the Th2Zna7-type rhombohedral structure is real-
ized. If the stacking sequence is, instead, ababab, then the ThzNi17-type hexagonal
structure is formed. The structures are illustrated in fig. 3.1, both of which are rep-
resented by hexagonal cells with two and three RzFe17 formulas for the hexagonal
and rhombohedral structures, respectively.

3. I.I. Structure modification by the GIM process


The modification of the R2Fe17 structures by interstitial nitrogen and carbon atoms in-
troduced by the GIM process is essentially to expand the unit cells, without changing
the rhombohedral or hexagonal symmetry of the parent compounds. This retaining
of the crystallographic symmetry is the main difference when compared to the arc-
melted carbides, where a structure transition from the hexagonal symmetry to the
316 H. FUJII and H. SUN

v v v _ I

a
v v v
a

R 06C R (~)2d 02b


Fe ~6c ~gd (~)18f O18h Fe ~ 4 f @6g @12J 4D12k

Z • 9e o18g Z • 6h 0121

Th2Znl7 Th2Nl17

Fig. 3.1. Crystal structuresof R2Fe17,left: rhombohedralTh2ZnlT-type;right: hexagonalTh2Ni17-type,


showing the rare earth sites (R), iron sites (Fe) and the interstitialsites (Z).

rhombohedral symmetry was observed with increasing carbon content (for detail see
the following section). Lattice parameters of nitrides and carbides of all the RzFel7
compounds are collected in table 3.1, in which also included are the carbonitrides
and carbonitrohydrides for some of the rare earths. As shown in table 3.1, the total
number of nitrogen and carbon atoms per formula unit is close to or slightly less than
3 in nearly all the interstitials. The expansion of the cell volume upon nitrogenation
or carbonation, or the combination of them, is 6-7% for all the R2Fea7 compounds
except for Ce, where the volume increase is more than 8%. It is also noticeable that
nitrides have a slightly larger volume than carbides in general. Figure 3.2 compares
the lattice parameters of the GIM nitrides R2Fel7N v and carbides R2Fel7Cv, the arc-
melted carbides R2FeivC and the host R2Fel7 compounds (2/3e and 2/3V for the
rhombohedral compounds are plotted). The normal variation in lattice parameters
associated with the lanthanide contraction is shown excepting the Ce compounds.
The anomalous position of Ce suggests a near tetravalent (4f°) configuration for Ce
in CezFe17, which changes to an intermediate valence configuration between 4f ° and
4f I in the nitrides and carbides.

3.1.2. Structure studies of the arc-melted and melt-spun carbides


The expansion effect of the unit cell volume due to interstitial carbon atoms intro-
duced by the arc-melting method is similar to those introduced by the GIM method,
INTERSTITIALLY MODIFIED INTERMETALLICS 317

TABLE 3.1
Crystallographic structure data and magnetic properties of R2Fel7 and their interstitial compounds. V is
the crystal cell volume calculated from V = -~a2c. Saturation magnetization as at room temperature
and 4.2 K are expressed in the unit of/zB/formula.

Compound a (]k) C (/k) V (•3) AV/V (%) Tc (K) as (#B/f.u.) Ref.


RT 4.2 K
Ce2Fel7 8.48 12.38 773.5 210 29.3 [13, 24]
Ce2Fe17N2.8 8.73 12.65 834 8.8 713 36.4 [1]
Ce2FelvN2.5 8.743 12.673 838.9 8.24 700 32.2 36.3 [2, 3]
Ce2FeI7N25 8.743 12.688 847.9 9.43 [4]
Ce2FeI7N3. 6 8.73 12.81 845.5 9.3 712 31.6 39.9 [13, 24]
Ce2FelTCu 8.73 12.56 829.5 8.3 589 33.8 [7]
CezFe17C2.8 8.74 12.65 836 8.0 716 [21]
Ce2Fe17C2.5 8.73 12.58 830 8.2 608 25.1 30.0 [22]
Ce2Fel7C2 8.72 12.64 589 [8, 9]
Ce2FeI7CxN/, 8.73 12.68 721 [8, 9]

Pr2Fel7 8.574 12.324 790.7 290 18.1 [1]


Pr2Fel7N2. 5 8.77 12.64 841.8 6.5 728 37.9 [1]
Pr2FelvN25 8.794 12.668 848.4 6.65 720 36.5 41.3 [2, 3]
Pr2FelTN2.86 8.795 12.659 847.9 6.68 [4]
PrzFelTCy 8.80 12.59 843.3 6.7 653 34.5 [7]
Pr2Fel7C2. 7 8.79 12.63 845 6.8 735 [21]
PrzFel7C2. 5 8.85 12.61 855 8.3 690 27.0 30.8 [22]
Pr2Fel7C 2 8.78 12.65 654 [8, 9]
PrzFelvCxNu 8.79 12.66 737 [8, 9]

Nd2Fel7 8.563 12.444 790.2 330 17.1 [1]


Nd2Fel7N2.3 8.76 12.63 838.8 6.2 732 40.5 [1]
Nd2Fel7N2.5 8.783 12.663 845.5 6.45 740 34.9 43.9 [2, 3]
Nd2Fel7N2. 4 8.776 12.661 844.5 5.85 [4]
Nd2FelvCu 8.79 12.60 842.1 6.6 659 35.1 [7]
NdzFel7C2.2 8.76 12.62 839 5.9 739 [21]
NdzFel7C2.5 8.80 12.60 845 6.7 662 29.0 31.8 [22]
Nd2Fel7C2 8.76 12.64 658 [8, 9]
NdzFel7CxN u 8.76 12.65 740 [8, 9]

Sm2Fel7 8.55 12.43 786.9 389 22.4 34.1 [13, 24]


Sm2Fe17N2.3 8.73 12.64 833.7 6.3 749 31.9 [1]
SmzFeI7N2. 5 8.741 12.666 838.2 6.36 750 32.2 38.2 [2, 3]
Sm2FelTN2. 2 8.730 12.630 834.1 6.01 [4]
Sm2Fel7N2 8.732 12.631 834.1 745 35.1 [5]
SmzFe17N3.1 8.74 12.65 6.2 752 34.0 38.1 [13, 24]
SmzFel7N2. 9 8.76 12.76 [14]
Sm2FelTN3.o 8.728 12.638 833.73 748 35.7 [15]
Sm2Fe17N2.7 8.778 12.74 [16]
Sm2Fel7Ns.2 8.838 12.82 [16]
SmzFelvN2.9 8.743 12.703 740 34.3 35.0 [17]
SmzFeI7N2. 6 8.71 12.61 35.5 [18]
Sm2Fel7 N2.94 8.7425 12.659 837.9 6.2 746 36.2 [19]
318 H. FUJII and H. SUN

TABLE 3.1
(Continued)

Compound a (,~) c (-~) V (,~3) AV/V (%) Tc (K) ~s (/zB/f.u.) Ref.


RT 4.2 K
Sm2Fel7Cy 8.75 12.57 833.0 6.2 668 34.5 [7]
Sm2Fel7C 2 8.749 12.595 5.9 670 26.6 30.1 [20]
Sm2Fe17C2.5 8.73 12.67 836 6.5 760 [21]
Sm2FelvC2.5 8.77 12.61 840 7.1 679 26.2 29.2 [22]
Sm2Fe17C2 8.73 12.65 680 [8, 9]
Sm2Fel7C:rNy 8.75 12.65 758 31.2 [8, 9]
Sm2Fe17CN u 8.712 12.604 828.5 778 32.3 35.1 [lO]
Sm2Fe17Ny 8.742 12.651 837.3 6.2 753 [11, 12]
Sm2Fe17Co.4Ny 8.736 12.650 836.0 759 [11, 12]
Sm2FelTCo.TN u 8.742 12.633 836.0 753 [11, 12]
Sm2FelTCo.9N u 8.765 12.683 843.8 760 36.3 [i1, 12, 15]
Sm2Fel7N3.oHo.8 8.739 12.733 842.12 748 36.3 [15]
Sm2FelTCo.sN2. 4 8.742 12.652 837.34 752 35.0 [15]
Sm2FelTCo.5N2.4Ho.8 8.735 12.715 840.14 752 35.2 [15]
Sm2FeÂTC2.6 8.744 12.572 832.48 673 29.3 [15]
Sm2FeI7C2.6H1.1 8.753 12.622 837.49 674 [15]
Sm2Fel7 C2.6No.1 8.747 12.584 833.85 678 [15]
Sm2FelTC2.6NoA HI.O 8.754 12.634 838.54 678 [15]

Gd2Fel7 8.508 12.432 779.4 477 10.4 [1]


Gd2Fel7N2. 4 8.69 12.66 827.6 6.2 758 26.7 [1]
Gd2Fe17N2.5 8.713 12.652 831.8 6.89 740 17.9 27.3 [2, 3]
Gd2Fe17N2.5 8.715 12.653 832.3 6.01 [4]
Gd2Fe17C u 8.70 12.61 825.8 5.9 711 20.1 [7]
Gd2FelTC2.5 8.68 12.69 828 5.9 763 [21]
Gd2Fe17C2 8.68 12.66 712 [8, 9]
Gd2FeI7CzN u 8.70 12.67 764 [8, 9]
Gd2FelTCNy 8.695 12.603 825.2 813 28.1 [lO]

Tb2Fe17 8.484 12.410 773.6 404 11.6 [1]


Tb2Fe17N2.3 8.66 12.66 823.1 6.4 733 22.4 [11
Tb2Fe17N2.5 8.683 12.666 827.0 6.54 730 23.6 25.0 [2, 3]
Tb2Fel7Cy 8.67 12.64 823.2 6.4 680 21.3 [7]
Tb2Fe17C2 8.65 12.66 680 [8, 9]
Tb2FelTCxNy 8.65 12.71 748 [8, 9]
Tb2Fel7CNy 8.678 12.602 821.9 778 24.3 [10]

Dy2Fel7 8.445 8.304 512.9 367 11.4 [1]


Dy2Fe17N2.8 8.64 8.45 545.9 6.4 725 27.1 [1]
Dy2Fe17N2.5 8.670 12.666 824.5 6.52 720 25.6 23.0 [2, 3]
Dy2Fel7Cy 8.65 8.42 545.2 6.3 674 17.1 [7]
Dy2FelTC2. 7 8.63 8.42 543 6.1 733 [21]
Dy2Fe17C2.5 8.69 8.46 553 8.5 681 17.9 14.2 [22]
Dy2Fel7C 2 8.63 8.42 683 [8, 9]
Dy2FelTCxNu 8.66 8.45 724 [8, 9]
Dy2Fe17CNv 8.643 8.461 547.4 758 23.1 [10]
INTERSTITIALLY MODIFIED INTERMETALLICS 319

TABLE 3.1
(Continued)

Compound a (A) e (.A) V (A 3) AV/V (%) Tc (K) as (#B/f.u.) Ref.


RT 4.2 K
Ho2Fel7 8.439 8.280 510.7 327 11.2 [1]
Ho2FelTN3.0 8.62 8.45 543.8 6.5 709 27.2 [1]
Ho2Fe17N2.5 8.632 8.472 546.7 6.56 710 25.2 24.5 [2, 3]
Ho2Fel7N2.1 8.609 8.480 544.3 6.27 [4]
Ho2Fel7Cy 8.61 8.43 540.8 5.9 672 17.5 [7]

Er2Fel7 8.421 8.272 508.0 296 7.4 [1]


Er2Fel7N2.7 8.61 8.46 542.8 6.9 697 31.7 [1]
Er2FelTN2.5 8.622 8.476 544.9 6.78 690 27.4 20.5 [2, 3]
Er2Fe17N2.45 8.584 8.469 540.5 5.84 [4]
Er2Fel7C u 8.60 8.41 538.5 6.0 663 17.9 [7]
Er2Fe17C2 8.630 8.438 6.7 675 [20]
Er2Fel7 C2.4 8.61 8.44 541 6.5 708 [21]
Er2FelTC2.5 8.68 8.46 553 7.3 671 22.0 12.8 [22]
Er2FelTCy 8.63 8.44 675 [8, 9]
Er2FelTCxN u 8.63 8.49 700 [8, 9]
Er2FelTCNy 8.615 8.460 543.8 723 27.4 [10[
Er2Fel7Ny 8.633 8.472 546.8 7.2 683 [11]
Er2Fel7Co.4Ny 8.643 8.469 547.9 712 [11]
Er2Fel7Co.6Ny 8.643 8.469 547.8 706 [11]
Er2FelTCo.sNy 8.651 8.480 549.6 709 [111
Er2Fe17Cl.0Ny 8.647 8.478 549.0 701 [111
Er2Fel7Cl.sNy 8.652 12.628 818.6 698 [11]

Tm2Fel7 8.397 8.276 505.4 260 [11


Tm2Fel7 N2.7 8.58 8.47 540.6 7.0 690 32.5 [1]
Tm2FeI7N2.5 8.583 8.482 541.1 6.59 690 32.9 30.2 [2, 31
Tm2Fel7C u 8.60 8.43 539.6 6.8 656 21.2 [7]
Tm2Fel7Ny 8.569 8.481 539.3 6.3 700 [11]
TmzFel7Co.zNy 8.584 8.486 541.5 701 [11]
TmzFeI7Co.4Nu 8.590 8.478 541.8 703 [111
Tm2Fe17 Co.6Ny 8.586 8.485 541.7 705 [11]
Tm2Fel7Co.8Nv 8.586 8.484 541.6 701 [11]
TmzFel 7C 1.oNy 8.584 8.485 541.4 703 [11]

Yb2Fe17 8.414 8.249 505.7 280 [23]


Yb2Fel7N2.8 8.5701 8.495 540.0 6.48 675 34.2 [2, 31

Lu2Fel7 8.390 8.249 502.9 255 [11


Lu2Fe17N2.7 8.57 8.48 539.4 7.1 678 35.2 [11
Lu2Fe17N2.5 8.576 8.475 539.8 6.68 675 39.0 [2, 31
Lu2FelTCy 8.57 8.42 535.4 6.3 657 36.4 [71

Y2Fel7 8.477 8.264 514.3 325 18.6 32.8 [1]


Y2Fel7 8.51 12.38 776.4 300 34.0 [24]
320 H. FUJII and H. SUN

TABLE 3.1
(Continued)

Compound a (h) c (A) V (~3) AV/V (%) Tc (K) O's (p,B/f.u.) Ref.
RT 4.2 K
YzFeI7N2.6 8.65 8.44 547.3 6.4 694 34.2 [11
Y2Fe17N2.7 8.637 8.465 690 38.5 [3]
Y2FeI7N2.8 8.646 8.484 549.2 6.70 [4]
YzFelvN3.1 8.67 12.69 826.1 6.4 701 39.8 [24]
YzFeI7Cy 8.66 8.40 546.1 6.2 668 35.8 [7]
Y2FeITC2 8.689 8.414 6.8 673 28.9 34.1 [20]
Y2Fel7C2.8 8.64 8.46 547 6.6 722 [21]
Y2FelTC2.5 8.65 8.40 544 6.2 678 27.9 32.5 [22]
Y2FelTC2 8.67 8.40 668 [8, 9]
Y2Fel7CzN u 8.66 8.48 717 [8, 9]
YzFel7CNy 8.635 8.415 543.4 723 35.8 [10]
Y2Fel7Co.6Ny 8.634 8.488 547.9 722 [11]
YzFe17C1.2Ny 8.673 12.606 821.2 724 [11]

Th2Fel7 8.572 I2.472 794 320 30.4 [6]


Th2Fe17Nu 8.798 12.703 853 747 38.1 [6]

References:
[1] Sun et al. (1990b) [13] Fujii et al. (1992)
[2] Buschow et al. (1990) [14] Machida et al. (1993)
[3] Liu J.P. et al. (1991) [15] Christodoulou and Takeshita (1993f)
[4] Isnard et al. (1992d) [16] Uchida H. et al. (1992)
[5] Katter et al. (1990) [17] Wei et al. (1993)
[6] Jacobs et al. (1991) [18] Mukai and Fujimoto (1992)
[7] Sun et al. (1992) [19] Katter et al. (1992a)
[8] Altounian et al. (1993) [20] Hu and Liu (1991)
[9] Chen et al. (1993) [21] Liao et al. (1992)
[10] Yang Y.C. et al. (1992a) [22] Tang et al. (1992)
[11] Kou et al. (1991a) [23] Buschow (1972)
[12] Kou et al. (1991b) [24] Fujii et al. (1995a)

but the overall expansion is much less in the arc-melted carbides (table 3.2), which
is obviously associated with the smaller amount of carbon atoms introduced in. As
mentioned in section 2, the solubility y of carbon in R 2 F e l T C v obtained from the
melt and subsequent high temperature annealing is limited to y ~< 1.6 for heavy rare
earths and y ~< 1.0 for light rare earths. From the study of the dependence of the
unit cell volume V on carbon content y for R = Sm (Popov et al. 1990, Gr6ssinger
et al. 1991, Wang and Hadjipanayis 1991a), Gd (Dirken et al. 1989), Ho (Haije et
al. 1990), Tm (Gubbens et al. 1989, Gr6ssinger et al. 1991), Er (Kou et al. 1991c),
Y (Coene et al. 1990, Sun et al. 1990a) and Th (Isnard et al. 1992e, Jacobs et al.
1991), a linear relationship between V and y was found until y ~ 1.
In the case where the host R2Fe17 takes the Th2Ni17 structure, there is a structure
transition from the hexagonal to the rhombohedral structure with increasing carbon
concentration (Sun et al. 1990a, Coene et al. 1990, Haije et al. 1990, Kou et al.
1991c). This structure transformation is reminiscent of the structural transformation
INTERSTITIALLY MODIFIED INTERMETALLICS 321

8.7
,<
._... 8 . 6 - o r~ .-~...~ ~ ~...~,
R ~ - / ~o '~'o.~
~'~ 0 - - 0 ~' ~0~ ~ D

8.4 ~ + ~
8.5-
A

o 8.3 o , . n l o ° . . . o° ,.. o°--o° o°__~o oo- -o


o\/ °
8.2 o

560 _li-i~_,
"-" 540 ""~ ~L~....~,,,~.~,,,~.
O~C[/ - 0 ~ ° ~ o... []~ 0
> 520
0... / 0"-0
~o
0 ~0~
500 °-°'-'-°
Y Pr Pm Eu Tb Ho Tm t u
Ce Nd Sm Gd Dy Er Yb
Fig. 3.2. Lattice parameters and cell volumes of R2FeI7, the arc-melted R2FelTC, the GIM R2FelTCy
and R2Fel7Ny (after Sun et al. 1990b and Zhong et al. 1990).

observed in R2Fel7 when passing in the opposite direction through the lanthanide
series (from Lu to Pr). In both sequences the hexagonal structure is observed for
relatively small lattice constants and the rhombohedral structure for relatively large
lattice constants. The lattice parameters as a function of carbon concentration y
for Y2FelTC v are shown in fig. 3.3, where a discontinuity in the lattice constants
near the transformation point can be seen. It is also clear that the change in crystal
structure is accompanied by an expansion and contraction of the lattice in the a and
c directions, respectively, with ahex < arhomb, but Chex > Crhomb at the transition point.
It should be noted that the value of y at which the transformation occurs was found
to be different by different authors. This is due to differences in heat treatment
condition and difficulties in the accurate controlling of the carbon content in the
2:17 structures. The structural and magnetic properties are very sensitive to the heat
treatment condition because the dissolution of carbon in the matrix depends strongly
on the annealing temperature and time (Sun et al. 1990a, Wang and Hadjipanayis
322 H. FUJII and H. SUN

TABLE 3.2
Crystallographic structure data and magnetic properties of the arc-melted carbides R2FeI7C. V is the
crystal cell volume calculated from V = '/~aZc. Saturation magnetization ~s at room temperature and
2
4.2 K are expressed in the unit of ~B/formula.

Compound a (A) c (,~) V (,~3) AV/V (%) Tc (K) as (/zB/f.u.) Ref.


RT 4.2 K
Ce2FeI7C 8.540 12.424 784.5 2.3 297 32.8 [1]
Pr2FeI7C 8.604 12.466 799.5 1.1 370 [1]
Nd2FeI7C 8.630 12.474 804 1.7 449 40.6 [1]
Nd2Fe17C 8.6299 12.4739 804.5 435 36.5 [2]
Sm2Fel7C 8.644 12.476 807 2.9 552 34.0 [1]
Sm2Fe17C 8.6297 12.4614 803.7 495 35.0 [2, 3]
Sm2Fe17C 8.624 12.459 516 31.0 35.1 [4]
Gd2Fe17C 8.627 12.470 804 3.1 582 23.6 [1]
Tb2Fe17C 8.602 12.462 798 3.1 537 19.0 [1]
Dy2Fe17C 8.585 12.454 795 3.3 515 17.0 [1]
Ho2Fe17C 8.572 12.453 792 3.3 504 17.4 [1]
Er2Fe17C 8.538 8.331 526 3.5 488 18.8 [1]
TmzFelvC 8.524 8.321 524 3.8 498 24.4 [1]
Lu2Fel7C 8.487 8.321 519 3.0 490 35.2 [1]
Y2Fe17C 8.589 12.448 795 3.1 502 35.5 [1]
ThzFel7C1.2 8.697 12.518 820 3.1 24.8 35.8 [5] ND
Th2Fel7C1.2 8.694 12.509 819 462 34.5 [6]
References:
[1] Zhong et al. (1990) [4] Popov et al. (1990)
[2] Weitzer et al. (1990) [5] Isnard et al. (1992e)
[3] Weitzer et al. (1991) [6] Jacobs et al. (1991)

1991a, b). In the R2Fe17C series, the stability range o f the rhombohedral structure
extends towards the Lu end and the structure transformation from ThzZnl7 to Th2Nil7
takes place at Er, instead o f at Gd in the pure R2Fel7 c o m p o u n d s (Zhong et al. 1990).
A novel type o f stacking, in which h o m o g e n e o u s sheets o f dumbbell pairs o f iron
atoms and h o m o g e n e o u s sheets o f rare earth atoms along particular ordering planes
are stacked at random, has been reported for Y2FeI7Cy when y is close to 0.6 at which
t h e structure transition occurs (Coene et al. 1990). These stacking faults can locally
change the R site s y m m e t r y (Gubbens et al. 1989) and m a y also form an intrinsic
barrier in these materials for the attainment o f high coercive forces (Buschow et al.
1990).
With increasing o f carbon content, a similar structure transformation from the
hexagonal to the rhombohedral symmetry was also found in the interstitial carbides
prepared by the melt-spinning method (see for example Cao et al. 1993a). On the
other hand such a phase transition was not observed in the G I M interstitial carbides
and nitrides because o f the low reaction temperature.
For R2Fe17 with the ThzZnl7 structure (R = Sin, Gd and Th), the expansion o f the
cell in the a-axis direction upon carbonation is much larger than that in the c-axis
direction, which could be related with the crystallographic site occupation o f the
interstitial atoms. In ThzFe17Cv, when y > 1.2, the appearance o f a BaCdal type
INTERSTITIALLY MODIFIED INTERMETALLICS 323

8.8 I I I
(a/
8.7 a
8.6
8.5 o ~° ~ ° ~°/°-°-°-°-°
8.4'
0
8.3 - - ° ~ ° ~ ' ~ ° " ° - ° - ° - ° ' °
8.2 c
8.1
8.0 I I I 600
550 (b)
Tc 550
o o
oso-o--o-
54O 50O
53O 450 ,.t
>
520 4O0
350
51C
I I I 30O
0.0 0.5 1.0 1.5 2.0

Carbon Content, y
Fig. 3.3. Carbon concentration dependence of (a) the lattice parameters and (b) Curie temperatures and
cell volumes of Y2FelTCv (after Sun et al. 1990a).

phase was reported and the fraction of this new phase was found to increase with
increasing carbon content (Jacobs et al. 1991).

3.1.3. Structure properties of the nitrides with intermediate nitrogen content


Unlike in the arc-melted carbides, a continuous variation of the nitrogen content
in the GIM nitrides was first reported not to be possible. However, more detailed
studies have later demonstrated that nitrides with nitrogen content in between 0 and
3 do exist (see section 2.3.3 and the references cited there in). There are a few
reports that the nitrogen content y can be as high as 6 or 8 in Sm2Fe17Nv (Iriyama
et al. 1992, Wei et al. 1993), where the best magnetic properties can be obtained
at y -- 3. Structural and magnetic properties of SmzFelTNu as a function of y have
been investigated by Katter et al. (1992a) and they found that the SmzFel7 nitrides
existed over the whole concentration range 0 ~< y ~< 3. The lattice parameters and
cell volume dependences on the nitrogen content y are redrawn in fig. 3.4. The
unit cell volume has increased by 6.2% at y = 3.0, but most of the expansion has
occurred at y < 1.99.
The gas-phase interstitial modification process can be simulated by the lattice-gas
model which has been considered to bear strong resemblances to the magnetic Ising
model (Skomski and Coey 1993a). The corresponding parameters are the critical
temperature Tcrit in the lattice-gas system and the magnetic ordering temperature
Tc in the Ising system. Terit describes a second-order phase transition; below Tcrit
there is the possibility of two coexisting phases, the nitrogen-poor a-phase and the
324 H. FUJII and H. SUN

8.80 , ~ ~ , ,

,<
8.70 / f ' f ...... "

8.60 i , / -
i

8.5o I I I I I

12.60 ,~"-"-*"-*~
,,< /
v
0 12.50 ~,....~,/"
12.40 I I I 1 I
820840[-F j./.~*~'-'--
v
800~....~ . . . .
>
780[-
760~- I I I I I
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Nitrogen Content, y
Fig. 3.4. Nitrogen concentrationy dependenceof lattice parameters a and c, and the unit cell volume V
for Sm2Fe]7Ny (after Katter et al. 1992a).

nitrogen-rich/3-phase. For Sm2Fel7Ny, Tcrit is not yet known, but it does not depend
on the gas pressure and there are strong evidences indicating that Terit < 400°C
(Fujii et al. 1994c). This indicates that above 400°C, Sm2FelvNy forms a gas-solid
solution phase rather than a two-phase mixture of a and/3 nitrides. The lattice-gas
model provides theoretical support for the existence of the nitrides with intermediate
nitrogen concentration.

3.1.4. Location of interstitial atoms


The expansion of the cell volume due to nitrogen or carbon additions has led to the
conclusion that the added atoms occupy interstitial hole sites in the 2:17 structure.
In order to understand the drastic changes of structural and magnetic properties
accompanied by interstitial modifications, various crystallographic studies have been
conducted. The interstitial carbides were studied first and later work was focused on
the nitrides after they had been discovered.
By X-ray and neutron diffraction studies, carbon atoms in the arc-melted carbides
RzFelvCy were first reported to be located on the 3a site, which is in between the
dumbbell iron atoms and which is too small to accommodate one carbon atom (Luo
et al. 1987a, b). It was later established by neutron diffraction studies that the carbon
atoms fill voids of nearly octahedral shape formed by a rectangle of Fe atoms and
two rare earth atoms at opposite comers (Helmholdt and Buschow 1989, Haije et
al. 1990), which are the 9e sites in the ThzZnl7-type structure in fig. 3.1. Similar
conclusions had been drawn earlier by Block and Jeitschko (1986 and 1987 ) from
INTERSTITIALLY MODIFIED INTERMETALLICS 325

the structure refinement of X-ray data on single crystal of R2Mn17C3_y. While


full occupation of the 9e (6h for the Th2Ni~7-type structure) sites by carbon atoms
corresponds to the formula R2Fel7C3, only less than half of the full occupancy has
been achieved in the arc-melted carbides.
Neutron diffraction results on the crystallographic position and occupancy of ni-
trogen atoms in the interstitial nitrides R2Fe~7Ny for various rare earths are sum-
marized in tables 3.3a and b. Nitrogen has a large neutron scattering length pi =
TABLE 3.3
The occupancy factor n and the atomic position parameter x of the interstitial sites occupied by in-
terstitial nitrogen atoms in (a) the rhombohedral Th2ZnlT-type R2FelTNy (R = Ce, Pr, Nd and Th)
and (b) the hexagonal Th2Ni~7-type Y2FelTNy compounds.
a

9e (0.5, 0, 0) 18g (x, 0, 0.5)


Compound n z n a (,~) e (A) Ref.
Nd2Fel7N2. 6 0.86(2) 8.7760(1) 12.6366(8) [1]
Nd2Fe17N2.56 0.85 8.762 12.631 [2]
Nd2FelvN2.52 0.60(2) 0.806(6) 0.12(1) 8.763(1) 12.644(2) [3, 41
NdzFe17Nk26 0.14(5) 0.805 0.14(5) 8.629(1) 12.512(2) [3, 4]
NdzFe17Nv 0.860(8) 8.776(1) 12.661(1) [5]
NdiFe17N3 0.984(8) 8.786(1) 12.676(1) [6]
NdzFelvN2.85 0.95(4) 8.7746(5) 12.6570(6) [7]
NdzFel7N2.91 0.95(4) 0.811(57) 0.01(2) 8.7773(4) 12.6602(5) [7]
NdzFelTN4.5 0.8 0.854 0.35 8.763 12.688 [8]

Ce2Fe17N3 0.992(12) 8.737(1) 12.702(1) [6]


Pr2Fe17N2.9 0.972 8.771(1) 12.629(1) [9]
ThzFelvN3 0.988(4) 8.8020 12.737 [10]

References:
[1] Ibberson et al. (1991) [6] Isnard et al. (1992b)
[2] Yang Y.C. et al. (1991a) [7] Kajitani et al. (1993)
[3] Jaswal et al. (1991) [8] Yan et al. (1993)
[4] Yelon and Hadjipanayis (1992) [9] Isnard et al. (1992c)
[5] Miraglia et al. (1991) [10] Isnard et al. (1993)

6h (x, 2x, 0.25) 12i (x, 0, 0)


Compound x n X n a (h) c (£) Ref.
Y2Fe16.5N2.9 0.8329(5) 0.77(1) 0.1465(19) 0.10(1) 8.6393(1) 8.4749(2) [11
Y2FelTN2.5 0.833 0.83 8.653 8.451 [2]
Y2Fe18N2 0.8319 0.67(2) [3]
YzFels.sN2 0.8292(13) 0.56(3) 8.6622(11) 8.4702(13) [4]

References:
[1] Ibberson et al. (1991) [3] Jaswal et al. (1991)
[2] Yang Y.C. et al. (1991a) [4] Yelon and Hadjipanayis (1992)
326 H. FUJII and H. SUN

0.930 x 10 -12 cm which makes it easy to determine the location of nitrogen ac-
curately. All the studies indicate that nitrogen predominantly occupies the 9e(6h)
octahedral sites and the best fits have been obtained with exclusive occupancy of
these sites in Pr2Fe17N2.9 (Isnard et al. 1992c). For Nd2Fe17, some authors also
suggested a partial occupation of a second interstitial site, the 18g sites (see table
3.3a). Jaswal et al. (1991) have reported that nitrogen fills the 18g site first, to its oc-
cupancy limit n = 1/6 and then fills the 9e site to its limit n = 2/3. They also found
that the cell volume increased almost linearly with increasing 9e site occupancy, but
not with the total N content. In fitting the neutron diffraction spectra, Kajitani et al.
(1993a, b) positioned a small amount of nitrogen atoms on the 18g site. However,
they suspected that these 18g-site interstitial atoms might be hydrogen instead of
nitrogen, since some hydrogen atoms coexisted with nitrogen in the samples because
the nitrogenation was performed in a NH3-H2 gas mixture.
A different model was given by Yan et al. (1993) who reported a nitrogen con-
tent as high as 4.5 atoms per formula of Nd2Fe17. From their structural analysis
these authors concluded that nitrogen filled the 9e and 18g sites simultaneously and
achieved a final occupancy of 0.80 and 0.35, respectively. The occupancy of ni-
trogen on the 18g site was limited to 50% of the nitrogen hexagon at alternating
vertexes and consequently a full occupation of 6 nitrogen atoms per formula was
suggested in the model. From the abnormally large thermal factor of nitrogen on
the 18g sites, the authors concluded that vibrating ellipsoids of nearest neighbors
overlapped, so that the nitrogen atoms should be able to move freely from one site
to another nearest-neighbour site. The cluster consisting of the nitrogen hexagon at
18g sites and two rare earth atoms at 6c sites above and below the nitrogen hexagon
was considered to be responsible for the metastability of the nitride, because the
nitrogen atoms at the 18g sites would have sufficient kinetic energy to move to other
sites at high temperatures.
For the Y2Fel7 compound, Jaswal et al. (1991) reported that the hexagonal host
compound showed considerable compositional variation as well as disorder associ-
ated with less than complete dumbbell substitution of one-third of the rare earth sites.
These features remained after nitrogenation and complicated the structure refinement
of Y2Fel7Ny. Nevertheless, N atoms were located at the 6h octahedral sites with 2/3
occupancy and a final composition of Y2Fe18N2. Another type of partial disorder
consisting of the replacement of some of the rare earth atoms at the 2b site by a pair
of Fe 4e atoms in the host Y2Fe17 was taken into account by Ibberson et al. (1991).
In the nitride, they placed 12% of the N atoms on the 12i site and this led to a final
formula of Y2Fe16.sN2.9 was deduced.
Neutron diffraction studies of the Sm compounds are not possible because of the
large absorption cross section of Sm atoms. The site occupancy of nitrogen atoms
in SmzFeITN u was studied by EXAFS experiments using the SmLni edge (~ =
0.184 nm) (Coey et al. 1991a, Capehart et al. 1991). The magnitude of the Fourier
transformation of the EXAFS is shown in fig. 3.5(a) and (b), where the experimental
(solid line) and calculated (dashed line) radial distribution of atoms surrounding
the Sm atom, before and after nitrogenation, were plotted. Figure 3.5(a) shows
only the first shell of iron neighbors with 7(2) atoms at a distance of 0.310(3) nm.
INTERSTITIALLY MODIFIED INTERMETALLICS 327

(a)

N
c

_c

0.0 0.2 0.4 0.6 0.8 1.0


Radlal dlstance (nm)
Fig. 3.5. Radial distribution functions deduced from Sm2FelTLIn edge EXAFS of (a) before and
(b) after nitrogenation (after Coeyet al. |991a).

" 18,-Fo /~ ,~ " ~, •


• 9e-z / / ~/

/,,°/
O~c-R ,~__ ~ ~/

~( o ,~____¢
/,,, // zo0
O 18h-Fe / p~''Q''.. /
9d-Fe / /

/- """". " '("o---3'/

Fig. 3.6. Atomic site position of z = 0 and z = 1/6 planes of the Th2Zn17 unit cell, where the interstitial
9e and 18g sites are showed.
328 H. FUJII and H. SUN

In fig. 3.5(b), the R-Fe distance has expanded to 0.316(3) nm, and another closer shell
appears, corresponding to 2(1) nitrogen atoms at 0.252(2) nm. This distance, which
equals to the sum of the samarium metallic radius (0.180 nm) and the nitrogen single
bond radius (0.074 nm), allows the nitrogen atoms on the 9e octahedral interstitial
site to be located.
Yang C.J. et al. (1993) reported a detailed X-ray diffraction study of the Sm2FelT-
nitride made by nitrogenation of the melt-spun Sm2Fel7 compound. The site occu-
pancy by nitrogen atoms were found to be the same as those derived from neutron
diffraction and EXAFS experiments. However, they reported that the volume ex-
pansion was less for the rapid quenched compound than the cast compound.
Figure 3.6 shows the atomic site position of the z = 0 and z = 1/6 planes of the
rhombohedral unit cell, where the interstitial 9e and 18g sites are included. As the
9e sites are in the same e-plane with rare earth sites and have the rare earth sites
as the nearest neighbors, the electronic environment of the rare earth atoms will be
modified severely upon introducing N and/or C atoms into the 2:17 structure and the
magnetocrystalline anisotropy is expected to be changed drastically.

3.2. Curie temperature and exchange interactions


The magnetic ordering temperature Tc (Curie temperature) is governed by three
kinds of exchange interactions: (1) the 3d-3d exchange interactions, which are the
direct exchange interaction due to overlapping of the 3d-electron wave functions
and these are strong enough to dominate Tc of the 3d rich intermetallic compounds;
(2) the 3d-4f exchange interactions, which couple the 3d and 4f moments; and
(3) the interactions between the rare earth spins, which are assumed to be weak
and negligible in comparison with the interactions mentioned above. Exchange
interactions in the R2Fe17 compounds are weak compared to those in the elemental
Fe. The Curie temperatures are around room temperature for R2Fel7, whereas Tc
for Fe is about 1043 K. This has been assumed to be due to the very short Fe-Fe
interatomic distances at the dumbbell sites (Givord and Lemaire 1974). It is well
known (the N6el-Slater curve) that 3d transition metal atoms interact with each
other positively, i.e. couple ferromagnetically, at large interatomic distances, but that
at too short distances they interact negatively or couple antiferromagnetically. In
the RzFel7 compounds the Fe-Fe distance at the dumbbell 4f or 6c sites is typically
about 2.4 A, which is shorter than 2.42 A below which negative exchange interaction
occurs. This weakens the overall positive exchange interactions and lowers the Curie
temperature.
As a direct effect of the volume expansion by interstitial modifications, the
Fe-Fe interatomic distances increase, which induces the enhancement of the ex-
change interactions and thus the Curie temperature increases enormously. Using
the mean field model, Anagnostou et al. (1994) fitted the temperature dependence
of the reduced magnetizations obtained from the M6ssbauer hyperfine fields and
estimated the exchange interactions between the four different Fe sites in YzFe17
and Y2Fe17Nu. The results have shown that the exchange interaction between the 4f
sites is strongly negative (antiferromagnetic) in Y2Fel7, and it increases and becomes
INTERSTITIALLY MODIFIED INTERMETALLICS 329

I I i I I I I

800
e ~ . e .~ ,b..~ e - - ' " - " e " - - ' - " e ~ e . ~ , R2Fel7Ny

600 \/ ..Fe,.C,
v • nn~ n""~" ran...,m~,..~_~.__ nn

400

"..........
200 R2Fel7

Y Pr Prn Eu Tb Ho Tm Lu
Ce Nd Sm Gd Dy Er Yb

Fig. 3.7. Curie temperature of R2Fel7, the arc-melted R2FelTC, the GIM R2Fe17Cy and R2Fe17N u
(after Sun et al. 1990b and Zhong et al. 1990) through the rare earth series.

120 - I ' ' ' I ' ' I ' ' ' I /,


'-

• O
100
A N

80

v
u 60 o

U
F'- 0
<1 40 • I ~
0

20
- #
o.-" Y2Fel7Zy
#
I l l l l l l , l l l t l l
0 2 4 6
A V I V (%)
Fig. 3.8. ATc/Tc as a function of AV/V for Y2Fel7 and its hydrides, carbides and nitride. The dashed
line was calculated from the pressure dependence of the Curie temperature d(Tc)/dp = - 4 7 K/GPa
(Nikitin et al. 1991) and the bulk modulus 13 = 128 GPa (Beuerle et al. 1991).
330 H. FUJII and H. SUN

weakly negative in Y 2 F e l 7 N v. Data on Curie temperatures for the host compounds


R2Fel7, the arc-melted carbides R2Fel7C, the GIM carbides R 2 F e l 7 C u and nitrides
RzFelTNy are listed in table 3.1 and 3.2 and plotted in fig. 3.7. The average en-
hancement of Tc are 150 K, 332 K and 386 K for the arc-melted R2Fe17C, the GIM
carbides and nitrides, respectively.

3.2.1. Relation between Curie temperature and the unit cell volume
The volume-dependence of Te for hydrides, carbides and nitrides of Y2Fel7 is plotted
in fig. 3.8. The dashed line was calculated by using the pressure dependence of Curie
temperature d(Tc)/dp = - 4 7 K/GPa (Nikitin et al. 1991) and the bulk modulus
B = 128 GPa (Beuerle et al. 1991). The common trend indicates that the volume
effect on Tc outweighs the chemical particularities of the individual interstitials.
Figure 3.9 plots Tc versus the corresponding unit cell volume for various interstitials
of Sm2Fe17, which includes the data obtained by Katter et al. (1992a) for the
intermediate nitrogen content interstitials Sm2Fe17N u (0 ~< y ~< 2.94), and data
obtained by Christodoulou and Takeshita (1993f) for hydrides, carbides, nitrides and
the combinations of them. Here a linear relationship between the two quantities is
demonstrated again. Data by Fujii and coworkers (Nagata and Fujii 1991, Fujii et
al. 1995a) have shown that the increase of Tc in the interstitial compounds is related
linearly to the increase of the lattice constant a rather than to the unit cell volume V.
The a-axis expansion induces a reduction of the hybridization between the Fe 3d
and R 5d electron states. This in turn reduces the density of states of the 3d-electron
at the Fermi energy and leads to the increase in Te according to the spin fluctuation
theory (Mohn and Wohlfarth 1987).
Most of the published results agree that the maximum interstitial nitrogen solubility
in RzFe17 is 3 nitrogen atoms per formula unit and that the occupancy is exclusively
on the 9e sites (Isnard et al. 1994). In this interstitial range (0 ~< y ~< 3), the Curie

600 , , ,

5OO

40O

,,t
v 300
(.~
1-
20O

100

780 800 820 840


Cell Volume V ( A3)
Fig. 3.9. Relationship between the Curie temperature Tc and the crystal cell volume V for the Sm2Fel7
nitrides, carbides, hydrides, carbonitrides, carbonhydrides, nitrohydrides and carbonitrohydrides. Data
were taken from Katter et al. (1992a) and Christodoulou and Takeshita (19931").
INTERSTITIALLY MODIFIED INTERMETALLICS 331

temperature increase can be well explained by the magnetovolume effect, as shown


in figs 3.8 and 3.9. However there are publications which claim that y could be as
high as 8, but the relation between Curie temperature and volume breaks down for
y > 3. Wei et al. (1993) reported that the strength of exchange interactions reaches
a maximum at y = 3 and then T~ decreased with further increase of y.

3.2.2. Estimation of the strength of exchange interactions


The exchange interactions can be analyzed by the molecular field model, which
is commonly used to describe the variation of the Curie temperature in the R-Fe
intermetallic series, under the assumption that the localized 3d-electron model is
applicable. The exchange interactions take place between all unpaired spins in the
3d-4f system and they are generally considered to be of the Heisenberg type.
Applying the two-sublattice molecular field model to the paramagnetic state (Be-
lorizky et al. 1987), the following expression can be obtained,

To=1 [TFe + TR -k- i ( T F e - TR)2 -I- 4T2Fe], (3.1)

where

rFe = nFeFeCFe, (3.2)

TR = ~2nRRCR, (3.3)

and

TRFe = I ' Y [ n R F e ~ : V/(Tc -- TFe)(Tc - TR). (3.4)


Here nij are the molecular field coefficients (with the same dimension as #o),
CR = NRg2j(j + 1)/~2/3kB, NR is the number of rare earth atoms per unit volume,
CFe = NFe4SFe(SFe + 1)#2/3kB, NFe is the number of Fe atoms per unit volume and
3' = 2(9 - 1)/9. Taking SFe = 1 and neglecting TR, Tc is given by,

Tc=1 (TFe q- V/T2e + 4rR2Fe) (3.5)

and, nFeFe and nRFe can be calculated from


TFe
nFeFe = CFe (3.6)

and

1 /Tc(Tc - TFe),
(3.7)
nRpe--I'Y~V ~C~e
respectively.
332 H. FUJII and H. SUN

TABLE 3.4
Molecular field coefficients n ~ e and nFeFe of R2Fel7, R2FeI7Ny and R2Fel7Cv with y ,,~ 2 to 3.

Compound np,Fe (/z0) nFeFe (#0)


Ce Pr Nd Sm Gd Tb Dy Ho Er Tm
R2Fe17 -- 345 355 328 251 227 227 215 204 -- 181
R2Fe17Nv -- 695 476 352 199 197 220 221 227 -- 515
R2Fel7Cy -- -- 87 131 158 123 127 149 124 -- 496

Taking the Curie temperature of the Lu compounds as Tve, nFeFe can be deduced
and then nRFe can be obtained by substituting the appropriate Tc data of each rare
earth into eqs (3.5), (3.6) and (3.7). Values of nFeFe and nRFe for R z F e I 7 and their
carbides and nitrides are listed in table 3.4. As can be seen from the table, the
Fe-Fe interaction is more than doubled for the nitrides and carbides compared with
the parent compounds, whereas the R-Fe interaction nRFe is slightly weakened for
the nitrides on average and is more signifcantly reduced in the case of carbides. This
feature is reflected in fig. 3.7, in which the Tc curve of R 2 F e l 7 C y is much flatter than
that of R2Fel7.
Another method for the estimation of nRFe is by high-field magnetization studies
(Liu J.E et al. 1991). At a critical field strength Beri, the antiparallel configuration
between the R-sublattice magnetization and the Fe-sublattice magnetization in the
heavy rare earth compounds is broken and the two sublattice moments start to bend
towards the parallel direction with each other. In this situation the quantity nRFe
can be derived from the slope of the M versus B curve above Bcri. Conclusions
achieved by this method are similar to those obtained by the molecular field method.
The reduction of the magnetic intersublattice coupling strength has also been
concluded from the inelastic neutron scattering study of Gd2Fel7 and its nitride
, Gd2FelTNu (Loewenhaupt et al. 1994), from the fitting of the temperature depen-
dence of 166Er and 169Tin hyperfine fields in Er2Fe17C and Tm2FeayC (Gubbens et
al. 1994) and from the fitting of the spin reorientation temperatures of Tm2FelvCy
in terms of the crystal field and the T m - F e exchange interactions (Zhao et al. 1993).

3.3. Magnetic anisotropy


In hard magnetic materials, the anisotropy energy originates from both the rare earth
and 3d sublattices. The 3d anisotropy can be deduced from compounds with non-
magnetic rare earth elements. The rare earth contribution to the magnetocrystalline
anisotropy is dominant at low temperatures whenever the 4f ions have non-zero
orbital moments, but it rapidly decreases with increasing temperature. In the case
when there are competitions between the rare earth and the 3d sublattice anisotropies,
temperature-induced spin reorientation phenomena can occur as a consequence of the
cancellation of the anisotropy contributions from the 4f and 3d sublattices.
The easy magnetization direction of the R 2 F e l 7 compounds lies in the basal plane
in the whole temperature range except for Tm2Fe17, where a change of the easy
magnetization direction from basal plane to e-axis occurred at around 80 K with
INTERSTITIALLYMODIFIEDINTERMETALLICS 333

decreasing temperature (Givord and Lemaire 1974, Gubbens et al. 1976). This
has been understood as the result of a competition between the easy-plane Fe and
the easy-axis Tm sublattices anisotropies. Interstitial modification by nitrogen or
carbon has led to a radical change of the magnetic anisotropy. In the interstitially
modified compounds, spin reorientation phenomena were observed not only for R =
Tm, but also for R = Er. In the case of Sm compounds, the room temperature
anisotropy was found to be of strongly uniaxial character instead of planar, as in
the host SmzFea7. This strong easy axis anisotropy combined with the reasonable
high Curie temperature and high magnetization, makes the Sm2Fea7 interstitials to
be very promising candidates for permanent magnet applications.

3.3.1. Theoretical background


The main contribution to the magnetic anisotropy in permanent magnet materials
is the crystal-field induced single-ion anisotropy on the magnetic rare earth-atoms.
The 4f moment of the rare-earth atoms will prefer a certain magnetization direction
which is determined by the properties of the rare-earth atoms themselves and the
crystal electric field at the particular crystallographic sites. In compounds consisting
of rare-earth and 3d metals, the easy magnetization direction is controlled by the 4f
sublattice magnetic anisotropy owing to the strong exchange interaction between 3d
and 4f moments.
The macroscopic anisotropy energy can be expanded phenomenologically as

EA : E E K~ sin~ 0 cos m ¢, (3.8)


n m

where 0 and ¢ are the polar and azimuthal angles of the magnetization with respect to
the [001] and [100] crystallographic directions, respectively. Considering the lowest
order term only, the above equation can be expressed as EA = KI sin 2 0. In the
absence of an external field, the spontaneous magnetization direction favours the
crystallographic c-axis when satisfying the following condition: ~EA/~O = 0 and
O2EA/O02 > 0, at 0 = 0. This means that a positive value of K1 will lead to an
easy-axis anisotropy.
On the other hand, within the ground state multiplet, the crystal field Hamiltonian
on the rare earth atom is written as

HCF = E E B'~Om~(J)' (3.9)


m

where the parameters Og~ are Stevens equivalent operators (Hutchings 1964) and
B ~ are the crystal-field parameters. The thermal averages of O~ vary initially as a
power n(n+ 1)/2 of the rare earth magnetization (Callen and Callen 1966). Hence the
higher the order of the anisotropy, the lower its contribution at higher temperature.
B ~ can be separated into terms related to the surrounding charges (A~) and terms
related to the 4f ions only (On(r~)),

B'~ = O~{rn)A~, (3.10)


334 H. FUJI1 and H. SUN

where A T are known as the crystal field coefficients, 0n is a constant depending on


the rare earth known as the Stevens factor (c~j,/3j, 7J for n = 2, 4, 6, respectively)
(Stevens 1952), and (r n) is the mean of the nth power of the 4f radius, which has
been calculated on the basis of Dirac-Fock studies of the electronic properties of the
trivalent rare earth ions (Freeman and Desclaux 1979). The second order Stevens
factor a j has a fixed value for a given rare earth; the sign is related to the asymmetry
of the charge cloud of the 4f electrons. The shape of these charge clouds could be
either like a pancake (o~j < 0, Ce, Pr, Nd, Tb, Dy, Ho) or like a cigar (c~j > 0, Sin,
Tm and Er). In the case of Gd, aa = 0, because the 4f charge cloud has spherical
symmetry.
For both the hexagonal and rhombohedral symmetries, the relation between the
macroscopic anisotropy parameter K1 and the crystal field-related parameters can be
given as (Lindgard and Danielsen 1975),

K (T = O) = - 3 d(r2)(2j2 j)AO ' (3.11)

considering only the ground state multiplet, which is usually a good approximation
for most of the rare earths. However Sm is a typical exception, where the involve-
ment of higher multiplets in the calculation is necessary (Sankar et al. 1975). The
above equation means that a positive value of K1k requires that aj and A° are of
opposite sign.
A ° represents the lowest order deviation from sphericity of the electrostatic po-
tential and it can be split into two terms; the contribution from the charges on other
atoms in the lattice, A°(lat), and the contribution from the charges of the valence
electron of the rare earth atoms themselves, A°2(val). Recent band structure calcula-
tions of crystal field parameters for rare earth intermetallic compounds have showed
that the asphericity of the valence electron charge density of the rare earth itself forms
the dominant contribution to the lowest order crystal field parameter A° (Zhong and
Ching 1989, Coehoorn et al. 1990).
The sign of A ° is negative in the pure 2:17 compounds. In the nitrides and
carbides, N and C atoms occupy the interstitial sites around rare earth atoms within
the basal plane (fig. 3.6), which causes the increase of the rare earth valence electron
charge density in the plane in order to match those of the interstitial atoms, and A°
decreases to more negative values (Coehoorn 1991). Consequently, for rare earths
whose a j are positive (Sm, Er and Tm), the easy axis anisotropy of the rare earth
sublattice is enhanced by interstitial modification. For the Sm interstitial compounds,
the easy axis Sm sublattice anisotropy predominates that of iron even above room
temperature. In the case of Er and Tm, the easy plane iron sublattice anisotropy is
still dominant at high temperatures, but the spin reorientation temperature increases
upon interstitial introduction.

3.3.2. Summary of experimental data


The easy magnetization direction at room temperature for all compounds in the
RzFe17 series lies in the basal plane. The easy magnetization direction at room
INTERSTITIALLY MODIFIED INTERMETALLICS 335

temperature can be deduced from X-ray diffraction patterns of magnetically-aligned


powders (fig. 3.10). When the alignment direction is parallel to the the scattering
vector of the X-ray beam, the enhanced (001) or (hk0) reflections indicate that the
easy magnetization direction is parallel and perpendicular to the crystallographic
c-axis, respectively. As shown in fig. 3.10, the Fe sublattice anisotropy retains the
same sign of the host R2Fe17 compounds in nitrides and carbides, in which the easy
magnetization direction lies in the basal plane in the whole magnetically ordered
regime. The Sm compounds are the only members in the interstitial nitrides or
carbides which exhibit easy e-axis anisotropy at room temperature.
There are many studies on the magnetic anisotropy of the SmzFcI7 nitrides and
carbides. In the arc-melted carbide Sm2FelvCv, the easy c-axis anisotropy at room
temperature was observed when y ~> 0.4 by X-ray diffraction studies on magnetically
aligned powder (GrSssinger et al. 1991), by magnetic measurements (Zhong et
al. 1990, Popov et al. 1990) and by 57Fe MOssbauer measurement (Ding and
Rosenberg 1991). When y < 0.3, SmzFe17Cv has easy plane anisotropy and when
0.3 < y < 0.45 it exhibits easy cone anisotropy (Popov et al. 1990). The room
temperature anisotropy field yoHa was found to be 5.3 T for y = 1.0 (Kou et al.
1990).
In the nitrides Sm2FelvN v, #oHa is also strongly related to the nitrogen content y
(Katter et al. 1992a). Due to different measuring methods and different maximum
nitrogen content attained by different authors, #0Ha at room temperature has been
reported to be in a wide range between 11 and 26 T. By fitting the magnetization

' ~oo I i~20' ' I I ' '

Y2Fel7 $m2Fe 17

600
600 440

olo61-7--T ---T-v T-v-r---r-


$m2Fe17N2,3
Y2Fe17N2,6

5" 3O0
?7

c
,,-4,, ;-i V
600

Y2Fe17C2.2
440
003
i , , ]
L
'
OO6
,
'
,
'
I
I
(pog,
' ' "
I
I

Sm2Fe 17C2,2
: I0012
~,--

30O

. . . . . . 600 00.3 , , .1_ .~_ .... I ,.009, I , ,


I I I

40 60 BO 40 60 BO
2e (') 2e (')
Fig. 3.10. X-ray diffraction patterns of magnetically aligned Y2Fe]7, Sm2Fel 7 and their nitrides and
carbides. Sm2Fel7 and all the Y compounds show easy-plane anisotropy, while SmzFelvN2.3 and
Sm2Fel7C2. 2 show easy-axis anisotropy.
336 H. FUJII and H. SUN

curves of magnetically aligned powders, Katter et al. (1992a) estimated the total
anisotropy constants K1 and K2 for the Sm2FelTNu series. Nitrogen concentration
dependences of #0Ha, K1 and K2 are shown in fig. 3.11. It can be seen that K1
increases almost linearly with nitrogen content y, whereas K2 is not much affected
when y > 0.8. The nitrogen concentration at which the easy magnetization direction
reaches the c-axis can be estimated from the change of the sign of K1, the value of
which has been found to be y = 0.55. An easy-cone concentration range was found
at y ,-~ 0.4.
In a study of the magnetic properties of Sm2Fe~7C=N v (x = 0, 0.4, 0.7 and 0.9,
x + y ,-~ 3), Kou et al. (1991a) reported that the anisotropy fields were higher for
x = 0.4, 0.7 and 0.9 than for x = 0, and they concluded that the effect of carbon
on the crystal electric field acting on the Sm ion was slightly higher than that of
nitrogen. On the other hand, Chen et al. (1993a, b) claimed that the effects of
nitrogen and carbon were almost the same. Most of the reported #0Ha values at
room temperature for the GIM carbides, carbonitrides and carbonitrohydrides were
around 15 T (Christodoulou and Takeshita 1993f, Wei et al. 1993, Sun et al. 1992,
Hu and Liu 1991).
The temperature dependence of the anisotropy field for the Sm interstitial com-
pounds has been studied by Kou et al. (1990), Katter et al. (1990), Chen et al.
(1993a, b), Miraglia et al. (1991). The easy magnetization direction (EMD) remains
parallel to the c-axis up to the Curie temperature and #oHa increases with decreasing
temperature. From high field magnetization data, Liu J.P. et al. (1991) derived that

30

~ . . @ f i
20

"-r"
.j.jo.--'~"~" °
10
j" J

0 ¢ I I t I,,,, [ I I I I

8 J
KI~ "
co
6
,=J
4 .J
.J K2
x,'o J 2 e - - e ~ ,
j ' ~ : 7 .... °
o
.2, r ~ •
, , , , I , , , , I
0 1 2
Nitrogen Content, y
Fig. 3.11. Nitrogen concentration dependences of the anisotropy field/z0Ha , anisotropy constants K]
and K2 at room temperature for Sm2FelTN u (after Katter et al. 1992a).
INTERSTITIALLY MODIFIED INTERMETALLICS 337

#oHa at 4.2 K is larger than 35 T in Sm2FelTN2.7. In a fitting of the high field mag-
netization processes by Kato et al. (1993), #0Ha of Sm2Fe17N3.o at 4.2 K was found
to be extremely large and the field of saturation along the magnetic hard direction
was estimated to be higher than 70 T. By using the singular point detection (SPD)
technique, Chen et al. (1993a, b) measured #0Ha of Sm2Fe17N2.3 and Sm2Fe17C2,
and Sm2Fe17C~Nv made by two different methods, in the temperature range from
370 K to 670 K. The data for the carbides and nitrides are replotted in fig. 3.12.
The other two members in the 2:17 series which are interesting for the study of
magnetic anisotropic properties are the compounds with R = Er and Tm, which
show spin reorientation phenomena. At low temperatures, the anisotropic properties
are dominated by the rare earth sublattice and the magnetization lies along the c-axis
direction. At higher temperatures, the iron sublattice anisotropy dominates and the
magnetization lies in the basal plane. The temperature at which the easy magne-
tization direction turns from the c-axis to the basal plane is the spin reorientation
temperature Tsr. It is often measured by thermomagnetic scan in a low magnetic
field, or by the temperature dependence of the a.c. susceptibility. It can also be de-
duced from analyzing the 57Fe M6ssbauer spectrum as will be described in section
3.4.2. Spin reorientation studies of various compounds made by different methods
are listed in table 3.5. Some of the data taken from table 3.5 are plotted in fig. 3.13.
In the arc-melted ErzFe17C v system, spin reorientation occurs when y >/0.8 and
Tsr increases with increasing y (see the references in table 3.5). Two spin reorien-

12 I [ I

oil
10
I

Sm2Fe,~TN2.3
k-
v

Sm2Fe1702

2
I I I
300 400 500 600 700
T (K)
Fig. 3.12. Anisotropy field/~0Ha as a function of temperature for the GIM Sm2Fel7C2 and Sm2FelTN2.3.
The values were obtained by the SPD technique and have not been corrected for the demagnetizing field
(after Chen et al. 1993a, b).
338 H. FUJII and H. SUN

TABLE 3.5
Investigation of the magnetic spin reorientation properties of the interstitially modified R2Fel7 (R =
Er and Tin) carbides and nitrides by various experimental techniques. M-T is the magnetization
versus temperature curve.

Compound Method Ref.


Er2Fel7Cy (0 ~ y ~ 1.5), arc-melted M/3ssbauer [1]
Er2Fel7Cy (0 ~ y ~ 2.0), arc-melted M-T [4]
Er2Fe17Cu (0 ~ y ~< 1.5), arc-melted a.c. susceptibility [11]
Er2Fe17Cy (0 ~< y ~ 3.0), melt-spun M-T, M6ssbauer [2, 3]

TmzFel7 Cu (0 ~< y ~ 1.4), arc-melted M-T [4]


Tm2FelvCv (0 ~< y ~ 1.0), arc-melted M-T [5]
TmzFel7Cv (0 ~< y ~< 1.4), arc-melted a.c. susceptibility [12]

Er2Fel7N2.7, GIM Mtissbauer [6, 7]


Er2Fel7N2. 7, GIM M-T, MOssbauer [8]
Er2Fel7N2.7, GIM a.c. susceptibility [9]

Tm2Fel7N2.7, GIM M-T, M0ssbauer [8]


Tm2Fel7N2.7, GIM a.c. susceptibility [9]

Er2Fel7C2.2, GIM M-T, M0ssbauer [10]


Er2Fel7C2, GIM a.c. susceptibility [14]
TmzFeI7C2.2, GIM M-T, Mtissbauer [10]

Er2Fel7CxNy (0 ~ x ~ 1.5) a.c susceptibility [13]


Tm2Fel7CxN u (0 ~ x ~ 1.0) a.c. susceptibility [13]
References:
[1] Zhou R.J. et al. (1992) [8] Hu B.E et al. (1990)
[2] Kong et al. (1992) [9] Liu J.P, et al. (1991)
[3] Kong et al. (1993c) [10] Qi et al. (1992)
[4] Ohno et al. (1993) [11] Kou et al. (1991c)
[5] Gubbens et al. (1989) [12] Grtissinger et al. (1991)
[6] Gubbens et al. (1991) [13] Kou et al. (1991a)
[7] Gubbens et al. (1992) [14] Hu B.E and Liu (1991)

tation temperatures were observed when y /> 1.0 due to the existence of both the
rhombohedral and hexagonal modifications in the specimen (Kou et al. 1991c). It
was found that Tsr(hex) > Tsr(rhomb), as the hexagonal form is more contracted in
the a-direction than the rhombohedral form. As can be seen from fig. 3.13, data ob-
tained by different researchers are quite scattered, which could be due to the difficulty
in the accurate determination of the concentration of the interstitial atoms.

3.3.3. Estimation of A ° of Sm2Fel7Ny


By fitting the experimental magnetic anisotropy field data on the basis o f a two-
sublattice model including the exchange and crystal-electric-field (CEF), the values
of the crystal field parameters A T can be deduced. Data obtained in this way for the
SmzFel7 interstitials are listed in table 3.6. All the calculations included not only the
ground state J = 5/2, but also the first and second excited states J = 7 / 2 and 9/2.
INTERSTITIALLY MODIFIED INTERMETALLICS 339

200 ' ' ' I . . . . I ' ' ' '

Er2FeleCy
150
w ._..L---~.
,¢,

I.-,-
100
/:
50
~'} ErzFe~TN2.7

I I j I I I I I t I I I I I
Tm2Fe17Cy O ~ A II

200 • °

~ ) Tm2Fel7N2.7
~'~~
t--- 150

IO0
:/1
50 I i i I i i i i

0 ] 2
Carbon Content, y
Fig. 3.13. The spin reorientation temperature TSR as a function of carbon concentration y for Er2Fel7Cy
and Tm2FelTCy. Data were taken from references in table 3.5. Values for Er2Fe17 and TmzFe17 nitrides
are also included.

TABLE 3.6
Crystal field parameters A~m of the Sm2Fel7 nitrides and carbides. /z0Ha (in unit of Tesla) is the
experimentally measured anisotropy field at room temperature and 0 K.

Compound A20 (Kao 2) A40 (Kao 4) A°6 (Kao 4) #oHa (T) Ref.
RT OK
SmzFe]TNa.94 -242 22 [1, 2]
Sm2FelTNu - 160 :t: 5 8.0 + 1.0 -3.0 + 1.0 14 [3]
SmzFelTNv -340 200 13.7 31.2 [4]
SmaFelTN3 -600 -20 26.0 > 70 [5]
Sm2Fe17C2.2 -134 :k 20 18.0 + 5.0 - 4 4- 2.0 13.5 [6]
References:
[1] Li and Cadogan (1991) [4] Zhao et al. (1991)
[2] Li and Coey (1992) [5] Kato et al. (1993)
[3] Li and Cadogan (1992a) [6] Li and Cadogan (1992b)

Different values c a l c u l a t e d b y different authors can at least partially be attributed


to the fitting o f different e x p e r i m e n t a l data. M a g n e t i z a t i o n curves o f m a g n e t i c a l l y
a l i g n e d Sm2Fe17N3.0 parallel and p e r p e n d i c u l a r to the easy m a g n e t i z a t i o n direction
at 4.2 K and 296 K are replotted in fig. 3.14, together with the fits, w h e r e the
i m p o r t a n c e o f the inclusion o f the excited J multiplets is d e m o n s t r a t e d (Kato et al.
1993).
340 H. FUJII and H. SUN
! I i I
(a) obs. 4.2K (c) talc. 4.2K
40
H//HaJ~n __ H//c-axis....,.......-"k~.~q/
/ ~ ~ / \ H _Lc-axis
--t
20
' ~ -- i;rCol~:i2gm:~;i:td2~;tiplets

} I i i
0 2~)0 0 200 400 600 800
I i

(b) obs. 296K (d) calc. 296K


40 /..,. ....... ..,,,

20 / HJ-Ha~ign A %=88o
HJ.c-axis
I ,, i I
O~ 200 200
H (kOe) H ( kOe )
Fig. 3.14. Observed and calculated magnetization curves of Sm2Fel7N3. 0. Solid lined in (c) and
(d) represent the results in which ground, first excited, and second excited J multiplets are taken into
account, while broken lines are those including the ground J mulfiplet only. The thinner solid line in
(d) denotes the case when the angle 0/~ between the c-axis and the field direction is 88°, simulating a
situation of incomplete alignment (after Kato et al. 1993).

The saturation magnetization of the hard direction was found to be lower than that
of the easy direction. This anisotropy resulted from the ferrimagnetic coupling of
the Sm and Fe moment when the external field was applied along the hard direction,
whereas they coupled ferromagnetically when the field was in the magnetic easy
direction (Kato et al. 1993, Zhao et al. 1991, Li and Cadogan 1991). The reason
for this field induced ferrimagnetism has been explained by the intermultiplet mixing
which originated primarily from the molecular field interaction rather than the CEF
interaction. As the relative contribution from the molecular field becomes larger
than that from the CEF at elevated temperatures, this ferrimagnetic coupling is more
pronounced at 296 K than at 4.2 K. Xu et al. (1993) analyzed the data obtained by
Iriyama et al. (1992) for Sm2Fe]7N v (0 < y < 6) and estimated the effect of different
interstitial nitrogen sites on A T. They concluded that nitrogen atoms on the 9e site
had a negative contribution to A °, and hypothetical nitrogen atoms on the 3b and
18g sites contributed a positive value to A ° and reduced the easy-axis anisotropy,
which was reported to be in agreement with the experimental data.
The observed spin reorientation temperatures of Tm2Fe]7Cv (0 ~< y ~ 2.2) was
fitted by Zhao et al. (1993) and the average A ° value of the two Tm sites was
derived (fig. 3.15). The absolute value of the average A ° increases with increasing
carbon content in a nearly linear relationship. Similar results have been obtained by
Li and Cadogan (1992b) for Sm2Fe]7C v and Sm2Fel7Ny.
INTERSTITIALLYMODIFIED INTERMETALLICS 341

500 . . . . , . . . . ,
Tm2FelTCy
400 ,~"
,/
/

300 J"

200 /
/
/"
1oo

O0 , , , , I , , , , I ,
I 2
Carbon Content, y
Fig. 3.15. Carbon concentration dependence of the average second-order CEF parameter A° for
Tm2Fe17Cy (after Zhao et al. 1993). A0 was obtained by fitting the data of the spin reorientation
temperatures.

3.4. MSssbauer and NMR studies


MSssbauer spectroscopy is an effective experimental technique for obtaining infor-
mation of the magnetic properties on an atomic scale. Both the 57Fe M6ssbauer
and some of the rare earth M6ssbauer effects have been studied for the interstitially
modified compounds.

3.4.1. 57Fe MOssbauer studies


As there are four different crystallographic sites for iron in both the hexagonal
and the rhombohedral-type R2Fe17 compounds, the observed spectrum must be a
superposition of at least four sextets. Point charge calculations showed that under
the combined effect of the dipolar field and quadrupole interaction, only the dumbbell
4f(6c) sites remain equivalent (Steiner and Haferl 1977). When the magnetization
direction is along the c-axis, the angle between the hyperfine field and the electric
field gradient is the same for all crystallographic equivalent sites so that there is no
additional splitting and the sites remain equivalent for both the structures. When the
magnetization direction is in the basal plane, for the rhombohedral structure, the 9d,
18h and 18f sites each splits into two groups with an intensity ratio 2:1, while in the
hexagonal structure the 6g and 12k sites split into two groups with an intensity ratio
2:1, and the 12j site splits into three groups with an intensity ratio 1:1:1. Since the
dipolar fields at two of the three groups of the 12j sites are very close, they can be
treated in the same way as the 18f site in the rhombohedral compounds by splitting
them into two groups with an intensity ratio 2:1. Thus the spectra of all the RzFelv
compounds can be fitted to seven independent sextets.
342 H. FUJII and H. SUN

In the case of the interstitial modified carbides and nitrides, it is possible that the
interstitial sites are not fully occupied. Then the 12j(18f) and 12k(18h) iron sites,
which have one neighboring octahedral interstitial site, can have either one or zero
interstitial neighbor. For most of the nitrides, as the occupancy of the interstitial
sites is nearly full, the probability for one neighbor is much higher than that for zero
neighbor. Hence it is reasonable to fit the spectra in the same way as for the host
compounds and the variation in their near-neighbor environments may be considered
as the reason for the broadening of the absorption lines. Ten subspectra were used
by Long et al. (1994) in the fitting of the Th2FelTN2.6 spectra because the easy
axis of magnetization was found to be in a general basal direction and not oriented
along one of the basal axes. For much lower interstitial content, especially in the
case of the arc-melted carbides, a further splitting of the subspectra arising from the
influence of the substoichiometric concentration has to be taken into account, which
complicated the fitting procedure. Some authors simply ignored this influences or
derived the parameters of the overall averaged hyperfine field and the distinctly
separate hyperfine field of the 4f(6c) dumbbell sites only (Zhou R.J. et al. 1991 and
1992).
As to the site assignment of the spectra, most of the analyses were based on the
hyperfine field and intensity considerations. The idea is that the strength of the hy-
perfine field on each site is predominantly determined by the number of iron and
rare earth near-neighbours of the site. The higher the number of iron-neighbours,
the larger is the hyperfine field, whereas the higher the number of rare earth neigh-
bors, the smaller is the hyperfine field. According to the above considerations,
Hu B.P. et al. (1991) decided that the hyperfine fields were in the order of 4f(6c)>
6g(9d)> 12j(18f)> 12k(18h), which agreed with by many other studies, while Kong
et al. (1993c) fitted their spectra in a different order of 4f(6c)>12k(18h)>12j(18f)>
6g(9d). Another way of spectra assignment has been used by Long et al. (1992),
who took isomer shift as a main clue for the assignment. The model was based on
the Wigner-Seitz size cell environment of each iron site and they also considered
the orientation of the magnetization and the magnetic moments as determined from
either neutron-diffraction measurements or band structure calculations. The data of
57Fe Mrssbauer hyperfine fields are summarized in table 3.7, where data for the
6g(9d), 12j(18h) and 12k(18f) sites are the weighted averages. It can be seen that
although the fitting and assignment procedures are different, the values of the overall
averaged hyperfine field (Bhf) does not differ too much. However, the individual
hyperfine field values at the various sites scatter a lot.
The overall average hyperfine field of RzFel7, RzFelvN v and R2Fel7Cy at 15 K
are plotted in fig. 3.16. The general feature is that the average hyperfine field (Bhf)
across the series increases by about 4 T in the nitrides, but it decreases slightly in the
carbides. These results could be understood by considering the different polarization
effects on Fe atoms by their nitrogen and carbon neighbors. The hyperfine field
in metals is largely due to the Fermi contact term Bs, which is proportional to the
unpaired spin density at the nucleus. The increase of the hyperfine field indicates
a larger polarization of s electrons. It is considered that the 4s band of Fe is more
highly polarized by nitrogen atoms in nitrides than by carbon atoms in carbides,
INTERSTITIALLY MODIFIED INTERMETALLICS 343

TABLE 3.7
57Fe M0ssbauer hyperfine field (in units of Tesla) of each crystallographic site (averaged over subspectra)
and the overall weighted average (Bhf) for the RzFe17 interstitial compounds at various temperatures.
The superscripts a, b and c correspond to different methods of making the carbonitrides as described in
section 2.3.4.

Compound 6g(9d) 12k(18h) 12j(18f) 4f(6c) (Bhf) T (K) Ref.


Ce.2FelTNu 36.7 31.4 34.2 38.3 34.1 15 [1]
Ce2FeI7C v 30.7 25.4 31.6 34.7 29.7 15 [2]
PrzFelvN u 35.6 30.8 33.3 37.7 33.3 15 [1]
Pr2Fe17Cv 33.5 25.1 28.5 35.4 29.0 15 [2]
Nd2Fe17Nv 36.1 30.5 33.3 38.8 33.5 15 [1]
Nd2Fe17Cy 33.7 25.1 29.0 35.2 29.2 15 [2]
Sm2Fel7N v 39.4 31.1 35.6 41.6 35.4 15 [1, 3, 4, 6]
SmzFet7Cy 36.2 26.2 30.2 36.5 30.6 15 [2]
SmzFe17 32.2 27.4 30.1 36.1 30.3 4.2 [3]
SmzFelyNo.4o 34.3 28.7 31.4 37.6 31.7 4.2 [3]
SmzFelvNo.81 37.4 30.3 34.1 41.9 34.3 4.2 [3]
Sm2Fe17N1.2o 38.2 31.4 35.4 42.3 35.3 4.2 [3]
Sm2Fe|7N1.99 39.5 31.8 35.7 42.1 35.7 4.2 [3]
Gd2Fe17Ny 36.7 31.5 34.8 38.8 34.5 15 [1]
Gd2Fe17C v 34.3 25.5 29.3 36.5 29.7 15 [2]
Gd2Fe17C2.o 25.5 31.1 32.1 36.7 31.1 12 [15]
GdzFe17 26.6 30.0 32.0 36.9 30.9 12 [15]
Gd2Fe17Co.5 27.4 31.2 32.8 37.0 31.8 12 [15]
Gd2FelvC1. 0 26.2 32.0 33.0 37.1 31.9 12 [15]
Gd2Fel7C1.5 26.0 31.5 32.5 36.6 31.4 12 [15]
Tb2Fe17Ny 37.2 31.6 34.6 39.5 34.6 15 [1]
Tb2Fe17C v 34.9 26.0 31.4 37.3 30.8 15 [2]
DyzFe17N u 37.3 32.6 35.3 40.1 35.3 15 [1]
Dy2Fel7C u 34.8 26.4 31.3 36.6 30.8 15 [2]
Ho2Fel7Nv 36.8 32.3 35.8 40.7 35.3 15 [1]
HozFe17Cy 35.1 26.4 30.9 36.5 30.7 15 [2]
Er2Fe17N v 40.7 33.5 35.9 42.5 36.8 15 [1]
ErzFe17C v 36.5 26.1 30.7 37.4 30.9 15 [2]
Er2Fel7C2.5 24.7 28.4 35.4 38.0 31.4 12 [14]
Er2Fel7 37.4 31.6 4.2 [13]
ErzFelvCo.5 37.5 32.3 4.2 [13]
Er2FeI7Co.8 41.4 33.0 4.2 [13]
Er2Fel7Cl.o 40.9 32.9 4.2 [13]
Er2Fe17C1.2 40.9 32.7 4.2 [13]
Er2Fe17C1.4 39.7 32.4 4.2 [13]
Er2Fe17C1.5 40.0 32.6 4.2 [13]
TmzFe17N u 39.9 33.1 35.3 41.7 36.2 15 [1]
Tm2FelvCu 36.4 25.3 31.1 37.4 30.7 15 [2]
LuzFe17Ny 37.1 31.7 35.5 40.4 35.0 15 [1, 6]
Lu2Fe17Cy 34.5 25.5 30.0 36.8 30.0 15 [2]
YzFe17Nv 36.8 31.6 35.2 40.2 34.8 15 [1, 4, 5]
YzFelvCy 34.7 26.2 30.7 36.2 30.5 15 [2, 5]
YzFelTH2.7 32.9 28.2 29.7 34.6 30.3 15 [5]
344 H. FUJII and H. SUN

TABLE 3.7
(Continued)

Compound 6g(9d) 12k(18h) 12j(18f) 4f(6c) (Bhf) ~/~(K) Ref.


Pr2Fel7N2. 6 37.4 31.5 31.8 35.9 33.2 85 [10]
Nd2Fel7N2. 6 37.6 32.8 30.4 36.0 33.3 78 [8, 9]
SmzFelvNv 39.7 31.1 35.7 41.7 35.5 77 [4, 6, 7]
Sm2Fel7 C:~N,~ 33.7 26.7 28.9 36.7 29.9 77 [7]
Sm2FelTC~Nv~ 33.2 28.2 31.6 39.5 31.6 77 [7]
Sm2FelTCxN~ 38.9 30.4 34.9 41.2 34.8 77 [7]
Sm2Fe17C2 33.2 25.7 27.9 37.6 29.2 77 [7]
Sm2Fe17H3.7 32.1 29.1 30.7 34.0 30.8 77 [7]
Er2Fel7C2.5 24.6 28.2 35.1 37.5 31.1 70 [14]
Lu2FelTNu 36.9 31.7 34.5 39.7 77.4 [6]
Y2FelTN2.3 35.3 31.4 34.9 40.1 34.3 77 [7]
Y2FelTC2 29.8 25.9 28.4 35.9 28.6 77 [7]
Y2FelTC~N~ 31.3 27.5 30.4 36.3 30.2 77 [7]
Y2Fel7CxNub 32.5 28.7 31.6 37.6 31.4 77 [7]
YzFelvC~N~ 34.9 30.8 33.8 39.2 33.6 77 [7]
Th2FeI7N2. 6 35.4 33.1 32.7 35.7 33.7 85 [11]
Th2FelTN2. 6 30.3 32.6 35.9 35.8 33.8 78 [12]

Ce2FelTNy 33.6 28.9 31.9 35.9 31.6 293 [1]


Ce2Fel7Cy 29.7 20.7 26.1 33.5 25.7 293 [2]
Pr2FelTNy 34.2 29.2 31.3 36.6 31.7 293 [1]
PrzFelvN2.6 35.5 30.2 29.9 34.2 31.5 295 [10]
PrzFelTCy 31.2 22.8 26.5 33.0 26.8 293 [2]
NdzFe17N u 34.1 28.7 31.2 36.9 31.5 293 [1]
Nd2Fel7N2. 6 35.9 31.4 29.0 34.3 31.7 295 [8, 9]
NdzFe17Cu 30.8 22.8 24.1 31.8 25.7 293 [2]
SmzFel7N~ 37.3 29.5 33.3 39.0 33.3 293 [1, 3, 4, 6]
Sm2Fe17 23.0 19.5 21.9 26.4 21.8 RT [3]
SmzFe17No.4o 26.5 21.4 24.2 29.1 24.2 RT [3]
Sm2FelTNo.81 30.6 23.9 27.6 36.7 27.9 RT [3]
Sm2Fe17Nl.20 33.6 26.3 30.7 37.7 30.5 RT [3]
Sm2FelvN1.99 36.8 29.0 32.9 38.9 32.9 RT [3]
Sm2FelvCu 32.8 24.1 25.0 34.1 27.1 293 [2]
Gd2FelTNy 34.6 29.6 32.0 36.7 32.2 293 [1]
Gd2Fe17C u 32.7 24.0 26.8 32.9 27.6 293 [2]
Gd2Fe17C2.o 23.8 26.9 30.3 33.6 28.3 293 [15]
Gd2Fel7 22.7 24.3 25.6 30.2 25.2 293 [15]
Gd2Fe17Co.5 24.1 26.0 27.6 32.0 26.9 293 [15]
Gd2Fe17Cl.o 24.5 27.0 29.0 32.7 28.0 293 [15]
Gd2Fe17Cl.5 23.9 27.0 29.5 32.8 28.1 293 [15]
Tb2Fel7N u 34.6 29.2 32.3 37.2 32.2 293 [1]
TbzFe17Cy 32.8 23.7 28.4 32.8 28.1 293 [2]
DyzFe17Ny 33.8 29.2 32.6 37.3 32.2 293 [1]
DyzFel7C u 31.7 23.8 28.4 33.7 28.0 293 [2]
Ho2FelTNu 33.8 29.0 32.2 37.2 31.9 293 [1]
Ho2Fe17Cy 31.7 23.5 28.4 33.0 27.8 293 [2]
Er2FelTNy 33.5 28.5 31.9 37.0 31.6 293 [1]
Er2Fel7Cy 30.4 22.8 26.9 32.6 26.7 293 [2]
INTERSTITIALLY MODIFIED INTERMETALLICS 345

TABLE 3.7
(Continued)

Compound 6g(9d) 12k(18h) 12j(18f) 4f(6c) (Bhf) T (K) Ref.


Er2Fe17C2.5 22.2 24.0 29.9 33.1 26.8 300 [14]
Tm2FelTNu 34.5 28.7 32.0 36.5 31.8 293 [1]
Tm2FelTCy 31.2 23.3 26.9 33.2 27.1 293 [2]
Lu2Fe]7N u 34.2 28.4 31.9 36.6 31.6 293 [1]
Lu2Fe17N u 32.7 28.1 31.3 35.5 300 [6]
Lu2FezTC v 31.8 23.2 27.5 33.5 27.4 293 [2]
Y2Fe]7Ny 34.0 28.5 32.2 36.9 31.8 293 [1, 5]
Y2Fe17Cu 30.5 23.1 26.1 33.3 26.7 293 [2, 5]
Y2Fe17H2.7 14.8 3.0 10.8 18.5 9.7 293 [5]
Th2FeITN2. 6 33.0 30.8 30.7 33.7 31.5 295 [11]

References:
[1] Hu B.R et al. (1991) [9] Pringle et al. (1992)
[2] Qi et al. (1991) [10] Long et al. (1993)
[3] Zhou R.J. et al. (1993) [11] Long et al. (1994)
[4] Kapusta et al. (1992) [12] Jacobs et al. (1991)
[5] Qi et al. (1992b) [13] Zhou R.J. et al. (1992)
[6] Zouganelis et al. (1991) [14] Kong et al. (1993c)
[7] Chen et al. (1993b) [15] Kong et al. (1993d)
[8] Long et al. (1992)

. o. R2Fel
. . 7 I . . . . I . . . .
38 - - x R2Fel7CY
• R2Fel7Ny • •
36
~- 3 4
o
~-32
m 30
28
(a)
26 --o
I!l!l
o R2FeI7
x R2Fel7Cy (b)
• R2Fel7NY
0.1
Q

E
E
v

0.0 x X X x N X--.
x x
x
oo oo o
o
-0.1
~ l l l l I l l l ~ l l l
CePrNd Sm Gd Tb Dy Ho Er Tm Lu

Fig. 3.16. Overall average (a) hyperfine fields and (b) isomer shifts of R2Fe]7, R2Fe17N u and
R2FelTCy at 15 K (after Hu B.R et al. 1991, Qi et al. 1991).
346 H. FUJII and H. SUN

and consequently the magnetic hyperfine field is higher in nitrides than in carbides.
Qi et al. (1992b) have attributed the difference of the hyperfine field for nitrides
and carbides to the transferred hyperfine field from interaction with neighboring
atoms, which is sensitive to the chemical nature of the interstitial impurity. The
importance of this transferred hyperfine field can also be seen from the fact that
the proportionality of the incremental moment and hyperfine field is not valid in
interstitially modified intermetallics.
57Fe Mrssbauer studies of interstitial compounds with intermediate interstitial con-
centration have been carried out for the arc-melted ErzFel7Cy (0 ~< y ~< 1.5) by
Zhou R.J. et al. (1992), for the arc-melted and melt-spun Gd2Fe17Cu by Kong et
al. (0 ~< y ~ 2.0) and for Sm2Fe17Nv (0 <~ y ~< 2.94) by Zhou R.J. et al. (1993).
Hyperfine field data for the above three series of compounds are included in table
3.7 and the data for Sm2Fe17Nv at 4.2 K are replotted in fig. 3.17. Also listed in
table 3.7 are the 77 K magnetic hyperfine fields of the carbonitrides of Y and Sm
compound made by Chen et al. (1993b), where the superscript a, b and c indicate
the different synthesizing routes, for detail see ref. [7] of table 3.7.
There is a general increase in the overall isomer shifts for the interstitially modified
R2Fe17 compounds. The increase is 0.12 mm/s in the nitrides, which is higher than
the value if considering the volume effect only. The situation in the carbides is
the opposite, the 0.05 mm/s increase in the average isomer shift is lower than the
contribution from the cell volume increase. This suggests that the interband charge

45 ' ~ 6c

40 9d _

I--
f ~ I--
• 18f
t
~'~ 35
m= 18h

30 ~ ~ ~ - ~ ~------------

25 I I
< Bhf >
I- 35 "' °
A
rn
V 30

0 1
I 2
I 3

N i t r o g e n Content, y

Fig. 3.17. Dependence of the hyperfine field at the four Fe sites and the average hyperfine field (Bhf)
on the N concentration for Sm2Fe17Ny at 4.2 K (after Rosenberg et al. 1993).
INTERSTITIALLY MODIFIED INTERMETALLICS 347

transfer occurs in the opposite sense in nitrides and carbides. As the isomer shift
increases with decreasing 4s occupation, the increase of 3d occupation will also
increase the isomer shift because the 4s electron density at the nucleus will be
reduced owing to the expansion of the 3d shell. Thus it was suggested (Qi et al.
1991) that there could be a 4s-+3d transfer in the nitrides and a smaller 3d--+4s
transfer in the carbides, which is consistent with the consideration of the average
electron density at the boundary of the Wigner-Seitz cell. The other possibility is a
greater 4s-+2p interatomic charge transfer in the nitrides, on account of the greater
electro-negativity of nitrogen than carbon. However, definite conclusion can not be
drawn from the M6ssbauer data alone.

3.4.2. Spin reorientation studies by 57Fe M6ssbauer measurements


Besides the Fermi contact term Bs, the magnetic hyperfine field Bhf also contains
an anisotropic orbital contribution Bo created by the electronic current around the
nucleus. When there is a temperature induced change of the magnetization direction
(the spin reorientation), there is a discontinuity in the temperature dependence of
the magnetic hyperfine field. The temperature at which the hyperfine field anomaly
happens corresponds to the spin reorientation temperature Tsr. Thus Tsr c a n be
determined by studying the temperature dependence of the 57Fe M6ssbauer hyperfine
field.
In the R2Fel7 compounds and their interstitial compounds, spin reorientation oc-
curs when R = Tm and Er. The hyperfine field discontinuity results mainly from
the reduction of the orbital contributions when the iron magnetic moments rotate
from the direction perpendicular to the c-axis to parallel to the c-axis with decreas-
ing temperature. The anomaly is most pronounced for the dumbbell 4f(6c) site.
An example is given in fig. 3.18, which shows the spin reorientation transition of
ErzFe17Cy. Tsr determined by this method has been proved to be in good agreement
with those obtained by a.c. susceptibility measurements and thermomagnetic scans.

3.4.3. Rare earth M6ssbauer spectroscopy and NMR studies


The principal component of the electric field gradient Vz~ at the nucleus can be de-
duced directly from M6ssbauer quadrupole splitting when the electric field gradient
tensor has axial symmetry with Vx~, = Vvv. The M6ssbauer spectroscopy of 155Gd
nuclei is of special interest because the spherical 4f shell of Gd does not itself con-
tribute to V~z at the nucleus. In other types of rare earth M6ssbauer spectroscopy,
for instance 166Er and 169Tin, the crystal field contribution to V~ can be deduced
by taking the difference of the total V~ and the free ion contribution (the asym-
metric 4f ion contribution). Besides rare earth M6ssbauer studies, NMR is another
experimental technique from which V~ can be obtained.
Within the point charge model, which is based on considering the electrostatic
charges due to surrounding ions as point charges and performing lattice summations
over a sufficient number of neighbors, the following relation was often used,

1 -- ")1oo
e V ~ z = - 4 C A °, with C - - - , (3.12)
1--o'2
348 H. FUJII and H. SUN

I I I

35 I~
O ~ Q
• ~ 4 f
i--
30
tn _, . . . .

25 ~ 2 k "
I I *1
34

I--
A 30
rn
V "~'0~0
26
I I I
0 100 200 300
T(K)
! I ! I i I ,

r-

.0

v
/
zJ
g

0
co

~xxxx~
! , I l I a I n l l
110 120 130 140
T(K)
Fig. 3.18. Hyperfine fields at the four Fe sites, the average hyperfine field (Bhf) and the a.c. magnetic
susceptibility of Er2FelvC2. 2 as a function of temperature(after Qi et al. 1992a, b). The spin reorientation
temperature is indicated.

where "Too is the Sternheimer antishielding factor and ~r2 is the screening constant.
If the semi-empirical value of C can be determined, then A ° can be calculated from
V~. Data of V~ and A ° found in the literature are summarized in table 3.8. It can
be seen that there is a nearly three-fold increase in V~ for the nitrides with respect
to the V~ value of the host compounds. If the ratio of A ° to V~, is preserved,
there is in turn a three-fold increase in the absolute value of A °. The value of V~
for Gd2ColTN v is somewhat larger than that for Gd2FelTNu. This could mean that
the rare earth sublattice anisotropy in the Co compounds at 4.2 K is larger than in
the corresponding Fe series (Mulder et al. 1992). In contrast to carbon or nitrogen
insertions, hydrogen was found to lower V~ at the Gd nuclei and reduce the crystal
field induced anisotropy.
INTERSTITIALLY MODIFIED INTERMETALLICS 349
TABLE 3.8
The electric field gradient V,z at the rare earth nucleus deduced from the M6ssbauer quadrupole
splitting and the second order crystal field parameter A ° estimated from Vzz for various R2FeI7 and
their interstitial compounds.

Compound Isotopes Vzz (1021V/m 2) A° (Kao 2) Ref.


Gd2Fel7 155Gd 4.3(1) -200 [1]
Gd2Fel7CI. 2 155Gd 9.3(3) -430 [1, 9]
Gd2FelTN3 155Gd 12.6(2) -580 [1]
Gd2Co17 155Gd 4.8 [2]
Gd2Co17N3 155Gd 14.9 [2]
Gd2Fe17 155Gd 4.02(13) -351 [3]
Gd2Fe17H3 155Gd 2.66(11) -233 [3]
Gd2FelTH5 155Gd N0 ,-~0 [3]
Er2Fel7 166Er -50 4- 100 [4, 51
Er2Fe17C 166Er 9.9 4. 1.5 -290 4. 50 [4, 5, 101
ErzFel7N2.7 166Er -400 -t- 50 [4, 5]
Tm2FeI7C 169Tm 9.7 4- 1.5 -300 4- 50 [6, 5, 10]
Tm2Fel7N2.7 169Tm -300 4- 50 [5]
Sm2Fe17 NMR 10.2 [7, 8]
Sm2Fe17N3 NMR 33.9 [7, 8]
References:
[1] Dirken et al. (1991) [6] Gubbens et al. (1989)
[2] Mulder et al. (1992) [7] Kapusta et al. (1992)
[3] Isnard et al. (1994b) [8] Kapusta et al. (1992c)
[4] Gubbens et al. (1991) [9] Dirken et al. (1989)
[5] Gubbens et al. (1992) [10] Gubbens et al. (1994)

The validity of the general proportionality relation between Vzz and A ° has been
questioned by Coehoorn and B u s c h o w (1991) from band structure calculations. Prob-
lems arise from that Vz~ (or the quadrupole splitting) measures the a s y m m e t r y of
the electric charge distribution at the rare earth nucleus, whereas A ° is a measure
of the a s y m m e t r y charge distribution experienced by the rare earth 4f electrons. Al-
though contributions from 6p and 5d electrons are of equal importance to A °, V**
is determined almost entirely by the p electron charge density. However, within a
series of structurally related compounds, experiments have shown that the A°/Vz~
(Kao2/1021Vm -2) ratio is m o r e or less a constant, which is - 4 6 + 3 for the 2:17
c o m p o u n d s (Dirken et al. 1991). It should be mentioned that A ° estimated in this
w a y has a relatively large error due to the uncertainty of the factors 7o~ and cr2. As
an example, unreasonable large value of A ° ~ - 1 0 0 0 K a o 2 has been estimated for
the Gd2FeITC v c o m p o u n d (Dirken et al. 1989, Jacobs et al. 1990).

3.5. Substitution effect


Similarly to the various substitution studies on the host R2Fel7 compounds, the
effect of substitutions for both the Fe sublattice and the R sublattice by other metal
or nonmetal elements on the structural and magnetic properties have been studied for
the R2Fe17 interstitial compounds. The results will be summarized in this section.
350 H. FUJII and H. SUN

3.5.1. Substitution of Co for Fe


The unit cell volume of R2Co17 is smaller than that of R2Fel7 and thus the spatial
size of the interstitial sites is smaller in the Co compounds, which makes it more
difficult to introduce interstitial atoms into the 2:17 lattice. Early studies on the
Y2(Fe1_xCo~)17N v series by Hurley and Coey (1991) have showed that the 2:17
nitride phase exists only when x < 0.85. They found that the suitable nitrogenation
temperature increased with the Co content x and at x > 0.85 the required nitrogena-
tion temperature became higher than the nitride phase decomposition temperature.
The amount of absorbed nitrogen atoms decreased with increasing Co content, which
changes from 2.6 for x = 0 to 1.6 for x = 0.8. Similar conclusion has be reported
by Xu and Shaheen (1993a, b) on the R2(Fel_xCo~)17Ny series with R = Ce, Pr
and Nd, where the 2:17 nitrides were formed in a limited range of x ~< 0.6. The
situation was more critical for the Ce compounds and reasonable pure nitride phase
was successfully synthesized only in the range of 0 ~< x ~< 0.3 (Xu and Shaheen
1993b).
However, later work by Katter et al. (1992b) has proved that by using lower nitro-
genation temperature (~< 450°C) and longer nitrogenation time the nitride phase can
be formed for the whole substitution range from x = 0 to 1.0 in S m 2 ( F e l - x C o x ) 1 7 N 2 . 7 ,
although the decomposition temperature of Sm2Fe17Nv is reduced by the substitu-
tion of Co and the diffusion of nitrogen in Sm2Co17 is slower than in S m 2 F e l 7 . In
fact, the nitrides R 2 C o l 7 N u of all the rare earth members have been synthesized by
Liu J.R et al. (1993) and their structural and magnetic properties were studied. Data
obtained by Liu et al. (1993), and other authors are summarized in table 3.9. After
nitrogenation, the Curie temperature is lower than that in the parent compounds.

TABLE 3.9
Crystallographic structure data and magnetic properties of R2ColTNy. V is the crystal cell volume
calculated from V = -~a2c. Saturation magnetization Ms and anisotropy field/t0Ha are data pertain
to room temperature and are expressed in units of Tesla.

R y a (A) c (]k) V (~3) A W E (%) Ms (T) Tc (K) #0Ha (T) Ref.


Ce 2.7 8.58 8.30 529.4 8.0 [1]
Pr 2.7 8.63 12.40 800.1 5.9 [1]
Nd 2.1 8.62 12.31 795.6 6.0 [1]
Sm 2.6 8.57 12.40 789.6 6.1 840 [1]
Sm 2.7 8.591 12.473 794.9 6.2 1.03 811 16.4 [2]
Sm 2.2 8.584 12.462 795.3 7.1 1.02 11.8 [3]
Gd 2.2 8.55 12.37 784.4 6.1 [1]
Tb 2.0 8.48 8.44 525.9 6.2 [1]
Dy 1.7 8.47 8.38 521.3 6.0 [1]
Ho 1.6 8.46 8.35 517.5 6.0 [1]
Er 2.4 8.42 8.35 513.7 5.8 [1]
Tm 2.5 8.42 8.35 512.2 6.4 [1]
Y 2.0 8.48 8.32 518.6 5.7 [1]

References:
[1] Liu J.R et al. (1993)
[2] Katter et al. (1992b)
[3] Hu B.R et al. (1992b)
INTERSTITIALLYMODIFIED INTERMETALLICS 351

This has been explained as a result of the weakened Co-Co exchange interaction
and the reduced Co moment. From the drastic increase of the high field differential
susceptibility at 4.2 K, Liu et al. suggested that the R-Co exchange interaction was
strongly reduced by nitrogenation.
Nitrogenation was found to have a strong influence on the magnetic anisotropies of
both the 3d and rare earth sublattices. The easy axis anisotropy range of 0.5 < z ~< 1
in Y2(Fel-xCox)17 was extended to 0.15 < z ~< 1 in Y2(Fel_xCo~)ITNu (Hurley
and Coey 1991). As Y is a nonmagnetic ion, this reflects the modification of the
interstitial nitrogen on the 3d anisotropy. Ce in Ce2Fel7 can roughly be considered
to be in a nonmagnetic state because of the strong hybridization between the Ce
4f-electron states and the ligand 3d-electron states. Experiments have demonstrated
that in Ce2(Fel-:cCox)17 the transition from easy plane to easy axis anisotropy with
increasing z is shifted from z ~ 0.5 to z ~ 0.15 upon nitrogenation, which is in
agreement with the result on the yttrium compounds (Xu and Shaheen 1993b). The
maximum of the concentration dependence of the room temperature saturation mag-
netization has been found to be located at z = 0.2 in Y2(Fel_=Coz)17Nu, which
is close to z = 0.35 in the host compounds (Hurley and Coey 1991). The com-
bined effects of nitrogenation on the anisotropy and magnetization make it possible
to maximize the saturation magnetization while simultaneously achieving uniaxial
anisotropy of the 3d sublattice by choosing appropriate z values.
Some of the results obtained by Katter et al. (1992b) in a systematic study of
the Sm2(Fe1_~Co=)17N v series are replotted in fig. 3.19. It is showed that Tc first

. . . . I . . . .
1200 Sm2(Fel.xCo,)17 o..._...--- -~
~1000 o~
v 800 .....,.~. o ~ e ~ °
v
" ~

600 o/ Srn2(Fe~-xCOx)lTNy
/
400,
I I / / I I I I [
Srn2(Fet.,eOx)lTNy
1.6

~ 1.2
--°~°~'~O~o~
0.8 I I I I ] L I I [
Sm2(Fe1.xCOx)
lTNy
1-
25. t o ~ o
" m

e ~ o ~ e ~
20
::~ 15
10 l l l l l l l r l
0.0 0.5 1.0
Nitrogen Content, y
Fig. 3.19. Composition dependence of the Curie temperatureTc, the saturation polarization Js and the
anisotropy field /~0Ha (determined by the SPD method) for Sm2(Fel_xCo=)17Ny (T = 293 K). Also
included are Tc of Sm2(Fel_=Co=)]7 (open circle) (after Katter et al. 1992b).
352 H. FUJII and H. SUN

increases with increasing Co content x reaching a maximum at x = 0.5 and then


decreases again, which is different from the monotonical increase with x in the host
series. Nitrides for the entire x range show a strong uniaxial anisotropy at room
temperature, whereas the easy axis of magnetization of the host compounds lies in
the basal plane for z < 0.45. Most excellent intrinsic magnetic properties for hard
magnetic applications were achieved for Sm2(Fe0.sCo0.2)lTN2.8 with Ors = 1.55 T,
#0Ha = 23.7 T at room temperature and Tc = 842 K.

3.5.2. Substitutions of elements other than Co for Fe


The effects of substitution for Fe of Si, Ga, A1, Ti, V, Ni and Nb etc. have been
studied by Hu B.E et al. (1992), Tang et al. (1992), Li X.W. et al. (1993), Valeanu
et al. (1994) and Middleton and Buschow (1994) on Rz(Fel-xMx)a7Ny compounds,
where M is one of the above mentioned elements. The largest stability range of the
2:17 nitride phase was reported for A1, with 0 ~< z ~< 0.4, for R = Sm (Li X.W. et
al. 1993) and with 0 <~ x < 0.6 for R = Ce (Middleton and Buschow 1994). For all
other M elements, only data with x < 0.2 have been reported.
Although many of the substitutions raise the Curie temperature of the host RzFe17
compounds, the general effect of substitution in RzFelvNu has been found to reduce
To. Exceptions are Ti and V, for which Hu B.E et al. (1992b) showed that Tc of
Sm2(Fe0.982Ti0.018)ivN2.3 is 764 K, being higher than that of Sm2Fel7N2.6 (743 K).
The Tc values of Smz(Fel_~V~)a7N v (x = 0.03 and 0.059) are almost the same as
in the un-substituted nitrides. It has been accepted that an increase in the unit cell
volume of the RzFeI7 compounds results in an increase in Tc due to the magneto-
volume effect. Nitrogenation might have maximized Tc and further substitution by
non-magnetic atoms can only act as magnetic dilution which leads to a decrease of
Tc, in spite of the additional unit cell expansion. No comments were given as to the
slight increase of Tc when M = Ti and V.
Saturation magnetizations and magnetic anisotropy fields are generally reduced by
substitutions. When M = Ti and V, either Ms or #0Ha were raised slightly. Un-
fortunately both parameters did not increase simultaneously for a given substitution
element.

3.5.3. Substitutions for rare earths


For the technical important Sm2Fe17Nu interstitial compound, replacements of Sm
by mischmetal and Ce (Huang et al. 1991b and c), by Nd (Huang et al. 1991c,
Liang et al. 1991, Yu et al. 1992, Katter et al. 1992c), by Tb (Huang et al. 1991b),
by Dy and Er (Tegus et al. 1992), and by Y (Huang et al. 1991b, Lu et al. 1992)
have been studied. The system (Sm1_xR~)2Fe17Nu exists for 0 ~ x ~< 1.0, whereas
for mischmetal the single phase region depends on the amount of La contained in
the mischmetal. As La2Fe17 does not exist, higher La concentration would results
in a narrower x range.
Modification of the Curie temperature by substitutions of other rare earths for Sm
is not significant, which is in the range of the Tc variation across the R2Fel7Ny se-
ries. Saturation magnetization Ms increases by Nd substitution owing to the higher
magnetic moment of the Nd atoms than the Sm atoms. Substitution of nonmagnetic
INTERSTITIALLYMODIFIED INTERMETALLICS 353

(Sm 1-xNdx)2(Fel-z C°z)17 N~2.7


1.0 ~l i ~ ~~ , , -
I

1.25
0.8 - ~ \\ ~ ~ - -
N

E 0.6 '-'~'~ ~ ~~ \~ 1 45
~
C
O 5
0 0.4 1.55.
O
o
; /
0.2
/ iI // I(-~
o
o 0.2 0.4 0.6 0.8 .0
Nd Content, x
Js(T) --- I.t0HA (T)
Fig. 3.20. Contour line diagram of the saturation polarization Js and the anisotropy field #0Ha for
(Sml_zNd~)2(Fel_zCoz)17Nv at room temperature (after Katter et al. 1992c).

Y also results in an increase of Ms, suggesting an antiferromagnetic Sm-Fe cou-


pling, as has been discussed in section 3.3.3. The value of Ms decreases with z
in (Sm]_~R~)zFeaTNu when R is a heavy rare earth, because of the ferrimagnetic
configuration between the heavy rare earth sublattices and the Fe sublattices.
The strong uniaxial anisotropy of Sm2Fe17N u is significantly reduced by substi-
tution of other rare earths R for Sm. When R = Y, #0Ha decreases linearly with
increasing Y concentration z. In (Sm1_~Nd=)2Fea7Nu, the easy axis anisotropy is
retained up to z ,-~ 0.6, while the uniaxial anisotropy can be preserved up to z > 0.8
in most of other rare earths. Substitution of Nd or mischmetal for Sm is of interest
because their price is much lower than that of Sm and thus there are possibilities
for the fabrication of inexpensive permanent magnets. Huang et al. showed that
(Smo.6Mm0.4)zFelvNv exhibited an anisotropy field of 6.8 T, which was close to
that of NdzFe14B, but the Curie temperature is 150°C higher than that of Nd2Fe]4B
(Huang et al. 1991c).
The (Sml_~Nd~)z(Fel_~Co~)N2.7 series were investigated by Katter et al. (1992c).
By substitution of Co for Fe, the decrease of the anisotropy field due to Nd sub-
stitution was partly balanced and the saturation magnetization and the Curie tem-
perature were further increased. The contour line diagram at room temperature for
(Sml_xNdx)z(Fel_zCoz)N2.7 is drawn in fig. 3.20. For (Sm0.vNd0.3)2(Fe0.8Coo.z)N2.7
intrinsic magnetic properties of/z0Ha = 14.8 T, Js = 1.57 T at room temperature
and Tc = 835 K were obtained.

4. Interstitial compounds of the l:12-type structure

Intermetallic compounds with the composition RFel2_xT:c have been of interest in


the field of magnetism for quite a few years. Although the binary compound RFe]2
354 H. FUJII and H. SUN

.,.-® ,
0 j'=' 0 ''=' I
I I I I°

o,,~:~ c~l,-~ I ~--~'v ~ I

R C) 2a
Fe(T) qD 8f (~)8i ~8j
N • 2b
Fig. 4.1. Unit cell of RFe12_=T=Ncompoundshaving the ThMn12-typebody centered tetragonal
structure.

does not exist, the 1:12 phase can be stabilized for T = Si, A1, Ti, V, Cr, Mo and W. It
has been well established that this series of compounds crystallizes in the tetragonal
ThMn12-type body-centered tetragonal structure (space group I4/mmm), which is
directly related to the hexagonal structure of the R2Fe17 compounds (Mose et al.
1988). As shown in fig. 4.1, there is one single crystallographic site (2a) occupied by
rare earth atoms and there are three different sites (8i, 8j and 8f) occupied by Fe and
T elements. Among various RFe12_=T= compounds, SmFe11Ti has been considered
to be a suitable candidate for potential permanent magnet applications. However, it
can not compete with Nd2Fe14B owing to the following practical reasons: Although
its Curie temperature is close to that of Nd2Fe14B, its saturation magnetization is
substantially lower than that of Nd2Fel4B, which leads to a much lower theoretical
maximum energy product. In addition Sm is a much more expensive rare earth than
Nd.
Stimulated by the success in the 2:17 series, interstitial modification studies were
naturally extended to the 1:12 materials. Similarly to the case of the 2:IT-type
intermetallic compounds, interstitial modifications lead to remarkable changes on the
magnetic properties also in the l:12-type materials. Upon introducing N or C atoms
into the tetragonal structure, the Curie temperature increases by about 200 K and
the second order crystal field parameter A ° changes its sign from weakly negative to
relatively strongly positive. This leads to a strong uniaxial anisotropy in compounds
containing Nd, for which the second order Stevens coefficients c~j is negative. In
this section, we will summarize the experimental results on interstitial modifications
o f the 1:12 compounds.

4.1. Location of N atoms in the tetragonal structure


It appears to be more difficult to make fully nitrogenated or carbonated 1:12 intersti-
tial compounds than 2:17 interstitial compounds. The partially interstitially-modified
INTERSTITIALLY MODIFIED INTERMETALLICS 355

materials consist of compounds with a range of interstitial atom contents. Conse-


quently there are significant spreads in the lattice parameters and the magnetic prop-
erties observed by various researchers. The structural and magnetic inhomogeneities
broaden the neutron diffraction peaks and reduce the reliability of the crystallographic
parameters deduced from them. Fortunately, reasonably high quality (homogeneous
phase) interstitial compounds can be obtained for T = Mo. Therefore most of the
neutron diffraction studies have been done for RFe12_~Mo~ (R = Nd and Y). The
lattice parameters and the occupancy of the 2b interstitial sites deduced for various
compounds are listed in table 4.1. Besides neutron diffraction, site occupation of the
nitrogen atoms in SmFex0MozNu has been studied by Psycharis et al. (1991) using
X-ray diffraction method.
All the studies led to the same conclusion that nitrogen atoms are located in the
octahedral 2b interstitial sites, which are equivalent to the 9e or 6h sites in the 2:17
structures. Full occupation of this 2b site leads to the composition RFe12_~T~Na.
Upon nitrogenation, the crystallographic symmetry of the host is retained, but the
lattice is expanded by more than 3%. Two nitrogen atoms become the nearest
neighbors of the rare earth site, which is responsible for the radical changes of the
magnetocrystalline anisotropy. Most of the results indicated that Ti and Mo shared
the 8i site with Fe, while Sun et al. (1993b) obtained a better fits of the diffraction
pattern when placing a small amount of Mo on the 8f sites. Vanadium atoms were
reported to be located on the 8j site by Yelon and Hadjipanayis (1992).
In many cases, the experimentally estimated nitrogen content is more than 1 ni-
trogen atom per formula unit. As decomposition takes place concurrently with ni-
trogenation, most of the authors believed that the extra amount of nitrogen is con-
tained in the decomposition products consisting of rare earth nitrides or iron nitrides,
whereas Suzuki et al. (1992c) suggested that not only the 2b site but also other
sites are occupied by nitrogen atoms. A neutron diffraction study of the nitride
NdFesCo3Ny performed by Fujii et al. (1995b) showed that the nitrogen occupancy

TABLE 4.1
Nitrogen occupancy y of the 2b interstitial site in interstitial nitrides with the tetragonal ThMnl2-
type structure, a, c and V are the lattice parameters and tetragonal cell volume. All the data
were deduced from room temperature neutron diffraction diagrams.

Compound a (,~) c (]~) V (]~3) Y Ref.


YFell TiNy 8.611 4.821 0.5338 [1]
YFel0V2Ny 8.5436 4.7834 349.16 0.44 [2]
YFel0.zMol.sNu 8.646 4.8519 362.69 0.8 [5]
YFel0.6Mol.4Nu 8.6648 4.8012 360.47 0.8 [5]
YFellMoN u 8.62925 4.79663 357.18 0.96 [4]
YFe9Mo3Ny 8.7022 4.82734 365.57 0.94 [4]
NdFel0Mo2Ny 8.6593 4.8295 362.13 0.48 [2]
NdFel0MozNy 8.660 4.869 365.2 0.49 [3]

References:
[1] Yang Y.C. et al. (19910 [4] Sun et al. (1993b)
[2] Yelon and Hadjipanayis (1992) [5] Tomey et al. (1993)
[3] Wang et al. (1993a)
356 H. FUJII and H. SUN

of the 2b site is only 0.5 although the nominal nitrogen content was y > 1.0. Con-
sidering the periodicity loss along the crystallographic c-axis when y > 1.0, Fujii et
al. suggested that the remaining N-atoms occupy some of the tetrahedral interstitial
sites near Ti, possibly the 16m or 8h site.

4.2. Structural and intrinsic magnetic properties


Interstitial modification of RFea2_xT~ with T = Si, Ti, Mo and V has been reported.
The structure and magnetic properties of the interstitially modified RFelz_zTxZy
(Z = N and C) compounds are summarized in table 4.2. Combining high temperature
annealing (1100°C) with water quenching, Fujii and coworkers (Fujii et al. 1992b,
Akayama et al. 1994) prepared for the first time the RFellTi compounds with
R = Ce and Pr. It had been reported so far to be difficult or impossible to synthesize
them. Also included in the table 4.2 are the data of RFel0Si2Cy compounds formed
by arc-melting. Although carbon atoms were introduced into the interstitial sites
of RFeloSi2 by melting, the amount was much less than that obtained by gas-solid
reactions and the influences were smaller (Xu X.F. et al. 1994).
The structure and the magnetic properties of RFe12_~T~Zv are sensitive not only
to the interstitial atom content y but also to the concentration z of the stabiliz-
ing element T, as shown in the studies on structure and magnetic properties of
NdFe12_xMoxNv (Wang et al. 1993b). As both :c and y can differ from the nominal
concentration due to the compositional changes occurring during sample prepara-
tions, the data listed in table 4.2 scatter over quite a wide range. Nevertheless, the
modification effects due to interstitial atom additions are significant and clear.

4.2.1. Curie temperature


The averaged increase in Curie temperature is about 200 K for the nitrides and 160 K
for the GIM carbides. The mechanism has been understood to be a volume effect,
as in the 2:17 series: the unit cell volume is expanded by interstitial modification,
which induced an enhancement of the Fe-Fe exchange interaction, and the Curie
temperature was greatly increased. An almost linear relationship between the Curie
temperature and the corresponding cell volume for YFeaaTiC v and RFe1~TiNv has
been reported by Li Z.W. et al. (1993) and Akayama et al. (1994), respectively.
Based on the mean-field theory, the Fe-Fe interaction was estimated to be 35%
stronger in RFe11TiNy than in RFelaTi, which corresponds to an increase from
5.5x 10 -22 J in the hosts to 7.4x 10 -22 J in the nitrides (Li Z.W. et al. 1992). Some of
the data taken from table 4.2 are plotted in fig. 4.2. The familiar variation of Tc across
the rare-earth series showing maximum at R = Gd is not always demonstrated in
the interstitial modified compounds. This could be related to the different interstitial
content for different rare earths.

4.2.2. Magnetic anisotropy


In the host RFel2-xTx, when z is small, the easy magnetization of the Fe sublattice
is along the crystallographic e-axis. If we consider only the lowest order crystal
field interaction, the rare-earth sublattice anisotropy prefers the e-axis for R = Sm,
INTERSTITIALLY MODIFIED INTERMETALLICS 357

n
r~

>.=
~Q

~ Q

t'xl ~ 0

~<

t'~ ,.

tr). w~ tt~ q'~ tm tr~


ZZZZrJ o d ~ ~ ~ ~ d ~
ZZ~ Z Z Z rJrJ ZZ~..
©

ee zzzzzz r~
358 H. FUJII and H. SUN

~ ~ ~ - ~ ~ - : : ~ P ~ @ ~

~ Z
~ ~ Z z ~Z

~zz~ ~ ~ ~ ~ ~ - ~
INTERSTITIALLY MODIFIED INTERMETALLICS 359

r-..: ,::5 .ff = o ' - =

oq. oq.~

~
o

z ~ ~z~z~z~z z Z ~ z ~ z z~
_ - ~ z z z¢qz z¢q ¢q ¢q ¢q
~ z ~ z~.~ r q O~ z c.i ~9_ ~ z¢q o, - iZ ~ z . f-~
~ z ~d ~ ~~ z z
,,-, ~ ~ ~ ~ . ~ ,..,~ ~ ' ~ , . a ";z ~'~ o~ o o
0 0 0 0 © 0 O 0 0 0 0 ~ 0 0 0 0 ~ ~ 0",
360 H. FUJII and H. SUN

O~ oo

X~ ..~

w
o ~ u~ o 0o un

Cxl
.4 t ~ l ~ 0 O t ~ O 0 [ ~ 0 0 [ ' ~ O 0 o O [ ' ~ O O P ' ~ t ~ l ~ t ~ ~'31¢h o

d M N M M N N N ~ N N d N d N M d d d~ d

NNNN~

z~ ~ z ~ ~ z ~~ z z~ ~ z z ~ ~ ~ oo

t
INTERSTITIALLY MODIFIED INTERMETALLICS 361

. . . . I . . . . I . . . . I

• RFe~ITi
800 • RFe~TiNy
* RFellTiCy
,~ RFeloM%Ny
• .-mm~m

700

k
600

500
"\/
400

, I , , , , [ . . . . I

Y Pr Pm Eu To Ho Tm Lu
Ce Nd Sm Gd Dy Er Yb

Fig. 4.2. Curie temperature Tc across the rare earth series for RFellTi (Hu B.P. et al. 1989 and
Akayama et al. 1994), RFea0MozN v (Anagnostou et al. 1991b), RFel]TiN v (Yang Y.C. et al. 1991d)
and RFellTiC v (Hurley and Coey 1992).

Er and Tm whose second order Stevens coefficients ad are positive because the
second order crystal field parameter A° is negative. For RFellTi, all the members
show easy axis anisotropy at room temperature due to the preponderant contribution
of the iron sublattice, except Pr and Tb (Hu B.E et al. 1989, Akayama et al. 1994).
At low temperatures, spin reorientation takes place in the compounds with R = Nd,
Tb, Dy as a result of the competition between the Fe and rare earth sublattices and
in the compound with R = Er also due to the effect of high order terms of the crystal
field parameters.
The Fe sublattice anisotropy remains uniaxial after interstitial modification, al-
though it is slightly weakened by carbonation (Hurley and Coey 1992). Drastic
changes take place in the rare earth sublattice anisotropy. In the 1:12 interstitial com-
pounds, the rare earth atoms have two nearest interstitial neighbors located along the
c-axis direction instead of within the plane configuration perpendicular to the c-axis
in the 2:17 interstitial compounds. While the in-plane interstitials have a strong nega-
tive contribution to A ° in the 2:17 compounds, interstitials along the c-axis contribute
positively to A2° in the 1:12 structure. Consequently, A2° is changed from negative in
the host compounds to positive in the interstitially modified compounds and there-
fore the rare earth sublattice anisotropy changes completely. In the 1:12 interstitial
compounds, Sm, Er and Tm have easy-plane anisotropy and all the other magnetic
rare earths should have easy-axis anisotropy. X-ray diffraction patterns on aligned
powders of SmFellTi and its nitride (Coey and Otani 1991) demonstrate clearly the
362 H. FUJII and H. SUN

effect of interstitial nitrogen atoms on the magnetic anisotropy (fig. 4.3). Only the
Sm compound shows easy-plane anisotropy at room temperature. Spin reorientations
at 45 K for ErFealTiNy and 50 K for ErFel0.sMol.sNv, have been reported by Yang
Y.C. et al. (1991c, 1993). It is likely that also the Tm compounds will exhibit
spin reorientations as a function of temperature or interstitial atom content when the
easy-axis anisotropy of the iron sublattice is outweighted by the opposing rare earth
contribution. However, no publications were found for the spin reorientation studies
of the Tm compounds.
Using the single-ion model, Yang Y.C. et al. (1991e) estimated A° = 85 Kao 2
for RFealTiNu in contrast to A° = - 3 0 Kao 2 for RFel~Ti. The value of A° in
NdFe11TiNywas calculated by Li and Cadogan (1992c) using their 'bonding charge
model' and A° = 170 Kao 2 was estimated for NdFe11TiN0.5. They also predicted an
A ° value of 370 Kao 2 for NdFe11TiN1.0, which corresponds to an anisotropy field
of 13 T at room temperature.
Technically important materials are the Nd compounds. In fact, most of the studies
have been devoted to them as can be seen from table 4.2. Although the Mo or V
containing materials might have higher anisotropy fields than the Ti containing com-
pounds, their magnetizations are low. Hence NdFellTiN v is still the most promising

r 1 W"'--

002

SmFellTi

004

,.s

40O

SmFellTiN0.8

40 60 80
Two Theta (degree)
Fig. 4.3. Cu Ks X-ray diffraction patterns of SmFellTi and SmFellTiN0.8 oriented in an applied
magnetic field of 1.2 T.
INTERSTITIALLY MODIFIED INTERMETALLICS 363

material. Of equal importance is PrFellTiN v as reported by Fujii et al. (1992b)


and Akayama et al. (1994). We can say, that when compared with Nd2Fe14B, their
anisotropy fields are about the same, their magnetizations are slightly lower, but their
Curie temperatures are much higher.

4.2.3. 57FeMOssbauer effect and magnetization studies


Although there are only three different crystallographic iron sites in the 1:12 structure,
the subspectra needed for fitting the 57Fe Mrssbauer spectra are more than three as
Fe in a given crystallographic site can have different nearest neighbor environments
because of the existence of the T atoms. The situation becomes more complicated
when nitrogen or carbon occupy the 2b interstitial sites partially, which is true in
most of the cases. Details of the fits and the number of the subspectra used by
various authors were different, but the order of Bhf(8i) > Bhf(8j) > Bhf(8f) was
generally adapted for both the host and interstitial compounds (Li Z.W. et al. 1992
and 1993, Qi et al. 1992c, Ge et al. 1992, Sun J.J. et al. 1993, Wang Y.Z. et al.
1993a).
The room temperature hyperfine fields of the nitrides and carbides of YFe11Ti are
listed in table 4.3. The agreement between data from different sources is not a very
good one, even for the overall averaged hyperfine fields. Probably the composition
deviation is the main reason for the data scattering. The 3d sublattice magnetization
of YFellTiZy (Z = N and C) was investigated in detail by Qi et al. (1992c).
Magnetization and 57Fe hyperfine fields were measured and compared with the results
of band structure calculations. The magnetization at 4.2 K increases by about 14%
for both the nitride and carbide, which was considered to be essentially due to a
narrowing of the 3d-band caused by the volume expansion. The chemical effect
of N or C was considered to reduce the iron moment slightly by hybridization,
especially on their nearest neighbor 8j-sites. The average hyperfine field increases
by 12% for the nitride, whereas it hardly changes for the carbide (1%) although the
volume expansion and moment increases are virtually identical for the nitride and
carbide. This was understood originating from differences in the contribution due to
the transferred hyperfine field from neighboring atoms.
TABLE 4.3
57Fe M0ssbauer hyperfine field (in units of Tesla) of each crystallographic site (averaged over
subspectra) and the overall weighted average (Bhf) for the YFe11Ti nitrides and carbides at room
temperature.

Compound 8i 8j 8f (Bhf) Ref.


YFe11TiNo.8 32.9 29.9 25.1 29.0 [1]
YFel 1TiNu 35.5 32.8 28.7 32.0 [2]
YFel 1TiNy 32.7 29.8 24.7 28.7 [3 ]
YFel 1TiCo.9 28.9 24.9 21.1 25.1 [ 1]
YFel 1TiCo.7 30.2 [4]

References:
[1] Qi et al. (1992c) [3] Ge et al. (1992)
[2] Li Z.W. et al. (1992) [4] Li Z.W. et al. (1993)
364 H. FUJII and H. SUN

4.3. Substitution studies


Effects of substitution of Co for Fe have been studied for RFell_~Co~TiNy (R = Nd,
Pr) (Suzuki et al. 1992c) and NdFel0.v_xCo~Til.3N u (Kim et al. 1992). The nitrogen
absorption ability decreases with increasing Co concentration because the crystal
lattice contracts as the Co content x increases. The nitrogen content y nearly vanishes
when x > 5. Similar properties have been observed in the 2:17 compounds. The
anisotropy field #oHa is not significantly influenced by Co substitution, but the
Curie temperature and magnetization increase. The best magnetic properties have
been obtained for NdFesCo3TiN1.3, for which Is = 1.50 T, #0Ha = 7.5 T at room
temperature and Tc > 600°C (Suzuki et al. 1992c).
Substitution of Dy for Nd in Ndl_~Dy~Fe11TiN v was studied by Kong et al.
(1993e). They showed that the nitrogen content y and the magnetic anisotropy field
at room temperature increase by Dy substitution for Nd. Compounds with a small
Dy concentration showed better intrinsic magnetic properties than the unsubstituted
nitride.

4.4. Interstitial modification study on compounds with various structures


Stimulated by the success of interstitial modification of intermetallic compounds with
the 2:17 or l:12-type structures, similar studies have been extended to materials with
other crystal structure types. Published research works on the interstitial modification
of various compounds are summarized in table 4.4.

4.4.1. R-rich compounds


It is difficult to avoid small amounts of Sm rich impurity phases (SmFe2 and SmFe3)
when preparing SmzFel7 because of the peritectic formation of the SmzFel7 com-
pound in the binary Sm-Fe phase diagram. The nitrogen absorption behaviors of
SmFe2 and SmFe3 have been investigated by several groups (Christodoulou and
Takeshita et al. 1993 a and 1993b, Ishizaka et al. 1993, Yau et al. 1993). Un-
fortunately, all the results showed that SmFe2 and SmFe3 reacted with nitrogen to
form the decomposition products SmN and a-Fe, without formation of any inter-
stitial nitrides. The produced soft magnetic a-Fe phase deteriorates the magnetic
properties of Sm2FelyNy magnets. Immediate decomposition by nitrogenation was
also observed in the Nd5Fea7 compound, forming NdN and a-Fe (Helmolt et al.
1993). It has been suggested that in intermetallic rare earth-iron compounds with
a high rare earth content, the formation of the metastable nitride is impeded by the
high Fe diffusivity, but the stable equilibrium phases are formed quickly.
Nitrogenation and carbonation of RFe2 with R = Y, Tb and Dy were studied
by Singleton et al. (1993). The nitrides and carbides were formed at low heat
treatment temperature (lower than 350°C) and the crystal lattices expanded slightly.
However, the Curie temperature hardly changed. Similar results were obtained by
Yamamoto I. et al. (1993) in the nitrogenation of YFe2, YFe3 and Y6Fe23. These
authors even observed a negative volume expansion for the 1:2 and 1:3 compounds
INTERSTITIALLYMODIFIED INTERMETALLICS 365

TABLE 4,4
Various rare earth-transition metal intermetallic compounds that have been subjected to the interstitial
modification studies.
Formula Composition Structure type Space group Ref.
RFe2 Sm, Y, Tb, Dy MgCu2,fcc Laves phase Fd3m [1--4]
RFe3 Sm, Y, Tb PuNi3, rhombohedral R3m [2, 4-7]
GeFe3 Ni3Sn, hcp (high temperature) P63/mmc [8]
Cu3Au, fcc (low temperature) Pm3m [8]
RsFe17 Nd hexagonal P63/mcm [9]
R6Fe23 Y Th6Mn23, fcc Fm3m [4]
RFe5 Y, Sm CaCus, hexagonal P6/mmm [10]
R2FeI4B Y, Nd, Dy tetragonal P42/mnm [11-14]
Nd2(Fe,Ti)19 Nd TbCuT, hexagonal P6/mmm [15]
RFeloSi Ce, Pr, Nd, Sm, Y BaCdlb tetragonal I41/amd [16]
La(COl_xFe~)t3 0 <~ z ~< 0.4 NaZnl3, fcc Fm3c [17]
LaFel3_zSi~ 1.5 <~ • ~< 2.4 NaZnl3,fcc Fm3c [18]
3.6 ~< x <~ 5.0 Ce2Nil7Si 9, bct I4/mcm [18]
References:
[1] Christodoulou and Takeshita (1993b) [10] Yang EM. et al. (1992)
[2] Ishizaka et al. (1993) [11] Zhang X.D. et al. (1992)
[3] Singleton et al. (1993) [12] Onneby et al. (1993)
[41 Yamamoto I. et al. (1993) [13] Leccabue et al. (1992)
[51 Yau et al. (1993) [14] Barrett et al. (1993)
[6] Cheng et at. (1993) [15] Cadogan et al. (1993)
[7] Christodoulou and Takeshita (1993a) [16] Singleton and Hadjipanayis (1993)
[8] Xu Y. et al. (1992) [17] Huang et al. (1992)
[9] Helmolt et al. (1993) [18] Tang et al. (1993c)

after nitrogenation, which could be related to the change of magnetostriction and


thermal expansion properties.
Nitrogenation of the highly magnetostrictive material TbFe3 was studied by Cheng
et al. (1993). Nitrogen absorption was found to be an amorphization process in which
the nitrogen atoms diffused into the TbFe3 lattice but did not enter the interstitial
sites of the structure. The amount of nitrogen absorbed in the amorphous phase was
a function of the annealing time.

4.4.2. Nd2Fe14B-type compounds


N d - F e - B magnets have drawbacks like the poor thermal stability and inferior cor-
rosion resistance. If similar effects on the Curie temperature as in 2:17 and 1:12
compounds could be realized in the 2:14:1 compounds by interstitial modification,
the thermal stability of Nd2Fel4B could be greatly improved. The only successful
formation of interstitial nitrides, YzFet4BNy and NdzFel4BN v (y N 0.3), has been
reported by Yang Y.C. and coworkers (Yang Y.C. 1991g, Zhang X.D. et al. 1992).
Upon nitrogenation, the tetragonal structure of the R2Fe14B compounds was retained,
but the unit cell volume slightly increased. While Tc increased by about 60 K, the
magnetocrystalline anisotropy and the saturation magnetization decreased somewhat.
Neutron diffraction study showed that nitrogen atoms occupy the 4f interstitial sites
in the tetragonal structure, and full occupation would correspond to the composition
RzFelaBN.
366 H. FUJII and H. SUN

However, the above work has been criticized by Barrett et al. (1993). In a detailed
study of the effects of nitrogen uptake in Y2FeI4B compound, they observed only
a decomposition of the starting alloy into c~-Fe and RX v. Conclusions drawn by
Barrett et al. (1993) in Y2Fel4B are in agreement with those obtained by ()nneby et
al. (1993) on NdzFel4B and by Leccabue et al. (1992) on DyzFe14B. Thus it is still
an open question whether RzFel4BNu quaternary phases exists or not.

4.4.3. CaCus-type related structures


Compounds with the composition R10Fe79Si11 (R = Y and Sm) and their nitrides
were studied by Yang EM. et al. (1992). The X-ray diffraction patterns were
indexed on the basis of the CaCu5 structure type. After nitrogenation, the unit
cell volume expanded by more than 6% and the Curie temperature increased by
108 K and 127 K for the Sm and Y compounds, respectively. Furthermore the easy
magnetization direction of the Sm compound changed from planar to uniaxial upon
nitrogenation with a room temperature anisotropy field of 12.5 T. Sm10Fev9Si11Ny
shows interesting intrinsic magnetic properties for permanent magnet applications.
A new ternary phase Nd2(Fe,Ti)19 has been reported by Cadogan et al. (1993).
X-ray diffraction and Mrssbauer data indicated that the 2-19 phase is related to the
hexagonal TbCu7 superlattice structure based on the CaCu5 structure. Nitrogenation
induced a volume expansion of 5.4% and a Curie temperature increase of about
200 K. The average 57Fe hyperfine field at room temperature became enhanced by
42% primarily due to the increase in Curie temperature. Later, the exact composition
of this phase was determined to be Nd3(Fe,Ti)29, which crystallizes in a monoclinic
structure of the space group P21/c (No.14) (Li H.S. et al. 1994, Hu Z. and Yelon
1994, Yang EM. et al. 1994).

4.4.4. Compounds with the NaZn13-type structure


Two systems have been studied for compounds of the 1:13 structure, namely the
La(Col_~Fex)13 (0 <~ x ~< 0.4) series (Huang et al. 1992) and the LaFel3_~Si~
(1.5 ~< x ~< 5.0) series (Tang et al. 1993a, b, c). La(COl_zFex)13 absorbs substantial
amounts of nitrogen and the unit cell volume increases by 8.3-8.9%. However, both
the Curie temperature and the magnetic moment decrease and therefore no beneficial
effects were obtained by nitrogenation. The LaFe13_xSi~ compounds were found
to have two types of crystal structure; the fcc NaZn13-type for x ~< 2.4 and the bct
Ce2NilTSi9-type for 3.5 ~< x ~< 5.0. For the cubic phase, nitrogenation expands the
unit cell volume by 6.2-10.8% and this cell expansion led to an increase of the
Curie temperature by 52 to 80 K. At the same time the spontaneous magnetization
increases by ~10%. On the other hand, the structural and magnetic properties of the
tetragonal phase were not significantly affected by nitrogenation.

4.4.5. RFeloSiCv and the Fe3GeNu compounds


Using the GIM process, Singleton and Hadjipanayis (1993) reported the formation
of the BaCdll-type structure compounds RFel0SiC u with R = Ce, Pr, Nd, Sm and Y,
where the C atoms were known to fill the octahedral voids in the tetragonal cell. This
INTERSTITIALLYMODIFIEDINTERMETALLICS 367

success demonstrated that the GIM method can become a new route for synthesizing
compounds not readily formed by standard metallurgical techniques.
The simple transition metal-based Fe3Ge and the nitrides were studied by Xu Y.
et al. (1992). Fe3Ge crystallizes in two different type of structures. The high
temperature phase exhibits a hcp crystal structure of the Ni3Sn-type with the space
group P63/mmc and the low temperature phase is the face centered cubic Cu3Au-type
structure with the space group Pm3m. The fcc phase absorbs nitrogen up to z = 0.24
in Fe3GeNy, the cubic structure being retained, but the effect of nitrogenation is to
decrease both Tc and the room temperature magnetization. The hcp phase transforms
to either the fcc or the tetragonal phase or a mixture of them upon nitrogenation.

5. Electronic band structure calculations

As described in sections 3 and 4, experimental results have clearly indicated that


the structural and magnetic properties of the rare earth-iron intermetallic compounds
can be strongly modified by the introduction of nitrogen or carbon atoms into the
interstitial crystallographic sites. Interstitial modifications of the binary RzFe17 and
the psuedo-binary RFelz_:~Tx are of particular interest and have been studied in great
details. In order to understand the basic physical origin of the strong influences of
the interstitial atoms on the various aspects of magnetic properties, electronic band
structure calculations have been carried out for the RzF17Z3 and RFel2_xTxZa (Z = C
and N) interstitial compounds by many research groups.

5.1. Calculations of the 2:17-type interstitial compounds R2Fel7Z3 (Z=N or C)


5.1.1. Compounds of the nonmagnetic rare earth element Y
The calculated magnetic moments on each atomic sites of YzFel7 and Y2Fe17Z3
(Z = C and N) and the methods used for the calculations are summarized in table
5.1a. For comparison, the experimental results on the magnetic moments are also
given there.
The non self-consistent spin-polarized electronic energy band calculations using
the orthogonal linear combination of atomic orbital (OLCAO) approximation were
carried out for hexagonal Y2Fel7 and its nitride YzFel7N3 (Li Y.E et al. 1991), with
the nitrogen atoms on the 6h-site. The calculated total moments for YzFel7 and the
nitride are qualitatively in agreement with the values obtained by the magnetization
measurement at 4.2 K. Although the calculation is not self-consistent, general feature
of the influence of the interstitial N on the electronic band structures is clarified. In
this calculation, it was stressed that the enhancements of the Fe average moments
and Curie temperatures mainly come from volume expansion due to interstitial mod-
ification.
Beuerle et al. (1991) first calculated self-consistently the local magnetic moments
on all Fe and Y sites in Y2Fel7 and its expanded form as would be obtained by
nitrogenation, but without nitrogen atoms on the interstitial site. The self-consistent
tight-binding, linear-muffin-tin-orbital method (LMTO) in the atomic-sphere approx-
imation (ASA), the so-called self-consistent LMTO-ASA method, was used and the
368 H. FUJH and H. SUN

TABLE 5. la
Calculated and measured values of the magnetic moments on different atomic sites and of the total
moment per formula unit/>r in Y2Fe17 and Y2FelTZ3 (Z = N or C) with the hexagonal Th2Ni17-type
structure. #z is the magnetic moment of the interstitial atom Z.

Compound lAR (lAB) lAFe(lAB) lAZ (lAB) tAT (lAB/f'u') Method Ref.
2b 2d 4f 12j 12k 6g
Y2Fe17 -0.41 -0.42 2.41 2.35 2.12 1.91 36.8 scLMTO-ASA[1]
-0.47 -0.45 2.53 2.25 2.00 1.95 35.5 scLMTO [2]
-0.25 -0.23 2.96 2.23 2.14 1.78 37.0 nscOLCAO [3]
Y2Fel7(exp.) 32.8 MOssbauer [4]
34.7 Magnetization [5]
Y2FelTN3 -0.20 -0.45 2.65 2.01 2.57 2.53 -0.06 40.3 scLMTO [2]
-0.06 -0.09 3.46 2 . 2 0 1.69 3.11 39.2 nscOLCAO [3]
2.50 2.08 2.32 2.66 39.4 scLMTO-ASA[7]
Y2Fel7Ny(exp.) 38.1 MOssbauer [4]
40.3 Magnetization [5]
Lu2Fe17(exp.) 2.47 2.07 1.87 2.23 35.3 M0ssbauer [6]
Lu2Fe17Nv(exp.) 2.69 2.35 2.17 2.51 40.0 M0ssbauer [6]
YzFe17C3 2.42 1.65 2.06 2.58 34.8 scLMTO-ASA[7]
Y2FelTCz(exp.) 36.1 M0ssbauer [4]
References:
[1] Beuerle et al. (1991) [5] Huang et at. (1990)
[2] Jaswal et at. (1991) [6] Zouganelis et al. (1991)
[3] Li Y.P. et al. (1991) [7] Beuerle and Fahnle (1992)
[4] Qi et al. (1992)

calculated results showed that the Fe moments on all the sites at 0 K increased
by about 10% due to the experimentally observed volume expansion. Extending
the above study, these authors calculated the local magnetic moments and hyperfine
fields on the Fe, Y and Z atoms in Y2Fe17 and Y2Fel7Z3 with Z = H, C or N
by means of the local-spin-density approximation and self-consistent LMTO-ASA
method (Beuerle and Fahnle 1992). They have found that there are two counteracting
effects due to the interstitial atoms, namely a geometrical effect (volume expansion
and local relaxation) which increases the Fe moments and the hyperfine fields, and
the hybridization effects of the Z atoms with the neighboring Fe atoms, which re-
duce the Fe moments and hyperfine fields. The calculated moments and average
hyperfine fields are slightly larger for the nitrogenated system, but they are smaller
for the carbonated system than those for the pure Y2Fe17. This is not in agreement
with experimental observations, as can be seen in table 5.1a. This may be related to
stoichiometric deviations of the carbon content in the compound. Since the carbon
content does not experimentally exceed y = 2.0 in Y2Fel7Cy, the geometrical ef-
fect rather than the hybridization effects may strongly contribute to the experimental
enhancement of #T- Theoretically, it has been deduced that, when y reaches 3.0,
the magnetic moments in Y2FeI7C3 becomes smaller than in the host compound
Y2Fe17 because of the strong hybridizations between C p- and Fe d-states. In their
calculations, the ratios of the atomic sphere radii were chosen as ry/rFe = 1.35 and
rz/rFe = 0.7.
INTERSTITIALLY MODIFIED INTERMETALLICS 369

Furthermore, F~ihnle and Beuerle (1993) suggested that fluorine could be the opti-
mum interstitial dopant atom for achieving large magnetic moments in R2Fe17. The
reason for this is that, when going from N to O or even to F, the potential around
the dopant atoms becomes steeper, the p-state appears at lower energies and the
hybridization with the Fe states becomes weaker. As a result, it was deduced that
the geometrical effect due to the increase of the unit cell volume becomes essential,
leading to a strong increase in magnetic moments and high Curie temperatures. In
addition, it has been pointed out that the strong electronegativity of the F atom pos-
sibly induces a highly aspherical valence charge density around the R atom, resulting
in large crystal electric field (CEF) parameters and strong rare earth contribution to
the magnetocrystalline anisotropy.
Based on the semi-relativistic LMTO method in the local-density and ASA ap-
proximations, Jaswal et al. (1991) and Jaswal (1992) conducted the self-consistent
spin-polarized electronic structure calculations for the hexagonal Y 2 F e l 7 and the
nitride Y2Fe17N3. The lattice constants which were used in the calculations were
the experimental obtained values, a = 8.48 (8.66) A and c = 8.26 (8.51) ,~ for
Y 2 F e l 7 ( Y 2 F e l 7 N 3 ) respectively. The ratio of the Wigner-Seitz radii was chosen
as ry : rFe : r N = 1.4 : 1.0 : 0.75 for the Y, Fe and N atoms. The spin-polarized

Y2Fe~7'
' '
/'
A ' A '
o v \

0. . . . . . .

_._

I * I * I I I
-10 -8 -6 -4 -2 0 2
Energy ( eV )

_ (~)) I I ' ^ / ' ~ 1I I

E
O Y2Fe~7N3 /~./ ~
> SpinUp / - \i
• 0

~ F eY SpinDown ~ V
o -1 ........N
D
, I I I I * I ~ I
-10 -8 -6 -4 -2 0 2
Energy ( eV )
Fig. 5.1. Spin-polarized partial density of states for (a) Y2Fel7 and (b) Y2FelTN3 calculated by using
the linear-muffin-tin-orbitals (LMTO) method in the scalar-relativistic approximation (after Jaswal et al.
1991, Jaswal 1992).
370 H. FUJII and H. SUN

partial densities of states (DOS) for Y2Fel7 and the nitride are given in figs 5.1 (a)
and (b), respectively. As expected, the DOS are dominated by the Fe d-bands near
Fermi level. Upon nitrogenation, the Fe d-bands shift to the higher binding energy
side owing to the reduction of the overlap between the Fe d-wave functions and
hybridizations between Fe d- and N p-states. This leads to a decrease in the values
of both the up and down-spin DOS at the Fermi level. The structure around 6 eV
in the DOS of Y2Fel7N3 mainly originates from the N p-states. Since the charge
transferred away from the N atoms is estimated to be about 0.25, being quite small,
there are no strong chemical bonding effects of N with its neighbors. The interesting
aspects of the calculated results on Y2Fe17 and Y2Fel7Z3 are summarized as follows:
i) The magnetic moments decrease for the Fe atoms that are nearest neighbors to
N atoms, owing to the strong hybridizations of the Fe 3d-states with N 2p-states.
ii) The magnetic moments of the Fe atoms which are farther from the N atoms
increase because of the reduction in overlap between the Fe 3d-states due to lat-
tice expansion. This effect also enhances the interatomic exchange interactions
between the Fe atoms and leads to an increase in the Curie temperature.
iii) There is an overall increase in the total magnetization upon nitrogenation, which
is in excellent agreement with the magnetization values measured at 4.2 K.
iv) Small negative moments are induced on the Y atoms. The reason is that the
Fe 3d-bands are on the average at higher binding energies than the Y 4d-bands.
Since the exchange splitting raises the energy of the minority Fe 3d-band with
respect to that of the majority band, an increase in the hybridization of the
minority Fe 3d-band with the Y 4d-bands occurs, and it enhances the DOS
of Y 4d-electrons with the same spin direction as the minority Fe 3d-band.
Thus, the Y moments are induced and are antiparallel to the Fe moments. This
also explains the fact that in all the R-T compounds the R spins with non-zero
4f-electrons always couple antiferromagnetically to the spins of the 3d transition
metal atoms when having a more than half filled 3d-band.
The Curie temperature is almost doubled by interstitially introducing Z atoms into
R2Fel7. It is of interest to discuss this problem using the energy band structures.
The quantitative predictions of the Curie temperatures are rather difficult because
we have not yet any reliable and available method to evaluate the effects of the
spin fluctuations. However, the general trends of the Curie temperature, the strength
of the coupling between the magnetic moments on each atomic site, can be well
described when using the results of the band structure calculations.
According to the spin fluctuation theory of Mohn and Wohlfarth (1987), the Curie
temperature Tc is given by

Tc o~ Mg/Xo, (5.1~
where Mo is the magnetic moment per atom at 0 K and Xo is the enhanced suscep-
tibility given by

X° 1 - 1 ( - -1 + -
1
- I
) . (5.2)
2# 2 2N¢(EF) 2NI.(Ez)
INTERSTITIALLYMODIFIEDINTERMETALLICS 371

Here, Ny(EF) and N~(EF) are the spin-up and spin down DOS at the Fermi level EF
and I is the Stoner parameter. Jaswal et al. (1991) estimated the increase in Curie
temperature upon nitrogenation for YzFel7 from the results of the band structure cal-
culations, leading to the ratio Tc(Y2Fe17N3)/Tc(YzFe17) = 2.34. This suggests that
an increase in M0 and a substantial decrease of N,(EF) and N+(EF) upon nitrogena-
tion are essential for the almost doubling of Tc. From these analyses, Jaswal et al.
have concluded that the Tc increase mainly originates from the reduction in overlap
between the Fe 3d-wave functions due to lattice expansion upon nitrogenation. It
seems likely that the spin-fluctuation theory is a reasonable model for understanding
the magnetism of this class of compounds at finite temperatures.

5.1.2. Compounds of the magnetic rare earth elements R = Nd, Sm and Gd


In a series of papers, Lai and coworkers (Gu and Lai 1992, Zeng et al. 1992 and
1993, Gu et al. 1993) have reported first principle calculations of the rhombohedral
R2FelTZ3. The electronic band structures of Nd2Fe17 and NdzFel7N3 with N atoms
on the 9e, 3b or 18g sites were calculated using the first principle spin-polarized
OLCAO method (Gu and Lai 1992, Gu et al. 1993). For the case that the N atoms
are located on the 9e sites, the total DOS, site-projected partial DOS and the magnetic
moments on the four non-equivalent Fe sites were given. The highest Fe moments are
found on the Fe 6c sites in NdzFe17 and on the Fe 9d sites in Nd2Fe17N3, while the
smallest ones are on the Fe 9d sites in NdzFe17 and on the Fe 18f sites in the nitride.
When the N atoms occupy the 3b or the 18g sites, the maximum local moments were
deduced to be on the Fe 6c sites and the minimum moments to be on the Fe 18h sites.
The result of these calculations gives the impression that the N atoms preferentially
occupy either the 3b or the 18g sites, but not the 9e sites. However, without doubt the
N atoms have been confirmed to occupy the 9e sites in the rhombohedral RzFe17N3.
Therefore, the results of calculated band structures of Nd2Fe17N3with the N atoms
on the 9e sites are suspect because self-consistent iteration is not performed in the
calculation. Electronic structures of R2Fel7Z3 (R = S m or Nd; Z = N or C) were
studied by the self-consistent local spin density functional method using several
cluster models (Zeng et al. 1992 and 1993). The distribution of electrons and the
magnetic moments on different Fe sites, Sm(Nd) and N(C) sites were calculated.
The calculated moments are given in table 5. lb, which are in fairly good agreement
with the experimental results obtained by Kajitani et al. (1993b), suggesting the
validity of the method employing cluster-treatments.
The work of Woods et al. (1993) is the only published photoemission study of
Sm2FelvNu for y = 0 and y ~ 2.6. These authors also calculated the electronic band
structures by the self-consistent spin-polarized LMTO method in the scalar relativis-
tic approximation, and compared the computational results with the experimental
results obtained by photoemission studies. The major features of the ultra-violet
photoemission spectra include the Fe 3d band with a strong peak at 0.8 eV and a
small peak at 2.9 eV below the Fermi energy, which are in good agreement with the
theoretical DOS calculation. There is a subtle difference in the experimental energy
distribution curves (EDC) of Sm2Fe17 and Sm2Fel7N2.6 near the Fermi level. The
systematic shift of the EDC to higher energy by 0.05 eV in the nitride is clearly
372 H. FUJII and H. SUN

TABLE 5. lb
Calculated and measured values of magnetic moments on different atomic sites and of the total moment
per formula unit/zT in R2Fel7 and R2FelTN3 (R = Nd, Sm and Gd) with the rhombohedral Th2ZnlT-type
structure.

Compound /ZR (/ZB) #Fe (/ZB) /zN (/ZB) /zT (#B/f.u.) Method Ref.
6c 6c 18f 18h 9d
Nd2Fe17 2.68 2.02 2.34 1.93 37.3 nscOLCAO [9]
1.22 2.54 2.33 2.58 2.01 43.1 scLMTO [10]
Nd2FelT(exp.) 3.0 2.6 1.5 1.4 0.7 30.7 neutron 14K [11]
Nd2FelTN3 0.48 2.48 1.50 2.19 2.66 0.07 36.3 nscOLCAO [9]
1.95 2.67 2.24 2.62 2.46 -0.11 45.4 scLMTO [10]
Nd2FelTNu(exp.) 3.7 3.0 2.2 2.5 2.4 48.8 neutron 14K [11]
3.8 3.4 2.2 1.9 2.1 45.3 neutron 4.2K [12]
Sm2Fel7 1.03 2.02 2.39 2.57 2.13 42.3 scLMTO [10]
Sm2FelT(exp.) 34.1 Magnetization [15]
Sm2FelTN3 2.11 2.61 2.27 2.58 2.41 -0.03 45.8 scLMTO [10]
Sm2Fe17N3(exp.) 38.1 Magnetization [15]
Gd2Fe17 -7.58 2.38 2.29 2.06 2.15 22.0 scLMTO [13]
-7.17 2.61 2.29 2.10 1.77 22.5 FLAPW [14]
-7.38 2.62 2.28 2.20 1.71 22.5 FLAPW [14]
Gd2FelT(exp.) 22.9 Magnetization [16]
Gd2Fe17N3 -7.53 2.43 2.15 2 . 4 4 2.41 0.06 24.2 scLMTO [13]
-7.07 2.69 2.13 2 . 3 8 2.44 0.05 25.6 scLMTO [14]
-7.21 2.75 2.12 2.38 2.48 0.05 25.6 FLAPW [14]
Gd2FelTN3(exp.) 27.3 Magnetization [17]
References:
[9] Gu and Lai (1992) [14] Yamaguchi and Asano (1991)
[10] Zeng et al. (1993) [15] Fujii et al. (1992)
[11] Kajitani et al. (1993b) [16] Verhoef (1991)
[12] Isnard et al. (1992b) [17] Liu J.P. et al. (1991)
[13] Coehoorn and Daalderop (1992)

found in the expanded binding energy scale, which agrees with the calculated D O S
at small binding energies.
Apart from the increase in the magnetization and Curie temperature, the third ex-
citing p h e n o m e n o n asso~ iated with interstitial modification is the strong enhancement
o f the magnetocrystalline anisotropy energy, which mainly comes from the crystal
electric field acting on 4 f electrons of the R site. A few attempts have been made
to calculate the crystal field parameters at the R site starting from a realistic charge
density distribution determined theoretically from band calculations.
The crystal electric field (CEF) parameter A ° and the electric field gradient V,z at
Gd sites in Gd2Fel7, Gd2Fe17N3 and Gd2Fe17C3 were first calculated by Coehoorn
and Daalderop (1992) using the augmented spherical wave method with atomic
sphere approximation ( A S W - A S A ) . The results o f their calculations of A°(val) c o m -
ing from valence electrons and Vzz are given in table 5.2. The calculated A°(val)
values show an increase upon doping with N and C. The trend is in fairly g o o d
agreement with the C E F estimations based on the experimental results which has
been discussed in sections 3.3.3 (table 3.6) and 4.2.2. Furthermore, they stressed
INTERSTITIALLYMODIFIEDINTERMETALLICS 373
TABLE5.2
Crystal electric fieldparameter A° (Kao2) at the Gd site deducedfrom electronic band
structure calculation(the FLAPWmethod) in GdFel2, GdEel2N, Gd2Fel7, Gd2FelTN3
and Gd2Fel7C3.
Compound GdFel2 GdFeI2N Gd2Fe17 Gd2Fel7N3 Gd2Fel7C3 Ref.
A~(val) -303 -475 -613 [1]
A~(val) 64 1454 -421 - 1174 [2]
A~(lat) -89 -439 135 226 [2]
A~(tot) -25 1015 -286 -948 [2]
References:
[1] Coehoornand Daalderop (1992)
[2] Yamaguchiand Asano (1994)

that especially the carbon interstitials lead to a stronger anisotropy than the nitrogen
interstitials, though the increase of magnetic moment in carbides is much lower than
in nitrides.
Yamaguchi and Asano (1994) have calculated the electronic structures and the
charge distributions of Gd2Fe17 and Gd2Fe17N3 by the self-consistent full-potential
linearized augmented plane wave method (FLAPW) within the framework of the
local spin density approximation (LSDA). In fig. 5.2, the contour maps of the electron
density of Gd2Fe17 and Gd2FelTN3 are drawn for planes parallel and perpendicular to
the c-axis. The electron density distribution around the Gd(6c) ion seems to be rather
uniform in the c-plane (fig. 5.2(a)), while in the plane parallel to the c-axis, there is
a low density area between the Gd(6c) ions (fig. 5.2(b)). This anisotropic electron
distribution produces an appreciably negative A° at the Gd site. When nitrogen is
introduced and occupies the interstitial 9e site, the electron density around the Gd(6c)
ion changes, as in figs 5.2 (c) and (d). Nitrogen strongly hybridizes with the Gd(6c)
atoms and increases the electron density around the Gd(6c) ion along the direction of
the N(9e) ion. Since the anisotropy of the density distribution is much larger than that
in Gd2Fe17, it gives rise to a larger negative A° at the Gd(6c) site. From the calculated
charge distributions, the second order crystal electric field parameters A° at the Gd-
site have been carefully determined, the results of which are included in table 5.2.
A°(val) is negative and large in Gd2Fe17, due to the valence electrons within its own
muffin-tin (MT) sphere at the Gd site. Though an appreciable part of the A°(val) is
cancelled by the positive A°(lat) contributions coming from other charges in both the
other MT spheres and the interstitial region, the calculated value of the total A° is still
negative and much larger than the experimental one determined in Sm2Fel7 (Li and
Cadogan 1991). In Gd2Fe17N3, the calculated value of the total A° is negative and
about 3 times that in Gd2Fel7, which is also much larger than the experimental one
for Sm2Fe17N3 (see table 3.6) (Kato et al. 1993). The origin of these discrepancies
is an open question, and requires further investigations of theoretical approaches to
the 4f-magnetic anisotropy. Yamaguchi and Asano (1994) also estimated the electric
field gradient V~z at the Gd nuclei from the charge distributions. Results are listed
in table 5.3 together with the value obtained by the analysis of the 155Gd M6ssbauer
spectrum. It can be seen that the calculated values of V~z are in good agreement
with the experimental ones.
374 H. FUJII and H. SUN
100 10.0

50.0 50.0

-I0.0 -10.0
"10.0 0.0 10.0 -10.0 0.0 10.0
a.u. a.u.

(a) (b)

10.0 1010

500
cd

-10.0 -10.0
-I0.0 0,0 10.0 -10.0 0,0 100
a.u. a.u.

(c) (d)
Fig. 5.2. Electron density distribution of Gd2Fel7 ((a) in a plane perpendicular to the c-axis, (b) in a
plane parallel to the c-axis) and of Gd2Fe17N3 ((c) in a plane perpendicular to the c-axis, (d) in a plane
parallel to the c-axis). The lowest contour value is 0.0125 a.u.-3, the contour spacing is 0.0125 a.u.-3,
and the highest contour value is 0.1 a.u.-3. The lowest density regions in (b) and (d) have densities
between 0 and 0.0125 a.u.-3. The density regions with densities higher than 0.1 a.u.-3 near each atomic
site are not shaded (after Yamaguchi and Asano 1994).

Finally, Yamaguchi and Asano calculated the electronic structures and the magnetic
properties o f Gd2Fel7 and the nitride by the self-consistent L M T O - A S A method.
In their calculations, the lattice constants were determined by minimizing the to-
tal energy. The increases in the lattice constants and the magnetic moments due
to introduction o f the N interstitials can be well explained when the ratio o f the
atomic sphere radius is chosen as rGd/•Fe = 1.35. The calculated magnetic moments
are included in table 5. tb, together with the values estimated by the self-consistent
F L A P W - L S D A method. The results are in g o o d agreement with each other, indicat-
ing that the choice of rGd/rFe = 1.35 in the self-consistent L M T O - A S A method is
INTERSTITIALLYMODIFIED INTERMETALLICS 375
TABLE 5.3
Calculated and experimentalelectric field gradient Vzz (1021 Vim2) at the Gd nucleus in
GdFe12, GdFe12N, Gd2Fe17 and Gd2Fel7N3.
Compound GdFe12 GdFeI2N Gd2Fe17 Gd2Fe17N3 Method Ref.
Vzz(cal) 1.1 15.5 5.8 11.0 FLAPW [1]
6.4 10.5 scLMTO-ASW[2]
Vzz(exp) 4.3 12.6 [3]
4.4 [4]
References:
[1] Yamaguchi and Asano (1994) [3] Dirken et al. (1991)
[2] Coehoorn and Daalderop; Daalderop et al. (1992) [4] van Steenwijk et al. (1977)

quite reasonable. The calculated Fe magnetic moments are in fairly good agreement
with those determined for Nd2Fe17N3 by the neutron diffraction studies (Isnard et
al. 1992b, Kajitani et al. 1993b).

5.2. Calculations of the l :12-type interstitial compounds


RFel2_xT~Zy (Z=N or C)
Electronic band structure calculations have been carried out for the l:12-type com-
pounds RFe12_~T~Z v with T = Ti and Mo, R -- Nd and Y, and also for the hypothet-
ical compound YFe12Z. Results are summarized in table 5.4 and will be discussed
here.

5.2.1. Calculations of RFelz_xT~N v with R = Y or Nd and T = T i


The spin-polarized electronic structures of YFe11Ti and YFel aTiN were calculated
by Li Y.R and Coey (1992) using the non self-consistent OLCAO approximation
and the calculations succeeded in describing some physical aspects of this system.
However, it has been pointed out later by others that the results possibly lack the
stability of the electronic system because of the non self-consistent character of
the method. The results of the calculation of Li Y.E and Coey showed that in
YFellTiN there are differences on the different sites as to the opposite effect of
hybridizations of Fe 3d-states with N 2p-states and the effect of volume expansion
due to interstitial nitrogen atoms on local magnetic moments. The competition of
these opposite effects leads to a slightly increase of the total iron moment per formula
from 20.3#B to 20.5#8.
Self-consistent spin-polarized band structure calculations have been performed by
the LMTO-ASA method with a ratio of atomic radii ry : rFe " rTi ---=1.35 : 1.0 : 11.1
in the frame of local spin density functional formalism by Sakuma (1992). To exam-
ine separately the lattice expansion effect due to introduction of N atoms, Sakuma
also calculated the electronic structures of YFe11TiN0 with the same atomic structure
and lattice constant as in YFe11TiN but with no N atoms. The result indicates that
the total moment of YFellTiN (23.6#B) is larger than that of YFe11TiNo (22.4# B)
and YFexlTi (21.7#B), which is different from the results of the non self-consistent
spin-polarized band calculations by Li Y.R and Coey (1992). The asphericity pa-
rameters Anp and And were estimated by Sakuma to gain insight into what caused
the increase of the moment due to the inclusion of N atoms in addition to the lattice
376 H. FUJII and H. SUN

TABLE 5.4
Calculated and measured values of magnetic moments on different atomic sites and of the total moment
per formula unit pq- in RFellT, RFel2 and their interstitial compounds RFeI1TN, RFel2N with T = Ti
and Mo. #z is the magnetic moment of the interstitial atom Z.

Compound #R (#B) /ZFe (/ZB) ]Az (/ZB) /zT (#B]f.u.)Method Ref.


2a Fe(8f) Fe(8i) T(8i) Fe(8j) Z(2b)
YFellTi 0.27 1.77 1.91 0.21 1.76 20.3 nsc OLCAO [1]
-0.46 1.70 2.37 -0.83 2.23 21.7 scLMTO [2]
YFell Ti(exp) 18.6 Magnetization [3]
YFe11TiN 0.19 1.73 2.31 0.15 1.73 0.10 20.5 nsc OLCAO [1]
-0.36 2.18 2.67 -0.83 1.98 0.04 23.6 sc LMTO [2]
YFe11TiN(exp) 21.8 Magnetization [3]
NdFell Ti 3.2 1.67 2.46 -0.86 2.32 26.9 scLMTO [4]
NdFe11Ti(exp) 25.7 Magnetization [5]
NdFe11TiN 3.3 1.93 2.55 -0.89 2.14 -0.02 28.1 sc LMTO [4]
NdFe11TiN(exp) 27.0 Magnetization [5]
YFe11Mo -0.04 1.64 2.45 -0.48 2.15 21.6 sc LMTO [6]
YFel 1Mo(exp) 2.15 2.36 - 1.0 2.42 26.0 neutron [7]
25.8 Magnetization [8]
YFellMON -0.41 2.21 2.62 -0.45 1.92 23.7 scLMTO [6]
YFe11MoN(exp) 2.38 2.43 - 1.0 2.32 27.4 neutron [7]
27.0 Magnetization [8]
YFe12 -0.43 1.80 2.45 2.40 26.2 scLMTO [9]
YFelzN -0.30 2.30 2.61 2.03 27.5 scLMTO [9]
YFeI2C -0.39 2.30 2.60 1.64 25.6 scLMTO [9]
References:
[l] Li Y.P. and Coey (1991) [6] Ishida et al. (1994)
[2] Sakuma (1992) [7] Sun H. et al. (1993a)
[3] Yang Y.C. et al. (1991) [8] Sun H. et al. (1993b)
[4] Jaswal (1993) [9] Asano et al. (1993)
[5] Akayama et al. (1994)

expansion. Here, Anp and And give the degree to which the p and d charge densities
are prolate or oblate (Coehoorn 1990). The results of the calculations show that
Anp > 0 and And > 0 for the nitride, suggesting that the valence orbitals elongate
along the c-axis. The asphericity, therefore, predicts that the N atoms attract elec-
trons from the Y atoms rather than from the Fe atoms and in turn release the Fe
atoms from bonding with the Y atoms. This effect leads to a further increase o f
the Fe magnetic moments in addition to the increase associated with the expansion
o f the lattice upon nitrogenation. As general features, it has been deduced that the
m o m e n t s on the Fe atoms which are nearest neighbors to the N atoms are lowered
upon nitrogenation, while the Fe atoms which are farthest from the N atoms have
the largest moments. The calculated total Fe magnetic moments after nitrogenation
are much larger than the experimental values obtained so far. This might be related
to the incomplete nitrogenation and to the fact that the occupancy of the 2b site
by nitrogen is only about 50% in YFelxTiNy. Experimental data by A k a y a m a et
al. (1994) have shown that the total m o m e n t per formula unit for NdFellTiNa.5
and N d F e l l T i are 27.0#B and 25.7#B, respectively. Assuming Nd behaves as a free
INTERSTITIALLYMODIFIEDINTERMETALLICS 377

ion with a moment of 3.6#B in NdFe11TiN1.5, the moment on the Fe-sublattice is


deduced to be 23.4#B, which is close to the above calculated value of 23.6#B for
YFe11TiN. The self-consistent spin-polarized electronic structure calculations have
been also carried out for NdFeI1Ti and NdFe11TiN0.5 by Jaswal (1993). The calcu-
lated magnetic moments are comparable with the experimental data by Akayama et
al. (1994) (see table 5.4).
Based on the spin-fluctuation theory of Mohn and Wohlfarth (1987), the enhance-
ment of Curie temperature by interstitial modification was discussed by Sakuma
(1992) and Jaswal (1993) independently. Using eqs (5.1) and (5.2) in section 5.1.1,
the ratio Tc(NdFeHTi)/Tc(NdFealTiNo.5) is estimated to be 1.32 from the band struc-
ture parameters. As the corresponding experimental value is 1.31, it has been con-
cluded that the Curie temperature enhancement upon nitrogenation is again primarily
due to the lowering of the DOS for both spin directions at the Fermi level.
Electronic structures of NdFea 1Ti and its nitride were studied by the photoemission
spectroscopy and also by the self-consistent spin-polarized band structure calculations
using the LMTO-LDA method (Fernando et al. 1993). We notice that there is an
overall agreement between the experimental and calculated DOS.

5.2.2. Calculations of RFel2_xT~N with R = Y and T = M o


The electronic structures of YFelz_xMo~ and their nitrides have been calculated
by Ishida et al. (1994) using the self-consistent LMTO-ASA method in the non-
relativistic approximation to examine the effect of N and Mo atoms on the magnetic
properties. The local DOS of the Fe 8f, 8i and 8j-atoms in YFellMO are similar to
those of YFel2 which will be discussed later.
The calculated magnetic moments are listed in table 5.4, which are not inconsistent
with the results of neutron diffraction study by Sun H. et al. (1993b) for YFeaIMo
and the nitride. Regardless the Mo content, the moments on the different Fe sites
increase in the order of #(Sf) < #(8j) < #(8i) for YFe12_~Mo~, and nitrogenation
leads to a decrease of the moment on the 8j site that is nearest to the N atoms, and
an increase of the moments on the sites 8i and 8f that are further from the N atoms.
These changes are explained by the strong hybridization between the N 2p-states
and the Fe d-states. The substitution of Mo atoms for Fe atoms is also reflected
in the values of the Fe moments as well. The calculated Fe moment on all sites
decrease gradually with increasing the Mo concentration, which originates from the
hybridization between the Mo d-states and Fe d-states near the Mo atoms. Similar
results have been obtained by Jaswal et al. (1990) in YFel0T2 (T = V or Cr). These
authors found that the moments of V or Cr couple antiparallel to the Fe moments,
leading to a reduction in the average moment per Fe atoms. The overall tendency
of the concentration dependence of the calculated moments in YFe12_~Mox and the
nitride is similar to the tendency of the experimental values obtained by Sun H. et al.
(1993a) for z < 0.2, but the discrepancy is large for x > 2. Besides the experimental
error, this discrepancy might be due to the assumption in the calculation that the Mo
atoms are located only at the 8i-site. In real compounds, it is possible that a small
amount of Mo atoms might locate at the 8f-site as well (Sun H. et al. 1993b).
378 H. FUJII and H. SUN

5.2.3. Calculation of the hypothetical RFe12Z


The results of electronic structures calculations for Y F e l 2 (Coehoorn 1990) and
YFeloT2 (T = Cr or V) (Jaswal et al. 1990) have indicated that the characteristic
features in the DOS are commonly conserved for all the YFe12_xT~systems. Thus,
it is valid and of interest to study the electronic band structures of RFe12Z (Z = N
or C) theoretically.
The self-consistent spin-polarized electronic structures of the hypothetical RFe12
and RFe12Z (R = Y, Ce or Gd; Z = N or C) have been calculated by Asano et al.
(1993) using the LMTO-ASA method within the frame work of the LSD approx-
imation in order to examine the effect of the Z atoms on the magnetic properties.
In the calculation, the ratio of the atomic radii was chosen as rR/rFe = 1.4 and the
total energies of RFe12 and their interstitial compounds were calculated as functions
of the lattice constants. The theoretical lattice constants determined by minimizing
the total energies were used as initial parameters for further calculations.
In figs 5.3 and 5.4, are shown the local DOS of the Fe d-states, Y s-, p-, d-,
f-states and N 2p-states in Y F e l 2 and YFe12N, respectively. For Y F e I 2 , the up-spin
bands in the Fe d-states (indicated by full line in fig. 5.3) are almost fully occupied
for all the three Fe sites, but the unoccupied peak in the down-spin states (shown by
dotted line in fig. 5.3) is smaller for the Fe 8f-sites than for the Fe 8i- or 8j-sites,
indicating/~Si o r #sj >/Z8f" From the local DOS of the s-, p- and d-states of Y atom,
we notice that the up-spin and the down-spin bands are separated, but the number

30 Fe(8j)(d) Y [2al(f)
EF

20
._. |
I0

.-&
Fe[Si}[d)
"~"30

~ 2o

~'\ ............
• i , "1"I
~

Fe(Sf)(d}
30

20 Y (2al (sl

N!
I0

-0.6 -0.4- -'0.2 0.0 0.2 -0.6 -0.4- -0.2 0.0 0.2
Energ/ (~d) Energy (Fb'd)
Fig. 5.3. Spin-polarized local density of states of the Fe(8f), Fe(8i) and Fe(8j) d-states and of the
Y s-, p-, d- and f-states in YFe12 calculated using the LMTO-ASA method in the non-relativistic
approximation (after Asano et al. 1993).
INTERSTITIALLY MODIFIED INTERMETALLICS 379

i /"
" i ~- Er 20 Y (2a)(f)

, ,/i lo Y C&ICd~ ,~

°
2~2o F~CO~Cdl '"f n~- 5 ¥ (2a) (p)

2.5 u ~,
,
• . .....

i}ij., 5 Y t~II~l
10 2.5
b

' -0.6 -'0.4- -0.2 0.0 0.2 -0.6 -0.4- -0.2 0.0 0.2
Energy (R/d) Energy (R~d)
Fig. 5.4. Spin-polarized local density of states of the Fe(8f), Fe(8i) and Fe(8j) d-states, N(2b) p-state and
of the Y s, p, d and f-states in YFe12N calculated using the LMTO-ASA method in the non-relativistic
approximation (after Asano et al. 1993).

of occupied states in the s- and p-bands is almost the same for both the spin states,
while the occupied states in the d-bands are more full for the down-spin than for the
up spin. This indicates that the spin polarization is small for the s- and p-states, but
the magnetic moment due to the d-spin polarization is appreciable and antiparallel
to the Fe moments, which leads to negative exchange interactions between the Fe
spins and Y spins.
For YFe12N, the general DOS shapes of the Fe 3d-states are similar to those of
YFe12. However, there are some essential differences as follows: The main peaks of
the N p-states are located below -0.6 Ryd, where the d-band tails of the Fe 8j-atoms
(the nearest of the N atoms) becomes larger than those of RFe12. This shows that
the Fe d-states of the 8j-sites well hybridize with the N 2p-states. Consequently,
the tail of the up-spin band above EF becomes larger for the Fe 8j-sites but smaller
for the Fe 8f- and 8i-sites upon nitrogenation. On the contrary, we notice that the
unoccupied DOS peaks of the Fe down-spin state becomes smaller for the Fe 8j-sites,
but larger for the Fe 8f-sites compared to those of RFe12. These variations of DOS
upon nitrogenation lead to a decrease in the moment on the Fe 8j-sites but to an
increase in those on the Fe 8f- and 8i-sites. Paying attention to the DOS of the N
p-states, we can see that there is a peak just below EF in the up-spin states that is
located at the shoulder in the Fe d-bands. There is a similar peak above EF located
at the center of unoccupied peaks of the down-spin Fe d-states (in the energy range
between +0.1 Ryd). These peaks mainly come from the antibonding states between
380 H. FUJII and H. SUN

the N p-states and Fe d-states. We notice, moreover, that the sharp peaks at EF
for the Fe 8i- and 8j-sites in YFe12 disappear in YFel2N, indicating a substantial
decrease of the DOS at EF upon nitrogenation. As described above, it has been
shown that the Curie temperature is enhanced by nitrogen uptake because the spin
fluctuation is restrained, according to Mohn and Wohlfarth (1987) model.
The calculated values of the magnetic moments on the different sites for RFel2
and RFel2Z (Z = N or C) are summarized in table 5.4. We notice that the calculated
moments of the Fe atoms become larger in the sequence the 8f-, 8j- and 8i-sites in
YFe12, while the Fe moment on the Fe 8f-sites is larger than that on the Fe 8j-sites in
YFe12N. This tendency is common to all the cases, irrespective of the rare earth R.
The band calculations also predict that the total magnetic moment per formula unit in
YFe12 increases from 26.2#B to 27.5#B upon nitrogenation, but decreases to 25.7#B
by carbonation. For YFel2C, the DOS peaks of the C p-states due to the antibonding
states shift to the higher energy region above EF. This makes the effect of moment
enhancement due to hybridization weak compared to the nitride. Therefore Asano et
al. (1993) claimed that N is the best among H, N, C and B to improve the magnetic
properties of RFe12. Unfortunately, since RFe~2 dose not exist, we can not directly
compare the calculated moments of RFe12 and RFe12N with experimental results.
However, we notice that the calculated moments are in fairly good agreement with
the extrapolated values to z = 0 of the moments of YFe12_xMox and YFe12_xMoxN
(Sun H. et al. 1993a).
In order to shed light on the appearance of uniaxial anisotropy in Pr- and Nd-
systems, Yamaguchi and Asano (1994) determined the CEF parameter A° at the Gd
site in GdFe12 and GdFe12N from the electronic structures and charge distributions
calculated using self-consistent FLAPW-SDA method. In figs 5.5(a) and (b), are
shown the contour maps of electron density distribution in GdFe12 and GdFe12N
parallel and perpendicular to the c-axis. The density distribution around the Gd(2a)
ions in GdFe12 seems to be rather uniform, reflecting the small value of A°. When
the N atoms are introduced and occupy the interstitial 2b sites between the Gd(2a)
ions, the electron distribution is changed as is seen in fig. 5.5(b). Here, we can see
that N is strongly bonded to the Gd(2a) atoms, which seems to be much stronger than
the Gd-Fe bonding. Those bonding states between the Gd and N atoms increase
the electron density around Gd along the direction of the N atoms. This type of
anisotropic density distribution around Gd produces the large positive value of A°.
The calculated values of A ° are shown in table 5.3 for GdFe12 and GdFe12N. In
GdFe12, the calculated net value of A°2 = A°(val) + A°(lat) is small. In GdFelzN,
while A°(val) is very large and positive, A°(lat) is negative and cancels an appreciable
part of A°(val). Nevertheless the net value of A° is positive and still large.

5.3. Summary
We can qualitatively understand the origin of the improvement of the magnetic prop-
erties in both the interstitially modified 2:17 and 1:12 systems from the calculated
band structures. However, there are obvious discrepancies between the experimental
results and the results of electronic band structure calculations. It is necessary to
INTERSTITIALLY MODIFIED INTERMETALLICS 381

10.0

:5 0.0

-10.0
-10.0 0.0 10.0
a.u.

(a)

10.0

:50.0
ai

-10.0
-10.0 0.0 10.0
a.u.
(b)

Fig. 5.5. Electron density distribution in a plane parallel to c-axis: (a) of GdFel2, (b) of GdFel2N. The
lowest contour value is 0.0125 a.u. -3, the contour spacing is 0.0125 a.u. -3, and the highest contour
value is 0.1 a.u. -3. The lowest density regions in (a) and (b) have the density between 0.0125 and 0.025
a.u. -3. The regions with densities higher than 0.1 a.u. -3 near each atomic site are not shaded (after
Yamaguchi and Asano 1994).
382 H. FUJII and H. SUN

perform more careful experimental studies using high-quality interstitially modified


samples. On the other hand, band structure calculations involve various atomic-
sphere approximations which can be not true in the real materials. We believe that
both the experiment and calculation are need to be improved. By quantitatively
comparison with more reliable experimental data, the validity of the various approx-
imations can be clarified and subsequently the electronic band structure calculation
method can be further sophisticated.

6. Applications

Owing to their excellent intrinsic magnetic properties, interstitially modified materi-


als have been regarded as promising candidates for permanent magnet applications.
Especially Sm2Fe17 in the 2:17 series and NdFel2_~T~ (T = Ti, V, Mo, etc.) in the
1:12 series are of great interest. Unfortunately these materials are metastable com-
pounds and disproportionate at temperatures higher than 600°C in general. Thus the
conventional high-temperature powder metallurgy is not applicable to the interstitial
materials. The problem now is to either improve their stability or develop a suitable
processing route for making bulk permanent magnets.
By partially substitution of other elements for Fe, the stability has been more or
less improved. However, the metastability still remains as a problem to be solved.
Various techniques have been applied to the interstitial compounds for producing both
isotropic and anisotropic magnets. An extremely high coercivity of more than 4 T has
been obtained by combined mechanical alloying and pressure-assisted zinc bonding
(Kuhrt et al. 1992). An energy product (BH)rnax of 206 kJ/m 3 has been realized in
an isotropic nanocrystalline exchange-coupled two-phase magnet composed of the
hard SmzFelyN v phase and the soft a-Fe phase, although the coercivity is still low
(Ding et al. 1993).
It should be noted here that the interstitial compounds are stable at room temper-
ature for a period of the order of 1015 years if we take into account of the diffusion
of the iron atoms only (Skomski and Coey 1993a, b). So, it is not necessary to
consider deterioration of the permanent magnetic properties due to aging at room
temperature.
In this section, studies concerning the stability improvement will be first summa-
rized, followed by different processing techniques and their achievements in making
permanent magnets. The studies on the magnetic domain structure and the coercivity
mechanism will also be reviewed. The emphasis is on the SmzFel7 nitride, for which
most of the progress has been made.

6.1. Improvement of the thermal stability


6.1.1. Effects of substitutions
Due to differences in the definitions involved with the decomposition and with
the various experimental techniques, substitution effects on the thermal stability
of Sm2Fe17-nitride have been reported to be slightly different by different authors.
Hu B.E et al. (1992) studied the thermal properties of Sm2(Fel_~M~)17Nu (M = Co,
INTERSTITIALLYMODIFIEDINTERMETALLICS 383

Ni, A1, Ti and V) by a thermopiezic analyzer (TPA) and using thermal magnetic anal-
ysis (TMA), and concluded that there was no significant effect of M substitutions
for Fe. Sugimoto et al. (1992a, b) investigated the decomposition temperature of
Sm2(Fel_xM~)lTNy with M = A1, Si, Ti, V, Cr, Mn, Fe, Co, Ni, Ga, Zr, Nb and Mo
by detecting the phase changes using X-ray diffraction and SEM-EDX techniques.
These authors have found that Cr is effective for improving the thermal stability, the
decomposition temperature of Sm2(Fe0.9Cr0.a)17Ny being about 100°C higher than
that of the pure Sm2Fe17N v phase. Zhou S.Z. et al. (1992) reported that the replace-
ment of Fe by Co (10 at.%), or Ga (1.7 at.%), or Zr (1.7 at.%) enhanced Ton, the
starting temperature of the decomposition being defined as the temperature at which
the magnetization starts to increase in the thermo-magnetic curve.

6.1.2. Effects of carbonation


There are two routes for synthesizing carbonitrides. The interstitial modified car-
bides Sm2FelvCx with x < 1.5 can be made first by melting. The carbonitrides
SmzFel7CxN u are then formed by the subsequent gas phase nitriding process (Kou
et al. 1991a, b, Yang Y.C. et al. 1992a). The other route is to synthesize car-
bonitrides by the gas phase interstitial modification process only. Nitrogenation is
done first, followed by carbonation in C2H2 for a short time (Chen et al. 1993a,
Altounian et al. 1993). Carbonitrides made by the second method were reported to
have better thermal stability than the pure nitrides. The N2 out-gassing temperature
was increased by 180 K mainly as a result of the low diffusion rate of nitrogen
in the outer-most carbon layer formed during the carbonation process. The other
advantage is that there is less segregation of c~-Fe in the carbonitrides made by the
second method, because of the shorter annealing time required for carbonation and
because the formation of cementite Fe3C. The magnetic properties of SmzFelTC~Ny
are similar to those of the SmzFel7N v, but with a somewhat better thermal stability.

6.2. Development of permanent magnets


Although the thermal stability can be improved to some extent, the metastability of
the nitrides (or carbonitrides) remains to be a problem and dictates that the magnets
have to be made by a low temperature process and thus the solution falls naturally on
bonded magnets produced from coercive powder. Anisotropic magnets can be made
from single crystal powders either by metal or epoxy bonding, while nanocrystalline
fine-structured powders produced by mechanical alloying, mechanical grinding, rapid
quenching or HDDR, are used for making isotropic magnets. Most of the published
data are still at the stage of making high coercive powders, that is, a stage pre-
ceding the manufacture of bulk magnets. The processing procedures of the various
techniques are shown in fig. 6.1 to 6.4.

6.2.1. Bonded anisotropic magnets based on the Sm2Fel7 nitride


The procedure for making anisotropic bonded magnets is as follows (fig. 6.1). The
host ingots are produced by melting of the starting materials (either arc-melting or
induction melting), which are then subjected to a homogenization treatment and are
384 H. FUJIIand H. SUN
melting
Sm, Fe metals I

meltingandtomogenization
SmzFe17compound

pulverization
I powdersI
nitroginatton

InitridepowdersI
furtherpulverization
l finepowders I
|
bonding and ma.~etacgil nment
I I
Fig. 6.1. Schematicdiagramof the meltingprocess.

crushed into powders with the size of a few 10 #m in diameter prior to the gas
phase interstitial modification process. Nitrogenations are performed in N2, NH3, or
mixtures ofN2, NH3 or H2 at temperatures typically of 450°C. The bulk magnets are
then made by epoxy bonding or metal bonding of the nitrided powders. A magnetic
field is applied during the bonding process for magnetic alignment. Sometimes, the
nitride powders are further pulverized down to a few #m in diameter before bonding.
In the case of metal bonding, a heat treatment performed at temperatures slightly
higher than the melting point of the bonding metal is necessary.
Every step in the processing procedure is important for the finally obtained hard
magnetic properties. First, a high purity Sm2Fe17 phase in the host compounds
used for nitrogenation is essential. As Sm is an element with a relative high vapor
pressure, Sm evaporates during melting, more or less depending on the melting
condition, and the Sm content in the melt becomes lower than that in the starting
mixture of the pure elements. Oxidation can also cause losses of Sm. Thus an excess
amount of Sm is needed for compensation of the losses. However, it is difficult to
INTERSTITIALLYMODIFIED INTERMETALLICS 385

apply the exact amount of Sm excess because of the difficulties in controlling the
Sm losses during melting. If not enough Sm is present, c~-Fe appears as a second
phase in the Sm2Fel7 compounds. This soft magnetic c~-Fe phase remains present
after nitrogenation and certainly is harmful to the hard magnetic properties of the
material. On the other hand, if too much Sm is present, SmFe2 or SmFe3 will appear
as impurity phases. They directly decompose into StuN and c~-Fe upon nitrogenation
(Christodoulou and Takeshita 1993a, b, Rodewald et al. 1993, Fukuno et al. 1992)
and again the soft magnetic c~-Fe phase appears. In order to avoid the harmful c~-
Fe phase, the composition of the Sm2Fe17 ingot after homogenization must match
the stoichiometric composition as exactly as possible. The nitrogenation process
is always accompanied by the decomposition process of the 2:17 phase into c~-Fe,
SmN and possible some other unknown phases. Thus carefully controlling of both
the alloy preparation and the nitrogenation process is important for reducing the
formation of the soft magnetic phases and in turn for making high quality magnetic
powders.
Nitrided powders of SmzFe17Ny having the highest value reported so far for the
maximum energy product of 244 kJ/m 3 were obtained by Suzuki et al. (1993).
The influence of the Sm content on the magnetic properties of Sm2Fe17N v was
studied by Fukuno et al. (1992) and these authors succeeded in obtaining (BH)max
values as high as 240 kJ/m 3 and a coercivity of 1.0 T for a Sm content of about
11 at.%, which is slightly higher than 10.5 at.% Sm in stoichiometric SmzFea7.
Fukuno et al. performed a pre-hydrogenation treatment at 250°C for 4 hours before
nitrogenation. The 250°C pre-hydrogenation treatment was reported to be effective
also by Rodewald et al. (1992). Nitrogenation in a mixed gas of NH3 and H2
resulted in a (/3H)max value of 170 kJ/m 3 for Sm2Fel7N v powders (Iriyama et al.
1992, Kobayashi et al. 1992).
The magnetic properties of the nitride powders are found to be closely related
with the particle size. The coercivity increases almost exponentially with decreasing
average particle size. The dependence of coercivity on particle size was studied in
detail for SmzFelvN v by Suzuki and Miura (1992) and for Smz(Fe0.8Co0.2)IvN3.03
by Yamamoto H. et al. (1993). Particle sizes in between 2-4 #m were considered
to be close to the size of single magnetic domain particles (Iriyama et al. 1992) and
the single domain theory was used to explain the sharp increase of the coercivity
upon further pulverization after nitrogenation. A sensitive effect of particle size on
coercivity was also demonstrated by Wendhausen et al. (1992). The authors found
that the squareness of the hysteresis loop could be improved just by controlling the
particle size in a certain range by sieving.
The nitrided powders can be processed into real magnets in bulk form by two kinds
of binders, epoxy and metal. Suzuki et al. (1993) obtained 2.5 wt.% epoxy bonded
SmzFe17Nv magnets with a (BH)max of 145 kJ/m 3 and a density of 6.1 g/cm 3. They
also studied the dependence of the magnetic properties on the compaction pressure
and found that pressures in the range of 800-1000 MPa were necessary for reaching
a high (BH)max. However, the coercivity decreased slightly with increasing pressure
for unknown reasons. The (BH)max value of 125 kJ/m 3 and a density of 5.93 g/cm 3
386 H. FUJII and H. SUN

were obtained in Sm2(Fe0.sCo0.2)17N3.03 by Yamamoto H. et al. (1993) by adding


3 wt.% of epoxy and applying a pressure of 784 MPa.
Non-magnetic metals with a melting point much lower than the decomposition
temperature of the nitride are considered to be suitable candidates that could be used
as binders for making bulk nitride magnets. Zn (melting point 419 °C), Bi (271 °C),
Sn (231 °C), A1 (620°C) and In (157°C) have been tried experimentally (Otani et al.
1991, Huang et al. 1991a). Zn was found to be the only suitable one and has been
studied most intensively. Sn is also of some use but it is much inferior to Zn (Huang
et al. 1991a, Rodewald et al. 1992). Bi, A1 and In were found to be unsuitable.
The bonding processes were performed mostly at temperatures slightly higher than
the melting point of the bonding metal. However, Rodewald et al. (1993) reported
that annealing of the mixture of Sm2Fel7Ny and Zn powder at temperatures above
the melting point of Zn (419°C) resulted in the decomposition of Sm2Fel7N v by the
reaction Sm2Fe17Ny + Zn --+ Sm2FeZn2 + c~-Fe + N2, and the magnetic properties
deteriorated due to the appearance of ~-Fe.
The effect of Zn content on coercivity, remanence and magnet density has been
investigated by Suzuki and Miura (1992), and Rodewald et al. (1993). Coercivity
values in excess of 2.5 T were reported by using up to 50 wt.% Zn powder (Suzuki
and Miura 1992). Because of the dilution of the magnetic phase by the non-magnetic
Zn metal, the remanence is low and the energy product is also low. Rodewald et
al. (1993) suggested that 10-15 wt.% Zn powder should be used in order to reach
a balance between the remanent polarization and the coercivity. In table 6.1 we
summarize the magnetic properties of the metal and resin bonded magnets, as well
as the magnetic properties of the nitride powders.
The mechanism of the strong positive effects of Zn bonding on the coercivity has
basically been understood on the basis of the formation of the secondary paramagnetic
Zn-Fe phase (Zn7Fe3 or Zn4Fe etc.) between the nitride particles. The surfaces
of the nitride particles were smoothed by the Zn-Fe phase and as a result, the
interparticle stray fields were reduced. At the same time, the deleterious o~-Fe could
be eliminated or at least be reduced. A similar effect was found for Sn and attributed
to the formation of SnFe (Huang et al. 1991a, Hu J.E et al. 1993b). On the other
hand, a microstructural study of the Zn bonded SmzFea7N3 magnet by transmission
and analytical electron microscopy revealed an alternative mechanism for the high
coercivity (Hiraga et al. 1993). The reaction between the matrix Sm2FeayNv phase
and Zn metal produced the Smz(Fe, Zn)17Nu (assumed to be non-magnetic) and
excess a-Fe. The main part of the produced c~-Fe combined with Zn and in turn the
Zn-Fe phase was formed. The remaining c~-Fe existed in the form of Fe-Zn primary
solid solution. The Sm2(Fe, Zn)lyN v coated the main nitride phase and prevented
contact between the ~-Fe and the main nitride phase, improving the coercivity.
The temperature coefficients of Br and #0iHc for the epoxy bonded magnets were
reported to be -0.07 ~ -0.08 %/K and -0.44 ~ -0.53 %/K, respectively (Suzuki et
al. 1993, Yamamoto H. et al. 1993), and the temperature coefficient of #0iHc for the
Zn bonded magnet was -0.45 %/K (Rodewald et al. 1992). In general, the temper-
ature stability was better than for NdFeB magnets, and also a better environmental
stability was reported by Suzuki et al. (1993).
INTERSTITIALLY MODIFIED INTERMETALLICS 387

Lu

ddd oo

r-~

~.~

~.~ 0

tal3 ~ " ,

ooo c5 c~

o.e
cq

/X
g

e~

o d d d d ~ d

~.~

"6 0 O0

ZZ ~ZZZZ ~ ~ZZ

ffffffffFffffffffffff
388 H. FUJII and H. SUN

6.2.2. Mechanical alloying of the 2:17 compound


The mechanical alloying method starts from elemental Sm and Fe powders and leads
to a two-phase mixture of amorphous Sm-Fe and c~-Fe after milling in a protected
atmosphere, which is Ar in most of the cases (fig. 6.2). The Sm2Fe17 phase with an
optimized microstructure forms during vacuum or argon annealing at around 700 °C.
Subsequent nitrogenation in the temperature range between 400-500°C transforms
the soft easy plane anisotropy Sm2Fe17 phase into the hard Sm2Fe17N v phase with
a very strong uniaxial anisotropy. Hard magnetic properties of the epoxy bonded
or cold pressed isotropic powders obtained by various authors are summarized in
table 6.2.
An extremely high coercivity of 4.36 T (Kuhrt et al. 1992) has been obtained
by mechanical alloying and subsequent bonding with 20 wt.% Zn at a high pres-
sure of 270 MPa. The remanence of 0.4 T was low because of the dilution by
the non-magnetic Zn and also because of incomplete densification in spite of the
high pressure applied. The high coercivity was understood as arising from pressure-
assisted formation of the paramagnetic Zn7Fe3 phase at the grain boundaries inside
the powder particles. The effect of compaction pressure on the coercivity and on
the magnet density was also studied. Although the density increased with increasing
compaction pressure, pressures much higher than 270 MPa induced the decomposi-
tion of the nitride phase and resulted in a lower remanence.
A method analoguous to mechanical alloying (MA) is mechanical grinding (MG).
Instead of using elemental Sm and Fe powders, the Sm-Fe ingot is first synthe-
sized by melting and then subjected to ball milling in an Ar atmosphere (Ding et
al. 1992b, Wang K.Y. et al. 1993, Majima et al. 1993). The as milled two-phase

mechanical alloying

Sm,Fepowders ]
mil~ng
cx-Fe+ Sm-Feamorphousphase I
annealing
nitrog;nation
finelycrystallized
nitride powders

bonding
l bulkmagnetI
Fig. 6.2. Schematicdiagram of the mechanical alloyingprocess.
INTERSTITIALLYMODIFIED INTERMETALLICS 389
TABLE 6.2
Hard magnetic properties achievedby nitrogenationor carbonationof the Sm-Fe compoundsproduced
by mechanical alloying(MA) or mechanical grounding (MG). #0iHc and Br are the intrinsic coercivity
and remnant magnetization, respectively. (BH)max is the maximum energy product.
Composition Method #0inc (T) Br (T) (BH)rnax(kJ/m3) Ref.
Sm12.sFe87.5 MA 2.94 0.71 86 [1, 2]
Sm13.sFe86.5 MA+Zn bonding 4.36 0.4 [3, 4]
Sm14.6Fe85.4 MA 3.13 0.51 [5]
Sm13Fe87 MA 2.95 0.624 101 [6]
Sm14Fe86 MA 3.61 89 [6]
Sm15Fe87 MG 3.13 0.7 83 [7]
SmlsFe85 MG anisotropic 1.94 0.785 167 [6, 8]
Sm13.5Fe86.5 MA carbide 2.32 0.61 58 [9, 4]
References:
[1] Schnitzke et al. (1990) [6] Ding et al. (1992c)
[2] Schultz et al. (1991) [7] Wang K.Y. et al. (1993)
[3] Kuhrt et al. (1992a) [8] Ding et al. (1992b)
[4] Kuhrt et al. (1993) [9] Kuhrt et al. (1992b)
[5] Ding et al. (1992a)

product containing a-Fe and amorphous Sm-Fe is similar to that obtained by MA.
The followed two-step heat treatment (annealing and nitrogenation) is also similar
and again the high coercivity is considered to be related to the ultrafine-grain mi-
crostructure. The advantage of MG when compared with MA is that the handling
of the S m - F e melted ingot is much easier than the handling of the Sm and Fe metal
powders. Ding et al. (1992b) reported the successful manufacture of anisotropic
powder by MG with hard magnetic properties of Br = 0.785 T, #0iHc --= 1.94 T and
(BH)max = 170 kJ/m 3. The key point is the control of the milling time which has to
be kept below a certain critical period, depressing the formation of the amorphous
phase. The latter would have transformed into randomly oriented nanocrystals during
the two-step heat treatment and resulted in isotropic magnets.
The effect of the milling atmosphere during MA and MG was studied by Calka et
al. (1992). While MA in neither N2 nor NH3 resulted in the formation of the 2:17
phase alloy, MG of pre-arc-melted Sm-Fe ingot in N2 or HH3 could be of interest.
An alternative way of MA is the chemical reduction via mechanical alloying as
reported by Liu Y.N. et al. (1992). The starting materials are Sm203, SmF3, Fe
and Ca, which are much cheaper and easier to deal with than the pure Sm and Fe
powders. The 2:17 nitride phase is formed together with CaO or CaF2 by milling
and the two-step heat treatment mentioned above. The maximum coercivity obtained
is 2.5 T. Apart from the task of developing magnetic anisotropy, which is common to
all other MA and MG materials, separation of the magnetic phase from the impurities
CaO or CaF2 is a main task in this kind of materials.

6.2.3. Rapid quenching of the 2:17 compounds


The single-wheel melt-spinning method is frequently used for rapid quenching (see
fig. 6.3). The pre-melted stoichiometric alloy is placed in a quartz crucible with an
orifice at the bottom. After being induction melted, the alloy is ejected by argon
390 H. FUJH and H. SUN

gas through the orifice onto a rotating Cu wheel, which is coated by Cr in some
cases. With a wheel surface velocity of around 30 m/s, a mixture of the modified
2:17 phase (possible the 1:7 phase) and an amorphous phase can be obtained. In
order to achieve hard magnetic properties, a two-step heat treatment after melt-
spinning, not only nitrogenation, but also a short time annealing (less than 1 hour)
in argon or vacuum at temperatures between 650°C and 750°C, has to be done. The
microstructures developed in the materials have subtle dependence on the spinning
condition, as well as on the annealing condition and the nitrogenation condition. As
the hard magnetic properties (the coercivity) are closely related to the microstructures,
the melt-spinning, annealing and nitrogenation conditions under which the best hard
magnetic properties can be obtained are quite different by different authors. Table 6.3
summarizes the hard magnetic properties reported in the literature.
The interstitial carbides included in table 6.3 were made by arc-melting and melt-
spinning, without the gas-phase interstitial modification process. Sm2FelnGa3C1.5
was melt-spun at 30 m/s and the as-spun material showed a high coercivity of 1.5 T
(Shen et al. 1993a), while Sugimoto et al. (1992b) obtained #0iHc = 1.26 T by

melt-spinning

I SmzFe17compound
melt-s inning
¢

l ribb )ns I
pulverization

E powders I
annelling
nitrogSnation

finelycrystallized
nitride powders

bon~ing .

I bulk magnet I
Fig. 6.3. Schematicdiagram of the melt-spinning process.
INTERSTITIALLYMODIFIED INTERMETALLICS 391
TABLE 6.3
Hard magnetic properties achieved in the nitrides and carbides of Sm-Fe compounds producedby
melt-spinning. /~oiHcand Br are the intrinsic coercivity and remnant magnetization, respectively.
(BH)max is the maximum energy product.
Composition /z0iHc (T) Br (T) (BH)max (kJ/m3) Ref.
Sm10.6Fe89.4Ny 0.616 0.86 69 [1]
Sml2Fe88Ny 2.10 0.73 65 [1]
SmlsFe85Nu 2.23 0.72 68 [2, 3]
Sm2.TFelTNu 1.77 95 [4]
Sm2(Feo.9Gao.1)17C1.6 0.85 [5]
Sm2(Fe0.85Ga0ACo0.05)17CI.6 1.26 62 [5]
Sm2Fe14Ga3Ck5 1.5 [6]
References:
[1] Katter et al. (1991) [4] Christodoulou and Takeshita (1993c)
[2] Pinkerton and Fuerst (1992) [5] Sugimoto et al. (1992b)
[3] Pinkerton and Fuerst (1993) [6] Shen et al. (1993a)

performing a 30 rain vacuum annealing at 600°C to 800°C after melt-spinning of


Smz(Fe0.9Gao.l)17C1.6 and Sm2(Fe0.85Ga0.1Co0.05)17C1.6.
Similarly to MA nitrides, pressure-assisted zinc bonding raises the coercivity of the
melt-spun nitrides. Kuhrt et al. (1992a) reported a coercivity increase of 75% from
1.17 T to 2.06 T in melt-spun and nitrided Sm2Fe17 by zinc bonding at a very high
pressure, whereas the remanence decreased from 0.7 T to 0.4 T at the same time.
Pinkerton and Fuerst (1992) reported that a pre-nitriding heat treatment on finely
grounded particles (<25 #m) was effective for improving the magnetic properties,
while heat treatment on coarser particles (<45 #m) under the same condition had
little effect.
Again, similarly to the MA nitrides, the hard magnetic properties are attributable to
the favorable fine-grained microstructure generated by melt spinning and annealing,
in addition to the excellent intrinsic magnetic properties of the 2:17 nitride. The
microstructure of the as-spun Sm-Fe ribbons (35 m/s) was studied by TEM by
Pinkerton and Fuerst (1993) who observed that the grain size was less than 40 nm
in diameter.

6.2.4. HDDR of the 2:17 compounds


HDDR starts with pre-melted Sm2Fel7-based alloys. The alloy absorbs hydrogen at
temperatures lower than 200 °C and the 2:17 hydride forms. Upon further increasing
the temperature in the H2 atmosphere up to about 500°C, the formed hydride de-
composes into Sm hydride and a-Fe by the following reaction, SmzFe17H6 + (2 + z -
6/2)H2 --+ 2SmH2+~+ 17c~-Fe. Heating the above formed SmH2 and a-Fe in vacuum
or argon atmosphere at around 750 °C results almost simultaneously in hydrogen des-
orption and recombination of the 2:17 phase, where the Sm hydride dissociates into
elemental Sm and H2 and where a subsequent reaction of the Sm and a-Fe reforms
the SmzFe17 phase. The hard magnetic properties are then obtained by nitrogenation
of the HDDR powder (fig. 6.4). The data of isotropic fixed powders are summarized
in table 6.4. The HDDR process leads to microstructures of very fine grains with
392 H. FUJH and H. SUN

HDDR

I SmzFe17compound [
pulverlzation
I owder I
"I°"
finelycrystallized2:17 phase
nitrog~ation
Initride
powders.,,
I
b°nling
Jbulkmagnet I
Fig. 6.4. Schematic diagram of the HDDR process.

TABLE 6.4
Hard magnetic properties achieved by nitrogenation of the Sm-Fe compounds produced by the
HDDR method. P,oiHc and/3r are the intrinsic coercivity and remnant magnetization, respectively.
(BH)max is the maximum energy product.

Composition #0iHc (T) Br (T) (BH)max (kJ/m3) Ref.


Sm2Fe]7 0.82 111 [1, 2]
Sm2Fei7 1.6 0.75 [3]
Sm2(Feo.95Cr0.o5)17 2.0 0.71 [3]
Sm2(Fe0,983Gao.ol7) 17 2.5 0.73 [3]
Sm2Fel7 0.87 [4, 5, 6]
Sm3.2Fel7 1.6 [4, 5, 6]
References:
[1] Christodoutou and Takeshita (1993d) [4] Nakamura (1992)
[2] Takeshita (1993) [5] Sugimoto et al. (1992c)
[3] Zhou S.Z. et al. (1992) [6] Okada (1992)

submicron size, that are held responsible for the hard magnetic properties (Takeshita
1993).

6.2.5. Other methods for manufacturing the 2:17-type magnets


An explosion technique was used by Hu B.E et al. (1993) for sintering of the
arc melted, nitrided and subsequently ball milled Sm2Fea7 powders. This tech-
nique mainly involved the application of a high pressure shock wave. The powders
were c o m p a c t e d to a high-density magnet by the shock wave. The achieved hard
INTERSTITIALLYMODIFIEDINTERMETALLICS 393

magnetic properties by the above authors are /3r = 0.83 T, # 0 i H c = 0.57 T and
(BH)max = 88 kJ/m 3 with density of 6.0-7.4 g/cm 3. The demagnetization curve of
the explosion sintered magnet has a better rectangularity than that of the resin bonded
magnet. Compared with NdFeB magnets, the explosion sintered Sm-Fe-N magnet
was reported to have better temperature stability, the temperature coefficients c~ of
remanence and ~ of coercivity being -0.076 %/K and -0.51%/K, respectively.
Nano-structured two-phase magnetic system, being composed of small soft mag-
netic grains which are strongly exchange-coupled to a hard magnetic phase, has
regained interest. The idea is to achieve an enhanced remanence due to the exis-
tence of the soft magnetic phase with high saturation magnetization and result in a
remarkable high energy product. The melt-spun NdzFeI4B and Fe3B (Coehoorn et
al. 1989, Kneller and Hawing, 1991), NdzFe14B and c~-Fe systems (Manaf et al.
1993a, b) were reported to show the above mentioned characteristics of the two-
phase magnet. Schrefl et al. (1993) simulated and calculated the dependence of the
coercivity on the size of the soft magnetic grains embedded in a hard magnetic matrix
for a two-dimentional system. The result showed that in order to induce magnetic
hardness from the hard magnetic phase into the soft magnetic grains, the size of the
magnetically soft regions has to be under 10 nm or smaller than twice the domain
wall width of the hard magnetic phase.
A remanence enhancement was reported by Ding et al. (1992c, 1993) for me-
chanically alloyed isotropic SmvFe93-nitride, with the hard magnetic phase being
SmzFea7Ny and the soft magnetic phase being c~-Fe. The magnetic properties re-
ported are Br -- 1.127 T,/z0iHc = 0.39 T and (BH)max = 207 kJ/m 3. TEM exam-
ination on the as-milled powders showed that nanocrystalline c~-Fe grains of about
5 nm size were embedded in an amorphous Sm-Fe matrix. If a higher coercivity
could be generated, a much higher (BH)max is expected.

6.2.6. Magnets based on the l:12-type interstitial nitrides


Compared with the 2:17-type nitrides, there are much less studies in the field of the
l:12-type nitrided materials. Hard magnetic properties obtained by various methods
are listed in table 6.5, the main interest being the Nd containing compounds.
Induction melted and nitrided NdFe9_xCo3Ti~Ny compounds were studied by
Suzuki et al. (1992a, b). The highest coercivity 0.59 T, was obtained by 50 wt.%
Zn bonding for x = 1.2. Good environmental resistance (high temperature and hu-
midity) was reported. Nitrides and carbides of Nd(Fe,M)ae alloys (M = Mo, V, Ti
or combination of Ti and V) were investigated by Tang et al. (1993a, b) using melt
spinning and annealing in order to achieve suitable microstructures for hard mag-
netic properties. Only the nitride NdFel0Mo2Nvshowed a relative high coercivity
(0.88 T), while nitrides and carbides of other compositions had much lower coerciv-
ities although the intrinsic magnetic properties were similar. Preliminary results of
hard magnetic properties including a coercivity of 0.3 T for PrFel0.sMOl.sN u were
obtained by Yang Y.C. et al. (1992b) by using the hydrogenation disproportionation
and desorption (HDD) method.
In an investigation of mechanical alloyed 1:12 compounds, Gong and Hadjipanayis
(1992 and 1993) reported an increase in coercivity by 0.2-0.3 T through replacement
394 H. FUJII and H. SUN

TABLE 6.5
Hard magnetic properties of the interstitial nitrides and carbides of the 1:12-type compounds pro-
cessed by various techniques. /.t,0inc and Br are the intrinsic coercivity and remnant magnetization,
respectively. (BH)max is the maximum energy product.
Composition Method /zoiHc (T) Br (T) (BH)max Ref.
NdFe7.8Co3Til.2N1 +6 Zn bonding 0.59 [1]
NdFel0Mo2Nu melt-spinning 0.88 [2, 3]
PrFelo.sMOl.sNy HDD 0.3 [4]
Nd10Fe75VlsNv MA 0.75 [5, 6]
Nd10Fe75Mo15Nu MA 0.85 [5, 6]
Nd10Fe75MOlsNv MA+A1 bonding 0.95 [5, 6]
Nd10Fe75MolsCy MA 0.4 [5, 6]
NdllFe74Mo15Nu MA 0.64 0.4 [7, 8]
References:
[1] Suzuki et al. (1992a, b) [5] Gong et al. (1992)
[2] Tang et al. (1993a) [6] Gong et al. (1993)
[3] Tang et al. (1993b) [7] Hirosawa et al. (1992)
[4] Yang Y.C. et al. (1992b) [8] Itsukaichi et al. (1993)

of Nd by 1.5 at.% Dy forming Nd8.sDy1.sFe82MosN u. The same authors also studied


the metal bonded MA materials and found that small amounts of A1 (0-5 wt.%)
increased the coercivity by 0.1-0.2 T. The use of Zn was found to be not suitable
due to the absence of the Fe-Zn phases, which were observed in the Sm2FelTNy
Zn-bonded magnets.

6.2. 7. Coercivity mechan&m


Kerr microscopy observations of the domain structure of SmzFel7Ny were reported
by Mukai and Fujimoto (1992) and by Hu J.F. et al. (1993a). In close correspon-
dence with the intrinsic magnetic properties, a sensitive nitrogen content dependence
was also found for the intrinsic magnetic domain parameters, such as the domain
wall energy 7, the average exchange constant A, the domain wall width 3B, and the
critical diameter of the single domain particles De. Kou et al. (1993) calculated
the magnetic domain parameters of Sm2Fe17N u from its Curie temperature and the
magnetocrystalline electric-field anisotropy constants K1 and K2. Values of the mag-
netic domain parameters obtained by various authors and various methods are listed
in table 6.6 for comparison, together with those of Nd2FeI4B. By calculating the
temperature dependence of the average exchange constant A, Hu J.F. et al. (1993b)
also studied the temperature dependence of the above mentioned domain parame-
ters in the temperature range from 200 K to 580 K. The results showed that with
increasing temperature, A, 7 and De decrease, while ~B increases slightly.
The coercivity mechanism in hydrogen decrepitated, metal bonded, explosion sin-
tered and mechanical alloyed Sm2Fe17Nx magnets were studied in a series publica-
tions by Hu J.E et al. and Kou et al. (Hu J.E et al. 1992, 1993b, 1993c, 1993d,
Kou et al. 1993) on the basis of the micromagnetic theory developed by Kronmtiller
(1978) and Friedberg and Paul (1975), all the conclusions supporting a nucleation
field controlled coercivity.
INTERSTITIALLYMODIFIED INTERMETALLICS 395
TABLE 6.6
Intrinsic magnetic domain parameters of the Sm2Fe17 nitride. In the table, are included those of
Nd2Fe14B. W is the observed domain width, 3' domain wall energy, A the average exchange constant,
De the critical diameter of single domain particles and 6B is the domain wall width.

Composition V (/zm) 7 (10-3 J/m2) A (10-11j/m) Dc (#m) 6B (nm) Ref.


SmzFe17Ny 0.85-1.6(1.0) 62.5-27.9(39) 2.98-0.59(1.16) 0.3-0.6(0.36) 6.03-2.68(3.75) [1]
SmEFel7N2.6 0.76 28 0.27 [2]
Sm2Fe17Ny 33 1.19 0.32 4.3 [3]
Nd2Fe14B 0.55 24 0.2 3.9 [4]
References:
[1] Hu J.E et al. (t993a) [3] Kou et al. (1993)
[2] Mukai and Fujimoto (1992) [4] Durst and Kronmtiller (1986)
TABLE 7.1
Comparison of the intrinsic magnetic properties at room temperature
for some well-known permanent magnets.

Compound Tc (K) Ms (T) /~0Ha (T)


SmCo5 1000 1.14 28
Sm2Co17 1193 1.25 6.0
Nd2Fel4B 588 1.60 8.0
Sm2FelTN3 749 1.54 26
NdFel 1TiN 729 1.45 12

7. Prospects

The discovery that metastable intermetallic compounds can be produced by a process


of gas-phase interstitial modification has significantly widened the range of rare
earth intermetallics with intrinsic magnetic properties suitable for use as permanent
magnets.
In order to have a sufficiently high magnetic hardness as well as a reasonable
temperature stability, high magnetization and high Curie temperature are essential
for new materials suitable for permanent magnet applications. Table 7.1 lists the
intrinsic magnetic properties at room temperature for some well known permanent
magnet materials. As can be seen from the table, Sm2Fe17N v is the first new magnetic
material since 1983 that has a chance of rivalling Nd2Fex4B in permanent magnet
applications. It combines a Curie temperature that is high for iron-rich intermetallics,
with a large uniaxial anisotropy field and a respectable spontaneous magnetization.
The theoretical upper limit on the attainable energy product is 472 kJ/m 3.
In fig. 7.1, the energy product of commercialized permanent magnets is plotted
as a function of cost per unit mass. Permanent magnets with different levels of
(BH)max have their own market according to different requirement in various kinds
of applications. Substantial progress has already been made in developing hard
magnets from the interstitial compounds. The latter materials are particularly inter-
esting for the applications where anisotropic bonded magnets with (BH)max around
200 kJ/m 3 are desired, as indicated by the shaded X area in fig. 7.1. The (BH)max of
Sm2Fe17N3 based nucleation-type magnet has reached 140-180 kJ/m 3 by improving
396 H. FUJII and H. SUN

400

30C ............................................................................
Nd-Fe-B ..............................
(s)

200
~E
"l-
rn
Sm-Co
lOO ................................................................................ (B) ......................................
t Nd-Fe-B
(B) i

0.2 0.4 0.6 0.8 1.0


Cost / g ( arb.unit )
Fig. 7.1. The maximumenergy product as a function of cost per unit mass for various permanent
magnets (provided by S. Suzuki).

the nitrogenation processes (Iriyama et al. 1992, Fukuno et al. 1992, Suzuki et al.
1993), but the temperature coefficient is still large compared to the Sm-Co bonded
magnets. Obviously the Sm-Fc-N magnets have to be further improved, especially
the pinning-type anisotropic magnets with nano-structurc are of much interest. The
appearance of a new candidate for permanent magnets provides us with new hopes
and also presents new challenges as to the realization of a material that can be
practically used. In solving the problems and making progresses, we believe that
better understanding of the basic magnetism in the interstitial modified compounds
and sophistication in the permanent magnet processing techniques are important
issues.

Acknowledgements
The authors would like to express their sincere thanks to Mr. M. Akayama, Mr. Y.
Miyazaki and Mr. K. Tatami for their kind assistance in the preparation of this
manuscript. Prof. S. Asano is thanked for providing us some of his original figures
for being used in this manuscript.
H. Fujii would like to take this chance to thank the NEDO, Japan, for the financial
support in the past three years of an International Joint Research Programme on
Development of Functions and Applications of Interstitially-Modified Rare Earth
Intermetatlics from 1991 to 1993. We are grateful to our colleagues of the NEDO
team, especially Prof. J.M.D. Coey, Prof. S. Asano and Dr. M. Sagawa, for the
kind cooperations and many helpful discussion.
INTERSTITIALLY MODIFIED INTERMETALLICS 397

References

Akayama, M., H. Fujii, K. Yamamoto and K. Cao, L., L.S. Kong and B.G. Shen, 1992a, J.
Tatami, 1994, J. Magn. Magn. Mater. 130, Phys.: Condens. Matter 4, L515.
99. Cao, L., L.S. Kong and B.G. Shen, 1992b, Phys.
Altounian, Z., X. Chen, L.X. Liao, D.H. Ryan Status Solidi A: 134, K69.
and J.O. StrOm-Olsen, 1993, J. Appl. Phys. Cao, L., L.S. Kong and B.G. Shen, 1993a, J.
73, 6017. Phys.: Condens. Matter 5, 2001.
Anagnostou, M., C. Christides and D. Niarchos, Cao, L., L.S. Kong and B.G. Shen, 1993b, J.
1991a, Solid State Commun. 78, 681. Appl. Phys. 73, 5887.
Anagnostou, M., C. Christides, M. Pissas and D. Capehart, T.W., R.K. Mishra and EE. Pinkerton,
Niarchos, 1991b, J. Appl. Phys. 70, 6012. 1991, Appl. Phys. Lett. 58, 1395.
Anagnostou, M.S., I. Panagiotopoulos, A. Chen, X., L.X. Liao, Z. Altounian, D.H. Ryan
Kostikas, D. Niarchos and G. Zouganelis, and J.O. Str6m-Olsen, 1992, J. Magn. Magn.
1994, J. Magn. Magn. Mater. 130, 57. Mater. 111, 130.
Asano, S., S. Ishida and S. Fujii, 1993, Physica Chen, X., Z. Altounian and D.H. Ryan, 1993a,
B 190, 155. J. Magn. Magn. Mater. 125, 169.
Barrett, R., D. Fruchart, J.L. Soubeyroux, R. Chen, X., D.H. Ryan, Z. Altounian and L.X.
Ferre, R. Fruchart and A. Stergiou, 1993, J. Liao, 1993b, J. Appl. Phys. 73, 6038.
Alloys Comp. 201, 29. Cheng, S.E, R. Segnan, J.R. Cullen, A.E. Clark
Beloritzky, E., M.A. Fr6my, J.P. Gavigan, D. and M.Q. Huang, 1993, J. Appl. Phys. 73,
Givord and H.S. Li, 1987, J. Appl. Phys. 61, 5733.
3971. Christodoulou, C.N. and T. Takeshita, 1993a, J.
Benz, M.G. and D.L. Martin, 1970, Appl. Phys. Alloys Comp. 191, 279.
Lett. 17, 196. Christodoulou, C.N. and T. Takeshita, 1993b, J.
Beuerle, T., P. Braun and M. Fahnle, 1991, J. Alloys Comp. 194, 31.
Magn. Magn. Mater. 94, Lll. Christodoulou, C.N. and T. Takeshita, 1993c, J.
Beuerle, T. and M. Ffihnle, 1992, Phys. Status Alloys Comp. 196, 161.
Solidi B: 174, 257. Christodoulou, C.N. and T. Takeshita, 1993d, J.
Block, G. and W. Jeitschko, 1986, Inorg. Chem. Alloys Comp. 196, 155.
25, 279. Christodoulou, C.N. and T. Takeshita, 1993e, J.
Block, G. and W. Jeitschko, 1987, J. Solid State Alloys Cornp. 202, 173.
Chem. 70, 271. Christodoulou, C.N. and T. Takeshita, 1993f, J.
Buschow, K.H.J., 1972, J. Less-Common Met. Alloys Comp. 198, 1.
26, 329. Coehoorn, R., 1990, Phys. Rev. B 41, 11790.
Buschow, K.H.J., 1977, Rep. Prog. Phys. 40, Coehoorn, R., 1991, J. Magn. Magn. Mater. 99,
1179. 55.
Buschow, K.H.J., 1991, Rep. Prog. Phys. 54, Coehoorn, R. and K.H.J. Buschow, 1991, J.
1123. Appl. Phys~ 69, 5590.
Buschow, K.H.J. and J.S. van Wieringen, 1970, Coehoorn, R. and G.H.D. Daalderop, 1992, J.
Phys. Status Solidi 42, 231. Magn. Magn. Mater. 104-107, 1081.
Buschow, K.H.J., P.C.P. Bouten and A.R. Coehoorn, R., D.B. de Mooij and C. de Waard,
Miedema, 1982, Rep. Prog. Phys. 45, 937. 1989, J. Magn. Magn. Mater. 80, 101.
Buschow, K.H.J., T.H. Jacobs and W. Coene, Coehoorn, R., K.H.J. Buschow, M.W. Dirken
1990, IEEE Trans. Magn. MAG-26, 1364. and R.C. Thiel, 1990, Phys. Rev. B 42,
Cadogan, J.M., R.K. Day, J.B. Dunlop and A. 4645.
Margarian, 1993, J. Alloys Comp. 201, L1. Coene, W., F. Hakkens, T.H. Jacobs, D.B.
Calka, A., P. Millet, W. Kaczmarek and J.S. de Mooij and K.H.J. Buschow, 1990, J.
Williams, 1992, in: Proc. 12th Workshop Less-Common. Met. 157, 255.
on RE Magnets and Their Appl. (Scott Four Coey, J.M.D. and H. Sun, 1990, J. Magn. Magn.
Colour Print, Perth, WA, Australia) p. 79. Mater. 87, L251.
Callen, H.B. and E. Callen, 1966, J. Phys. Chem. Coey, J.M.D. and Y. Otani, 1991, J. Magn. Soc.
Solids 27, 1271. Jpn 15, 769 (in Japanese).
398 H. FUJII and H. SUN

Coey, J.M.D., J.E Lawler, H. Sun and J.E.M. Fruchart, D., O. Isnard, S. Miraglia, L. Pontonnier,
Allan, 1991a, J. Appl. Phys. 69, 3007. J.L. Soubeyroux and R. Fruchart, 1994, J.
Coey, J.M.D., H. Sun, Y. Otani and D.P.E Alloys Comp. 203, 157.
Hurley, 1991b, J. Magn. Magn. Mater. 98, Fujii, H., K. Tatami, M. Akayama and K.
76. Yamamoto, 1992a, in: Proc. 6th Int. Conf.
Coey, J.M.D., R. Skomski and S. Wirth, 1992, on Ferrites, ICF6, Tokyo and Kyoto, Japan
IEEE Trans. Magn. MAG-28, 2332. (The Japan Society of Powder and Powder
Colucci, C.C., S. Gama and EA.O. Cabral, 1992, Metallurgy, Tokyo) p. 1081.
1EEE Trans. Magn. MAG-28, 2578. Fujii, H., Y. Yamamoto, K. Tatami and M.
Colucci, C.C., S. Gama, L.C. Labald and C.A. Akayama, 1992b, in: Proc. 2nd Int. Symp.
Ribeiro, 1993a, J. Magn. Magn. Mater. 125, on Physics of Magnetic Materials, Vol. 2
161. (International Academic Publishers, Beijing)
Colucci, C.C., S. Gama and C.A. Ribeiro, 1993b, p. 624.
J. Alloys Comp. 194, 181. Fujii, H., K. Tatami, M. Akayama, H. Nagata
Croat, J.J., J.E Herbst, R.W. Lee and F.E. and K. Nakao, 1994a, Advanced Mate-
Pinkerton, 1984, J. Appl. Phys. 55, 2078. rials '93 I/B, Trans. Mat. Res. Soc. Jpn B
Das, D.K., 1969, IEEE Trans. Magn. MAG-5, 14, 1069.
214. Fujii, H. and K. Tatami, 1994b, unpublished
de Mooij, D.B. and K.H.J. Buschow, 1988, J. results.
Less-Common Met. 142, 349. Fujii, H., M. Akayama, K. Nakao and K. Tatami,
Ding, J. and M. Rosenberg, 1991, J. Less-Common 1995a, J. Alloys Comp. 219, 10.
Fujii, H., Y. Miyazaki, K. Tatami, H. Sun, Y.
Met. 168, 335.
Morri, M. Akayama and S. Funahashi, 1995b,
Ding, J., R. Street and P.G. McCormick, 1992a,
J. Magn. Magn. Mater. 140-144, 1089.
J. Magn. Magn. Mater. 115, 211.
Fukuno, A., C. Ishizaka and T. Yoneyama, 1991,
Ding, J., P.G. McCormick and R. Street, 1992b,
J. Appl. Phys. 70, 6021.
Appl. Phys. Lett. 61, 2721.
Fukuno, A., C. Ishizaka, H. Chihara and T.
Ding, J., P.G. McCormick and R. Street, 1992c, Yoneyama, 1992a, IEEE Trans. Magn.
J. Alloys Comp. 189, 83. MAG-28, 2575
Ding, J., P.G. McCormick and R. Street, 1993, Fukuno, A., C. Ishizaka and T. Yoneyama, 1992,
J. Magn. Magn. Mater. 124, 1. in: Proc. 12th Workshop on RE Magnets and
Dirken, M.W., R.C. Thiel, L.J. de J0ngh, Their Appl. (Scott Four Colour Print, Perth,
T.H. Jacobs and K.H.J. Buschow, 1989, J. WA, Australia) p. 60.
Less-Common Met. !55, 339. Ge, S.L., Q. Pan, G.Q. Yang, J.S. Zhang, Y.
Dirken, M.W., R.C. Thiel, R. Coehoorn, T.H. Yang, W. Tao, Y.C. Yang and L.S. Kong,
Jacobs and K.H.J. Buschow, 1991; J. Magn. 1992, Solid State Commun. 83, 487.
Magn. Mater. 94, L15. Givord, D. and R. Lemaire, 1974, IEEE Trans.
Durst, K.D. and H. Kronmialler, 1986, J. Magn. Magn. MAG-10, 109.
Magn. Mater. 59, 86. Gong, Wei and G.C. Hadjipanayis, 1992, IEEE
Endoh, M., M. Iwata and M. Tokunaga, 1991, J. Trans. Magn. MAG-28, 2563.
Appl. Phys. 70, 6030. Gong, Wei and G.C. Hadjipanayis, 1993, J. Appl.
Endoh, M., K. Nakamura and H. Mikami, 1992, Phys. 73, 6245.
IEEE Trans. Magn. MAG-28, 2560. Gr~3ssinger,R., X.C. Kou, T.H. Jacobs and K.H.J.
Fahnle, M. and T. Beuerle, 1993, Phys. Status Buschow, 1991, J. Appl. Phys. 69, 5596.
Solidi B: 177, K95. Gu, Z.Q. and W.Y. Lai, 1992, J. Appl. Phys. 71,
Fernando, A.S., J.P. Woods, S.S. Jaswal, B.M. 8911.
Patterson, D. Welipitiya, A.S. Nozareth and Gu, Z.Q., W.Y. Lai, X.F. Zhong and W.Y. Ching,
D.J. Sellmyer, 1993, J. Appl. Phys. 73, 6919. 1993, J. Appl. Phys. 73, 6928.
Florio, J.V., N.C. Baenzinger and R.E. Rundle, Gubbens, P.C.M., A.M. van der Kraan and K.H.J.
1956, Acta. Cryst. 9, 355. Buschow, 1976, Phys. Status Solidi A: 34,
Freeman, A.J. and J.R Desclaux, 1979, J. Magn. 729.
Magn. Mater. 12, 11. Gubbens, P.C.M., A.M. van der Kraan, T.H.
Friedberg, R. and D.J. Paul, 1975, Phys. Rev. Jacobs and K.H.J. Buschow, 1989, J. Magn.
Lett. 34, 1234. Magn. Mater. 80, 265.
INTERSTITIALLY MODIFIED INTERMETALLICS 399

Gubbens, EC.M., A.A. Moolenaar, G.J. Boender, Hu, J.E, X.C. Kou, T. Dragon, H. Kronmtiller
A.M. van der Kraan, T.H. Jacobs and K.H.J. and B.E Hu, 1993d, Phys. Status Solidi A:
Buschow, 1991, J. Magn. Magn. Mater. 97, 139, 199.
69. Hu, Z. and W.B. Yelon, 1994, Solid State
Gubbens, EC.M., A.A. Moolenaar, T.H. Jacobs Commun. 91, 223.
and K.H.J. Buschow, 1992, J. Magn. Magn. Huang, M.Q., B.M. Ma, W.E. Wallace and S.G.
Mater. 104-107, 1113. Sankar, 1990, Proc. 6th Int. Syrup. on Magn.
Gubbens, P.C.M., A.A. Moolenaar and K.H.J. Anisotropy and Coercivity in Rare Earth
Buschow, 1994, J. Alloys Comp. 203, 199. Transition Metal Alloys (Carnegie-Mellon
Gueramian, M., A. Bezinge, K. Yvon and J. Univ., Pittsburgh, PA) p. 204.
Muller, 1987, Solid State Commun. 64, 639. Huang, M.Q., L.Y. Zhang, B.M. Ma, Y. Zheng,
Haije, W.G., T.H. Jacobs and K.H.J. Buschow, J.M. Elbicki, W.E. Wallace and S.G. Sankar,
1990, J. Less-Common Met. 163, 353. 1991a, J. Appl. Phys. 70, 6027.
Huang, M.Q., Y. Zheng, K. Miller, J.M. Elbicki,
Handstein, A., P.A.P. Wendhausen, D. Eckert
S.G. Sankar and W.E. Wallace, 1991b, J.
and K.-H. MUller, 1992, IEEE Trans. Magn.
Appl. Phys. 70, 6024.
MAG-28, 2596. Huang, M.Q., Y. Zheng, K. Miller, J.M. Elbicki,
Helmholdt, R.B. and K.H.J. Buschow, 1989, J. S.G. Sankar, W.E. Wallace and R. Obermyer,
Less-Common Met. 155, 15. 1991c, J. Magn. Magn. Mater. 102, 91.
Helmolt, R.V., J. Wecker, C. Kuhrt, L. Schultz Huang, M.Q., Y. Zheng, K. Miller, J.M. Elbicki,
and K. Samwer, 1993, IEEE Trans. Magn. W.E. Wallace and S.G. Sankar, 1992, IEEE
MAG-29, 2842. Trans. Magn. MAG-28, 2859.
Hiraga, K., K. Okamoto and T. Irigama, 1993, Hurley, D.EE and J.M.D. Coey, 1991, J. Magn.
Mater. Trans. JIM-34, 569. Magn. Mater. 99, 229.
Hirosawa, S., K. MaNta, T. Ikegami and M. Hurley, D.EE and J.M.D. Coey, 1992, J. Phys.:
Umemoto, 1992, in: Proc. 7th Int. Symp. Condens. Matter 4, 5573.
on Magnetic Anisotropy and Coercivity in Hutchings, ET., 1964, Solid State Phys.
RE-TM Alloys (Scott Four Colour Print, (Academic Press, New York) 16, 227.
Perth, WA, Australia) p. 389. Ibberson, R.M., O. Moze, T.H. Jacobs and K.H.J.
Hu, B.P., H.S. Li, J.P. Gavigan and J.M.D. Coey, Buschow, 1991, J. Phys.: Condens. Matter 3,
1989, J. Phys.: Condens. Matter 1, 755. 1219.
Hu, B.P. and G.C. Liu, 1991, Solid State Iriyama, T., K. Kobayashi, N. Imaoka, T.
Commun. 79, 785. Fukuda, H. Kato and Y. Nakagawa, 1992,
Hu, B.P., H.S. Li, H. Sun, J.E Lawler and J.M.D. IEEE Trans. Magn. MAG-28, 2326.
Coey, 1990, Solid State Commun. 76, 587. Ishida, S., S. Asano and S. Fujii, 1994, Physica
Hu, B.P., H.S. Li, H. Sun and J.M.D. Coey, B 193, 66.
1991, J. Phys.: Condens. Matter 3, 3983. Ishizaka, C., T. Yoneyama and A. Fukuno, 1993,
Hu, B.P., X.L. Rao, J.M. Xu, G.C. Liu, E Cao, IEEE Trans. Magn. MAG-29, 2833.
X.L. Dong, H. Li, L. Yin and Z.R. Zhao, Isnard, O., J.L. Soubeyroux, S. Miraglia, D.
Fruchart, L.M. Garcia and J. Bartolom6,
1992, J. Magn. Magn. Mater. 114, 138.
1992a, Physica B 180-181, 624.
Hu, B.P., X.L. Rao, J.M. Xu, G.C. Liu, Y.Z.
Isnard, O., S. Miraglia, J.L. Soubeyroux and D.
Wang, X.L. Dong, D.X. Zhang and M. Cai, Fruchart, 1992b, J. Alloys Comp. 190, 129.
1993, J. Appl. Phys. 74, 489. Isnard, O., S. Miraglia, J.L. Soubeyroux, D.
Hu, J.E, X.C. Kou, H. Kronmtiller and S.Z. Fruchart and J. Pannetier, 1992c, Phys. Rev.
Zhou, 1992, Phys. Status Solidi A: 134, 499. B 45, 2920.
Hu, J.E, T. Dragon, M.-L. Sartorelli and H. Isnard, O., J.L. Soubeyroux, D. Fruchart, T.H.
Kronmtiller, 1993a, Phys. Status Solidi A: Jacobs and K.H.J. Buschow, 1992e, J. Alloys
136, 207. Comp. 186, 135.
Hu, J.E, I. Kleinschroth, R. Reisser, H. Isnard, O., S. Miraglia, J.L. Soubeyroux, D.
Kronmliller and S.Z. Zhou, 1993b, Phys. Fruchart, J. Deportes and K.H.J. Buschow,
Status Solidi A: 138, 257. 1993, J. Phys.: Condens. Matter 5, 5481.
Hu, J.E, H. Gerth, A. Forkl, G. Martinek, X.C. Isnard, O., S. Miraglia, D. Fruchart, J. Deportes
Kou and H. Kronmtiller, 1993c, Phys. Status and E L'Heritier, 1994, J. Magn. Magn.
Solidi A: 137, 227. Mater. 131, 76.
400 H. FUJII and H. SUN

Isnard, O., E Vulliet, A. Blaise, J.E Sanchez, Kim, R.K. and M. Takahashi, 1972, Appl. Phys.
S. Miraglia and D. Fruchart, 1994b, J. Magn. Lett. 2O, 492.
Magn. Mater. 131, 83. Kim, Y.B., H.T. Kim, K.W. Lee, C.S. Kim
Itsukaichi, T., M. Umemoto, I. Okane and S. and T.K. Kim, 1992, IEEE Trans. Magn.
Hirosawa, 1993, J. Alloys Comp. 193, 262. MAG-28, 2566.
Jacobs, T.H., M.W. Dirken, R.C. Thiel, L.J. de Kneller, E.E and R. Hawig, 1991, IEEE Trans.
Jongh and K.H.J. Buschow, 1990, J. Magn. Magn. MAG-27, 3588.
Magn. Mater. 83, 293. Kobayashi, K., T. Iriyama, N. Imaoka, T. Suzuki,
Jacobs, T.H., G.J. Long, O.A. Pringle, E H. Kato and Y. Nakagawa, 1992, in: Proc.
Grandjean and K.H.J. Buschow, 1991, J. 12th Workshop on RE Magnets and Their
Appl. Phys. 70, 5983. Appl. (Scott Four Colour Print, Perth, WA,
Jaswal, S.S., 1992, IEEE Trans. Magn. Australia) p. 32.
MAG-28, 2322. Kong, L.S., L. Cao and B.G. Shen, 1992, J.
Jaswal, S.S., 1993, Phys. Rev. B 48, 6156. Magn. Magn. Mater. 115, L137.
Jaswal, S.S., Y.G. Ren and D.J. Sellmyer, 1990, Kong, L.S., L. Cao and B.G. Shen, 1993a, Phys.
J. Appl. Phys. 67, 4564. Status Solidi A: 135, K75.
Jaswal, S.S., W.B. Yelon, G.C. Hadjipanayis, Kong, L.S., L. Cao and B.G. Shen, 1993b, J.
Y.Z. Wang and D.J. Sellmyer, 1991, Phys. Alloys Comp. 191, 301.
Rev. Lett. 67, 644. Kong, L.S., B.G. Shen, L. Cao, H.Y. Gong and
Kajitani, T., Y. Morri, S. Funahashi, T. Iriyama, Y.L. Chen, 1993c, J. Phys.: Condens. Matter
K. Kobayashi, H. Kato, Y. Nakagawa and K. 5, 2415.
Hiraya, 1993a, J. Appl. Phys. 73, 6032. Kong, L.S., L. Cao, B.G. Shen, J.G. Zhao, H.Y.
Kajitani, T., Y. Morri, S. Funahashi, T. Iriyama, Gong and Y.L. Chen, 1993d, J. Alloys Comp.
K. Kobayashi, H. Kato, Y. Nakagawa and K. 196, 183.
Hiraga, 1993b, in: Proc. 15th Int. Symp. on Kong, L.S., L. Cao and B.G. Shen, 1993e, J.
Adv. Nuclear Energy Research-Neutrons as Magn. Magn. Mater. 124, 301.
Microscopic Probes-1993 (JAERI, Tokai), p. Kou, X.C., R. Gr6ssinger, T.H. Jacobs and K.H.J.
455. Buschow, 1990, J. Magn. Magn. Mater. 88,
Kapusta, Cz., R.J. Zhou, M. Rosenberg, P.C. 1.
Riedi and K.H.J. Buschow, 1992, J. Alloys Kou, X.C., R. Grtissinger, M. Katter, J. Wecker,
Comp. 178, 139. L. Schultz, T.H. Jacobs and K.H.J. Buschow,
Kapusta, Cz., M. Rosenberg, H. Figiel, T.H. 1991a, J. Appl. Phys. 70, 2272.
Jacobs and K.H.J. Buschow, 1992b, J. Magn. Kou, X.C., R. Gr6ssinger, X. Li, J.P. Liu, ER.
Magn. Mater. 104--107, 1331. de Boer, M. Katter, J. Wecker, L. Schultz,
Kapusta, Cz., M. Rosenberg, R.G. Graham, T.H. Jacobs and K.H.J. Buschow, 1991b, J.
P.C. Riedi, T.H. Jacobs and K.H.J. Buschow, Appl. Phys. 70, 6015.
1992c, J. Magn. Magn. Mater. 104-107, Kou, X.C., R. Gr0ssinger, T.H. Jacobs and K.H.J.
1333. Buschow, 1991c, Physica B 168, 181.
Kato, H., M. Yamada, G. Kido, Y. Nakagawa, Kou, X.C., W.J. Qiang, H. Kronmtiller and L.
T. Iriyama and K. Kobayashi, 1993, J. Appl. Schultz, 1993, J. Appl. Phys. 74, 6791.
Phys. 73, 6931. Kronm~ller, H., 1978, J. Magn. Magn. Mater.
Katter, M., J. Wecker, L. Schultz and R. 7, 341.
Gr6ssinger, 1990, J. Magn. Magn. Mater. Kuhrt, C., K. O'Donnell, M. Katter, ]. Wecker,
92, L14. K. Schuitzke and L. Schultz, 1992a, Appl.
Katter, M., J. Wecker and L. Schultz, 1991, J. Phys. Lett. 60, 3316.
Appl. Phys. 70, 3188. Kuhrt, C., M. Katter, J. Wecker, K. Schnitzke
Katter, M., J. Wecker, C. Kuhrt and L. Schultz, and L. Schultz, 1992b, Appl. Phys. Lett. 60,
1992a, J. Magn. Magn. Mater. 117, 419. 2029.
Katter, M., J. Wecker, C. Kuhrt, L. Schultz and Kuhrt, C., K. Schnitzke and L. Schultz, 1993, J.
R. Gr0ssinger, 1992b, J. Magn. Magn. Mater Appl. Phys. 73, 6026.
114, 35. Leccabue, E, B.E. Watts, R. Panizzieri, G.
Katter, M., J. Wecker, C. Kuhrt, L. Schultz, X. Bocelli, G. Calestani, L. Dimesso, A. Deriu
C. Kou and R. GrOssinger, 1992c, J. Magn. and D. Carrillo, 1992, J. Magn. Magn. Mater.
Magu. Mater 111, 293. 114, 291.
INTERSTITIALLY MODIFIED INTERMETALLICS 401

Li, H.S. and J.M. Cadogan, 1991, Solid State Long, G.J., O.A. Pringle, E Grandjean, T.H.
Commun. 80, 905. Jacobs and K.H.J. Buschow, 1994, J. Appl.
Li, H.S. and J.M. Cadogan, 1992a, J. Magn. Phys. 75, 2598.
Magn. Mater. 103, 53. Lu, Y., O. Tegus, Q.A. Li, N. Tang, M.J. Yu,
Li, H.S. and J.M. Cadogan, 1992b, Solid State R.W. Zhao, J.P. Kuang, EM. Yang, G.E Zhou,
Commun. 82, 121. X. Li and ER. de Boer, 1992, Physica B 177,
Li, H.S. and J.M. Cadogan, 1992c, J. Magn. 243.
Magn. Mater. 109, L153. Luo, S., G.W. Zhang, Z.H. Liu, X.D. Pei, W.
Li, H.S. and J.M.D. Coey, 1992, J. Magn. Magn. Jiang and W.W. Ho, 1987a, J. Magn. Magn.
Mater. 115, 152. Mater. 70, 311.
Li, H.S., J.M. Cadogan, R.L. Davis, A. Luo, S., Z.H. Liu, G.W. Zhang, X.D. Pei, W.
Margarian and J.B. Dunlop, 1994, Solid State Jiang and W.W. Ho, 1987b, IEEE Trans.
Commun. 90, 487. Magn. MAG-23, 3095.
Li, X.W., N. Tang, Z.H. Lu, T.Y. Zhao, W.G. Machida, K., E. Yamamoto and G. Adachi, 1993,
Lin, R.W. Zhao and EM. Yang, 1993, J. Appl. J. Alloys Comp. 193, 271.
Phys. 73, 5890. Majima, K., N. Niimi, S. Katsuyama and H.
Li, Y.P., H.S. Li and J.M.D. Coey, 1991, Phys. Nagai, 1993, J. Alloys Comp. 193, 268.
Status Solidi B: 166, K107. Manaf, A., R.A. Buckley and H.A. Davies,
Li, Y.P. and J.M.D. Coey, 1992, Solid State 1993a, J. Magn. Magn. Mater. 128, 302.
Commun. 81, 447. Manaf, A., M. A1-Khafaji, P.Z. Zhang, H.A.
Li, Z.W., X.Z. Zhou and A.H. Morrish, 1992, J. Davies, R.A. Buckley and W.M. Rainforth,
Phys.: Condens. Matter 4, 10409. 1993b, J. Magn. Magn. Mater. 128, 307.
Li, Z.W., X.Z. Zhou and A.H. Morrish, 1993, J. Middleton, D.P. and K.H.J. Buschow, 1994, J.
Phys.: Condens. Matter 5, 3027. Alloys Comp. 203, 217.
Liang, J.Z., M.J. Yu, N. Tang, Y.L. Liu, J.E Miraglia, S., J.L. Soubeyroux, C. Kolbeck, O.
Liu, S.Q. Ji and G.M. Chen, 1991, J. Magn. Isnard, D. Fruchart and M. Guillot, 1991, J.
Magn. Mater. 102, 217.
Less-Common Met. 171, 51.
Liao, L.X., Z. Altonunian and D.H. Ryan, 1991,
Mishra, R.K., G. Thomas, T. Yoneyama, A.
J. Appl. Phys. 70, 6006.
Fukuno and T. Ojima, 1981, J. Appl. Phys.
Liao, L.X., X. Chen, Z. Altonunian and D.H.
52, 2517.
Ryan, 1992, Appl. Phys. Lett. 60, 129.
Mohn, P. and E.P. Wohlfarth, 1987, J. Phys. F
Lindgard, P.A. and O. Danielsen, 1975, Phys.
16, 2421.
Rev. B 11, 351.
Mose, D., L. Pareti, M. Solzi and W.I.F. David,
Liu, J.P., K. Bakker, ER. de Boer, T.H. Jacobs,
1988, Solid State Commun. 66, 465.
D.B. de Mooij and K.H.J. Buschow, 1991, J.
Less-Common Met. 170, 109. Mukai~ T. and T. Fujimoto, 1992, J. Magn.
Liu, J.P., J.H.V.J. Brabers, A.J.M. Winkelman, Magn. Mater. 103, 165.
A.A. Menovsky, ER. de Boer and K.H.J. Mulder, EM., R.C. Thiel, R. Coehoorn, T.H.
Buschow, 1993, J. Alloys Comp. 200, L3. Jacobs and K.H.J. Buschow, 1992, J. Magn.
Liu, N.C., H.H. Stadelmaier and G. Schneider, Magn. Mater. 117, 413.
1987, J. Appl. Phys. 61, 3574. Mailer, K.-H., G. Leitner, W. Pitschke, P.A.P.
Liu, Y.N., M.P. Dallimore, T. Alonso and P.G. Wendhausen, A. Handstein and D. Eckert,
McCormick, 1992, in: Proc. 12th Workshop 1992, Phys. Status Solidi A: 133, K37.
on RE Magnets and Their Appl. (Scott Four Nagata, H. and H. Fujii, 1991, Jpn. J. Appl.
Colour Print, Perth, WA, Australia) p. 25. Phys. 30, L367.
Loewenhaupt, M., P. Tils, D.P. Middleton, K.H.J. Nakamura, H., S. Sugimoto, M. Okada and M.
Buschow and R. Eccleston, 1994, J. Magn. Homma, 1992, Mater. Chem. Phys. 32, 280.
Magn. Mater. 129, L151. Nikitin, S.A., A.M. Tishin, M.D. Kuz'min and
Long, G.J., O.A. Pringle, E Grandjean and Yu.I. Spichkin, 1991, Phys. Lett. A 153, 155.
K.H.J. Buschow, 1992, J. Appl. Phys. 72, Ohno, K., T. Saito, K. Shinagawa and T.
4845. Tsushima, 1993, IEEE Trans. Magn.
Long, G.J., O.A. Pringle, E Grandjean, W.B. MAG-29, 2854.
Yelon and K.H.J. Buschow, 1993, J. Appl. Ojima, T., S. Tomizawa, T. Yoneyama and T.
Phys. 74, 504. Hori, 1977, Jpn. J. Appl. Phys. 16, 691.
402 H. FUJII and H. SUN

Okada, M., S. Sugimoto and M. Homma, 1992, Shen, B.G., L.S. Kong, EW. Wang and L. Cao,
in: Proc. 6th Int. Conf. on Ferrites (ICF6), 1993a, Appl. Phys. Lett. 63, 2288.
Tokyo and Kyoto, Japan (The Japan Society Shen, B.G., F.W. Wang, L.S. Kong, L. Cao and
of Powder and Powder Metallurgy, Tokyo) p. H.Q. Guo, 1993b, J. Magn. Magn. Mater.
1087. 127, L267.
Onneby, C., T. DebRoy and S. Seetharaman, Singleton, E.W. and G.C. Hadjipanayis, 1993, J.
1993, J. Magn. Magn. Mater. 127, 307. Magn. Magn. Mater. 128, L21.
Otani, Y., A. Moukarika, H. Sun and J.M.D. Singleton., E.W., Z.X. Tang, G.C. Hadjipanayis
Coey, 1991, J. Appl. Phys. 69, 6735. and V. Papaefthymiou, 1993, IEEE Trans.
Pinkerton, F.E. and C.D. Fuerst, 1992, Appl. Magn. MAG-29, 2836.
Phys. Lett. 60, 2558. Skomski, R. and J.M.D. Coey, 1993a, J. Appl.
Pinkerton, F.E. and C.D. Fuerst, 1993, J. Mater. Phys. 73, 7602.
Eng. Perform. 2, 219. Skomski, R. and J.M.D. Coey, 1993b, J. Mater.
Popov, A.G., E.V. Belozerov, A.G. Kuchin, A.S. Eng. Perform. 2, 241.
Ermolenko, G.M. Makarova, V.S. Gaviko and Skomski, R., C. Murray, S. Brennan and J.M.D.
V.I. Khrabrov, 1990, Phys. Status Solidi A: Coey, 1993, J. Appl. Phys. 73, 6940.
121, K i l l . Steenwijk, F.J. van, H.T. Lefever, R.C. Thiel and
Pringle, O.A., G.J. Long, E Grandjean and K.H.J. Buschow, 1977, Physica B 92, 52.
K.H.J. Buschow, 1992, J. Magn. Magn. Steiner, W. and R. Haferl, 1977, Phys. Status
Mater. 104-107, 1123. Solidi A: 42, 739.
Psycharis, V., M. Anagnostou, C. Christides and Stevens, K.W.H., 1952, Proc. Phys. Soc. A 65,
D. Niarchos, 1991, J. Appl, Phys. 70, 6122. 109.
Qi, Q.N., H. Sun and J.M.D. Coey, 1991, Strnat, K.J., 1967, Cobalt 36, 133.
Sugimoto, S., H. Nakamura, M. Okada and M.
Hyperfine Interact. 68, 27.
Homma, 1992a, in: Proc. 12th Workshop
Qi, Q.N:, M.D. Kuz'min, H. Sun and J.M.D.
on RE Magnets and Their Appl. (Scott Four
Coey, 1992a, J. Alloys Comp. 182, 313.
Colour Print, Perth, WA, Australia) p. 218.
Qi, Q.N., H. Sun, R. Skomski and J.M.D. Coey,
Sugimoto, S., K. Kurihara, H. Nakamura, M.
1992b, Phys. Rev. B 45, 12278.
Okada and M. Homma, 1992b. Mater. Trans.
Qi, Q.N., Y.P. Li and J.M.D. Coey, 1992c, J.
JIM-33, 146.
Phys.: Condens. Matter 4, 8209.
Sugimoto, S., H. Nakamura, M. Okada and M.
Rodewald, W., M. Velicescu, B. Wall and G.W. Homma, 1992c, in: Proc. 12th Workshop
Reppel, 1992, in: Proc. 12th Workshop on on RE Magnets and Their Appl. (Scott Fotu-
RE Magnets and Their Appl. (Scott Four Colour Print, Perth, WA, Australia) p. 372.
Colour Print, Perth, WA, Australia) p. 191. Sugita, Y., K. Mitsuoka, M. Komuro, H.
Rodewald, W., B. Wall, M. Katter, M. Velicescu Hoshiya, Y. Kosono and M. Hanazono, 1991,
and P. Schrey, 1993, J. Appl. Phys. 73, 5899. J. Appl. Phys. 70, 5977.
Rosenberg, M., R.J. Zhou, M. Katter, L. Schultz Sun, H., B.P. Hu, H.S. Li and J.M.D. Coey,
and G. Filoti, 1993, J. Appl. Phys. 73, 6035. 1990a, Solid State Comrnun. 74, 727.
Sagawa, M., S. Fujimura, H. Yamarnoto, Y. Sun, H., J.M.D. Coey, Y. Otani and D.P.F.
Matsuura and K. Hiraga, 1984, IEEE Trans. Hurley, 1990b, J. Phys.: Condens. Matter 2,
Magn. MAG-20, 1584. 6465.
Sakuma, A., 1992, J. Phys. Soc. Jpn 61, 4119. Sun, H., Y. Otani and J.M.D. Coey, 1992, J.
Sankar, S.G~, V.U.S. Rao, E. Segal, W.E. Magn. Magn. Mater. 104-107, 1439.
Wallace, W.G.D. Frederick and H.J. Garrett, Sun, H., M. Akayama, K. Tatami and H. Fujii,
1975, Phys. Rev. B 11, 435. 1993a, Physica B 183, 33.
Schnitzke, K., L. Schultz, J. Wecker and M. Sun, H., Y. Morii, H. Fujii, M. Akayama and S.
Katter, 1990, Appl. Phys. Lett. 57, 2853. Funahashi, 1993b, Phys. Rev. B 48, 13333.
Schrefl, T., H. Kronm011erand J. Fidler, 1993, J. Sun, J.J., D.S. Xue, B.G. Shen and ES. Li, 1993,
Magn. Magn. Mater. 127, L273. Phys. Status Solidi A: 136, K55.
Schultz, L., K. Schnitzke, J. Wecker and M. Suzuki, S. and T. Miura, 1992, IEEE Trans.
Katter, 1991, Mater. Sci. Eng. A 133, 143. Magn. MAG-28, 994.
Shen, B.G., L.S. Kong and L. Cao, 1992, Solid Suzuki, S., N. Inoue and T. Miura, 1992a, in:
State Commun. 83, 753. Proc. 12th Workshop on RE Magnets and
INTERSTITIALLY MODIFIED INTERMETALLICS 403

Their Appl. (Scott Four Colour Print, Perth, Wang, Y.Z., B.E Hu, X.L. Rao, G.C. Liu, L. Yin,
WA, Australia) p. 227. W.Y. Lai, W. Gong and G.C. Hadjipanayis,
Suzuki, S., N. Inoue and T. Miura, 1992b, IEEE 1993b, J. Appl. Phys. 73, 6251.
Trans. Magn. MAG-28, 2005. Wang, K.Y., Y.Z. Wang, L. Yin, L. Song, X.L.
Suzuki, S., T. Miura and M. Kawasaki, 1993, Rao, G.C. Liu and B.E Hu, 1993, Solid State
IEEE Trans. Magn. MAG-29, 2815. Commun. 88, 521.
Takeshita, T., 1993, J. Alloys Comp. 193, 231. Wei, Y.N., K. Sun, Y.B. Fen, J.X. Zhang, B.P.
Tang, N., Y.L. Liu, M.L Yu, Y. Lu, O. Tegus, Hu, Y.Z. Wang, X.L. Rao and G.C. Liu, 1993,
Q.A. Li, S.Q. Ji and F.M. Yang, 1992, J. J. Alloys Comp. 194, 9.
Magn. Magn. Mater. 104-107, 1086. Weitzer, F., K. Hiebl and P. Rogl, 1991, J. Appl.
Tang, Z.X., E.W. Singleton and G. Hadjipanayis, Phys. 69, 7215.
1992, 1EEE Trans. Magn. MAG-28, 2572. Wendhausen., P.A.P., D. Deckert, A. Handstein
Tang, Z.X., E.W. Singleton and G. Hadjipanayis, and K.-H. MUller, 1992, Phys. Status Solidi
1993a, J. Appl. Phys. 73, 6254. A: 129, K45.
Tang, Z.X., G . C . Hadjipanayis and V. Wiesinger, G. and G. Hilscher, 1991, in:
Papaefthymiou, 1993b, J. Alloys Comp. 194, Handbook of Magnetic Materials, Vol. 6
87. (North-Holland, Amsterdam) p. 511.
Tang, Z.X., X.H. Deng, G.C. Hadjipanayis, Woods, J.P., A.S. Fernando, S.S. Jaswal, B.M.
V. Papaefthymiou and D.J. Sellmyer, 1993c, Patterson, D. Welipitiya and D.J. Sellmyer,
IEEE Trans. Magn. MAG-29, 2839. 1993, J. Appl. Phys. 73, 6913.
Tegus, O., Y. Lu, N. Tang, J.X. Wu, M.G. Yu, Xu, X. and S.A. Shaheen, 1993a, J. Appl. Phys.
Q.A. Li, R.W. Zhao, J. Yuan and EM. Yang, 73, 1892.
1992, IEEE Trans. Magn. MAG-28, 2581. Xu, X. and S.A. Shaheen, 1993b, J. Appl. Phys.
Tomey, E., O. Isnard, A. Fagna, C. Desmoulins, 73, 5896.
S. Miraglia, J.L. Soubeyroux and D. Fruchart, Xu, X.E, Y.X. Sun, Z.X. Liu and C. Lin, 1994,
1993, J. Alloys Comp. 191, 233. Solid State Commun. 89, 409.
Uchida, H., H.-H. Uchida, T. Yanagisawa, Xu, Y., J.M. Elbicki, W.E. Wallace, S. Simizu
H. Kaneko, U. Koike, K. Kamada, 5i. and S.G. Sankar, 1992, IEEE Trans. Magn.
Matsumura, T. Noguchi and T. Kurino, 1992, MAG-28, 2569.
J. Alloys Comp. 184, L5. Xu, Y., T. Ba and Y.L. Liu, 1993, J. Appl. Phys.
Uchida, H.-H., H. Uchida, T. Yanagisawa, S. 73, 6937.
Kise, T. Suzuki, Y. Matsumura, U. Koike, K. Yamaguchi, M. and S. Asano, 1994, J. Phys.
Kamada, T. Kurino and H. Kaneko, 1993, J. Soc. Jpn 63, 1071.
Alloys Comp. 196, 71. Yamamoto, H., J. Iwasawa, T. Kumanbara and T.
Valeanu, M., N. Plugaru and E. Burzo, 1994, Kojima, 1993, IEEE Trans. Magn. MAG-29,
Solid State Commun. 89, 519. 2845.
Verhoef, R., 1991, PhD Thesis, Amsterdam. Yamamoto, I., E Sugaya, S. Takabatake, M.
Wallace, W.E., 1978, in: Topics in Applied Yamaguchi, T. Goto and M.I. Bartashevich,
Physics, Vol. 28: Hydrogen in Metals I, eds 1993, Technical Report I SSP A2654.
G. Alefeld and J. V~lkl (Springer, Berlin) p. Yan, Q.W., P.L. Zhang, Y.N. Wei, K. Sun, B.P.
169. Hu, Y.Z. Wang, G.C. Liu, C. Gau and Y.E
Wallace, W.E., 1985, Prog. Solid State Chem. Chen, 1993, Phys. Rev. B 48, 2878.
16, 127. Yang, C.J., W.Y. Lee and H.S. Shin, 1993, J.
Wang, Y.Z. and G.C. Hadjipanayis, 1991a, J. Appl. Phys. 74, 6824.
Appl. Phys. 69, 5565. Yang, EM., Q.N. Li, Y. Lu, N. Tang, O. Tegus,
Wang, Y.Z. and G.C. Hadjipanayis, 1991b, J. M.J. Yu, R.W. Zhao, B.G. Shen and L.Y.
Appl. Phys. 70, 6009. Yang, 1992, J. Magn. Magn. Mater. 114,
Wang, Y.Z., G.C. Hadjipanayis, A. Kim, D.J. 255.
Sellmyer and W.B. Yelon, 1992, J. Magn. Yang, EM., B. Nasunjilegal, H.Y. Pan, J.L.
Magn. Mater. 104-107, 1132. Wang, R.W. Zhao, B.P. Hu, Y.Z. Wang, H.S.
Wang, Y.Z., G.C. Hadjipanayis, Z.X. Tang, W.B. Li and J.M. Cadogan, 1994, J. Magn. Magn.
Yelon, V. Papaefthymiou, A. Moukarika and Mater. 135, 298.
D.J. Sellmyer, 1993a, J. Magn. Magn. Mater. Yang, J., S.Z. Dong and Y.C. Yang, 1994, J.
119, 41. Appl. Phys. 75, 3013.
404 H. FUJII and H. SUN

Yang, Y.C., X.D. Zhang, L.S. Kong, Q. Pan, J.L. and Their Applications (University of Dayton,
Yang, Y.E Ding, B.S. Zhang, C.T. Ye and L. OH) p. 406.
Jin, 1991a, J. Appl. Phys. 70, 6018. Yu, M.J., N. Tang, Y.L. Liu, O. Tegus, Y. Lu,
Yang, Y.C., X.D. Zhang, L.S. Kong, Q. Pan and J.E Kuang, EM. Yang, X. Li, G.E Zhou and
S.L. Ge, 1991b, Solid State Commun. 78, ER. de Boer, 1992, Physica B 177, 238.
317. Zarechnyuk, O.S. and EI. Kripyakevich, 1962,
Yang, Y.C., X.D. Zhang, L.S. Kong, Q. Pan and Kristallografiya 7, 543.
S.L. Ge, 1991c, Appl. Phys. Lett. 58, 2042. Zeng, Z., Q.Q. Zheng, W.Y. Lai and C.Y. Pan,
Yang, Y.C., X.D. Zhang, S.L. Ge, Q. Pan, L.S. 1992, J. Magn. Magn. Mater. 104-107,
Kong, H.L. Li, J.L. Yang, B.S. Zhang, Y.E 1157.
Ding and C.T. Ye, 1991d, J. Appl. Phys. 70, Zeng, Z., Q.Q. Zheng, W.Y. Lai and C.Y. Pan,
6001. 1993, J. Appl. Phys. 73, 6916.
Yang, Y.C., X.D. Pei, H.L. Li, X.D. Zhang, L.S. Zhang, X.D., Q. Pan, S.L. Ge, Y.C. Yang, J.L.
Kong, Q. Pan and M.H. Zhang, 1991e, J. Yang, Y.E Ding, B.S. Zhang, C.T. Ye and L.
Appl. Phys. 70, 6574. Jin, 1992, Solid State Commun. 83, 231.
Yang, Y.C., X.D. Zhang, L.S. Kong, Q. Pan, S.L. Zhao, T.S., X.C. Kou, R. GrOssinger and H.R.
Ge, J.L. Yang, Y.E Ding, B.S. Zhang, C.T. Kirchmayr, 1991, Phys. Rev. B 44, 2846.
Ye and L. Jin, 1991f, Solid State Commun. Zhao, T.S., X.C. Kou, R. GrOssinger and H.R.
78, 313. Kirchmayr, 1993, J. Appl. Phys. 73, 6041.
Yang, Y.C., X.D. Zhang, L.S. Kong, Q. Pan, Y.T. Zhong, X.E and W.Y. Ching, 1989, Phys. Rev.
Hou, S. Huang, L. Yang and S.L. Ge, 1991g, B 39, 12018.
J. Less-Common Met. 170, 37. Zhong, X.E, R.J. Radwanski, ER. de Boer, T.H.
Yang, Y.C., Q. Pan, X.D. Zhang and S.L. Ge, Jacobs and K.H.J. Buschow, 1990, J. Magn.
1992a, J. Appl. Phys. 72, 2989. Magn. Mater. 86, 333.
Yang, Y.C., Q. Pan, X.D. Zhang, J. Yang, M.H. Zhou, R.J., Th. Sinnemann, M. Rosenberg and
Zhang and S.L. Ge, 1992b, Appl. Phys. Lett. K.H.J. Buschow, 1991, J. Less-Common Met.
61, 2723. 171, 263.
Yang, Y.C., X.D. Zhang, Q. Pan and L.S. Kong, Zhou, R.J., Cz. Kapusta, M. Rosenberg and
1992c, J. Magn. Magn. Mater. 104-107, K.H.J. Buschow, 1992, J. Alloys Comp. 184,
1353. 235.
Yang, Y.C., Q. Pan, X.D. Zhang, M.H. Zhang, Zhou, R.J., M. Rosenberg, M. Katter and L.
C.L. Yang, Y. Li, S.L. Ge and B.E Zhang, Schultz, 1993, J. Magn. Magn. Mater. 118,
1993, J. Appl. Phys. 74, 4066. 110.
Yau, J.M., K.H. Cheng, C.H. Lin and T.S. Chin, Zhou, S.Z., J. Yang, M.C. Zhang, D.Q. Ma,
1993, IEEE Trans. Magn. MAG-29, 2851. EB. Li and R. Wang, 1992, in: Proc.
Yelon, W.B. and G.C. Hadjipanayis, 1992, IEEE 12th Workshop on RE Magnets and Their
Trans. Magn. MAG-28, 2316. Appl. (Scott Four Colour Print, Perth, WA,
Yoneyama, T., S. Tomizawa, T. Hori and T. Australia) p. 44.
Ojima, 1978, in: Proc. 3rd Int. Workshop Zouganelis, G., M. Anagnostou and D. Niarchos,
on Rare Earth-Cobalt Permanent Magnets 1991, Solid State Commun. 77, 11.
chapter 4

FIELD INDUCED PHASE


TRANSITIONS IN FERRIMAGNETS

A.K. Zvezdin
General Physics Institute
Russian Academy of Sciences
Vavilov st. 38
117942 Moscow
Russia

Handbook of Magnetic Materials, Vol. 9


Edited by K. H.J. Buschow
©1995 Elsevier Science B.V. All rights reserved

405
CONTENTS

1. Introduction ................................................................ 408


2. H - T phase diagram of isotropic ferrimagnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
3. Non-cotlinear structures in a weakly anisotropic ferrimagnets. Basic equations . . . . . . . . . . 419
3.1. Therlnodynamical potential of the non-equilibrium state . . . . . . . . . . . . . . . . . . . . . . . . 419
3.2. Magnetic anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421
3.3. Extreme conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 422
3.4. Low magnetic field approximation ......................................... 423
4. H - T phase diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424
4.1. Uniaxial magnetic anisotropy ............................................. 424
4.2. Cubic anisotropy ....................................................... 430
5. Field induced phase transitions in rare-earth-ferrite garnets . . . . . . . . . . . . . . . . . . . . . . . . . . 439
6. Single crystal ferrite garnet films ............................................... 450
7. Some general features of field induced phase transitions ............................ 454
8. Magnetization and susceptibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 456
8.1. Differential susceptibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 456
8.2. The temperature hysteresis of the magnetization. Hall and Faraday effects ........ 462
9. Thermal properties in the vicinity of the spin-reorientation phase transitions ............ 465
9.1. Magnetocaloric effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 465
9.2. Specific heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 466
10. Magnetoelastic anomalies ..................................................... 469
10.1. Magnetostriction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 469
10.2. Thermal expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 472
10.3. Young's modulus, sound velocity change (AE-effect) and sound absorption . . . . . . . . 473
11. Non-collinear phases and domain structure ....................................... 475
11.1. 'Break' of symmetry in the canted phase and formation of domain structures with
twins, triplets and quadruplets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475
11.2. Nucleation of new phases from domain walls. The hysteresisless first order phase
transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 477
11.3. Canted phase domains in ferrite garnet single crystals and films . . . . . . . . . . . . . . . . . 478

406
FIELD I N D U C E D P H A S E TRANSITIONS IN FERRIMAGNETS 407

12. Hexagonal ferrimagnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 482


12.1. Free energy and equilibrium conditions ..................................... 482
12.2. Phase diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485
13. Non-collinear magnetic structures in the intermetallic compounds DyCo 5 .............. 491
13.1. Crystal and magnetic structure ............................................ 491
13.2. Domains in the basal plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 491
13.3. Magnetization and magnetostriction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493
13.4. Magnetizations of the sublattices, exchange field and anisotropy constant . . . . . . . . . . 495
13.5. H - T phase diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 497
14. Spin-flop and spin-reorientation phase transitions in the anisotropic ferrimagnet HoCo3Ni2
with Tcomp = TSR1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499
14.1. Spin reorientation in HoCo3Ni2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499
14.2. The H - T phase diagrams ................................................ 500
14.3. The magnetization measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502
14.4. Magnetostriction and spin-flop transitions ................................... 503
15. Surface anisotropy effects and surface phase transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506
16. Phase transition at the local defect. Dislocations and FIPT's in Gd3Fe50~2 . . . . . . . . . . . . . 508
17. Free-powder samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 512
18. Spin-flop transitions in itinerant metamagnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 521
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 532
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 532
19. Appendix. Microscopic calculation of the thermodynamic potential of the non-equilibrium
state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 532
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 537
1. Introduction

This review is concerned with high magnetic field magnetization processes in fer-
rimagnetic compounds of the f-d type, i.e. in compounds consisting of rare-earth
and transition metal elements of the Fe group. Attention will be mainly focused on
such peculiarities of these processes which are connected with the variation of mu-
tual orientation of their magnetic sublattices and which reveal themselves through
characteristic kinks and jumps in magnetization curves. According to the present
views, magnetic symmetry of a crystal changes when the orientation of magnetic
sublattices spontaneously varies. In other words, a phase transition accompanied by
a change of the magnetic symmetry takes place. In the present terminology such
phase transitions are referred to as field induced phase transitions (FIPT) which may
be of 1st or 2nd order.
Yet this notion comprises a more general class of magnetic phase transitions (Date
1990, Franse 1990). It is evident that such phase transitions are accompanied by an
anomalous behavior of several physical properties and characteristics of ferrimagnets
(susceptibility, thermal properties, magnetostriction, magnetoelastic anomalies, Hall-
effect and magnetoresistance, magnetooptics etc.).
The role of the magnetic anisotropy in field induced phase transitions and in
corresponding phase diagrams has to be discussed in this review. It is natural that the
range of properties related to a compensation point of ferrimagnets (in the vicinity of
which the effects of magnetic anisotropy are most conspicuous) is worth considering
here in detail as well.
Field induced phase transitions are traditional subjects for the physics of mag-
netic phenomena. The first examples attracting considerable attention were spin-flop
transitions. The concept of these transitions was proposed by Nrel (1936) and they
were found later in CuClz-H20 by the Dutch group (Poulis et al. 1951). In these
transitions the collinear phase turns into a canted phase and further into a ferromag-
netic one in an increasing field. A distinctive feature of a spin-flop transition in an
ideal Nrel antiferromagnet is that the differential susceptibility of the material in the
angular phase is not dependent on the magnetic field and temperature. Later similar
phase transitions were investigated in ferrimagnets whose behavior tumed out more
sophisticated and interesting to be studied.
The ferrimagnetic structure may happen to be non-collinear in a certain range
of magnetic fields and temperature. Such structures are the result of a competition
among negative exchange interactions within the sublattices, that force their magnetic
moments to be antiparallel to each other. They may also arise as a consequence of
the interaction of sublattices with the applied field.

408
FIELD INDUCED PHASETRANSITIONSIN FERRIMAGNETS 409

That non-collinear magnetic structures may exist was first demonstrated theoret-
ically by Tyablikov (1956, 1958, 1965). He also defined the conditions of their
existence in the case of isotropic ferrimagnets and calculated anomalies of the mag-
netization, arising in the vicinity of such transitions. Further theoretical analysis
of non-collinear structures in two- or three-sublattice ferrimagnets was carried out
by Gusev (1959), Schltimann (1960), Gusev and Pakhomov (1963). An important
contribution to this problem was made by Clark and Callen (1968). These authors
gave a summary of earlier results and applied them to the study of rare-earth-ferrite
garnet compounds. Due to a negligible exchange interaction between the rare earth
and Fe sublattices, the induced non-collinear structures in garnets were shown to
occur at experimentally attainable values of the field. Perhaps the first experimental
manifestations of these phenomena were observed in rare-earth-iron garnets (Rode
and Vedyaev 1963, Kharchenko et al. 1968). We shall consider the basic theoretical
and experimental results in section 2.
At a compensation point T c a ferrimagnet behaves as an antiferromagnet. For
the ideal N6el ferrimagnet the critical fields of transition in a canted phase are
approximately equal to zero. To comprehend the physical properties of a ferrimagnet
and to obtain its phase diagrams, it is of prime importance to take the real magnetic
anisotropy of the material into account. Consequently, various interesting phase
diagrams result for ferrimagnets close to the compensation temperature. They include
the lines of 1st and 2nd order phase transitions, critical points of the liquid-vapor
type, and tricritical points, where the 1st order phase transition line goes over to
the 2nd one on the line separating the phases, etc. (sections 3-7). Thus a big
number of the anomalies detected experimentally in the vicinity of the compensation
temperature can be explained. These points are considered in sections 8-10.
So far the field induced phase transitions have been investigated more closely in
rare-earth-ferrite garnets where the critical fields of transition into the canted phase
are not too large. Gadolinium-iron garnet was studied in great detail owing to the
fact that the application of Ndel's model is well justified since the Gd +3 ion, whose
ground state is an orbital singlet 8S, has a comparatively small magnetic anisotropy.
The rare-earth sublattices in other rare-earth-ferrite garnets possess a substantial
magnetic anisotropy. Phase diagrams and the character of phase transitions differ
essentially from those of Ndel ferrimagnets. A forthcoming paper will be devoted
to these aspects.
There are several studies in which domain structures have been investigated in
the context of field induced phase transitions (Zvezdin and Matveev 1972a, b, Dik-
stein et al. 1980). Later such domains were experimentally observed in ferrite-
garnets in strong magnetic fields (Kharchenko et al. 1974, Lisovskii and Schapovalov
1974), and in epitaxial ferrite-garnet films (Lisovskii et al. 1976 a, c, Gnatchenko
and Kharchenko 1977, Dikstein et al. 1980) and in the YIG in megagauss fields
(Druzhinin et al. 1981) (see section 11). The discovery of domains in strong mag-
netic fields essentially enriches the knowledge of domain forming in magnets.
410 A.K. ZVEZDIN

For the last few years considerable attention was devoted to intermetallic com-
pounds RE-TM, where RE -- rare earth, and TM = transition metal (such as, for
instance, RCos, R2Fe17, R2Col7, R2Fe14B, and REe2 etc.; Buschow 1980, 1988,
Radwanski et al. 1989a, de Boer et al. 1989). Of various interesting phenomena,
inherent in these materials, field induced phase transitions from a collinear into the
canted phase are of prime importance because the critical fields of these transitions
allow us to estimate directly the value of an exchange interaction between rare-earth
and transition metal sublattices (Franse et al. 1990b). Until now polycrystal and
powder samples have been the main objects for measurements. With progress in
both single crystal growth and strong magnetic field production, crystal field effects
and the role of magnetic anisotropy will be obviously of particular interest. Berezin
et al. (1980) experimentally explored monocrystals of the type DyCos+a. Some
general features of phase diagrams of these intermetallic ferrimagnets with hexag-
onal symmetry and of rare-earth-ferrite garnets were found. This is considered in
sections 12 and 13.
Section 14 is devoted to the field-induced phase transitions for which a unique
situation occurs when the compensation To and the spontaneous spin-reorientation
Ts~ temperatures practically coincide (Zawadzki et al. 1993, 1994). Investigation
of this interesting subject gives an opportunity to understand clearly the interrelation
between the magnetic anisotropy and the f-d exchange.
Theoretical activity in the latter was devoted to the elucidation of the role of the
surface and the role of point-like, linear, and other defects of a crystal lattice in
processes that determine phase transitions. The study of a spin-flop transition in
dislocation-containing single crystals may be quite important from this viewpoint.
Similar study was made by Vlasko-Vlasov et al. (1983), Vlasko-Vlasov and Inden-
born (1984). These problems are considered in sections 15 and 16.
Sections 17 and 18 are devoted to field-induced phase transitions in free-powder
samples and itinerant metamagnets. Microscopic calculations of the thermodynamic
potential are given in the Appendix.
In the final section we emphasize the applicability of the material presented in
this survey. Namely, we operate herein with weak anisotropy ferrimagnets for which
the Ndel model "works" sufficiently well. Innovative trends concerning field in-
duced phase transitions investigated for strong anisotropy ferrimagnets and itinerant
ferrimagnets with an unstable d-sublattice are briefly discussed.
Two-sublattice, more precisely, two-subnet amorphous systems with an antiferro-
magnetic interaction between the subnets and ferrimagnet superlattices with antifer-
romagnetic interaction between layers have been also investigated. They are still
under investigation.
Though each of these systems possesses its own characteristics, a conceptual con-
nection with the FIPT is definitely present in all of these systems.
Referring to the question about the term 'field induced phase transition' let us note
that a broader class of phase transitions is described by this name:
• Ferrimagnetic - canted - ferromagnetic transitions which are under consideration
here. Occasionally these are referred to as spin-flop transitions by analogy with
antiferromagnets;
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 411

• Metamagnetic transitions in itinerant ferrimagnets of the YCo2 or RCo2 type


(Wohlfarth and Rhodes 1962, see also experimental work by Goto et al. 1989);
• Crossover level transitions which may be due to a field-induced crossing of
ground state levels of f- or d-ions. It should be stressed that this crossing is
accompanied by the transformation of the magnetic structure of a crystal, i.e. a
magnetic analog of the Jahn-Teller effect takes place (Zvezdin et al. 1976);
• First order magnetic processes (FOMP) which can be due to a competition
among crystal-field contributions of different order to the free energy (see, e.g.,
Asti and Bolzoni 1980, Asti 1990).
We would like to emphasize that there are no distinct boundaries between these
field-induced phase transitions. Moreover, it is often possible to find some features,
inherent in one of the type of transitions, in the others as well. For example, the
mechanism of spin-flop transitions considered in this survey is closely connected
with crossover type transitions. Such interrelation of different mechanisms of field-
induced phase transitions is of great significance for a strongly anisotropic ferrimag-
nets. So in (HoY)3FesO15 iron garnets the transformation of ferrimagnetic structure
into ferromagnetic structure does not proceed smoothly via the canted phase (as it
does, e.g., in Gd3Fe5O12) but follows the pattern of jumps of the magnetization,
originating from the crossing of the levels of Ho +3 ions (Demidov and Levitin 1977,
Zvezdin et al. 1977). A similar example is (TbY)3Fe5012 (Lagutin et al. 1990a, b).

2. H - T phase diagram of isotropie ferrimagnets

Let us begin by considering the basic regularities of non-collinear magnetic structures


in a ferrimagnet. For this purpose we can use the mean field approximation. The
equations, describing the state of the two sublattices of isotropic ferrimagnet in this
approximation can be represented in the form

]~1 = /-11 M~, (2.1)


Ht

M2 = __/~2 M2, (2.2)


//2

where ~q and ~r2 are the sublattice magnetizations and Mx and M2 are their absolute
values. The relations

/t~ = / ~ - A-/~2, (2.3)


& = / t - .X2v~rI (2.4)
412 A.K. ZVEZDIN

describe the effective fields acting on the 1 and 2 sublattices and A is the exchange
interaction parameter. Equations (2.1) and (2.2) denote that the sublattices magneti-
zations are oriented along effective field directions under thermodynamic equilibrium.
Generally speaking in eqs (2.1)_. and (2.2) one needs to take into account the intra
sublattice molecular fields AllM1 a n d )~22]~2 . However, it is clear that taking them
into account leads only to renormalization of the parameters Xi. Furthermore, in the
case most important for us, i.e. the case of 3d-4f compounds when All >> A >> A22,
the validity of the approximation under consideration here is practically valid.
In the free ion approximation the functions MI(T, H1) and Mz(T,/-/2) are reduced
to the well-known Brillouin functions. Let us consider the change of the vectors
with the increasing field H at a transition from theferrimagnetic ordered state
(M1 is antiparallel to M2) to the ferromagnetic state (M1 is parallel to 2~r2). First
of all we define the canted (angular) solutions. The equations for the magnetization
components perpendicular to magnetic field are of the form:

(M1)± + / ~ x I ( M 2 ) . L = 0,

) ~ x 2 ( M 1 ) ± q- ( M 2 ) ± = O,

where
M~
i = 1,2.
X i - Hi'
These equations have the trivial solution ( M 0 . = 0 corresponding to the collinear
phase non-trivial solution, (Mi)± ¢ 0, is realized at the condition

1 -- )~2/~lX 2 = O. (2.5)

The equations for the components parallel to t h e / t are given by

(M1)II + AxI(M2)II = x 1 H , (2.6)


~x2(M1)II + (M2)II = X2H. (2.7)

Relations following from these equations are

(1 - AZx1x2)(M1)II = XI(1 - AXz)H, (2.8)


(1 -- A2X1X2)(M2)It = X2(1 - / ~ x 1 ) H . (2.9)

Hence it follows that


1 1
X1 = - (2.10)
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 413

M2

Fig. 2.1. The orientations of the sublattice magnetizations kT/1, K/2 relative to the external magnetic
field/t.

in a canted phase, i.e. H1 = AM1 and H2 = AM2.


Let 01 and 02 be the angles defining the deviations o f ml and 2hr2 from/~. It is
obvious t h a t / 1 and the vectors/~ri are in the same plane (see fig. 2.1). These angles
can be derived from eqs (2.3) and (2.4). Here

1/2
H1 = ( H 2 q- ,~2M22 - 2HAM2 cos 02) , (2.11)

H2 = (H 2 + )~2M12 - 2HAM1 cos 01) 1/2, (2.12)

and MI(T) and M2(T) are known functions. By solving these equations we find

(-A2(M22 - M12) + H 2)
cos 01 = , (2.13)
2HAM1
( - A 2 ( M 1 2 - M2 2) + H 2)
cos 02 : (2.14)
2H.XM2

The overall magnetization in a canted (angular) phase can be defined from eqs
(2.1) and (2.2) directly; summarizing them and with regard to X1 = )~2 = 1//~ we
obtain

+& : !17.
A
414 A.K. ZVEZDIN

0~,02

HI //2 /-/

Fig. 2.2. Field dependence of the sublattice rotation angles for isotropic ferrimagnet.

Equations (2.13), (2.14) describe a continuous rotation of the magnetization from


the ferrimagnetic to the ferromagnetic phase (fig. 2.2). The critical fields restricting
the canted phase can be found from the eqs (2.13), (2.14) under assumption 0 = 7r
and 0 = 0.

Hcl = )~IM1 - M2 l, (2.15)

He2 = A(MI + M2). (2.16)

Thus three phases exist (see fig. 2.3):


1. Ferrimagnetic phase: [0a - 02[ = zr,
2. Ferromagnetic phase: 01 = 02 = 0,
3. Canted phase: 0i are defined by equations (2.13), (2.14).
The boundaries between the phases determined by the expressions (2.15), (2.16)
are the 2nd order phase transition lines.
Let us discuss these phase transitions from a different point of view. We con-
sider RE-TM ferrimagnets (for example rare-earth-ferrite gamets R 3 F e 5 0 1 2 , o r
RnTm intermetallics, where T -- transition metal) in which the magnetization of the
d-sublattice Md does not depend on the magnetic field, and the magnetization of the
f-sublattice Mf is determined by the action of the effective field:

qeff = -- a rd.

The absolute value of Heff decreases to 0 when H goes to H = AM (see the line OP
at fig. 2.3) and increases when H > AMd. The value H = AMd is the crossover point.
Here we have the field-induced crossing of levels of the f-ions, i.e. the ground state
of the f-ions is degenerate at the line OP. It leads to the instability of the collinear
magnetic structure when the temperature decreases, i.e. to the phase transition into
the angled (canted) phase at T < T* (fig. 2.4).
It is clear that the free energy of the system decreases due to magnetic splitting of
the energy levels and increases due to deflection of the ~rd from magnetic field /~.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 415

H H

MR'~.,,,4Mr ~ 0
0
\
\
\
\
o
rk "* T T
a) b)
Fig. 2.3. Phase H - T diagrams of isotropic ferrimagnets: a) with a compensation point and b) without
a compensation point. On the line O P the magnetization of one of the sublattices ('paramagnet'
sublattice) is equal to zero.

i\,\ x I
I ./" "
!/ "\i

H H
lit

XMd He, He2


a) b)
Fig. 2.4. Dependence of the ground state levels on the magnetic field (schematically): a) T > T*,
b) T < T * .

This field induced magnetic transition that is related to a crossing of energy levels
can be considered as a "magnetic Jahn-Teller" phase transition (Zvezdin et al. 1976,
1977), because there is a direct and profound analogy of this effect with the well-
known Jahn-Teller effect and related phase transitions. We emphasize that crossing
of the energy levels is accompanied by a transformation of the magnetic structure
(in the case that a canting of the spins is initiated) in full accordance with concept
of the cooperative Jahn-Teller effect.
At T < T* the collinear structure is unstable in the whole region adjacent to the
H = AMd line in the H - T diagram. This means that the instability occurs already
before the degeneracy of the ground state is lifted.
416 A.K. ZVEZDIN

The theoretical H - T diagrams of isotropic ferrimagnets shown in fig. 2.3 are in


good agreement with experimental diagrams for the polycrystal samples of rare-
earth-ferrite garnets. These compounds can be discussed in terms of two magnetic
sublattices: 'Fe' and 'rare earth'. A more detailed description of the structure and
magnetic properties of the rare-earth-ferrite garnets is given in section 5. The mag-
netic moments of the Fe sublattice are coupled together by a strong exchange in-
teraction (Hexch ~ 106 Oe). The RE-Fe interaction is weaker (Hexch ~ 105 Oe),
and the much more weak f - f interaction plays an essential role only at the lowest
temperatures. Proceeding from this, it can be assumed that magnetization of the Fe
sublattice is independent of the external magnetic field (at H << 106 Oe) and the
state of the rare-earth sublattice. The intersublattice exchange interaction between
the rare-earth moments may also be neglected. In this approximation the rare-earth
sublattice represents an 'ideal paramagnet', placed in external field and exchange
field.
Calculations show that at low temperatures the critical fields of transitions into
noncollinear phases are of the order of 105-106 Oe. The possibility to induce the
noncollinear magnetic structures by the field is comparatively high, close to compen-
sation point. Figure 2.5 illustrates experimental magnetic phase diagrams of some
rare-earth-ferrite garnets obtained from the measurements of various physical proper-
ties (Levitin and Popov 1975). It also shows theoretical dependencies for isothermal
and adiabatic regimes of the measurements. Satisfactory accordance of experimental
and theoretical dependencies of the critical fields is observed.
A lot of experimental work was done to study field induced phase transitions in
the rare-earth-transition-metal compounds (for instance, on RFe2, RT3, R2T7, RCos,
R2Co7B3, R6Fe23, R2T17, R2Fe14B, R2FelaC, R2Fel7C, RT12, RCol2B6, R2(FeMn)laC).
Most of the experiments were performed on fine (single crystal) powder particles,
being free to orient themselves in the applied magnetic field. Since they are free
to rotate inside the sample holder, the particles orient their total magnetic moment
parallel to the external field. This approach enables in all cases to circumvent the
difficulties connected with magnetic anisotropy and to use the simple isotropic model
for experimental data treatment (for details see the papers by Radwanski et al. 1989,
Franse et al. 1990, de Boer et al. 1990, Liu et al. 1994).
It is of interest to study the complete magnetization curve up to full saturation.
The required magnetic fields, however, are typically too large to make such a study
feasible. Matsuuva et al. (1979) have measured the magnetization of a crystal of
Mn(CH3COO)2.4H20 in an external field up to 400 kOe at 1.1 K. This system is
ferrimagnetic below He1, ferromagnetic above Hc2 and a spin-canted state appears at
He1 < H < Hc2, where He1 = 125 kOe and Hc2 = 288 kOe for/~ parallel to c-axis.
This is a clear example of a full magnetization process in a Heisenberg ferrimagnet.
Another clear example of a full magnetization process was reported by Gurtovoy
et al. (1980) in (GdY)3FesO12 iron garnet (for details see section 8).
Full magnetization curves were also measured by Demidov and Levitin (1977) in
(HoY)3Fe5012 garnets in fields up to 300 kOe and by Lagutin and Dmitriev (1990),
Lagutin and Druzhinina (1990) in (TbY)3Fe5012 garnets in magnetic fields up to
2 MOe. In contrast with a continuous magnetization process, they observed a series
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 417

150 -o~ , 150 o,

100 100 dCl

50 50
Gd3FesO12 Tb3FesO12

I I i I 0 I I I I
260 280 300 T K 220 240 260 T, K

200 ",,, 200 "'°",,,,,,o I I

150 o~ '~,, 150


0 ~

,,
100 m', 100 V ~,~ i

50 50 '.
Ho3FesO12

0 0 t t i i
190 210 230 T, K 100 120 140 T, K
Fig. 2.5. Magnetic phase diagrams of polycrystalline samples of rare-earth-ferrite garnets: light circles
correspond to longitudinal magnetostriction measurements, black circles - to transversal magnetostriction
measurements, light squares - to magnetization measurements, dark squares - to Faraday effect mea-
surements. Theoretical dependencies for regimes of isothermal (solid lines) and adiabatic (dotted lines)
measurements are presented; after Levitin and Popov (1975).

of j u m p s in the magnetization curves during the transition from the ferrimagnet to the
ferromagnet state. However, (HoY)3Fe50]2 and (TbY)3Fe50]2 are strong anisotropic
ferrimagnetics and demand a special consideration.
In the studies of Guillot and Le Gall (1976), Miura et al. (1977), Pavlovskii et
al. (1979, 1992), Druzhinin et al. (1981, 1992) the Faraday effect was used. The
critical fields of the spin-flop transitions from the collinear ferrimagnetic phase to the
canted phase in yttrium-iron, y t t r i u m - i r o n - g a l l i u m (Y3Fe5012 and Y3(FeGa)5012)
418 A.K. ZVEZDIN

IO

6
\

~4 \
\
\
\
2 \
\
0
I I I I I \
0.2 0.4 0.6 0.8 1.0
r = TK/600
Fig. 2.6. H - T phase diagram of Y3Fe5012 iron garnet. The diagram has been calculated by the use
of the exchange parameters fitted to Faraday rotation measurements in explosion generated magnetic
fields up to 10 MG; after Druzhinin et al. (1981).

and in other garnets are in the megagauss field range. In Y3FesO12, according to
Druzhinin et al. (1981), the critical field at 300 K is 2.6 MOe (fig. 2.6). These
authors have also observed a strong scattering of light in this field region. They
connect the origin of this effect with the appearance of a domain structure in the
canted phase (see section 11).
Great attention was paid to intermetallic compounds of rare earth and d-metals in
which the d-subsystems represent itinerant metamagnets and in which the f-subsystem
could be described in the framework of the localized electron model. Typical fer-
rimagnets of this family are the materials RCo2, (YR)Co2 and (YR)(CoA1)2 (for
details see, e.g., Levitin and Marcosyan 1988), where magnetic field induced transi-
tions in the electron d-subsystem from the paramagnetic to ferromagnetic state were
observed. Such metamagnetic phase transitions were predicted by Wohlfarth and
Rhodes (1962).
To explore magnetization processes and magnetic field induced phase transitions
in such ferrimagnets, it is necessary to abandon the 'rigid' d-sublattice approximation
used above and to include into the theory the possibility not only of changes of the
sublattice magnetization direction but also variations of its magnitude. The phase
transitions and phase diagrams for such system were investigated theoretically in the
framework of the isotropic approximation (Zvezdin 1993, Zvezdin and Evangelista
1995).
The new feature of the H - T diagrams is the occurrence of tricritical points at
which three phases can coexist: paramagnetic, ferrimagnetic and canted. In contrast
with the well-studied tricritical points in DyA1G and FeF2, three real phases can
coexist in these systems near tricritical points. Thus there is an analogy between
these tricritical points and those found in 3He-4He mixtures.
FIELD INDUCEDPHASETRANSITIONSIN FERRIMAGNETS 419

Of interest is also the possibility of a jump-like transition from the paramagnetic


into the canted phase in the isotropic ferrimagnet system. Another uncommon feature
of these diagrams is the occurrence of a reentrant-like phase transition (Zvezdin and
Utochkin 1992).

3. Non-collinear structures in a weakly anisotropic ferrimagnets.


Basic equations

3.1. Thermodynamical potential of the non-equilibrium state


The model outlined in the previous section reflects the principal features of non-
collinear magnetic structures in ferrimagnets and the character of their phase dia-
grams. The main weakness of this model is the neglect of magnetic anisotropy. The
anisotropy effects are of principal importance in the vicinity of the compensation
temperature Tc where the critical fields for the isotropic model become equal to zero
(fig. 2.3). Furthermore we shall see below that the presence of anisotropy has a
strong influence on the character of sublattices rotation in the non-collinear phase
and on the type of phase transitions.
Anisotropy effects on the phase diagram of ferrimagnets can be represented qual-
itatively (for uniaxial anisotropy) in the following way. It is evident that at Hllg (g
is the easy axis direction) anisotropy hinders the rotational process. Therefore, close
to Tc the rotation of the sublattices (meaning that the canted phase appears) may be
expected to start at the threshold field

Hc -= v/H-AHexch,
where HA is an anisotropy field, Hexch - the intersublattice exchange field. As
a result, the first order phase transition takes place in the vicinity of Tc and is
accompanied by different other types of anomalies here.
At H_l_r~ as well as in antiferromagnets the canted phase exists starting from the
zero field values. At T = Tc the ferrimagnet behaves as antiferromagnet, i.e. the
field increasing results in a monotone spin-flop trend of sublattices. Ferrimagnet
remains in the canted phase when H changes its value from zero to a value in the
ferromagnet phase.
Slightly below and above Tc the transition from the ferrimagnetic to the ferro-
magnetic phase could proceed in two steps when the magnetic field increases. In
low magnetic fields the spontaneous magnetic moment of the ferrimagnet compound
declines from the easy axis and can attain an intermediate position between H and 4.
In this case the magnetic moments of the sublattices become non-collinear, though
the angle between their directions has to be close to 7r. Such a low-field canted
phase exists up to the value of the external field that makes the overall magnetic
moment to be parallel to /~. The ferrimagnet remains in the collinear phase upon
further increase of the field up to the values at which the ferrimagnetic configu-
ration becomes energetically unstable due to the competition between the external
and exchange fields, and the system passes into the canted phase where further field
increase leads to a spin flop.
420 A.K. ZVEZDIN

Investigation of the H - T phase diagram for anisotropic ferrimagnets is a cum-


bersome and complicated problem requiring numerical computations in a number
of cases. In order to clarify qualitative peculiarities of the phenomena associated
with the anisotropy and noncollinear structures in rare-earth ferrimagnets we first
discuss a simplified model. When describing this model and its consequences we
shall follow the results already reported by Goranskii and Zvezdin (1969b), Zvezdin
and Matveev (1972a, b).
When studying anisotropic ferrimagnets it is more convenient to employ the ther-
modynamic approach instead of using the molecular field eqs (2.1) and (2.2) directly.
We take as basic model an f-d ferrimagnet, considered in the previous paragraph,
where the magnetization of one of the sublattices (let it be the d-sublattice) is sup-
posed to be saturated by the intrasublattice exchange field and where the second
(rare-earth) sublattice is an ideal paramagnet placed in the external field and the
exchange field created by the d-sublattice.
The thermodynamic potential for the system under consideration can be written in
the form (Goranskii and Zvezdin 1969b, Zvezdin and Matveev 1972a, b)

= -MdH , - f0 Hen Me(x) dz + K, (3.1)

where the first term is the Zeeman energy associated with the interaction of the d-
sublattice with magnetization Md and the external field/7. The second term is the
thermodynamic potential of the paramagnetic rare-earth ions in the effective field
Heff which is equal to

/7eff = / 7 - AMd. (3.2)

A particular dependence of Mf on Heff does not play an essential role here. In


some cases we shall approximate it by a corresponding Brillouin function.
The last term in (3.1) is the anisotropy energy. It should be stressed that mag-
netic anisotropy in the considered model is taken to be low in comparison with the
exchange energies

IKI << ~M~.


Only when this condition holds can the energy of magnetic anisotropy of rare earth
ions be represented as an additional term in the thermodynamic potential.
In a general case the anisotropy energy depends on the magnetization direction
in both sublattices. We restrict ourselves to the approximation where this energy K
depends only on the magnetization direction of one sublattice, K = K()~rd).
In connection with this assumption we note the following. In the paper of Zvezdin
and Matveev (1972) it was shown that the general form of the anisotropy energy, if it
depends on the magnetization directions in both sublattices due to the paramagnetic
character of the f-sublattice and the smallness of the anisotropy field compared to
the exchange field, can be reduced to the above mentioned form on condition that
H << AMd.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 421

But this weak field region is of importance mainly when considering anisotropy
effects (K << AM~). On the contrary, in strong fields (H ~ AMd) the anisotropy
effects in the weakly anisotropic ferrimagnet are small.
Although each of the thermodynamic potential terms in (3.1) has an obvious phys-
ical origin in the framework of the discussed model, they can be also justified mi-
croscopically (see Appendix section and the paper by Zvezdin and Popkov (1980)).
Formula (3.1) rep~sents the thermodynamic potential of the non-equilibrium state,
depending only on Md. It is assumed here that the paramagnet sublattice is in equi-
librium with the effective field. In such form the thermodynamic potential can be em-
ployed for the analysis of equilibrium properties and low-frequency non-equilibrium
processes (whose characteristic frequencies are much smaller than the frequencies of
precession of paramagnetic ions in the effective field). For studying high frequency
phenomena the thermodynamic potential has to be taken in a more general form
which is dependent on the both sublattices magnetization (see, for example, Zvezdin
and Popkov 1974, Popkov 1976b).

3.2. Magnetic anisotropy


The explicit form of the function K(2~ra) is defined by symmetry.
Let i,j,k be unit vectors of the Cartesian reference frame such that klIH, and let
0, ~ be the polar and azimuthal angles determined in the usual way.

3.2.1. Uniaxial anisotropy


F o r / t l l g (g is an easy axis direction) and/~[1~ we have

K(O, ~) = - K c o s 2 0, K > 0. (3.3)


At H_l_g in coordinate system (g, [h, g], h), where h = H/H), K(O, (y) is

K(O, ~) = - K COS2 ~ sin 2 0, K > 0. (3.4)

So the general formula for uniaxial magnetic anisotropy may be written as

K(O, ~) = -KF(O, ~),

where the explicit form of the function f(O, ~) follows from the eqs (3.3) and (3.4).
It is apparent that there is no pure uniaxial magnetic anisotropy in crystals. It
needs in any case to take into account the dependence of the anisotropy energy on
the azimuthal angle.
For a crystal with tetragonal and hexagonal structure the following expressions
for K(O, ~) are used:

K(O, ~) = K1 sin 2 0 + K2 sin 4 0 + K3 sin 6 0 + / ( 4 sin 4 0 sin &p,

for the tetragonal structure, and

K(O, ~) + K1 sin 2 0 + K2 sin 4 0 +/t" 3 sin 6 0 +/~24 sin 6 0 cos 6%


422 A.K. ZVEZDIN

for the hexagonal structure, where 0 is the polar angle between the tetragonal or
hexagonal axis and the magnetization of sublattice; ~ is the azimuthal angle between
the component of the magnetization in the basal plane and the [100] axis.
The magnetic anisotropy of rhombohedral crystals can be represented in the form
K = -KI C0S 2 0 -- K2 cos 4 0 + / ( 3 cos 0 sin 2 0 cos 3g),

where 0, ~o are the polar and azimuthal angles, if the z-axis is parallel with the c-axis
of the crystal.
The important case of hexagonal anisotropy is considered in detail in section 12.

3.2.2. Cubic anisotropy


The anisotropy energy in this case is defined by the expression
2 2 2 2 2 2
= Kl(Cq a 2 + % % K20el ct2o~3.
We will restrict ourselves to only the first invariant of this formula, i.e. we put
/(2 = 0 so that

K(O, ~) = -Klf(O, ~). (3.5)


Let us define the function f(O, ~) as:

f(O, qo) = - ( 1 / 4 ) ( s i n 2 20 + sin 4 0 sin 2 2qo), (3.6)

if/7111001] in the coordinate system (i,~, k) = ([100], [010], [001]);

f(O, ~o) = cos 0 sin 3 0 sin 3qo - (1/3) cos 4 0 - (1/4) sin 4 0, (3.7)
3

if/711 [111] in a coordinate system (i, 7, k~) = ([1 i0], [115], [111]);
f(O, ~) = (1/4)(sin 2 0 cos 2 g) - cos 2 0) + sin 2 0 sin 2 qo(1 - sin 2 0 sin 2 ~), (3.8)

ifI7111110] in a coordinate frame (i, j, k~) = ([110], [001], [110]).


For K < 0 the directions of the easy axes are along directions of the type [111]
and for K > 0 along [100]. In rare-earth-ferrite garnets the constant K is commonly
negative.

3.3. Extreme conditions


The extreme conditions of the thermodynamic potential relative to 0 and ~p have the
form

d~ O#
- - = sinOMdH(1 - A~(O,~a) + - - = O, (3.9)
dO O0
d~ OK
- - - - O, (3.10)
d~ 8~
FIELD INDUCED PHASE TRANSITIONSIN FERRIMAGNETS 423

where ~ is

Mf(Heff)
~,(0, ~) - (3.11)
Heff

The analysis of these equations and the determination of the conditions for their
stable solutions in particular cases will be discussed in sections 4, 12, 14, and 18.

3.4. Low magnetic field approximation


The most important area of anisotropy effects pertains to H << AMd. The thermo-
dynamic potential (3.1) and eqs (3.9), (3.10) can be represented in a more simple
form. Then up to second order terms in H/~Md we get

q5 = - ~ I H M d - K l f ( O , qg) + Q cos 2 0, (3.12)

where

77 = 1 - AXo, Q = -A~'MH 2,

~,=( d _
Xo = ~ ( H = 0), <0.

For T --+ Tc

1
X0-+ ~ ,

1/A - Xo
X t --+
),Ma
so that in the vicinity of Tc we have,

M°_-__ u o , (3.13)

H2 1
(3.14)

where Mf° is the magnetization of rare-earth sublattice at H = 0 and Xo is its


susceptibility.
In such approximations the problem of the determination of the ferrimagnet equa-
tion of state is reduced to the well-known equivalent problem of a uniform fen'o-
magnet reversal of magnetization Mt = 7/Md. It should be stressed that the magnetic
field induces here an additional effective anisotropy of the 'easy plane' type

aKeff = Q cos 2 0.
424 A.K. ZVEZDIN

The energy expressed in eq. (3.14) has a simple physical interpretation. As well
as in antiferromagnets, here 1/), is the transverse susceptibility of a system and X0
is the longitudinal susceptibility, i.e.

X± "~ l/A, XI[ ~ X0.

Since usually XII << X± the energy Q is a prerequisite (with enhancing H) for spin
reversal in a state with a maximum susceptibility value, i.e. in the canted phase.
The second and third terms in (3.12) are typical for spin-flop transitions in antifer-
romagnets. The competition between these terms defines the threshold field of such
transitions. Specific properties due to ferrimagnetism are reflected in the first term
of eq. (3.12).
In this approximation, x(O) can be replaced by

~(0) = Xo + x'HcosO, (3.15)

where X0 and X' are defined below eq. (3.12).


From the above reasoning it is convenient to express formula (3.12) for the ther-
modynamic potential in the following form

()C± -- Xll)(/~- - (/~l*)~)2


¢, = - - M°(gri) - K~I(O, ~) + const, (3.16)
2

w h e r e / ' = Md/Md. There is a clear analogy of this formula and similar expression
for the free energy of antiferromagnets (see, e.g., Belov et al. 1979) that can have a
definite heuristic significance.
Notice that at T = 0 the thermodynamic potential (3.1) takes the form

= - - m d ~r -- M O ( H 2 -Jr-)`2M2 - 2HAMo cos 0) 1/2 + K(O, ~),

where Mf° is the magnetization of the f-sublattice at T = O.

4. H - T p h a s e d i a g r a m s

4.1. Uniaxial magnetic anisotropy


4.1.1. Case 1: It& E A (EA is the easy axis of magnetization)
In that case the phase diagram has a most simple form. Substituting formula (see
subsection 3.2.1 and eq. (3.3))

df
-- = 2sin0 cos0, (4.1)
dO
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 425

H
A

"~
- ~3 :rr/4 II~

lm

ro r
Fig. 4.1. Phase diagram of a uniaxial ferrimagnet in a field perpendicular to the easy axis. The dashed
curves are isoclines 0 = O(H,T). The solid curves pertain to the second order phase transitions; after
Goranskii and Zvezdin (1969b).

into eqs (3.9), (3.10) we get two solutions 0 = 0 and 0 = 7r o f these equations cor-
responding to collinear phases. Boundaries o f the stability region o f these collinear
phases (in fig. 4.1 it is AA' and BB' lines) are determined by the equations

2K
1-A~(0) ---0, (4.2)
MdH
2K
1 - A~(Tr) + - 0. (4.3)
MdH

The canted phase is defined by

1 - A~(0) - 2- K cos 0 cos 2 F = 0, (4.4)


Man
2 K sin 2 0 cos ~ sin ~ = 0. (4.5)

It is easy to verify that the values T = 0 and T = 7r form the m i n i m u m conditions


o f a t h e r m o d y n a m i c potential so that the values of a corresponding thermodynamic
potential are equal. Such 'degeneracy' of solutions is a characteristic property o f the
angular structures, It can physically be realized by dividing a crystal into domains
when transitions to a canted phase occur (see section 11).
426 A.K. ZVEZDIN

The curves A A t and B B ~ lines in fig. 4.1 correspond to the phase transitions of
the second order. It is not difficult to check this by expanding F into a series of A0
in the vicinity of a the corresponding collinear phase. For example, close to AA' the
expansion of F in A0 has the form
¢' = ¢'o + a(H, T) 02 +/3(H, T) 04 + . . - , (4.6)
where a(H, T) = 0 and/3(H, T) > 0 on A X ( it is clear here that A0 = 0).
The expansion of F close to the curve B B p has an analogous form with the only
difference that the expansion parameter (order parameter) in that case is A0 --- ~- - 0.
In a small field region where H << AMd, equations for the curves AA' and B B t can
be represented in the following form (in first order of H/AMd):
T - Tc 2K
T ~ - ,~fx'IH + Md----H' (4.7)

where the T signs correspond to the curves A W and B B ~, respectively. Here we


have used the relation
1 - ,~Xo(T) ,,~ (T - Tc)/Tc, (4.8)
which is valid if IT - Tel << Tc; X0 and X' are defined by formulas (3.13).
A distinctive feature of phase diagram (fig. 4.1) in that case is the presence of
'narrow throat' joining the low field and high field areas of the canted phase. The
'throat' coordinates (Tth, H*) are

Tth,~Tc, H* = ~/ 2K
MA[X'I" (4.9)

Its semi-width is
AT 2~ / 2K/~IX'] (4.10)
rc V Md
In the low-field region the phase transitions are such that the continuous spin-
reorientation is accompanied by two phase transitions of the second order. In this
region the angle between the sublattices is a little bit different from ~. In the high
field region the transition FI ++ C ++ FE takes place with an essential sublattices
'inflection', i.e. the angle between the sublattices substantially differs from ~.
If the paramagnet sublattice is close to saturation,
Mf
Ix'l ~ - - (4.11)
a2M~
Substituting this value into (4.9), (4.10) and taking into account that in the 'throat'
region &~" ~ 1 yield

H* ~ (HAHex) ~/2, A T ~ 2(HAHex)I/2, (4.12)


Tc
where Hex = AMd. If the paramagnetic sublattice is far from saturation the magnitude
of H* strongly increases and AT/Tc decreases.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 427

4.1.2. Case 2." /~ll


EA (EA is the easy axis of magnetization)
The boundaries o f the collinear phases {0 = 0} and {0 = 7r} are defined in that case
by the equations

2K
1 - A~(0)+ -0, (4.13)
MdH

2K
1 - A~(~) - - - 0. (4.14)
MdH

Curves A X and B B ~ in the phase diagram (fig. 4.2) correspond to these bound-
aries.
The system possesses axial symmetry. Therefore the azimuthal angle ~ at 0 = 0
is arbitrary here.
The canted phase is defined by equation

2K
1 - A~(0) + - - cos 0 = 0. (4.15)
MdH

The main peculiarities in the corresponding phase diagram are the following. There
is area in phase diagram that is restricted by the curve X Q R . . . B ~ where both

H
A

\P/ 05"
H* ~t MF~ , MR ~ MF~

AL.j- R

7 T
Fig. 4.2. Phase diagram of an uniaxial ferrimagnet in a field parallel to the easy axis. BP, AR are
curves of the 2nd kind of phase transitions, AIR, PB ~ are curves representing the loss of the collinear
phase stability. PQ is a curve representing the loss of metastable canted phase stability, PRTc is
a curve describing the first kind of phase transition and P is the tricritical point; after Zvezdin and
Matveev (1972a).
428 A.K. ZVEZDIN

collinear phases are stable. RT is the curve of the first order phase transition where

~ ( 0 = 0) = ~ ( 0 = ~).

As in the case of /7_l_g, where g is parallel with the easy axis, we write the
thermodynamic potential as a series expansion in/tO approaching the AA' and BB'
lines, the points where the coefficient at 04 in the formula (4.6) (i.e. fl(H, T)) go
to zero (point Q and P) appear on these curves; /3 > 0 above those points, i.e.
transition from collinear phases to the canted phase become the second order phase
transition.
According to modem terminology these points are referred to as tricritical points.
Familiar examples of similar critical objects are tricritical points in the anisotropic
antiferromagnets Dy3A15Oa2 and FeF2 (see, e.g., Landau et al. 1971, Blume et
al. 1974, Giordano and Wolf 1977). In the vicinity of the tricritical point P the
thermodynamic potential expansion should be considered up to the sixth order term.
The coefficient of this latter term can be shown to be positive.
In the approximation quadratic of H/.~Md (see eq. 3.14 of section 3) the points
P and Q coincide. In this case

/3(H,T) = 2 K H .2 - 1 (4.16)

in a vicinity on the curves A X and BB'. H* in (4.16) is defined by formula (4.9).


The line corresponding to the first order phase transition between the collinear
phase {0 = ~r} and the canted phase (line PR in fig. 4.2) can be derived by
eliminating 0 from the system of equations

~/i(00) = ~/i(Tr), (4.17)

= 0. (4.18)
-~ 0=00

The canted phase loses its stability on curve PQ and the system transforms {0 = or}
in a jump-like way. The jump of the angle 0 increases from A0 = 0 at point P to
A0 = 7r at point Q. The field- and temperature-dependence of the PQ curve can be
found from the system of equations

0=0o

820 0=0o

This curve belongs to the family of curves O(H,T) = const.


FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 429

The intersection points of the curves A A ' and B B ~ associated with the loss of
collinear phase stability has the coordinates

H*=IHa Tth ~ Tc. (4.21)


~qX'l '

The equation of the curves A A ~ and B B ' can be represented in the low-field limit
by the formula

T - Tc 2K
F q T~ - ~Ix'IH- Md-----H'

where the ~: signs correspond to the curves A A ~ and B B ~, respectively.


The magnitudes of the magnetic field defining the points P and Q slightly differ
from H*. In particular, if we take

Hex = AMa = 2.5 x 105 Oe, K = 7 x 103 erg/cm 3, Tc = 300 K.

(parameters of gadolinium-ferrite garnet) then

He - H* : H* - HQ ~ 400 Oe.

It should be borne in mind that in the vicinity of the point H* the effective
anisotropy energy (3.12) changes its sign. Some additional details of the phase
diagrams in the vicinity of the 'narrow throat' were studied by Baryakhtar et al.
(1976).
To gain greater insight into effects of the magnetic anisotropy on the critical fields,
let us consider the case T = 0. It follows immediately from the eqs (4.2) and (4.3)
that

2K Mf
Hcl = )~(Mf - Md) T
M f - Md m~

2K mf
He2 ~-/~(mf -}- md) 4-
mf + md md'
where the signs - / + corresponds to the case H]]EA (H_I_EA).
Notice that in the center of the region of the canted phase (more precisely at
the line O(H, T) = 7r/2) there is no influence of the magnetic anisotropy on the
magnetization curve. Here, the susceptibility of ferrimagnet should be equal to 1/A
similar to the isotropic case. This follows immediately from the eq. (4.4) (see also
section 8).
430 A.K. ZVEZDIN

4.2. C u b i c a n i s o t r o p y

4.2.1. C a s e 1 : / t ] ] [ 0 0 1 ] , K1 < 0
The phase diagram for this case is shown in fig. 4.3. Three phases are present: two
collinear (A and B) and one canted (C) phases. Lines A A ~ and B B ~ are second-
order transition lines. The equations for the curves A A ~ and B B ~ are (4.2), (4.3) at
K = - K 1 . The canted phase C is determined by the equations

K1
1 - A;~(0) + cos 0 (2 cos 20 + sin 2 0 sin 2 2~) = 0, (4.22)
mdH
K 1 sin 4 0 sin 2~ cos 2~ = 0. (4.23)

The second of these equations and the stability conditions simultaneously define
the equilibrium values of the angle ~:

= ~/4, 3~/4, 5~/4, 7~/4,

i.e. the rotation of sublattices takes place in the (110) plane (~ = 7r/4, 5~r/4) or in
the (150) plane (~ = 37r/4, 77r/4).
A fourfold degeneracy of the tilted phase follows from the symmetry of the system:
the [001] axis is fourth order axis.
The canted phase splits into two different canted phases: (C, C t) in the low
field region. It means that in this range the solution of eq. (4.22), i.e. the function

H
' A •

,1o == ,i =,o)ot ,

M To T
Fig. 4.3. Phase diagram of a cubic fe~imagnet for HH[001], KI < 0. A A ~, B B I are curves of the
2nd l~nd phase transitions from the collinear phases A and B into the canted phase C. OTc is a curve
of the first-order phase transition between the canted phases. O is the critical point; after Zvezdin and
Matveev (1972a).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 431

O(H, T), is not unique and the rotation of the sublattices takes place discontinuously.
For phase C one finds 0 < 0 < ~r/2 and for phase C' 7r/2 < 0 < 7r.
The area restricted by the curves O M and O M ~ is the area of coexistence of the
phases C and C . These curves are defined by the system of the following equations
K1
1 - A~(0) - - - cos 0 (3 cos 2 0 - 1) = 0, (4.24)
MaN

- MdH)~x'(O) + K1 sin0(9cos2 0 - 1) = 0. (4.25)

In the approximation H / A M d (See (3.12)) the equations of the curves O M and


O M ' becomes
T-To 2 IK1] ( 2H2) (4.26)
~-c --T~Md--~ 1 - H,---~ •
The coordinates of point O in the plane H - T are
HO = H* T ° ~ Tc,

where H* is the coordinate of the 'narrow throat' in the phase diagram (see formula
(4.9)).
The curve TcO is the locus of the first order transitions between the phases C and
C'. There is a jump in angle from zero in point O up to A0 = 2 arccos(2/3) for
H-+0.
The critical point O is the limiting point of the curve on which there is an equi-
librium of two low field phases. For high magnetic fields (H > H ° ) the crystal is
homogeneous and for H < H ° it is divided into domains with different orientations
of the magnetic sublattices. From this viewpoint the critical point O is analogous to
the critical point of vapor-liquid phase transitions. Analogous to the density in that
case is the rotation angle. There are many peculiarities in the behavior in the system
near the critical point: anomalies of specific heat, propagation in the sound and scat-
tering of light. (For details see sections 7-10.) Alben (1970a, b) was the first who
paid attention to the possibility that this critical point existed in ytterbium-ferrite
garnet.

4.2.2. Case 2: / I l l [ I l l ] , K1 < 0


The phase diagram for this case is shown in fig. 4.4. Four phases exist: two collinear
phases, A and/3 and two canted phases C and D. The canted phases are three-fold
degenerated (it is clear since axis [111] is of the third order).
Curves A A ' and/3/3' refer to the stability loss of the collinear phases A and B.
They are defined by the equations
4t(1
1 - ),;~(0) - - - 0, (4.27)
3Mall

4K1
1 - A~(Tr) + - - - 0. (4.28)
3Mall
432 A.K. ZVEZDIN

H H
1

2 ©

B
D

MR**Mve I
v

ro r r

a) b)
Fig. 4.4. Phase diagram of a cubic ferrimagnet for HIl[lll], /41 < 0. a). Curves of the first order
phase transitions: Tc(1), To(2), To(3). b). Curves describing the loss of stability: phase A-AA', phase
C-CC' and F F ' , phase D-GG' and DD ~, phase B-BB'; after Zvezdin and Matveev (1972a).

The canted phases are determined by

K1 ( 4 cos2 0 - sin 2 0 cos 0 -


1 - A:~(0) - Md----H

- - - x/2 sin3 0sin3~ + v/2cos2 0 sin 0 sin 3cp)


- z 0, (4.29)
3
x/2K1 cos 0 sin 3 0 cos 3~ = 0. (4.30)

Equations (4.30), combined with the stability conditions, define the equilibrium
values of the angle ~:

PhaseC: ~=vr/2, 7rr/6, llrr/6, 0 < 0 < rr/2;

P h a s e D : ~ = 31r/2, ~r/6, 57r/6, rr/2 < 0 < rr.


The crystallographic planes (110), (011), (101) are easily seen to correspond to
the angles

~o = rr/2, 3rr/2; ~ = 77r/6, rr/6; ~ = 1Dr/6, 5r~/6,


FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 433

respectively.
The equilibrium values of O(H, T) at a given value of the azimuthal angle ~ are
defined by eq. (4.29).
The boundaries of the stability for the canted phases C and D do not coincide
with the corresponding stability curves of the collinear phases. These boundaries
are determined by the curves C C and F U for phase C and by the curves D D ' and
GG ~ for phase D. In that case all phase transitions A +4 C ++ D ++ B are of the
first order. This follows from the fact that terms of the third order in A0 are present
in the thermodynamic potential (3.1), (3.7).
A stereographic projection of the trajectories of the d-sublattice moment 2~rd at
phase transitions A - C - D - B is shown in fig. 4.5.
Let us bring forward analytical formulas for the characteristic curves of the phase
diagrams (fig. 4.4), which are valid for H << AMd. In the vicinity of A A ' the series
expansion of the thermodynamic potential in 0 has a form

r -- T c 0 2 V/2 04
= Mall ~ 2 3 -IKll03 + IKII((H/H*)2 - 3) ~- + - . . ,

where H* is defined by the formula (4.9) at K = IK1 t.


Along the curve A A ' the magnitude TA(H) is defined by eq. (4.7) for K =
-(2/3) IKll.

[0011

[0101

Fig. 4.5. Trajectory of the vector 3~td on the unit sphere for the transitions A-C-D-B, /~11[111].
Section AC corresponds to phase C, DB to phase D, the points A, B to phases A and B. A jump
of the vector M takes place between the points G' and D.
434 A.K. ZVEZDIN

The equation of the curve of the 1st order phase transitions (3To) for H > v/-3H *
is

--TA + T 41K~I
TA 9 M d H ( ( H / H * ) 2 - 3)

The jump in angle 0 along this curve is equal to

AO=
3 ( ( H / H * ) 2 - 3)

These results are valid for the area close to B B ' as well (with the replacements
(--TA + T) --+ ( - T + TB), 0 -+ 7r- 0 ). For H < v~H*, the coefficient of 04 changes
its sign. Therefore, it is necessary to take it into account in the analysis of terms of
higher order in 0.
The curves F F ' and GG' are defined by equations

v~lK, I
1 - t~'(7r/2) -4- - - - O,
3Mall

or

T-Tc ( H )2 v~IKI[
Tc ~- ~ T 3Md-------H

The jumps in angle 0 are equal to

AO:--(H/H*) 2 atH>H*.
3

Let us consider the asymptotes of the first order phase transitions curves as at
H --+ 0 (fig. 4.4a):

T - Tc 4H
To(3):
T~ 3~Md'

Tc(2): T = Tc,

T - Tc 4H
r~(3):
T~ 3),Md "
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 435

4.2.3. Case 3: /-IN[ll0], K < 0


The phase diagram is shown in fig. 4.6. There are five phases: two collinear phases
(A a n d / 3 ) and three canted phases (G, D, E).
The stability curves o f the collinear phases (AA' and /3/3') are defined by the
equations

K1
1 - ;,~(o) + - o,
MdH

K1
1-A~'(zr) - - - 0 .
MdH
The canted phases C and D are described by equations

K1
1 - ),~(0) + - - cos 0 (3 COS 2 0 - - 2) = 0,
MdH
qv = ±¢r/2.

For phase C one has 0 < 0 < 0o and for phase D 7r - 0o < 0 < ~r, where 0o is
defined by equation

sin 2 0o = 3/5.

3 A

rl/

Te 9 T
Fig. 4.6. Phase diagram of a cubic ferrimagnet for /~[l[ll0], t£1 < 0. AA', BB', 2-0, 3-0' are
the curves of the first order phase transitions; 7-0-8, 6 - 0 ' - 9 are curves describing the boundaries of
the existence range of the metastable phase; O, O ~ are the tricritical points; after Zvezdin and Popkov
(1977).
436 A.K. ZVEZDIN

A A I and B B ~ are curves pertaining to the second order phase transitions A - C and
B - D . The stability of the canted phases C and D are described by the curves 29
and 36, which are defined by the equations

25/2K1
1 - ),~(00) - 0,
53/2MdH
25/2K1
1 - A~(Tr - 00) + -- 0.
5312MdH

The canted phase exists for

sin20 > 3/5.

It follows from eqs (3.8), (3.9), (3.10) that

sin 2 (p = sin -2 0 - 2/3.

After substituting this formula into eq. (3.9), we obtain an equation for angle 0 in
phase E:

4K1
1 - A~'(0) + - - cos 0 (1 - 4 COS 2 0 ) = 0.
3MdH

The stability condition of this solution is of a form

Heff Heff ~Heff - (4/3)K1(1 - 12cos20) > 0.

It is seen from the last inequality that for sufficiently high values of H when

H > H1 = V/5/2H *

the stability condition is satisfied so that transitions C - E and D - E are continuous


in this region (transition curves 2 - D ' and 3 - D ) H* is defined by formula (4.9).
For H < H1 the transitions C - D and D - E are of the first order (curves OTc,
O'Tc). Points

O(H = H1, T = Tc - 0, 43AT),

O'(H = H1, T = Tc + 0, 43AT)


are tricritical points, AT is determined by formula (4.10).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 437

~1 ira]

C
[112]

Fig. 4.7. Trajectoryof the vector /17/aon the unit sphere at the transitions A - G - E - D - B , ~11[110].
The points A and B correspond to A and B phases, the trajectorysections A G and AGt - to phase G",
GD and G~D ~ - to phase E, D B and D~B ~ to phase D (see fig. 4.6).

The curves 7 - 0 - 8 and 6 - 0 ' - 9 in fig. 4.6 define areas where metastable phases
exist. These curves are defined by equations of the type (4.19, 4.20) (for more details
see Zvezdin and Popkov 1977).
In fig. 4.7 a stereographic projection of the trajectories at the sequential phase
transitions A - C - E - D - E is given (see fig. 4.6)

4.2.4. C a s e 4: HI[[001],
K1 > 0
The character of the phase diagrams changes substantially when the anisotropy con-
stant changes its sign. This will be illustrated by means of an example where
/~ll[001]. In both cases (K1 > 0 and K1 < 0) the phase diagrams have a most
simple form for s u c h / 1 orientation.
The phase diagram for the case/(1 > 0 is represented in fig. 4.8. Three phases
exist: two collinear (A and B) and one canted (C) phase.
The boundaries of the collinear phase stability regions are determined from eqs
(4.2), (4.3) (reference frame is identical to the one used in subsection 4.2.1). They
define lines A A ' and B B ' in fig. 4.8. The canted phase is described by the equations
2Ka
1 - A~(0) -- cos 0 (1 - 2 COS 2 0 ) : 0, (4.32)
MdH
= 0, 7r/2, 7r, 37r/2.
The equilibrium values of the azimuthal angle define the rotational in the planes
(010) and (001). It is easy to verify that the second derivative of • with respect to
~o satisfies the condition ~ t! > 0 here.
438 A.K. ZVEZDIN

/-/ ~M~o

n,
M~ 0 4o

0 '
r~ r
Fig. 4.8. Phase diagram of a cubic ferrimagnet for ErH[001], K1 > 0; after Zvezdin and Popkov
(1977).

The condition for a stable canted phase is

()~MdH)2( Heft ) +2Kl(1-6cOs20)>O"


OHeffi~Mf

The canted phase is seen from this equation to be stable in all ranges of 0 (from
0 up to re) for sufficiently high values of H > H1, where H1 ~ x/-5H*.
For H > HI the transitions A - C and B - C are the second order phase transitions
(lines AO, BO'). For H < H1 the stability regions of the phases A, B and C
overlap. The curves O T and O'T represent the first order phase transitions from
A - C and from B - C .
Points O and O' are tricritical. Their coordinates are

O(H = HI, T = Tc - (1/2)AT),

O'(H = H1, T = Tc + (1/2)AT),


where AT is defined by formula (4.10).
The curves 1-O-A' and 2 - O ' - B ' form the boundaries of the stability region of
the canted phase which are determined by simultaneous solution of eq. (4.32) and
by ~"o0 = 0.
Overlapping between stability range of the canted phase and the metastable regions,
the collinear phases 0 = re and 0 = 0 take place for

H < H2 = x/T-/2H*.
FIELD INDUCED PHASETRANSITIONSIN FERRIMAGNETS 439

M~ e

S r~ r

Fig. 4.9. Phasediagramof a cubic ferrimagnetfor/~ll[ll 1], K > 0; afterZvezdinand Popkov(1977).

In fig. 4.9 the phase diagram of a cubic ferrrimagnet is presented for the case of
K1 > 0 and Htl [111]. All phase transitions described here as well as those described
in 4.2.2 are the first order phase transitions.
Popkov (1976a, b) investigated phase diagrams of rhombohedrical ferrimagnetic
crystals and films and discussed the problems connected with pressure effects.

5. Field induced phase transitions in rare-earth-ferrite garnets

There is a number of investigations devoted to the study of the field induced phase
transitions and phase diagrams for the magnetic crystals with cubic and uniaxial
magnetic anisotropy. In most of these, rare-earth-ferrite garnets and ferrite garnet
single crystal films were investigated.
Single crystals of ferrite garnets have natural cubic magnetic anisotropy (Ndel
1954, Pauthenet 1958a, b, Pearson 1962); the films have an induced uniaxial mag-
netic anisotropy (Le Craw et al. 1971, Callen 1971, Rosencwaig et al. 1971,
Akselrad and Callen 1971, Stacy and Rooymans 1971, Gyorgy et al. 1971, Kurtzig
and Hagedorn 1971).
Let us start by considering the single crystals of ferrite garnets. The next paragraph
will be devoted to ferrite garnet films.
Rare-earth-ferrite garnets with R3Fe5012 as a chemical formula (R is a rare earth
element) are isomorphic in structure to the natural mineral orthosylicate garnet
Gd3A12(SiO4)3 which has a cubic crystal structure (O~° - Ia3d space group, Geller
and Gilleo 1957)). Magnetic garnets have been synthesized in the fifties (Forestier
440 A.K. ZVEZDIN

and Guiot-Gullion 1950, 1952, Bertaut and Forrat 1956, Geller and Gilleo 1958a, b,
1960, Geller 1960).
The ferrite garnet unit cell consists of 8 formula units: 64 cations (40 Fe +3 ions
and 24 R +3 ions) and 96 anions (oxygen ions). The ions of Fe and rare earth
are arranged in the node-to-node sets of the oxygen matrix. When characterized
according to their coordinate number they occupy tetrahedral [d], and octahedral (a)
and dodecahedral {c} positions (Geller 1960). The R +3 ions are in {c} positions, 24
Fe +3 ions are in [d] positions and 16 Fe +3 ions are in (a) positions. Therefore the
ferrite garnet formula is often written in the form {R 3+3}(Fe 3+3)[Fez+3]O12.
The magnetic ions of the same type located in equivalent cell nodes have the same
moment direction and form a magnetic sublattice. Therefore the three sublattice
model is in common usage to describe the properties of these garnets (N6el 1954).
The overall exchange interactions in the sublattices and the exchange interactions
between the sublattices are of the antiferromagnetic type: the strongest one is the
a-d interaction (the effective exchange field H a d ~ 2 x 106 Oe) as a result of which
the Fe +3 ions in (a) and [d] positions are oriented in an antiparallel to each other.
The exchange interaction of the rare earth with the Fe +3 ions is one order of mag-
nitude lower than the Fe(a)-Fe[dl interaction and the main contribution to the rare
earth - iron exchange is the c-d interaction (Hc-d --~ (1-4) x 105 Oe). Therefore,
a non-compensated magnetic moment arises (..~ 5#B/per formula unit at 0 K). It
should be noted that the rare earth sublattice consists of six different sublattices,
which is due to the orientation of the surrounding crystal field. However, this is
important only for rare earth ions with a nonzero orbital moment in the ground state.
This point will be considered in the second part of this review. Generally speak-
ing, the (a) and [d] positions in the garnet structure should also be subdivided into
two types of different nonequivalent positions. But by virtue of the fact that the
ions Fe +3 are scarcely affected by the crystal field, this inequality plays a consid-
erably lower role than the inequality of the {c} positions, occupied by rare-earth
ions.
The magnetization vector of the rare earth sublattice is antiparallel to the magne-
tization vector of the Fe sublattices. The intra sublattice interactions in the a and d
sublattices are not large compared with the a-d interaction (Hc--d ~ (1-4) x 105 Oe).
Furthermore, the exchange interaction within the c sublattice is very small, so for
T > 10 K the rare earth ions can be considered as a system of paramagnetic ions
placed in a strong effective field produced by the Fe ions (Pauthenet 1958a, b, Ale-
onard 1960, Anderson 1964, Clark and Callen 1968).
The predominant a-d exchange interaction between the Fe sublattices is destroyed
at the Curie temperature Tc which is approximately equal to 560 K for all rare-
earth-ferrite garnet materials (Pauthenet 1958a, b).
The temperature dependence of the magnetic moments of the rare earth and Fe
sublattices is different which leads to the existence of a compensation temperature
Tc (Pauthenet 1958a, b) for many of the magnetic garnets.
At T < Tc the rare earth sublattice magnetization prevails and at T > Tc the net
magnetization of the Fe ions dominates.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 441

In this section we consider only Gd3FesO]2 iron garnet since this material most
closely corresponds to the above considered theoretical model of weak anisotropic
ferrimagnets.
Actually, the ground state of the Gd +3 ion is 8S, i.e. it has L = 0, S = 7/2.
Therefore, the spin-orbit interaction is only a small perturbation for this ion and the
anisotropy energy is one or two orders of magnitude less than the Gd-Fe exchange
energy. This is the very condition for the applicability of the weak anisotropy model.
Some other rare earth garnets will be considered in the second part of the review.
A great number of papers are devoted to investigation of induced noncollinear
magnetic structures in ferrite garnets by using the Faraday effect (Kharchenko et
al. 1968, 1974, 1975a, b, Bernasconi and Kuse 1971, Gnatchenko and Kharchenko
1976, Lisovskii et al. 1976a-c, 1975, Smirnova et al. 1970).
The simplest compound, also from optical aspect, is Gd-ferrite garnet. There-
fore, the major part of investigations by means of the Faraday effect on induced
noncollinear magnetic structures have been carried out on this ferrimagnet. The

% min
100 296 K

80

60

40

20
7

0 $ H" 285.3
84
-20

-40

-60

-80

-100
I i I I I

0 10 20 30 40 50
Fig. 5.1. Magnetic field dependencies of the Faraday rotation in Gd3FesO12 in the vicinity of the
compensation temperature; (X = 6328 ,~; the broken line shows the Faraday rotation of the optical
system; after Kharchenko et al. 1975a, b).
442 A.K. ZVEZDIN

~o,deg

1.0

0 i i
k?o
' i i i i r l

-l.0

1.0
~~ttttqttttt $ ~ " 7 T 1
$j ~ 1~6 kOe
0 I t l l l l / l l l l l [ l l l l lI l I I

-1.0

1.0

I I I I I I I I I I I I I I I I P

-1.0
-~~ _ . t . . . , . I . ~ ~

1.0

-1.0
i

280
i '
2
- - "

285
: i i i

290
i r i i i

295
T,K

Fig. 5.2. The temperature dependencies of the Faraday rotation in Gd3Fe5012 in the vicinity of the
compensation temperature; after Kharchenko et al. (1975a, b).

dipolar contribution of the Gd sublattice to the circular birefringence is negligibly


small in Gd ferrite garnet. If we consider the two Fe sublattices as one, the de-
pendence of the Faraday rotation angle ~b on the direction of sublattices magnetic
moment may be represented in the lonagitudinal geometry in the form (Kharchenko
et al. 1975a, b) (magnetic intensity H is collinear to the light propagation direc-
tion f:):

)9 = ~0 COS 0Fe -k FH.

Here )90 is a spontaneous Faraday rotation of the total Fe sublattice. The term
FI-I accounts for the effect of the magnetic field on the excited energy states of
the crystal and this enables us determine the angle 0Fe between the direction of the
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 443

ZII °e

0±0
t ttt'~tlt#lw'~ t r~m~ t~lSlk 5.17
260 285 290 T, K
Fig. 5.3. The ellipticity of the circular polarized light in the canted phase of Gd3FesOl2, B[I[lll];
after Kharchenko et al. (1975a, b).

H, kOe

20 I I
II
II
15 • •

lO ,'o

t I I l I
275 280 285 290 295 T, K

Fig. 5.4. Phase diagram of the Gd-ferrite garnet for /~11[100]. Thick solid lines are the calculated
boundaries of the existence regions of stable and metastable phases, (o) are experimental points of the
transition into uncollinear phase, (o) and thin lines indicate the observed existence region of the magnetic
inhomogeneities, the dashed line is extrapolation; after Kharchenko et al. (1975a, b).

m a g n e t i c m o m e n t o f the F e sublattice and the direction o f the field H b y using the


e x p e r i m e n t a l data o f F a r a d a y rotation.
T h e d e p e n d e n c e o f the F a r a d a y rotation at t e m p e r a t u r e close to the c o m p e n s a t i o n
t e m p e r a t u r e is shown in figs 5.1 and 5.2. The e x p e r i m e n t a l d e p e n d e n c i e s are in
444 A.K. ZVEZDIN

H, kOe
15

10

283 284 285 286 287 T, K

"I
a)
H, kOe

1 2

3'~" • • x,,~ 4,
• ~3
I I I I I 1
2 1 0 1 2 T-Tc, K
b)
Fig. 5.5. Phase diagram of Gd-ferrite garnet for HII[lll]. a) Experimental boundaries of the ex-
istence region of the magnetic phase (corresponding magnetic phases are indicated in the circles).
b) Theoretically calculated stability curves (points - experiment); after Kharchenko et al. (1975a, b).

a c c o r d a n c e with theoretical curves c a l c u l a t e d for a three-sublatticed model.


I n f o r m a t i o n a b o u t the sublattices rotation can b e d e r i v e d b y e x p l o r i n g the birefrin-
g e n c e effects ( K h a r c h e n k o et al. 1975a, b, Pisarev et al. 1969, 1971, G r j e g o r j e v s k i y
and Pisarev 1973). F o r the case f~ll~ it was shown b y K h a r c h e n k o et al. (1975a, b)
that the linear b i r e f r i n g e n c e in G d ferrite garnet is equal to

A n = A n • sin 2 0Fe,
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 445

ATmin

H*V~15 " O=Tr /I iI ~ 0=0

Mad 4 /d i

Mre o/ A' ~ \x o Cd
/ //MGd- \\N

0 N I tel I Fe "N' I
270 290 TK 310
T,K
Fig. 5.6. H - T phase diagram of Gd3FesOl2 iron garnet for/211[100]. The thin solid curves A A ~ and
B B ' are theoretical curves of the 2nd order phase transitions, calculated according formulas (4.2), (4.3).
The thick solid curves A X and B B ' are experimental curves, obtained from the measurements (o) - the
specific heat, (×) - the sound absorption at a frequency 30 MHz with the wave vector ffl/2; OTe is a
curve representing first order phase transitions, O N and O N ' are the boundaries of the existence region
of the metastable phase, O is the critical point; after Kamilov et al. (1975).

H, kOe

200 +

100 + +

0 I I
50 100 150 T, K
Fig. 5.7. Magnetic phase diagrams of a Ho3Fe3O12 iron garnet single crystal (o) obtained for/211[111],
(e) obtained for/2111110], (+) - polycrystalline data; after Hug (1972).
446 A.K. ZVEZDIN

M,,/zB/molecule M,, p~/molecule

2 I H, kOe
8 1.5"- I
0 20 40~"'-

o
~2
o
Y t I
100
I
/ 0
I
22

~
6

0
1
J
I
100
t I
200

I
f 4
/
I
2

I I I I I I I I
100 200 0 100 200
f 6 ' I
~ -

i ~ 4 J

I 1 P I I I I t
0 100 200 0 100 200
H, kOe H, kOe
a) b)
Fig. 5.8. Experimental (a) and theoretical (b) plots of the magnetization of Hoo.41Y2.59FesO12 gar-
net against the field: solid curves - for /~11[111], dashed curves - for nll[ll0], dash-dot curves
- for ~11[100]. The insets show MII(H) in weak fields; after Silant'ev et al. (1980).

where An0 is the spontaneous linear birefringence in the direction perpendicular to


the optical crystal axis, and 0Fe is the angle of the Fe-sublattice moment orientation
relative to the field direction. Thus, the development of a noncollinear structure and
concomitant change of the Fe sublattice direction results in the change of birefrin-
gence as well. Figure 5.3 shows a typical example of the temperature dependence
of the birefringence in the vicinity of the compensation point.
The magnetic phase diagram of Gd-ferrite garnet (Kharchenko et al. 1974, 1975a,
b, Gnatchenko and Kharchenko 1976, Kamilov and Schachschaev 1972, Kamilov et
al. 1975) has been studied in detail on single crystals.
These investigations show that the theory presented above can qualitatively ex-
plain the main features of the Gd ferrite garnet phase diagram in the vicinity of
the compensation temperature. This includes the character of temperature depen-
H E L D INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 447

Mjr, 7~--~ ,
molecule

10

f
/ -

I I I I I I I I I
0 100 200 0 100 200 H, k O e
a) b)
Fig. 5.9. Experimental a) and theoretical b) plots of the magnetization of HOl.05Y1.95FesOI2 iron
garnet against the field: solid curves - for/t11[111], dashed curves - for n l l [ l l 0 ] , dash-dot curves -
for Hll[100]; after Silant'ev et al. (1980).

dencies of the critical fields, the number and the order of different magnetic phase
degenerations and the type of the phase transitions. However, a quantitative agree-
ment between the experimental and theoretical data in the two-sublatticed model
is not reached. This is due to the fact that even small changes of magnetiza-
tion of the total Fe sublattice in a field, owing to the finite value of the antifer-
romagnet exchange interaction between Fe sublattices, will result in a substantial
change of the critical field value when the temperature of compensation is ap-
proached.
Kharchenko et al. (1975a, b) calculated several phase diagrams of cubic ferrimag-
nets with three sublattices. Qualitatively these are similar to those considered above,
the only difference being renormalization of characteristic points and temperature
dependencies of critical fields of the phase diagrams.
Phase diagrams constructed on the basis of theoretical formulas reported by Khar-
chenko et al. (1975a, b) are satisfactorily consistent with the experimental data
for Gd-ferrite garnet (figs 5.5, 5.4, see also figs 4.3 and 4.4) though complete
agreement is not reached. Apparently this is due to the influence of the mag-
netoelastic energy on phase diagram (Kharchenko et al. 1975a, b, Gnatchenko
and Kharchenko 1976), dividing of a sample into domains (see section 11 be-
low).
Kamilov et al. (1975) investigated the specific heat and sound propagation in
G d 3 F e s O 1 2 n e a r the compensation temperature. Figure 5.6 shows the H - T phase
diagram of this garnet for HI1[001] according to the experimental results of these
authors (for details see sections 9 and 10).
448 A.K. ZVEZDIN

H, kOe IV H, ~Oe
200 - 200

150 A ~ Q
150

II
lOO 100

I
50

~r

I I I i~ F I 1 4 1 I I I
0 10 20 30 T,K 0 10 20 30 40 ~K
a) b)
H, kOe
200 I]I

150

100

50

I I I Jr I I
0 20 40 T, K
c)
Fig. 5.10. H - T - x phase diagrams for (HoY)IG when x = 0.67: a) HII[lll]; b) /~l[[ll0]. The solid
lines: theory. The open circles were obtained with increase of the field during the measurement process,
the dark circles with decrease of the field; after Babushkin et al. (1983).

Some features of the phase diagrams of a Ho ferrite garnet single crystal were
obtained for a field orientation along the [111] and [110] axes (fig. 5.7) (Hug 1972).
It is seen that these diagrams are in good qualitative agreement with the theoretical
phase diagrams presented in section 4.
Detailed investigations of the FIPT in (HoY)3FesO]2 and (TbY)3Fe5012 iron gar-
nets were made by Levitin and Demidov (1977), Zvezdin et al. (1977), Silant'ev
et al. (1980), Babushkin et al. (1983), Lagutin and Dmitriev (1990), Lagutin and
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 449

H, kOe ~ 9 ~ ? ~ H, kOe

200 ~ 200

150 150
L
100 ~ ~ ~ ~ ~, 100

/
i i r 1t"NI/ i ~t I K/t~ r i i i
0 20 To40 60 30 50 60 70
a) T,K b) LK

H, kOe
200

100 ~Q ' ~ Tcr~M'


5o ~I

I r$ B II t ,1. I I
20 30 Te 50 60 70 T, K
e)
Fig. 5.11. H - T - z phase diagrams for (HoY)IG when z = 1.05: a) HII[lll]; b) /qll[ll0]. The solid
lines: theory. The open circles were obtained with increase of the field during the measurementprocess,
the dark circles with decrease of the field; after Babushkin et al. (1983).

Druzhinina (1990) and Lagutin (1993). These ferrimagnets have a very strong (Ising-
like) magnetic anisotropy. It seems possible to find some common features between
the phase diagrams of the strongly and weakly anisotropic ferrimagnets. However,
strong anisotropy drastically changes phase transitions and phase diagrams. A dis-
tinctive feature of the strong anisotropic case is that the transitions follow the pattern
of jumps of the magnetization, i.e. the first order transitions. These drastic distinc-
tions can be seen in figs 5.8-5.11 in which the magnetization curves and H - T phase
diagrams of the (HoY)3Fe5012 garnets are displayed.
450 A.K. ZVEZDIN

There is an elaborate investigation of the phase diagrams of the Yb-ferrite garnet


(Alben 1970a, b, Feron et al. 1971, 1972, 1974). However this garnet is a strongly
anisotropic ferrimagnet and it also demands special considerations.

6. Single crystal ferrite garnet films

There are several investigations of field induced phase transitions in single crystal
ferrite garnet films with uniaxial magnetic anisotropy (Avaeva et al. 1975, Antonov
et al. 1976a, b, Gnatchenko et al. 1977, Dikstein et al. 1980, 1983, 1984, Lisovskii
et al. 1976a-c, 1980). Such films are prepared by means of liquid phase epitaxy on
Gd-Ga garnet substrates (see, e.g., Eshenfelder 1981).
The structural properties of these films are well characterized in terms of the
described above model. These films possess high optical transparency which makes
it easy to perform optical investigations. However, their complex composition, and
their intrinsic non-uniform strain are the cause of unstable compensation temperatures
over the film thickness which creates additional peculiarities in the phase diagrams.
An important manifestation of such inhomogeneities in these films is the existence
of so-called compensational domain walls near the compensation point (Hansen and
Krumme 1973, Krumme and Hansen 1973). Such walls complicate the picture of
the phase transitions.
It should be noted that investigations of field induced phase transitions particularly
are the most effective method for determining the compensation temperature profile
along the film thickness and other parameters.
As an example we show in fig. 6.1 the phase diagram of a Y2.6Gd0.4Fe3.9Gal.lO12
film grown from the liquid phase on to a Gd-Ga garnet substrate which had been cut
parallel to the (111) plane (Gnatchenko et al. 1977). The film thickness and compen-
sation temperature were of the order of 6 #m and 180 K, respectively. The critical
temperatures (or fields) represented in the phase diagram have been determined by
recording the temperature at which various magnetic phases arise and vanish in the
field.
The temperature dependencies of the Faraday effect presented in fig. 6.2 have been
employed as well. From well defined linear rotation angles studied as a function of
the temperature near the compensation point Tc the appearance of the canted phase
was determined. The observations show that the sample in weak fields (lower than
the threshold field Htn ~ 0.7-0.85 kOe in the case) can exist only in two states with
spins to be collinear with magnetic field.
Further processing of these results indicated the presence of a considerable gradient
of the compensation temperature over the thickness of the film, the averaged value
of which is equal to 2 × 103 K/mm.
The temperatures T( and T~ in figs 6.1 and 6.2 are defined as the temperatures
at which corresponding collinear phases (low and high temperature phases) become
unstable at least in one of the film layers.
Since investigated films possess a compensation temperature gradient along the
thickness direction, transitions from the collinear phases into the canted phase show
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 451

H, kOe

10

170 180 190 T, K


Fig. 6.1. The H - T phase diagram of an epitaxial film of Y2.6Gdo.4Fe3.9Gal.lO12in the vicinity of the
compensation temperature (H < 15 KOe). The easy axis of magnetization is parallel to the normal ff of
this film and ~rlla. The symbols (o) correspond to appearance and disappearance of the low-temperature
coUinear phase during heating of the sample. The symbols (o) - to appearance and disappearance of
the high-temperature collinear phase during cooling. The data (e) and (o) were determined visually,
the data (z~) were determined by Faraday effect measurements, the rectangles correspond to threshold
fields; the solid lines correspond to theory; after Gnatchenko et al. (1977).

corresponding inhomogeneous behavior. At T[ and T~ the whole sample switched


into the canted phase.
Different phases distributions in the film thickness correspond to different temper-
atures and magnetic field ranges. For H > H* the film has adopted the canted phase
with the orientations of the sublattice magnetizations varying along the thickness
direction.
In the temperature ranges from T~ to T~' and from T~ to T~' when going from one
range to another, the collinear and canted states arise alternatively. For He < H < H*
the magnetic structure of the film represents a mixture of collinear and canted phases.
For H < He only the two collinear phases (low and high temperature) exist and the
transition between them is of the first order.
Figure 6.3 shows phase diagram of a Y2.3Gd0.4Fe3.9Gal.lO12 film in a wide field
range up to 60 kOe. The critical fields here are obtained by means of extrapolations
of the observable temperature dependencies of the manetooptical rotation angle.
Examples of such extrapolations are depicted on curves 2 and 3 of fig. 6.2. The
452 A.K. ZVEZDIN

1.0 ~o,deg

0.5

0
0/ i i i/ ~ i
1160170,//190200
-0.5
t _1.0 ~ _ _ . ~ 2 1
-1.0
I I I I I I I
160 170 180 190 200 210 220 T, K
Fig. 6.2. The temperature dependencies of the Faraday rotation of an epitaxial film of
Y2.6Gdo.4Fe3.9GaI.IO12in different magnetic fields: 1 - H = 1 kOe, 2 - 10 kOe, 3 - 45 kOe; after
Gnatchenko et al. (1977)

H, kOe
60 \ \ 12 / /1

,0 " ,\\ ///


/
40 ,\ //
',\ //
20 c~ ~ a \ , / d

0 f
-30 -20 -10 0 10 20 (r-ro),K
Fig. 6.3. The 'reconstructed' high-field phase diagram of the epitaxial film Y2.6Gdo.4Fe3.9Gal.1012
with Tc = 183 K; broken lines correspond to different theoretical models; after Gnatchenko et al.
(1977).

critical fields are defined by the bend points o f the extrapolation curves and they
correspond to the following film parameters averaged over the thickness: Tc = 183 K
and Hc = 2.5 kOe. Let's note that Gnatchenko et al. (1977) e m p l o y e d three
sublattices model o f Nrel ferrimagnet for processing the experimental data.
Lisovskii (1980) and Dikstein et al. (1980) have studied F I P T in single crystal
films with the aim to elucidate the effects o f its layered nature. Experiments were
produced in the high magnetic stationary fields up to 150 kOe. A number o f the films
FIELD INDUCEDPHASETRANSITIONSIN FERRIMAGNETS 453

with different thickness and crystallographic orientations having a different number


of layers (from 5 up to 20 ) were studied.
Figure 6.4 shows the H - T phase diagram of the film (YGdYbBi)3(FeA1)5OI2
made up from 19 layers at ~lln where ~ is the easy axes (see also fig. 4.2). The
compensation temperature changes in the layers 3-18 from 201 K up to 207 K (the
layers 1, 2 and 19 don't have compensation temperature).
External boundaries of the canted phase are shown in fig. 6.4 by the solid lines
(H > 36 kOe). Broken lines show schematically the same boundaries for the layers
3 and 18.
The pattern of vertical lines (compare with the curve P T c in fig. 4.2) at this
diagram is caused by the distribution of the compensation temperature in the layers.
This pattern was observed by means of domain structure transformations.
The threshold field, where three phases - two collinear and one canted - coexist,
is spread out here into the shaded area in the diagram (in the region of H* ~ 34.8-
36 kOe and T ~ 201-207 K).
In the studied films of (YGdYbBi)3(FeA1)5012 one has Ha ~ 2 K u / M s "~ 300 Oe,
HE '-~ 106 Oe and H* ~ 30 kOe which agrees sufficiently well with the experimental
data.
Figure 6.5 shows H - T phase diagram of a film of (YGdYbBi)3(FeA1)5012 for
/1_LEA (easy axis) which should be compared with the theoretical diagram shown
on fig. 4.1 (see for details section 11). A peculiar feature of the film geometry
here is that the second order transitions from collinear state into the canted phase

H, kOe I I

80
I

II\
i
//// / / /
/
!

60 ~,,. \\ /// /

~
20 4 "a~" 0~"i~ , ~ * . "

0 I I I I ~ ! I I I I I

192 196 200 204 208 212 216 T, K


Fig. 6.4. H - T phase diagram of a ferrimagnetic film of (Y,Gd,Yb,Bi)3(Fe,AI)5012 made up from 19
layers in a magnetic field Erll,~, where a is the normal to the film. The easy axis of magnetization is
parallel with ~ in this film. High field solid lines are the external boundaries of the canted phase (at
H > 36 kOe). Broken lines are the same for the layers 3 and 18. The pattern of vertical lines is caused
by the distribution of the compensationtemperaturein the layers (comparewith fig. 4.2); after Lisovskii
(1980).
454 A.K. ZVEZDIN

H, kOe D, n l m -1

150 300

100 ~ _

50 -100 ~_

0
200 250 300 350 400 T, K

Fig. 6.5. The H - T phase diagram of an epitaxial iron garnet film of (Y,Gd,Yb,Bi)3(Fe,A1)5012at H_I_g,
where ~ is the normal to the film being parallel with the easy axis of magnetization of this film. The
lines He(T) (A) are the boundaries of the canted phase (see fig. 4.1). Temperaturedependencies of the
inverse period of the domain structure Do I at H = 0 are displayed together with the inverse critical
period of the domain structure Dc I (o) and the critical field Hc (A) of the film; after Dikstein et al.
(1980).

follow behavior of the 'soft mode' (Dikstein 1991). The soft mode in this case is
the spin-density wave transformed into an ordinary domain structure away from the
transition point.
There are interesting investigations of H - T phase diagrams of amorphous D y - C o
films in the vicinity of the compensation point in stationary magnetic fields up to
150 kOe (Fisch et al. 1986, Khrustalev et al. 1989, 1993a, b). The topology of
these phase diagrams qualitatively well agrees with the theoretical ones, considered
in section 4.1. However, there are some features inherent in the amorphous nature
of these films (e.g., the possible existence of an asperomagnetic state).

7. Some general features of field induced phase transitions

The phase transitions considered above are typical transitions with a magnetic sym-
metry change. For instance in the first case of section 4.2 (H_I_EA (easy axis)) during
the I-III transition the symmetry is relative to a rotation around the z-axis with an
angle 7r, i.e. it is the symmetry element C~ (in values). This symmetry element
is absent in phase III but the double number of equilibrium states in comparison
with phase I is present here: ~ = 0 and ~ -- rr, which are transferred from one
to the other by the 'broken' symmetry element. Identical values of the free energy
('degeneration') correspond to these two states.
It is accompanied with a division of the sample into domains of a low symmetry
phase. In the case Hllg (section 4.1.2) distortion of the continuous symmetry C ~
takes place during the transition into the canted phase. Such distinction of a 'break' of
FIELD INDUCEDPHASETRANSITIONSIN FERRIMAGNETS 455

the continuous or discrete symmetry elements are especially important for dynamical
properties of the system.
A change of sign of coefficient a(H, T) in the Landau expansion of the free energy
(see, e.g., section 4.1) causes peculiarities in the behavior of many physical properties
near the transition point:
• the susceptibility goes to c~. In this case the susceptibility describes the response
of an order parameter to the thermodynamically conjugated field (see section
8).
• the occurrence of anomalies thermodynamical magnitudes quantities such as
kinks, jumps and )~-curves. The heat capacity, specific heat, magnetocaloric
effect (see section 9), Young modulus and sound velocities, magnetostriction
(section 10), and magnetooptical phenomena.
• the conversion of the order parameter oscillation frequency to zero (soft mode)
and the hindering of its relaxation.
• the increasing order parameter fluctuations and their correlation radius
• the expansion of domain walls and the rearrangement of domain structure in the
sample
Many theoretical studies devoted to the field induced phase transitions have been
carried out using the mean field theory (or using equivalent approximations as the
Landau theory). These theories when constructing the free energy of a system
neglect, to some extent, the fluctuations of the order parameter. In the region of
the transition temperature, i.e. in the region where the system stability is lost, the
fluctuations increase strongly and these theories become inapplicable.
A characteristic feature of the studied phase transitions is that the Landau theory
can be used for their description with practically no limitations. The region of
inapplicability becomes extremely narrow AT ~ 10-6-10 -8 K. This is a consequence
of the fact that the fluctuations that occur in the region of the transition have a very
large value of the correlation radius.
We will consider this problem using the example of the orientation transition
investigated above. To study critical fluctuations we use thermodynamical potential
(3.1), taking additionally into account the energy of magnetic nonuniformities. For
small values of 0 we have
F = f dv{~O 2 q-/304 + A(grad0)2), (7.1)

where the constant o~(T) can be represented in the form ~(T) = at, t = (T - Tc)/Tc.
Suppose the fluctuation in the angle 0 is 30(~, and its Fourier transform is 30~. It
is well known that the mean square fluctuations are given by
T , t>0;
1~0~12 = V(at + Aq 2)
T , t<O.
Y(2ant] + Aq 2)
456 A.K. ZVEZDIN

An inverse Fourier transformation of this equation gives the correlation function of


the fluctuation of the order parameter

9(r3 = 60(0) 0(r) - T e_r/p,


47tAr

where

( A / a t ) /z t>0;
p = (A/Zat)U2, t<0,

is the correlation radius, which tends to o0 at the phase transition point.


The unlimited increase in the uniform fluctuations when approaching the phase-
transition point indicates that the Landau theory is not applicable in the immediate
vicinity of the transition point. The Levanyuk-Ginsburg criterion, which defines the
region in which the Landau theory is applicable, has the form

3)2 TcZ/ 2
t>>¢= A3a
(the temperature is measured in energy units).
An estimate of the value of ( using this formula for typical values of the parameters
p / a ~ 0.1, Tc ~ 100 K, A ~ 4 x 10 -7 erg/cm, and a ,~ 104 erg/cm3 gives ( ~ 10 -8.
For comparison we give the value of the corresponding quantity for a transition at
the Curie point (c g 10-1-10-2.
We will give once more the characteristic values of the correlation radii for field
induced phase transitions Pzi and transitions at the Curie point Pc (with the same
parameters)

3 × 103 (3-5)
PFI ~' - - A, PC ~ - A.

The 'orientation' fluctuations are more long-wave than the fluctuations near the Curie
point. This explains, in particular, their small contribution to the free energy (the
statistical weight of the long-wave fluctuations is small, while the short-wave fluc-
tuations are strongly suppressed).

8. Magnetization and susceptibility

8.1. Differential susceptibility


The magnetization of ferrimagnets in collinear and noncollinear phases is defined
in a usual way as

Mt - -- - ~(O)H + md (1 -- A~(0)) cos 0, (8.1)


OH
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 457

where ~ is determined by formula (3.1) and O(H,T) is taken from eqs (3.9), (3.10).
In the collinear phases 0 = 0, 7r we have

Mt = Md - Mf sign(AMa - H), 0 = 0,

Mt = Mf - Md, O = Tr.

In order to obtain the value of Mt in a canted phase let's express function x(O) in
terms of A, K, f(O) from the section 3.2 and substitute x(O) into (8.1). Then we
obtain

Mf-Md ( f° "] for0:Tr,


H K
Mr= ~+(H-AMdcos0)~ \sin0J for0<0<~r,

Md - Mr sign(H - AMd) for 0 = 0.

This formula is simple for the case when the transition into the canted phase takes
place in a strong magnetic field H > AMd and the influence of anisotropy can be
neglected.
Then
{ mf- md f o r H < Hcl,
M~= A for Hcl ( H < Hc2 ,

mf+md for H > Hc2,

where Hcl = )~(mf - md) , He2 = /~(mf q- md) (see section 2). The terms of the
order K/HMd << 1 are neglected here.
The longitudinal differential susceptibility is equal to

dmf for 0 = 0, 7r,


dH
(8.2)
for0<0<r.
; 1 + MOOd'-H H sinO ,/J

In the case of the strong critical fields considered above we have Xd = 1/A in
the canted phase. The susceptibility Xd displays a jump at the transition from the
collinear into the canted phase

1
AXd- A Xf.

Thus, the isotherms of the magnetization Mr(H) show to a kink and the longitu-
dinal susceptibility a jump at transitions into noncollinear phases. This fact can be
458 A.K. ZVEZDIN

M I I I I

21OK//

f iH1DY3Fe5012
I I I I
0 50 100 150 200 H, kOe

M I I I I

100

Ho3Fe5012
I I I I
0 50 100 150 200 I4 kOe

6j
M I I I [

7 ~ IH1 _

/// Er3FesO12

0 50 100 150 200 /4 kOe


Fig. 8.1. Field dependenceof the magnetizationof Dy3Fe5012iron garnets; after Levitin and Popov
(1975).

employed for the experimental determination of the critical fields. Particularly, by


such methods the critical fields in ferrite garnets of Yb, Gd, Ho, Dy, Er (Clark and
Callen 1968, Feron et al. 1971, 1972, 1974, Levitin and Popov 1975, Fillion 1974)
and in some hexaferrites (Sannikov and Perekalina 1969) have been found.
In fig. 8.1 the fields dependence of the magnetization in ferrite garnets of Dy, Ho
and Er is shown. The linear parts Mt(H) describe the magnetization in the collinear
phase and the curved parts describe the magnetization in the noncollinear phase.
Bend points (or the kinks) in the magnetization curves determine the values of the
transition fields from collinear into the noncollinear phases.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 459

A large body of research was devoted to the high field magnetization of ferrimag-
netic f - d intermetallic compounds. Many results were obtained by using oriented-
p o w d e r samples because single-crystalline materials are not always available (see,
for instance, Sinnema et al. 1984, Buschow et al. 1985, 1989, de Boer et al. 1987,
1988a, b, 1990, Radwanski et al. 1989a-c, Verhoef et al, 1989, 1990a-c, Z h o n g
et al. 1990a-c, Date 1990, Franse 1990, Liu et al. 1991a-c, Zhou et al. 1992a,
b). Figure 8.2 shows, e.g., the typical curves for several 2 - 1 7 c o m p o u n d s clearly
displaying the transitions from the collinear into the canted phase (Verhoef 1990).
The effect of anisotropy on the Mt and especially on the Xd behavior can be
essential in the low field part of the phase diagrams and can give rise to additional
anomalies. The region of canted phase narrowing (the 'narrow throat' of phase
diagram, see, e.g., fig. 4.1) is concerned to be the most 'dangerous' on that count. In
this area the angle rapidly changes with variation of the field and temperature which
results in anomalies in Xd.
In fig. 8.3 temperature dependencies of Xa calculated by formula (8.2) for and ,qll~
/~_l_g when H << AMd are depicted. The following analytical formulas correspond
to them. F o r m u l a (8.2) for Xd when 2K/Md << H << ),Md can be rewritten in the
form
1 2K .
___
H2- H 2
(3+ H2 '/
cos01 (8.3)

~ 1 H 2 + H .2 3- H2 cos 20 for/~_l_g,

where H* is defined by formula (4.9). For cos 0 we used eqs (4.4) and (4.7) for
/~IIEA and H_LEA, respectively.

100 ~ I ~ I D t
........ ~'7"
.,.;:'~,,''" ....

80 ~
Q . g l . e . t 3 . ~.t~.fi]..Ei.E]..[]......~......l~.~ } .......... ~ . , , _ t n ................ e ............ [].................,..~.....

"~ 60 • ,.o" ..............""' '~.........,'"" '..........


_:::::::::::::::::::::::::::::::::::::::::::::::::::::::::::..............~,............. .........~..........
40- .,~.............. ....."~'..........
a ~x..a..z~.~..,--~-~..a.-~..-...~.....-~..~...........,s.......... ,~............. aH°2C° 17
_~.e.,o.~,~,~,.~,.e,,~.~,.--.-.c,.-...-e,,e-.........~, ....-~>'0..............
- vHo2Col4Fe 3
20 - °Er3C°17
c~Er2Fel7
ODy2COl7
0 I i I I I i I
0 10 20 30 40
B [T]
Fig. 8.2. High-field magnetization curves at 4.2 K of several R2M17 single-crystalline spheres that are
free to orient themselves in the applied magnetic field; after Verhoef (1990).
460 A.K. ZVEZDIN

Xd

x IIEA

± \ IEA
,
0 Te T

Fig. 8.3. Temperaturedependence of the differentialmagnetic susceptibilitycalculatedaccordingto the


formulas (8.2), (8.3); after Bisliev et al. (1973).

Notice that the canted phase exists only when H > H* and HI]g. Therefore it is
reasonable to consider the case H > H* in the expression for Xdll"
The solution of these equations for AM >> H > H* can be represented in the form

T- Tc H m d H .2
for/1H EA,
Tc 2 K H 2 _ H .2
cos 0 = T - Tc H M d H .2 (8.4)
- - . for/~_I_EA.
Tc 2 K H 2 + H .2

It should be noted that the anisotropy part of to Xd changes its sign at H = H * / v ~ .


Formulas (8.3), (8.4), and fig. 8.3 form a representation of the effect of the
anisotropy on the longitudinal differential susceptibility behavior Xd. We do not
present here analogous (but more cumbersome) formulas for Xd for a case of cubic
anisotropy. Let us note only that in this case anomalies are possible not solely on the
boundaries separating canted and collinear phases but inside canted phases as well.
For example, Xd -+ oo in the vicinity of the critical point 0. In fig. 4.3 for HH[001]
when H tends to a critical point from above (since dO/dH ~ c~ in that case).
Susceptibilities Xd(T) of intermetallic compounds ErFe3, HoFe3 measured on the
polycrystal samples (Bisliev et al. 1973) are shown in fig. 8.4. The anomalous
behavior of Xd(T) in the vicinity of the compensation point is attributed to the
noncollinear structures in this field. It should be noted that magnetic anisotropy
in intermetallic compounds is much larger than the anisotropy of rare-earth-ferrite
garnets (close to Tc it accounts for about ~ 106 erg/cm 3) (Clark et al. 1974).
Therefore the effects of anisotropy on magnetic phase diagrams and the anomalies
of physical properties are of greater significance in these compounds than in rare-
earth-ferrite garnets.
Opposite, to some extent, is the example represented in fig. 8.5 (Gurtovoy et al.
1980). They studied the differential magnetic susceptibility of the Y3-xGd~FesO12
iron garnet system (0.01 ~< z ~ 0.2) in magnetic fields up to 50 T at temperatures
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 461

Xd,10-3 cma/g
/ 3 ",,,x\ (3) 1.0

(2) (1,4)[ a ~~ x / x~ . 4

0.1 -0.15 0.5

0.05- 0.1 0

0 -0.05

0
300 350 400 450 500 T, K
Fig. 8.4. Temperaturedependence of the differentialmagnetic susceptibilityof HoFe3 in the vicinity
of the compensation temperature for: 1) H = 11 kOe, 2) 9 kOe and 3) 1.5 kOe; and 4) of ErFe3 for
H = 11 kOe; after Bisliev et al. (1973).

X.IO 4

~__.4

I MFe
MGd 'v,/
41 jl

3"
10 20 30 40 50 H, T
Fig. 8.5. The field dependence of the differential magnetizationof (GdY)3FesO12 iron garnet; (o) -
experiment, solid lines - theory; after Gurtovoyet al. (1980).

between T = 186 K and 4.2 K. Transitions from the ferrimagnetic to the noncollinear
phase and from the noncollinear to the ferromagnetic phase are observed.
These examples illustrate that the 'universal law' Xd = 1/A for the canted phase
is appropriate in the strict sense only in the case of isotropic ferrimagnet. In section
462 A.K. ZVEZDIN

16 an example will be given where this law also holds in the anisotropic situation
(in the cone-canted phase).

8.2. The temperature hysteresis of the magnetization. Hall and Faraday effects
A temperature hysteresis of various physical quantities may arise during a transition
into the noncollinear phase. Diagrams depicted in figs 4.1-4.4, 4.6, 4.8, 4.9 show
that the transitions between some ferrimagnetic phases become first order phase tran-
sitions when the anisotropy is included. This can result in a temperature hysteresis
of the Faraday effect and the magnetization, which is attributed to the presence
of a phase coexistence area in the vicinity of the first order phase transition line.
Such effects become the strongest in weak magnetic fields close to the compensation
temperature Tc.
Let us consider the ferrimagnet magnetization behavior close to Tc in the scope of
model described above. Let HrlEA. It is seen from the phase diagram (fig. 4.2) that
a first order phase transition occurs on curve TcP (Belov and Nikitin 1970, Zvezdin
and Matveev 1972a, b).
As the temperature decreases, the transition I-II proceeds in a jump like on the
curve A'Q and reverse transition proceeds on the curve PBq Consequently the
overall magnetization in these phases depends on T in two different ways:

Md -- Mf(AMd - H) = -Ms(T) + xfH for 0 = 0,


Mt = ]. Mf(AMd + H) - Md Ms(T) + xfH for 0 = ~r,

where Ms(T) = Mf(AMd) - Md is saturation magnetization of the ferrimagnet for


H = 0, Mf(AMd + H) is magnetization of the rare earth sublattice in an effective field
AMd + H, Xf is its susceptibility. Taking into account that Ms(T) can be represented
in the vicinity of the compensation temperature Tc in the form

T-To
Ms(T) = Md - - ,
Tc
we obtain (Zvezdin and Matveev 1972)

IT-Tel
--Md -I- xfH for T > TI(H),
Tc
Mt = IT_ Tc[
Md Tc + XfH for T < T2(H),

where Tl(H) and T2(H) are determined by the stability lines A'Q and B'P in fig. 4.2.
In the case of weak magnetic fields they can be approximately defined in the fol-
lowing way:

HMd
T1,2(H) = Te 1 q: 2K //
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 463

Ms
\ J

J T

Fig. 8.6. Temperature hysteresis of the magnetization of an uniaxial crystal for HIIEA (easy axis) in
the vicinity of the compensation temperature; after Zvezdin and Matveev (1972b).

cr~ Gcm3/g

0.8

0.6

0.4

0.2
\ \
I I ^

480 t -\ 500 d 520 ~K

-0.2 ~- ~ "o"

Fig. 8.7. Temperature dependence of the magnetization of the ErFe2 compound in the vicinity of the
compensation temperature; after Belov et al. (1972b).

or more precisely by equations (4.13), (4.14).


The butterfly-like temperature hysteresis loops are shown in figs 8.5 and 8.6.
Such loops have been observed experimentally in the temperature dependence of the
remanent magnetization in ErFe2 near Tc (fig. 8.7, Belov et al. 1972b). Notice
that the very unusual two-peak temperature dependence of the coercive force of
464 A.K. ZVEZDIN

He, kOe
C
+MR +Mo

a) i i t---~l, . . . . ! t i
10 20 i J30 r, K
I i J

' I x.x~ '

Fig. 8.8. The temperature dependence of the coercive force a) and magnetization b) of the Er0.sDY0.mFeO 3
in the vicinity of the compensation temperature Tc ,-~ 25 K; after Belov et al. (1979).

polycrystalline Gd3Fe5012 iron garnet observed by Belov and Ped'ko (1960) can be
explained by similar model descriptions (Goranskii and Zvezdin 1969a).
The butterfly-like hysteresis curves are rare in nature. Figure 8.8 shows very
distinctive butterfly-like hysteresis of the magnetization near the compensation point
(Tc ~ 25 K) in the Erbium and Terbium orthoferrites ErFeO3, TbFeO3 (Derkachenko
et al. 1974, 1984, Belov et al. 1979).
There is an analogy of the hysteresis of magnetization considered here and the
butterfly-like hysteresis of the linear effects in the antiferromagnets and weak fer-
romagnetics (linear magnetostriction and piezomagnetism (Borovik-Romanov 1959,
1960, Zvezdin et al. 1985), linear magnetoelectric effect, linear birefringence (Khar-
chenko et al. 1978, Rudashevsky et al. 1977, Merkulov et al. 1981)). All these
effects are caused by the hysteresis of the antiferromagnetic vector/~.
A similar vector in the case of ferrimagnets considered here is

/~ =- ]~ff - f~d.
In f-d ferrimagnets the orientation of the vector/~ is fully determined by Md (see
section 3). Therefore the vector /~rd can be used for the analysis of behavior of
different physical properties in the vicinity of Pc.
The hysteresis of the vector/~ or Md in ferrimagnets near the compensation point
leads to many anomalies in the behavior of the physical values which are proportional
to L, e.g., of the kinetic effects. This is the situation with the galvanomagnetic effects
(Hall effect and magnetoresistance) in the vicinity of the compensation point. These
effect have been studied in ferrites (Belov et al. 1960b, 1961), in amorphous rare-
earth-Co(Fe) alloys (Asomoza et al. 1977, McCuire et al. 1977, Okamoto et al.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 465

1974, Ratajczak and Goscianska 1980), in the intermetallic compounds MnsGe2


(Levina et al. 1963, Novogrudskii and Fakidov 1964, Vlasov et al. 1976, 1980,
1982), and in amorphous Dy-Co films (Khrustalev et al. 1989, 1993a, b).
The temperature hysteresis and other features of the Faraday-effect in ferrite gar-
nets in the vicinity of Tc have been investigated by Smimova et al. (1970) (see also
Belov et al. 1979, Krinchik and Chetkin 1969, Eremenko and Kharchenko 1979).
The effects connected with this first order phase transition in amorphous films in
the vicinity of the compensation temperature have been studied by Makarov et al.
(1980), Kandaurova et al. (1982, 1985), Fisch et al. (1986), Popov et al. (1990).
Theoretical papers by Turov et al. (1964), Schavrov and Turov (1963), Schavrov
1965, Turov (1987), Zvezdin and Matveev (1972a, b) are devoted to this problem.

9. Thermal properties in the vicinity of the spin-reorientation phase


transitions

9.1. Magnetocaloric effect


The entropy of ferrimagnetic system (in the scope of the model studied) can be
expressed by

S- ~¢T - OT
O ~ fo He" Mf(x)dx, (9.1)

where the quantity ¢ is given by formula (3.1), and where O(H,T) is defined by
eq. (3.9). The assumption of a saturated d-sublattice has been made here, i.e.
OMd/OT = 0 and ~ / ~ 0 = 0 in equilibrium conditions.
Equation (9.1) in conjunction with (3.9) determines the isoentropic region in the
(H, T) plane and henceforth the magnetocaloric effect, i.e. the variation of the sample
temperature during adiabatic magnetization. It is defined usually by the isoentropy
slope, i.e. by the value

(OS/~T)H CH
( OMH ~ ,

where C H is the specific heat of the s_ystemfor H = const, MH is the projection of


an overall magnetic moment on the H direction.
Substituting the value MH = M(H) from (6.4) into the above formula, we get

T DMf for 0 = 7r,


CH1 8T

CH2 AHMa dT (H - AMd cos O) --sinf°O for 0 < 0 < 7r,


T ~Mf H - A M a
for 0 = O.
CH3 ~T I H - AMd]
466 A.K. ZVEZDIN

These formulas and (3.9) and (3.10) simultaneously determine the magnitude of
(dT/dH)s. Furthermore, it is easily seen from (9.3), to be equal to zero in the canted
phase in the isotropic case since K = 0. It follows at once from the fact that in this
case the magnetization Mt = H/A is independent of T (see section 2). Notice that
Belov et al. (1972a) have obtained more detailed formulas for ferrimagnets with
cubic anisotropy.
By contrast, (dT/dH)s > 0 in the low temperature phase {0 = 7r), i.e. an
increase of the temperature of the sample during magnetization takes place whereas
(dT/dH) < 0 in the phase {0 = 0}, i.e. cooling takes place. This property has an
evident physical meaning. The external field is directed in parallel with the exchange
field acting on H rare earth ion in the phase {0 = 7r) so that the Zeeman splitting of
its ground multiplet is enhanced with increasing field. As a result more of the lower
levels become occupied (entropy of the system decreases). The energy released at
such a transition heats up the sample. The situation becomes reversed in the phase
{0 = 0}.
A characteristic property of the second order phase transitions is the kink in the
isoentropic curves, i.e. the jump of (dT/dH)s at the boundary between the phases.
Experimental data for some compounds are shown in fig. 9.1 (Belov et al. 1970a).
It is seen that (dT/dH)s = 0 with a good accuracy in some phase of Gd-ferrite
garnets. This value perceptibly differs from zero in the noncollinear phase of Ho
and Dy ferrite garnets. This is attributed to the large value of the anisotropy energy
inherent in these materials.
The sign change of (dT/dH)s is also well-defined at the transition between the
phases {0 = 0} and {0 = 7r} in fig. 7.1. Similar dependencies of T(H) have been
observed in Gd-ferrite garnet, Yb-ferrite garnet and in the mixed Gd/Yb-ferrite
garnet as well in the work by Clark and Callen (1969). The isoentropic curves show
discontinuity on the boundary of the first order phase transitions.

9.2. Specific heat


The magnetic part of specific heat may be derived by differentiating formula (9.1):

dS
C H : T ( - ~ ) H. (9.3)

In order to obtain analytical expressions let us define a particular form of the


dependence of the magnetization of the rare-earth sublattice on temperature and
magnetic field

(9.4)

Such an approximation is valid in the vicinity of the compensation temperature


(and also above) and can be considered as an expansion of the Brillouin function.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 467

A7 10-2 K A7 10-2 K ~ 220.04 K


,.~_~:,- 228.19
5.0 Gd3Fe5012 rr°"o 283.21 K 2.0 Dy3Fo,o12f 220.34
~/j~-E85.n ~ : ~ : ~ 220.47
, ¢ , ~ ~ 285.47
2.5 285.65
285.89
288.07
1.0
~
~
220.60

~-.~'_ _ _1~_ _ _ _ ~ _ H, KOe 0 I I I

- ~ 288.47 H, KOe
- - ~ 7 °~~ _ 288.71
-2.5 r 287.12 -1.0
~ 287.32
-5.0 -2.0
a) b) .72

AT, 10-2 K
a/137.57 K
7.5 Ho3FesO12/

1.9

2.5 x~_ ~,.~-o~ooo 138.65


138.78
0 ~ " ~ 138.87
10 15 KOe
-~-~--o~,<>~a~ 138.96
-2.5
~ noq39.13

-5.0

-7.5 %xab'Q'~-,Ri ~ ~

c)
Fig. 9.1. Magnetocaloriceffect in a) Gd3FesOl2, b) Dy3FesO12, c) Ho3FesO12 in the vicinity of the
compensation temperature; after Belov et al. (1970a).

For example, we have a = 3/7, b = 67/1029, /zf = 7/ZB, Mg = 21#B for Gd-ferrite
garnet. Substituting (9.4) into (9.1) yields

aMf°~,f
S - - - (2H)~Mo cos e - H 2 - k2M~). (9.5)
2T 2
468 A.K. ZVEZDIN

Only the first term of the expansion in (9.4) is presented here. The error allowable
in such an approximation for Gd3FesO12 at T ~ Tc is lower than 5%. Replacing this
value in the formula for the specific heat Cr~ we obtain

Mf°#f
a- - (H -/~md) 2, for 0 = 0,
T2
M%
a - - (H 2 2HAMdcos 0 + AZM~+
CH= T2 \

+AMdHT dcos______O0) for 0 < 0 <


dT J '
M%f
a - - (H + AMd) a, for 0 = 7r.
T2

The magnitude O(H,T) is determined here by the equation of state (3.9) for the
corresponding canted phase.
Let us consider for example the case of cubic anisotropy HIll100]. The specific
heat changes in a jump like during the transition from the collinear phases into the
canted phases on the curves AA' and BB' (lines of the second order phase transition,
see fig. 4.1)

a2 M ° H2
AC = - - , (9.6)
2b /zf H a + 4 H , 2

where H* is given by formula (4.9).


Formula (9.6) predicts extremely curious behavior of the specific heat CH in a
canted phase. Schematically the CH(T) behavior at various values of H is represented
in fig. 9.2 (HII[001]).

cn
_ _ ~ H > H*

H - H*

T
Fig. 9.2. Schematic representation of the temperature dependence of the magnetic contribution of the
specific heat of a ferrimagnet of cubic symmetry in the vicinity of the compensation temperature for
HIll001] (see fig. 4.3); after Belov et al. (1979).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 469

Cp, 1N
~"
-2 g.K
ca]

w~ ttl
,,' , , , H=20.4k0e
coo ' d It I¢¢s %
14 "'"'"'"'"'""'""'"'"'"'"'""'" ,., ,.........,,,,,.,,,,,,v,.....
,,, ,; /,
°¢r # I~
It
.,~. ,,,# i t , 16.4
14 .......... "'". . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ' .......... ' " .......
,,,",I ~', .,.~ '1fl. t. . . ,l ,,' *, 6.0
,i¢•II' 11o11%6 ¢tllllllll I• .¢ I IB fill III" llllllli•llllllllllllll~l
14 "'" ..... " " " ......
16 ' '\
1,
s, ,"'
l ,~Ii
,~ ,1 ,.,,s ¢S I I~ 2.0
i o I 0 Ii IIIDII t I ii ~1111
Ii i iii I II I#I011111III • • |llle•l•
• •,~ 'tl°o •1•o¢p i • " lllll Ill
14 ........... ". . . . . . 0
14
I I I I
280 290 300 310 T, K
Fig. 9.3. Temperaturedependencies of the specific heat of the Gd3Fe5012 at different magnetic fields,
grll[100]; after Kamilov et al. (1975).

The peak of the specific heat in the vicinity of the critical point O is of interest
(see fig. 4.1). When this point is approached we have
d cos 0
- - --+ OO.
dT
Let us note that the peaks of the specific heat in the vicinity of the critical curves
and critical points represented in fig. 9.2 have not a fluctuational nature but have to be
attributed to strong temperature dependence of the order parameter O(T). Generally,
the region situated close to the phase diagrams 'throat' is of the most interest from
the point of view of observation of unusual specific heat dependencies (and other
physical parameters) since there is the strongest O(T) dependence.
Anomalies of the specific heat in Gd ferrite garnet for HI[[100 ] in the vicinity
of Tc have been observed by Kamilov and Schachschaev (1972), Kamilov et al.
(1975), as shown in fig. 9.3. The data are in qualitative conformity with the phase
diagram represented in fig. 4.3. We note that the comparatively weak magnetic fields
considered in these investigations are the most difficult ones for theoretical analyzes
since domain structures play an essential role here.
Kamilov et al. noted that (dT/dH)s ~ 0 in the canted phase. This experimental
result is also consistent with the theory.

10. Magnetoelastic anomalies

I0.I. Magnetostriction
The noncollinear magnetic structures in rare-earth-ferrite garnets are accompanied
by anomalies of the magnetostriction (Belov et al. 1969, 1970b, 1972a, Levitin
470 A.K.ZVEZDIN

et al. 1970, Popov 1971, Levitin and Popov 1975). If the magnetostriction is mea-
sured along the field direction then only longitudinal component of the anisotropic
magnetostriction and the bulk magnetostriction (magnetostriction of paraprocess)
contribute to the measured value in the collinear phases (if the magnetic moments
of the sublattices are oriented parallel or antiparallel with field direction).
Transverse components of the anisotropic magnetostriction also arise. Since lon-
gitudinal and transverse components of magnetostriction have the opposite signs, the
derivative of anisotropic magnetostriction, with respect to the field, changes its sign
during the transition from the collinear phase into the noncollinear one. Besides, the
bulk magnetostriction of the paraprocess changes in anomalous way as well during
formation of the noncollinear structure since it is dependent on the magnetization of
the sublattices and as it was shown above, the magnetization of the rare earth sub-
lattice depends in a different way on the field for collinear and noncollinear phases.
As a consequence of this anomalous magnetostriction, its variation in ferrite garnets
during the transition into the noncollinear phase has a much sharper character than
the variation of the magnetization, and minimums (or maximums) appear in the field
dependence of the magnetostriction curve at fields corresponding to the critical fields
of the transitions.
Let us consider in more details the field dependence of the magnetostriction of fer-
rite garnets at the transition into noncollinear phase. The anisotropic magnetostriction
is supposed to be a single-ion effect and the bulk magnetostriction is to be attributed
to exchange interaction between the rare-earth and Fe sublattices. Then the overall
magnetostriction of ferrite garnet polycrystal can be represented by means of usual
relations for magnetostriction (Belov et al. 1979)

1
All - 21 /~Fe(3COS2~)Fe_ 1)+~ /~R(BCos2~R- 1)+
(10.1)
+ aA(MRMFeCOS(0R+ 0Fe)+ MRMFe).

Here the first two terms describe anisotropic magnetostriction of Fe and rare-earth
sublattices, ~bFeand ~bR are the angles associated with the directions of the sublattice
magnetizations and the direction of the magnetostrictive deformation measurements.
It is necessary to take into account the field dependence of magnetostriction con-
stant AR of rare earth sublattice attributed to the paraprocess. In the single ion
approximation such dependence has a form of

AR = A°]5/2[L-I(MR/M°)],
where i5/2(x) is the reduced Bessel function, L -1 is the reciprocal Langevin function.
The third term in expression (10.1) is the bulk magnetostriction of the paraprocess.
The coefficient ax is related to the dependence of the exchange interaction between
the sublattices A upon the tension strains cri~ as follows
dA dA dA
d~xx d~yy d~zz
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 471

OR and 0Fe are the angles between the sublattices magnetizations and the field, which
are equal to 0 or 7r in the collinear phases and can be found from relations given in
section 2.
In sufficiently strong fields when the isotropic ferrimagnet approximation is valid
one may write

(HclHc2 -[- H 2)
COS0R =
2H MR MFeA '
(-HclHc2 + H 2)
COS 0Fe =
2HMRMFe~

If the magnetostriction is measured along the field direction then ~Ve = 0Fe, ~R =
OR and

,~111 : /
T 111
1 1
/~Fe(3 COS0Fe -- 1) + ~ -~R(3 COSOR -- 1)+
2
+ a,X(MRMFe COS(0R + 0Fe) -[- MRMFe).

It is easy to obtain the field dependence of the various contributions to the magne-
tostriction (Levitin and Popov 1975) by means of presented above formulas. Let us
note that the magnetic field dependencies of the rare-earth sublattice magnetostriction
in the ferrimagnetic phase above and below the compensation temperature are differ-
ent. This difference is attributed to the decreasing rare earth sublattice magnetization
by the field above the compensation temperature and to the increasing magnetization
below the compensation temperature.
Figures 10.1, 10.2 show field dependencies of longitudinal magnetostriction in
Gd, Tb, Dy and Ho-ferrite garnets close to the compensation temperature in fields
up to 250 kOe (Levitin and Popov 1975). The magnetostriction of explored garnets
anomalously depends on the field in this temperature range: maxima (and minima)
appear in the A1/l(H) curves for some values of the fields; these fields correspond
to the critical fields of transition into noncollinear phases.
Similar field dependencies of the magnetostriction in the vicinity of Tc have been
observed for the intermetallic compound ErFe3 (Nikitin et al. 1975).
Popov (1971) provides numerical computations of the theoretical field dependence
of the magnetostriction in Ho-ferrite garnet (formulas (10.1), (10.3)) when field-
induced noncollinear magnetic structure arise. Comparison between the computa-
tional results and the experimental field dependence of Al/l is made in fig. 10.2.
Taking into account approximate character of calculations the obtained agreement
between theoretical and experimental dependencies Al/l(H) can be considered as
satisfactory.
472 A.K. ZVEZDIN

(A///)n (A///)n
I I t I
292 K ' 36K
(xl0 -6) 4 (×10-6) 40
~ 289
0 240
~286
30

20
-16
10
-24 v Tb3FesO12
0 100 200 h kOe 100
' o'
2 0 H kOe
(AV0H (A///)II
i i
' ' ' ~24 t
(xl0 -6) 20 (×10 -6 ) 20
0 0

-40 -40

-80 -80

-120 DY3Fe5OI~ -120


~//Ho3Fe5012N~~ "
I I I T

0 100 200/-J kOe 0 100 200 H, kOe


Fig. 10.1. The longitudinal magnetostrictionof the rare-earth-iron garnets in the vicinity of the com-
pensation temperature(after Belov et al. 1970b, Levitin and Popov 1975).

10.2. Thermal expansion


The appearance of noncollinear magnetic structures leads also to anomalies of the
thermal expansion coefficient, which is magnetostrictional deformations (Nikitin et
al. 1975). The existence of these anomalies follows directly from the expression for
the magnetostriction (10.2) and this is due to the fact that only the magnetostriction
constants ~Fe, ,~n, a;~ vary with temperature in the collinear phases while also 0R
and 0Fe vary with temperature in the noncollinear phases. Figure 10.3 illustrates the
thermal expansion of Dy-ferrite garnet along the [111] axis. It is seen that thermal
expansion has no peculiarities at H = 0. However, the anomalies in A~/,k and
~(T) are observed to become sharper with increasing field strength for a sample
placed into a magnetic field. These anomalies are connected with the formation of
noncollinear magnetic structure in the field near the temperature of magnetic corn-
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 473

0 50 100 150 200 H, kOe


I I I I
(A///)II
Ho3FesO12
T = 110 K
-50

-100

-150

(xl0 ~ ) (-200)
I I P I

Fig. 10.2. The theoretical (1) and experimental (2) magnetic field dependencies of the longitudinal
magnetostriction of Ho3FesO12 iron garnet at T = 110 K (after Popov 1971),

AHI
1 or, 10-6 K-1
(xl0 -6 ) 100 20

50 10
4
e/, 0
210rC/,fl Tk 220 230 T,K
2~

-50
Fig. 10.3. The temperature dependencies of the thermal expansion of DY3FesO12 iron garnet along [111]
in the vicinity of the compensation temperature (To) at different magnetic fields: 1 - 0, 2 - 50, 3 - 15 kOe;
4 - temperature dependence of the coefficient of thermal expansion at H = 15 kOe; after Nikitin et al.
(1975).

pensation. S i m i l a r a n o m a l i e s h a v e been o b s e r v e d in the vicinity o f the c o m p e n s a t i o n


p o i n t o f ErFe3 (Nikitin et al. 1975, fig. 10.4).

10.3. Young's modulus, sound velocity change (AE-effect) and sound absorption
A d d i t i o n a l d e f o r m a t i o n s caused b y magnetostriction arise in the m a g n e t i c a l l y or-
d e r e d c o m p o u n d s u n d e r effect o f the tension strains. It leads to Y o u n g ' s m o d u -
lus c h a n g e s at transitions into the m a g n e t i c a l l y ordered state and to d e p e n d e n c e o f
Y o u n g ' s m o d u l u s on external fields (AE-effect). A n o m a l i e s o f Y o u n g ' s m o d u l u s
ought to a p p e a r in ferrimagnets as well since external strains affect not only the
474 A.K. ZVEZDIN

150 1

100

30

x 50 20 "7

lO

£230 .2¢-" 230 T,K 10

~50 P /
Fig. 10.4. Temperature dependence of the thermal expansion of the ErFe3 intermetallics in the vicinity
of the compensation temperature (Tc) (1 - H = 0, 2 - H = 50 kOe, 3 - H = 15 kOe) and the thermal
expansion coefficient (4 - H = 15 kOe, 5 - H = 50 kOe); after Nikitin et al. (1975).

AE/E
I I I I
O

(xl0 -3 ) 6
o - 100 K _ ° ~ ''m~
4

-2

0 50 100 150 200 H, kOe


Fig. 10.5. Magnetic field dependence ofthe AE-effect ofthe Ho3Fe5012 iron g~net; aher Levitin and
Popov (1975).

magnitude o f the sublattice magnetizations in the collinear phases but also their di-
rections in non-collinear phases. Measurements of AE-effect in Ho-ferrite garnet
in the fields up to 220 kOe (Levitin and P o p o v 1975) corroborate Young's modulus
j u m p existence during the transition into non-collinear phase (Fig. 10.5).
K a m i l o v et al. (1975) observed the anomalies of the sound propagation in the
G d - i r o n garnet near the compensation temperature which are in reasonable agree-
ment with the H - T phase diagrams o f this material. Figure 10.6 shows the sound
absorption coefficient in the Gd3FesO12 at frequency 30 M H z for the magnetic fields
H = 0, 2, 4, 10 k O e (~qll[001]).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 475

or, dR/cm a, dR/cm

• 3.8 k
3.8 - :: H=akOe
o
°°o ~oo e °ee°
-, ' H = 2 kOe
- ,~,~,
o ; : 3.7
oo 3.6
3.6 k9 oo~o ~ °

0 0 • 4~
3.8 ~ oH=0 --':
GO0 0000 0
3.4 L9 0 I q I I 3.4 3i51 f I r
10 20 30 40 0 10 20 30
T, oc T, oc
Fig. 10.6. The temperature dependence of the coefficient of sound absorption in the Gd3FesO12 garnet
at the frequency 30 MHz; 0"_LH[[[1001;after Kamilov et al. (1975).

Similar anomalies were obtained also for the sound velocity, for which there are
the pattern of the minima near the compensation temperature (see also figs 4.3, 5.6,
and 9.3).
Notice that magnetostriction of the rare-earth sublattice depends strongly on the
magnetic fields in the region of temperatures and fields where magnetization is satu-
rated (Vedernikov et al. 1988, Kadomtseva et al. 1989). This effect should be taken
into account for adequate determination of critical fields.

11. Non-collinear phases and domain structure

There are two specific phenomena in the context of field induced phase transitions:
i) splitting of a sample into twinned domains in the canted phase, and ii) domain
wall expansion and nucleation of the new phase from the domain wall during the
first order phase transition. The latter mechanism may be responsible for the fact
that the hysteresis of these transitions is often absent or very small.

11.1. 'Break' of symmetry in the canted phase and formation of domain structures
with twins, triplets and quadruplets
A noteworthy feature of the phase diagrams treated above is the presence of original
'degeneracy' of the canted phases. It means that in every case considered there exist
several solutions of the thermodynamic potential minimization problem correspond-
ing to one and the same value of this potential. Thus, for H l l [ l l l ] the degeneracy
is equal to three since three physically equivalent rotational planes are presented
here:

= ~r/2 (37r/2), ~ = 7~r/6 (7r/6), ~ = 117r/6 (57r/6)


476 A.K. ZVEZDIN

for canted phases D(C) in fig. 4.4(a) (K < 0).


In a case when/~U[100], the degeneracy ratio is equal to four, w h e n / t l l [ l l 0 ] it
is equal to two.
The degeneracy does not depend on the starting model but is a result of the
system's symmetry. In the c a s e / I l l [ I l l ] , the [111] axis corresponds to the third
order symmetry axis in the symmetrical (collinear) phase. During transition into an
angular phase the symmetry decreases. The symmetry element C is absent in the
canted phase. This broken symmetry element has transferred equivalent solutions
(rotational planes) from one to another after symmetry has been restored.
The interesting situation arises when HII [110]. In this case the 'break' of symmetry
element C2 i.e. rotation around the [110] axis takes place at a transition from the
collinear phases A and B into the canted phases C and D and the two-fold degeneracy
of the angular phases develops.
In going from the canted phases C and D into phase E one more symmetry
element - the mirror plane disappeares. It is also evident that the number of phases
becomes doubled here.
The degeneracy of a given phase may be revealed by the fact that the crystal
becomes divided into domains (twines, triplets, quadruplets). Let us consider the
character of such domain structure for the case/~11[110] (see fig. 4.6). Let's fix the
magnetic field and change the temperature, going over all phases A, C, E, D, B
sequentially.
At the transition A - C domains of the type (~ = 7r/2) and (qo = -7r/2) appear.
With increasing temperature 0 decreases from 7r down to the value 0 = - arcsin v/3-/5
after which each of the domains of phase C in the phase E splits into two parts with
=~+ and~=~_.
The domain walls strongly expands close to the point of phase transition between
phases C and E. As this point is moved away the distinction between the split
domains is enhanced, i.e the difference between ~+ and ~_ increases and attains
its maximum value in the center of phase E, after which the magnetizations in the
separated domains 'tend' to each other again and at the of E - D transition point the
split domains merge together.
The maximum expansion of the domain walls occurs close to the critical points O
and O'.
Two types of domains (0, ~ = 7r/2) and (0, ~ = -7r/2), exist in the phase D.
The difference between these domains vanishes at the D - B transition point and they
collapse in the collinear phase.
This picture of domain structure transformations during transitions over angular
phases practically does not depend on model assumptions, and is defined by the
symmetry properties of the system. Finer details, particular those pertaining to
domain wall behavior may be obtained by the methods of qualitative theory of
differential the equations (see below).
FIELD INDUCEDPHASETRANSITIONSIN FERRIMAGNETS 477

11.2. Nucleation of new phases from domain walls. The hysteresisless first order
phase transition
Let us note that first order phase transitions here can be accomplished in a non-
hysteretic way. The mechanism of such transition is continuous growth of a new
phase from the domain wall.
This mechanism has been put forward by Mitsek et al. (1969), Mitsek and Sere-
bryanik (1976) for the spin-flop transitions and later by Belov et al. 1974, 1975,
1976 to explain the absence of the temperature hysteresis at the first order transitions
in DyFeO3 and (YTb)3FesO12. Similar behavior of the domain walls at the first
order phase transitions was observed also in RE-TM intermetallics (see for details
the review by Asti 1990). There are direct visual observations of this picture of the
first order phase transitions (Dillon et al. 1974, King and Paquett 1973, Belyaeva et
al. 1977, Lisovskii 1980, Gnatchenko 1989, Szewczyk et al. 1985, Szymczak et al.
1983, 1987).
Mathematically we can illustrate this process easier by using an example diagram
as represented in fig. 4.8 (/~]l [100]).
Let us consider the series of transitions A - C - B at a fixed value of field H < / / 3 .
In an area lying under curves OA', O'B' the phases A and B can coexist in the form
of domains separated by 180 ° domain wall. It is easy to show that rotation of the
angle 0 in such walls occurs in (010) and (001) oriented planes. We shall follow the
work by Zvezdin and Popkov (1977) in the further analysis.
Free energy allowed for the exchange energy has a form

= f { [(grad 0)2 + sin 2 O(grad ~)2] + F(O, ~0)} dV, (11.1)

where A is an exchange stiffness constant. ~(0) is thermodynamic potential deter-


mined by formula (3.1).
The first integral of the differential Euler-Lagrange equations defining the O(ec)
dependence in a domain wall leads to the following equation (ec is coordinate in a
perpendicular to domain wall direction).

dO/dz = +A-1/2(~(0) - ~0) 1/2, (11.2)

where ~0 is equilibrium value of ~(0) at given H, T and qo = const. This equation


can be integrated at once but to clear up the main features of phase transformations
it is sufficient to restrict ourselves by a qualitative analysis of the equation obtained.
Figure 11.1 illustrates the series of the integral curves represented by the equation
(11.2) at different values of T. Obviously, the singular points on these curves where
dO/dec = 0 correspond to domains and the whole curve between these points conforms
to domain wall. Minima in the curves dO(ec)/dec are consistent with the bends of
domain wall where rotation of O(ec) slows down. This slowing down signals the
origination of the new phase from the wall. The domain of new phase come into
existence at the first order point (e.g., the curve O'To in fig. 4.8). The integral curve
at fig. 11. lb) nearly touches the abscissa axis at this point.
478 A.K. ZVEZDIN

--7l" 7"t" 0 -zr 01 zr 0

--71" 71" 0

Fig. 11.1. Integral curves in the plane 0~,0 governed by equation (11.2). They describe the domain
walls depending on the temperature at H = const (Hill001], KI > 0). The points of contact with the
axis 0~= 0 correspondto the domains, the lines connectingthese points describe the domain walls; after
Belov et al. (1979).

It is seen from the fig. 11.1 that bends in the domain wall grow when the tem-
perature decreases and approaches to the first order phase transitions (the transition
A --+ C in fig. 4.8). In the canted phase C domains of the 'old' phase A transform
into bends of domain walls separating the twinned canted domains (fig. 11.1(b)).
Corresponding bends of these domain walls convert into domains of phase B at
transition C -+ B and the canted domains collapse. Obviously, there is a continuous
conversion of domains, i.e. the process is completely reversible.
We have explored the transition along the line H = const. The discussion holds
completely for the other transition trajectories (T = const, for instance).

11.3. Canted phase domains in ferrite garnet single crystals and films
The first indirect experimental indication of the possibility of domains existing during
the transition into the noncollinear phase induced by the applied field has been given
by Kharchenko et al. (1968). The jumps and hysteresis phenomena were revealed
in the field dependence of the Faraday effect in gadolinium-ferrite garnet during the
transition into noncollinear phase. It was shown that observed peculiarities cannot
be explained by a rotation of the overall ferrite magnetic moment.
The domain structure arising at a transition into the noncollinear phase in D y -
ferrite garnet was visualized by use of the Faraday effect by Lisovskii and Schapo-
valov (1974).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 479

•~ " j:

. . . . . . . . , ~ - . ~

Fig. 11.2. Domain structure in a plate of a Gd3FesO12 single crystal (the dark and bright areas are the
domains of different phases): a) H = 7 kOe, T = 284.8 K, b) 285.1 K, c) 285.2 K, d) 285,3 K; after
Kharchenko et al. (1974).

The domain structure in Gd-ferrite garnet in high fields with/~ll[100] has been
observed by Kharchenko et al. (1974), Gnatchenko and Kharchenko (1976) at various
temperatures (fig. 11.2). In this case sample segregates into a mixture of the high and
low temperature noncollinear phases close to the point of magnetic compensation.
Suitable materials for experimental investigations of the domain structure in the
canted phase are the epitaxial ferrite garnets films. Dikshtein et al. (1980) inves-
tigated epitaxial films of magnetic garnets (Y, Gd,Yb,Bi)3(Fe,A1)5012, of thickness
5-15 #m, grown on substrates of Gd3GasO12 cut along (111). These authors have
shown that in the thin films thermodynamically stable domain structure can exist
in strong magnetic fields, up to the flip field of the magnetization. The domain
structure of the films was observed in polarized light, by means of a microscope,
using the Faraday effect. The films were placed in a magnetic field directed approx-
imately parallel to the chosen plane. The light was propagated along [111] axis.
The source of the magnetic field was an electromagnet of the 'Solenoid' installation
of the General Physics Institute of the Academy of Sciences, Russia, which allows
experiments to be performed in stationary magnetic fields up to 150 kOe (Veselago
et al. 1968).
480 A.K. ZVEZDIN

Notice that the magnetic anisotropy of films of magnetic garnets differs from
uniaxial (cubic and rhombic components also present; see, for example, Eshenfelder
(1981)), therefore a second-order phase transition, for a chosen orientation of the
film, occurs within a narrow interval of angles between H and the developed surface
of the film (Dikshtein et al. 1980).
The results of the experiments are shown in figs 6.5 and 11.3 for one of the films
investigated, of 5 # m thickness, with compensation point Tc = 310 K and Curie
temperature 420 K. The easy magnetization axis in the film was inclined to the
normal (the [111] axis) by 1 deg. The uniaxial anisotropy field, at any temperature
in the range 80 K < T < 420 K exceeded the saturation magnetization 47rMs and
therefore the domain structure that existed in the film was of the 'open' type (Kittel
1949).
At H = 0, within the temperature intervals T < 250 K and 375 K < T < 420 K,
an ordinary maze (spike type) domain structure was observed in the film. With
increase of the magnetic field, which was oriented so that the disappearance of the
domain occurred via the second-order phase transition, the period of the domain
structure decreased according to a linear law (see fig. 11.3). At the instant of the
disappearance of the Faraday rotation between domains with opposite sings of 37/
the period remains finite. Near the compensation point when H = 0, a single-domain

9
T=212K

3 I I I ~
5 6 7 8 H, kOe
Fig. 11.3. The magnetic field dependenceof the period of the domain structurein the film; after Dikstein
et al. (1980).

D c = ~ D e ~-

Fig. 11.4. Distribution of the magnetization in the uniaxial film of two sublattice ferrimagnets in the
vicinity of the second-orderphase transition with the magnetic field H_I_EA(easy axis); after Lisovskii
(1980).
FIELD INDUCEDPHASETRANSITIONSIN FERRIMAGNETS 481

interval (250 K < T < 375 K) was observed but with increasing of H, a domain
structure was generated near the critical values of the magnetic field He.
In fig. 6.5 there are plotted temperature variations of the critical field Hc and of
the reciprocal critical dimension of the domains in the critical field D~-~ = D-:(Hc)
and in the zero field D -1 = D-l(0). It is evident that the form of the curves Hc(T)
agrees well with the form of the theoretical diagram of fig. 4.1.
Figure 11.4 shows schematically the distribution of the sublattice magnetizations
in the vicinity of the second-order transition in the films studied.
When approaching to the compensation point Dc and Hc increase; the single-
domain interval with respect to 'high-field' domains is 45 K (290 K < T < 335 K).
For T = 290 K, coarse domains (~ 100 #m) were observed for H = 110 kOe.
For observing a domain structure in strong magnetic fields, it is necessary to
maintain carefully the conditions required for the occurrence of a second-order phase
transition (or of fist-order transition close to it). When appreciable departures from
these conditions occur, a domain structure is not observed at all, and an increase of
the field leads only to a replacement of one phase by the other via motion of the
interphase boundaries.
It is important to emphasize that in these experiments a domain structure in uniaxial
ferrimagnets in strong magnetic fields was observed during the phase transitions,
which are not accompanied by a jump of the resultant magnetization. This domain
structure can, occasionally, be thermodynamically stable (Khalturin 1976, Dikstein
et al. 1980).
Dikstein et al. (1983, 1984) discovered that regular domain structures in the
ferrite garnet films may become amorphous near the second-order phase transition.
Moreover, the process of amorphization is independent of crystal defects and exists
even in perfect, defect-free films. These observations can be understood, when the
possible formation of dislocation- and disclination-like magnetic defects are taken
into consideration. This idea has been confirmed by direct observations of the domain
structure at the points of second-order phase transitions.
By contrast with the common domain structures in ferrimagnets which are ad-
vantageous energetically, domains arising at a transition into noncollinear phase are
not energetically advantageous. Presumably the reasons for noncollinear ferrimag-
netic ordering during phase transition in the vicinity of compensation point ought
to be analogous to those giving rise to the appearance of antiferromagnet domain
structures. Particular domain structure can arise in a real crystal with defects, in-
homogeneities, internal strains and other impurities as a result of which one of the
domains becomes more preferable than others in that crystal region. Though the na-
ture of domain structure in ferrimagnets in strong field is not completely understood,
its discovery itself essentially extends our knowledge of criteria of domain formation
in magnets. The presence of metastable or energetically degenerate states in a certain
range of temperatures and magnetic fields should be considered as a main criterion
but not the smallness of the external field. Nowadays lots of objects are already
known (not only ferrimagnets) where similar domains can exist over wide ranges of
strong magnetic fields and temperatures.
482 A.K. ZVEZDIN

12. Hexagonal ferrimagnets

12.1. Free energy and equilibrium conditions


This paragraph is concerned with the phase transitions and the phase diagrams of
hexagonal ferrimagnets near the compensation point Te. The basal plane is assumed
to be that of easy magnetization. In this plane there are three easy and three hard
axes. We shall consider the two cases where a) an external field H is parallel with
one of the easy axes, and b) with one of the hard axes.
The case in which the hexagonal axis is the easy one has actually been examined
in section 4.1 (uniaxial anisotropy). I f / t is parallel with the hexagonal axis the
phase diagram is the same as for uniaxial anisotropy where H is perpendicular to
the easy axis (or easy plane).
The most important objects for which the developed theory may be applicable
are hexagonal compounds of d-f type, such as RCos, R2Fel7, R2Co17, Rz(FeMn)17,
RzFe14B and so on (Deryagin 1976, Kirchmayr and Poldy 1979, Buschow 1980,
1988, Sinnema et al. 1987, Givord et al. 1988, Yamada et al. 1988, Radwanski and
Franse 1989, Franse et al. 1990). There are compounds DyCo5 and TbCo5 in which
the basal plane of the crystal is the easy plane at low temperatures.
Compounds of this class are usually described in the two (d and f) sublattice
approximation. The magnetic moments of the d-sublattice are coupled via a strong
exchange interaction, Hex ~ 107 0 e . This produces the high Curie temperatures
(of the order of 1000 K) of these magnets. The d-f interaction is weaker Hex
106 0 e , and the f - f interaction is much weaker. On this basis we can say that the
magnetization of the d-sublattice does not depend on the value of the external field
or on the state of the f-sublattice. The exchange interaction between rare-earth ions
can also be neglected. In that approximation the rare-earth subsystem can be treated
as an 'ideal paramagnet' in an external field and in the exchange field generated by
d ions.
The presence of rare-earth atoms in these compounds leads to an appreciable
anisotropy in the basal plane at low temperatures including the magnetic compen-
sation temperatures (120-150 K). This leads us to expect here a strong influence of
the hexagonal anisotropy on the occurrence of field induced non collinear magnetic
structures.
Although the formula for the free energy of ferrimagnets has been discussed above
(section 3) we shall present the main arguments again to emphasize by this important
example the area of the application of the theory and its asymptotic behavior (small
parameters, etc.)
On the basis of the properties of the above model we shall say that the magnetiza-
tion of the d-sublattice does not depend on the directions of magnetization of d- and
f-sublattices, i.e. the thermodynamic potential of the d-subsystem can be defined by
the single vector "7 oriented along the magnetization of the d-sublattice. The rare
earth sublattice must be in equilibrium with the d-sublattice, whatever the direction
of "7. So the thermodynamic potential of the system can be written as

~("7) = qSd("7) + ~bf("7). (12.1)


FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 483

The potential ~d('7) is the sum of the magnetic energy -(lhrd/~ ) and the anisotropy
energy, i.e.

d~d('~) = - ( M d H ) + Kd('7) = -MdHCOS ~ - (1/6)K cos 6(~ + a),

where ~ is an azimuthal angle which defines the direction of ~ in the basal plane
relative to H. We are making use of the fact that/~ and/l~rd lie in the basal plane;
a is the angle in the basal plane, between one of the easy axes and/~.
The second term in (12.1) ~f('7) is the equilibrium thermodynamic potential of the
f ions in any effective field/teff

/'teff = /r~ _}_/-IM'

in which /~M = --A/~rd is a molecular field due to the d-sublattice experienced by


the of f ions. We shall divide ~f('7) into two parts

~f('7) = ~S(Heff) - (1/6)Kf cos 6(~f + a),

• S(Heff) is the axially symmetrical part of the function ~f('7); the second term is the
anisotropy energy of the f-sublattice in the basal plane, qaf is the azimuthal angle of
the vector/7.
Evidently ~S(Heff) does not depend on a because of the axial symmetry. Function
qSf(Heff) c a n be given as

~sS(geff) = - f0 He~ Mf(x) dx,

where Mf(Heff) is the magnetization of the f-ion when H_l_~', in the axial symmetry
approximation for the crystalline field. The explicit form of the function Mf(Heff) is
of no great significance. For qualitative conclusions we shall approximate it by the
Brillouin function (see, e.g., Li et al. 1988).
Thus

4)(~) = - M d H cos ~ - Mr(x) d x -


fo ~ (12.2)
- (1/6)Kd cos6(~ + a) - (1/6)Kf COS6(qof + a).

Here Heff = ( H 2 + H 2 - 2HHM COS qo)1/2, mf(x) is a known function of x. The


functions ~f, ~ are defined as follows (fig. 12.1):

HM
sin ~f -- sin ~, (12.3)
Heff
H - H M COS ~o (12.4)
COS ~ f = neff
484 A.K. ZVEZDIN

e.a.

Fig. 12.1. Orientation H , 37/d and -Oeff relative to the easy axis (e.a.) in the basal plane of a hexagonal
magnet.

Minimizing (12.2) with allowance for eqs (12.3) and (12.4) we can now determine
the equilibrium phases in the ranges in which they exist. However, we shall simplify
eq. (12.2) further. Expanding (12.3) and (12.4) in powers of H/HM we get

H
~ f = 71" ÷ ~ ÷ HM sin~p. (12.5)

The non-collinearity of the d and f sublattices depends on the angle ~b = qof - 71 - ~9


(see fig. 12.1). When H/HM << 1, z9 << min(~f, cp), so in the last term we can
substitute ~f = 7r + ~ and then

~5(~p) = - M d H cos T - fo n ~ Mf(x) dx - (1/6)K cos 6(~p + a), (12.6)

where K = Kd + Kf is constant reflecting the basis anisotropy of the crystal. This is


a reasonable approach when K << )~M~ because in this system the part of the H - T
phase diagram which is essentially dependent on anisotropy is concentrated in the
range H << HM (see below).
If necessary, the equilibrium values of 9~ and ~f can be found in the range H << HM
by the method of successive approximation with the small parameter

Here we have actually used the lowest approximation of this parameter. In RCo5
one has ),Md2 ~ 10 9 erg/cm 3 and if K ~ 10 5 erg/cm 3, then ~ ~ 10 -2.
The minimization of eq. (12.6) thus defines ~(H,T), and eq. (12.5) defines
qof(H, T) and the difference in sublattice orientations ~b. The minimum condition for
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 485

• (~) is

- - = M o l l sin~(1 - A~(Heff)) + K sin6(~ + a) = 0, (12.7)


0~
~}2~
= MdHcOs ~(1 - A~(Heff)) + 6 K cos 6(~ + a ) -

(12.8)
H2 H 2 d~
sin ~ >/0.
~ff d~ff

Here

Mr(Neff)
~ ( g e f f ) -- _ _ (12.9)
neff

12.2. Phase diagrams


12.2.1. Case 1: HIIEA (easy axis)
The analysis of (12.7) and (12.8) for a = mr/3, n = 0, 1,2 . . . . . reveals three phases:
two collinear phases A and B, and one canted phase C (fig. 12.2). The stability

H1 -- ~ / J

0 To T
Fig. 12.2. Magnetic phase diagram of a hexagonal ferrimagnet with a compensation point; H is parallel
with the easy axis in the basal plane of the crystal. Letters in boxes indicate phases of the magnet. The
curves A A t, B B t etc. indicate where the stability of the corresponding is lost. K and K ~ are tricritical
points, (Tc, H1) is a critical point; after Zvezdin and Popkov (1980).
486 A.K. ZVEZDIN

curves of the collinear phases A (~ --= 0) and B (~ = 70 are described by

1 -- AX(HM q: H) -4-6K/MdH = O.
The stability curves of the various phases have been indicated in the figure by AA p
and BB' (see fig. 12.2). These lines cross at a point with coordinates

Tl = Tc,
( 6K ~1/2 ( 6KA ) 1/2, (12.10)
nl = H---~X~j ~ (1SA--Xf
where
aMf
Xf= ~H

We shall find the canted phases by expanding (7) over H/HM. It is then described
in the form (0 < ~ < 70

~HMd H2(1 - AXe) 1


cosqo = - cosqo(4cos2 ~ - 1 ) ( 3 - 4cos3 ~) (12.11)
6K 6KA 3
in which

r / = 1 - AXo; )~o = x ( H = 0).

This equation is best investigated graphically (fig. 12.3). The abscissa of the
points of intersection of the lines

fl(x) = (1/3)x2(4x 2 - 1)(3 - 4x 2)

~flxl
+1

-11\ / ' - ' \ o " . . . (


\ /i \j X

-1
Fig. 12.3. Diagram explaining the solution of eq. (12.11).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 487

a) cos ~o [] b) cos 9 []

J Im

[] []

[] []
Fig. 12.4. Temperature dependence of cos ~,(T) for H parallel with the easy axis: a) - H > //3; b) -
H1 < H < / / 3 ; c) - H2 < H < H1; d) - H < HI. Letters in boxes indicate the phases corresponding
to the diagram in fig. 12.2; after Zvezdin and Popkov (1980).

and

HMdrl H2( 1 -- )~Xf)


f2(z) -- x,
6K 6KA

where z = c o s ~ is the solution of eq. (12.11) for a given H and T. The slope
of f2(z) and the points at which it crosses the axis z vary with the H and T. For
instance, if T = const and the field H is increased from H = 0 we obtain the
following: five intersection points, of which three satisfy the stability conditions;
then three points two of which are stable; and finally one point. Some solutions of
c o s ~ ( H , t ) are shown in the diagrams of fig. 12.4. The canted phases are twice
degenerate. Equation (12.14) obviously has two energetically equivalent solutions:

0 < ¢Pl(H,T) < zr and ~2(H,T) = 27r- ~I(H,T).

This degeneracy is a consequence of the symmetry, because the easy axis is a


second order axis. It is a characteristic feature of canted structures. Physically, it
is manifested by the fact that the crystal breaks up into domains on changing to the
canted phase.
In sufficiently large fields the transitions A - C and B - C are second order phase
transitions. The expansion of the thermodynamic potential in power of 3qo = p - ~0
where ~0 = 0 near the curves A K and B K ' is actually given by

1 1
~b ~__ ~ 0 q- a,(~gg) 2 -I- b(aT) 4 + .... (12.12)
488 A.K. ZVEZDIN
in which

a = H M d [1 - Axf(Heff)] sign cos T0 + 6K;

3H2H 2
b_ - - x}(Heff) - 216K - HMa(1 -/~xf(Heff ) sign cos ~0,
H ff
where signx = z / l z [. a ( H , T ) = 0 on the curves A A ' and B B ' (see (12.9)). On
the segments A K and B K ' of these curves b(H, T) > 0. According to the theory of
Landau this means that those are the second-order phase transitions.
The points K and K ' are tricritical. Their T, H coordinates are found from the
equations

a(H,T)=O; b(H,T)=O

The approximate solution of these equations is

70KA )1/2
(12.13)

( K \ 1/2 )
7 = Tc 1 + (1 - xf)) sign cos o . (12.14)

The ranges of collinear and canted phases overlap when H < //3. So those are
first-order phase transitions they occur on the curves of K T and K T ' which depend
on the conditions of equality of the thermodynamic potential of the boundary phases.
The curves of K C and I£~C ~ bound the range of metastability of the canted phases.
These lines are described by (leaving out ~p(H, T))

d~ d2~i
-- = 0; = 0. (12.15)
d~ d~ 2

When H < H1 a further ambiguity appears near ~ = 7r/2, 37r/2 on the curve
cos ~(H, T) (see fig. 12.4 b), d)) with corresponding jumps in ~. This means that
the canted phase C is split into two different canted phases C' and C". In phase C~:
0 < ~ < 7r/2 and 37r/2 < ~ < 27r, but in C": 7r/2 < ~ < 7r and 7r < ~ < 37r/2.
A first-order transition C ~ to C" occurs on curve of TcO in fig. 12.4 a). The
equation of the curve TcO: F(~c,) = F(~c,,). Curves O M and O M ~ in fig. 12.2
are the of stability curves of phases C and C"; they are defined by (12.19). The
point O in fig. 12.4 is a critical point of the vapor-liquid type.
When H < H2 (see fig. 12.2) the metastable ranges of the collinear phases so
greatly increase that a direct transition can occur between them (see fig. 12.4 d)).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 489

We can derive one more asymptote of the first-order phase transition curves where
H --+ O. The lines KTc and K'Tc:

T-To 3 H
Tc 4 AMd"

Line OTc

T - Tc 3
T~ 8

12.2.2. Case 2: KrIIHA (HA is the axis of hard magnetization (a = re~6 + nre/3;
n = O, 1,2 .... ))
Assuming that ~ = re/6 + nre/4 in the thermodynamic potential, we can see that the
only difference in comparison with that examined above (for H parallel to with the
easy axis) is the sign of the anisotropy constant K. The phase diagram is shown
in fig. 12.5. There are three basic phases, two collinear phases, A{~ = 0} and
B { ~ = re}, and one canted phase, C. The stability curves A(AA') and B ( B B ' )
are described by eq. (9) in which K must be substituted for - K . Here we have
second-order phase transitions curves because the coefficient b in the expansion of
(12.12) is positive everywhere.
In the canted phase the angle is defined by eq. (12.11) in which - K has been
substituted for K . The dependence in the canted phase is unambiguous if H >
H~ but ambiguous if H < H~', as we can see from the graphical method (see

i"/ "i
r,,..m

HI*

H:*
P : ~---" ~ 0 VP'~ -- ~ - 0 ' _
ro r

Fig. 12.5. Magnetic phase diagram of a hexagonal ferrimagnet with H parallel to the hard axis; after
Zvezdin and Popkov (1980).
490 A.K. ZVEZDIN

fig. 12.3). When H < H~, the first-order phase transitions C'-C" and C " - C ' " may
occur along MTc and M'Tc. The coordinates of the critical points M and M ' are
approximately

7 ( 6K,~ ~1/2
H = H i ~ ~ 1 - )~Xf,] '

T=Tc
{ 21
1 :t: ~-d
2K
(1 -- AXf)
],2} •

The equations of the lines MTc and M'Tc have the form of

T - Tc v/3 H
- - - + - - (1 - Axf).
Tc 4 tMd

MO, MO' and MP, M P ~ are the curves along which the stability of the adjacent
canted phases C ~, C " and C " is broken. They are given analytically by a system of
the (12.15) type.
The ranges of coexistence of the phases C" and C m overlap considerably in very
weak fields H < H~. In the canted phase, there may be direct phase transition when
the temperature changes avoiding the intermediate canted phase C", as it is shown
in fig. 12.6 c).
The anisotropy constants in rare-earth materials usually increase rapidly with tem-
perature and magnetic field (where T < Tc). Here we have assumed that K -- const.
This is sufficient for determining qualitatively the behavior of hexagonal ferrimagnets
in a magnetic field near the compensation point, including the number and sequence
of phases, topology of phase diagrams, order of magnitude of critical fields in the
region of the compensation point, type of phase transition.
Without affecting the qualitative pattern of phase transitions near Tc, a sudden
change in the anisotropy energy with temperature will make the phase diagram

a) cos ~o b) c) cos ~o
[] ,, [] L__ []
1

[] [] []

Fig. 12.6. Plots of cos ~(T), where H is parallel with the hard axis: a) - H > H~ ; b) - H~ < H < H~ ;
c) - H < H~. Letters in boxes indicate the phases corresponding to the diagram in fig. 12.5; after
Zvezdin and Popkov (1980).
FIELD INDUCEDPHASETRANSITIONSIN FERRIMAGNETS 491

highly asymmetrical, relative to the compensation point. In particular, the critical


points M in fig. 12.5 and K in fig. 12.2 will be displaced in downward direction on
the temperature scale. Preliminary evaluations show that the shift may be so great
that the transformation to a high-field canted phase can no longer be achieved by a
second-order phase transition.

13. Non-collinear magnetic structures in the intermetallie compounds DyCo5

13.1. Crystal and magnetic structure


Suitable objects for experimental investigation of field induced non-collinear mag-
netic structures in hexagonal ferrimagnets are the intermetallic compounds DyCo5
and TbCos, which possess anisotropy of the 'easy plane' type at below room tem-
perature.
The influence of hexagonal anisotropy on the occurrence of non-collinear mag-
netic structures, induced by an external magnetic field, has been investigated in the
intermetallic compound DyCo5 near the compensation temperature (To = 124 K) by
Berezin et al. (1980). Let us consider the main results of this study.
The compound DyCo5 possesses hexagonal structure of the CaCu5 type (space
group P6/mmm), in which some of the Dy atoms have been replaced by pairs ('dumb-
bells') of Co atoms (Dworschak and Khan, 1974). The cobalt excess in this case
made it possible to obtain a single-phase compound, whereas the alloy of the integral
composition DyCo5 contained the Dy2Co7 phase (Velge and Buschow 1968).
In a structure of the CaCu5 type the Co atoms occupy two non-equivalent positions
with, according to neutron-diffraction data, slightly different values of the magnetic
moments (Ermolenko et al. 1975). We shall, however, neglect this difference and
treat DyCos.3 as a ferrimagnet that consists of two (Dy and Co) magnetic sublattices.
Characteristic of all RCo5 compounds are the huge values of the uniaxial anisotropy
constants, often comparable with the R-Co exchange interaction, which may be of
different signs (Ermolenko et al. 1975, Druzhinin et al. 1977). In DyCos.3, the neg-
ative anisotropy of the Dy sublattice leads to the result that at temperatures below
320 K, the magnetic moments of the Dy sublattices lie in the basal plane, despite
the fact that for the Co sublattice the favored orientation of the spins is along the
hexagonal axis. The resultant anisotropy at low temperatures according to Berezin et
al. (1980), reaches 108 erg/cm 3. This prevents departure of the magnetic moments
of the sublattices from the basal plane when the crystal is magnetized along any
direction in this plane.

13.2. Domains in the basal plane


The high degree of symmetry in the case of hexagonal anisotropy results in the
existence of three equivalent axes of easy magnetization in the basal plane. Therefore
in the demagnetized state, there can exist in the specimen six types of domains,
oriented at the angle of 60 ° with one another, which must be taken into account in
the analysis of the experimental data.
492 A.K. ZVEZDIN

The presence of several easy axes and hard directions in the basal leads also to
the result that in the non-collinear phase, during rotation of the magnetic moments
of the sublattices in this plane, they should lag at the easy directions and pass more
rapidly through the hard directions. Then wave-like singularities appear in the field-
dependencies of the magnetization and magnetostriction.
Near the compensation temperature, the magnetic moments of the sublattices may
pass through the hard direction discontinuously; that is, a phase transition of first
order occurs. For an external field applied along an easy direction, a jump of the
magnetic moments may be observed on transition from the collinear phase to the
non-collinear (then the moments of the sublattices are 'thrown over' to the nearest
easy axes, oriented at an angle of 60 ° with respect to that along which the specimen
was magnetized). Along a hard direction, transitions of the first order are possible
only in the non-collinear phase (from the easy axes oriented at 30 ° to the field
direction to the easy axis perpendicular to the field). The diagrams in the upper part
of fig. 13.1 may explain this.

tH~~,May b

Mc~~~,- a
era, G.cm3/g %, G-cm3/g

20 20 ~ 1
1
10 I0 2

0
5 5
0
5 5
0 0 5
5
0 0
10 J 10
5 5
0 I I I 0
60 120 180 60 120 180
H, kOe H, kOe
a) b)
Fig. 13.1. Experimental field dependencies of the magnetization of DyCo5. 3 near the compensation
temperature, a) field directed along a-axis: 1 - 86 K; 2 - 105 K; 3 - 120 K; 4 - 152 K; 5 - 163 K; b)
field directed along g: 1 - 78 K; 2 - 97 K; 3 - 115 K; 4 - 157 K; 5 - 166 K. The dotted lines show
magnetization curves without allowance for the occurrence of a non-collinear magnetic structure; after
Berezin et al. (1980).
FIELD INDUCEDPHASETRANSITIONSIN FERRIMAGNETS 493

13.3. Magnetization and magnetostriction


The lower part of fig. 13.1 shows the experimental variation of magnetization with
the intensity of the external field, for temperatures above and below the compensation
point, in the fields directed along an easy axis a) and along a hard axis b) in the
basal plane. It is seen from the figure that after technical saturation of the specimen,
there is linear increase of magnetization, caused primarily by the paraprocess in the
Dy sublattice (curves 1, 2 , 4 and 5 in fig. 13.1).
With increasing field, the character of the curves changes: magnetization starts to
increase considerably more rapidly on the whole; one can see alternating sections of
rapid and slow increases of magnetization ('waves').
At temperatures close to the compensation temperature (curves 3 in fig. 13.1)
considerable increase of magnetization occurs over a whole field interval. It is also
seen that the determination of the critical transition field is subjected to a large
error.
In order to confirm independently the presence of rotation of the magnetic moments
of the sublattices in the basal plane, and in order to determine the critical field
more accurately, measurements were made of the magnetostriction along the a- and
b-axes (in the latter case, the field was directed perpendicular to the measurement
direction).
Characteristic field dependencies of the )~aa and )~ab are shown in fig. 13.2. It
is seen from the figure that singularities due to the occurrence of the noncollinear
phase and under the influence of hexagonal anisotropy show up considerably more
clearly in the field-dependence curves of magnetostriction than in those of magneti-
zation.
Consider the curves at 4.2 K in an increasing magnetic field (curves 1 in fig. 13.2;
the direction of the field is shown by an arrow). In weak fields (up to 25 kOe), the
change of magnetostriction along the axis a is caused by domain processes; along
axis b by both the displacement of boundaries and rotation of the magnetic moments
along the hard axis. On further increase of the field (in the non collinear phase),
saturation is observed for both directions (the slight linear drop of magnetostriction
is due, as was shown by measurements that we made on YCos, primarily to isotropic
magnetostriction of the paraprocess in the Co sublattice).
In fields exceeding a certain critical value Hal, there is an abrupt change of char-
acter of the field dependence of magnetostriction (along the b-axis, for example, it
changes sign), resulting from appreciable departure of the magnetic moments of the
sublattices from the direction of the field.
Along the b-axis in the non-collinear phase (H > H1) there can be also seen
wavelike anomalies, similar to those observed on the field-dependence curves of
magnetization.
With increase of temperature, the critical field of the transition decreases; near
the magnetic compensation temperature, no saturation is observed (curves 2 in fig.
13.2). On further increase of the temperature, a transition from the collinear phase to
the non collinear is detected again; the wavelike anomalies become less pronounced
(curve 3 in fig. 13.2). At high temperature, a noticeable role is played by isotropic
494 A.K. ZVEZDIN

~au~a 1 T I I

H~'r I
[
[
I
I
I I

U i
240 v 7, ° , ,
H, kOe

a) b)
Fig. 13.2. Experimental field dependencies of the longitudinal magnetostriction )~a and transversal
magnetostriction A~b of DyCos. 3 in a field applied along axes a and b respectively in the basal plane:
1 - 4.2 K; 2 - 124 K; 3 - 162 K. The arrows show the direction of change of the external rid& H~r
is the critical field for transition to the noncollinear phase during increase, Hc~ - during decrease of the
external field; after Berezin et al. (1980).

magnetostriction of the paraprocess; this leads to a linear increase of magnetostriction


with increase of field in the collinear phase (curve 3, fig. 13.2).
The magnetostriction measurements revealed still another peculiarity that was
scarcely noticeable on the field-dependence curves of magnetization; namely, appre-
ciable hysteresis, that is not coincidence of the values of the strain during increase
and subsequent decrease of the external field. Correspondingly, the critical fields of
the transition were strongly dependent on the past history of the magnetic state of
the specimen. The hysteresis phenomena decrease with the rise of temperature and
disappear completely at temperatures above 170 K.
The variation of the critical transition fields and of the anisotropy fields (the H - T
diagram) is shown in fig. 13.3, where open circles show the critical fields during
increase, filled circles - during decrease of the external field. It is seen from the
figure that near the compensation temperature (Tc), the collinear phase is absent for
both directions; the critical fields increase with the distance from the compensation
point. For the field dependencies of magnetostriction and at 4.2, 124 and 162 K,
and also for the magnetic phase diagrams along the a- and b-axes, the corresponding
theoretical relations were plotted.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 495

H, kOe I I I [

%,,,,>0,/
nlla /
180

120

60

irol r ,
50 100 150 200 T K
H, kOe ~ i i

180 Hll b A

120600 D~I CD~I~o,

1 K
Fig. 13.3. Experimentaland theoretical phase diagrams of DyCos.3 in a field directed along the easy
axis a and along the hard axis b in the basal plane: (o) - experimental critical field (H~r) and anisotropy
field during increase of the external field; (e) - critical field (Hc"r)during decrease of the external field;
(zx) - the critical fields coincide. The heavy lines are experimentalphase diagrams; the thin solid lines
are theoretical lines of phase transitions of the second kind; the dotted lines are theoretical stability
curves of the collinearphase, the dashed-dottedcurves are those of the non-collinear(high-field)phase;
after Berezin et al. (1980).

13.4. Magnetizations of the sublattices, exchange field and anisotropy constant


The thermodynamic potential (12.1) and equations of critical fields (12.9) and (12.11 )
were used to calculate the phase diagrams. The parameters that occur in eq. (12.1)
were specified as follows. Values of the magnetic moment of the Co sublattice as a
temperature function were taken from the paper of Nowik and Wernick (1965); with
allowance for the fact that our specimens had a slightly different structural formula,
the value of magnetization M was taken Moo = 8.2# B at 4.2 K. The magnetic
moment of the Dy sublattice was found as the sum (below the compensation point)
or difference (above the compensation point) of the spontaneous moment and the
moment of the Co sublattice.
As it has already been mentioned, the value of the intersublattice exchange in-
teraction in an isotropic ferrimagnet can be found from the measurements of the
496 A.K. ZVEZDIN

susceptibility in the non-collinear phase.


In the presence of hexagonal anisotropy, as is easy to show that the susceptibility
in the non-collinear phase at the compensation temperature should be 1/A in a field
applied along either the a- or the b-axis, if the angle of 'bending under' of the
magnetic moments of the sublattices is not too large.
Actually, it follows from the measurements of magnetization at 124 K that the
susceptibility in both cases, within the limits of accuracy of the experiment, is the
same and equal to (0.92-t-0.07)x 10 -3 hence HM = (9004-50) kOe. This value agrees
well with the value HM = 900 kOe obtained by Ermolenko et al. (1975) on the basis
of the analysis of the temperature variation of spontaneous magnetization of DyCo5
in the spin-reorientation range. This value disagrees with the value HM = 1570 kOe
determined from Mrssbauer measurements (Nowik and Wernick 1965).
The anisotropy constant K6 was determined at low temperatures from measure-
ments of the anisotropy field when the crystal was magnetized along the hard direc-
tion (/£6 = (42 -4- 7) x 104 erg/cm3 at 4.2 K), and at high temperatures from torque
curves in the basal plane.
On the basis of experimental magnetization data, allowance was also made for the
dependence of the values of the magnetic moments on the external field. Using the
formula for the value of magnetostriction of a hexagonal crystal in the basal plane,
Aaa = 1/2A'~'2(T) cos 2qODy,

"~ab = --'~aa
(here A"Y,2is the magnetostriction constant responsible for the strains of the crystal
in the basal plane) one can plot the theoretical Aaa(H) and Aab(H) relations. It is
assumed here that at low temperatures magnetostriction is caused principally by the
Dy sublattice (see below).
The angle FCo was found from the relation (12.3)
HM sin ~Co
sin ~Oy -- Heff

The theoretical field dependence curves of magnetostriction are given in fig. 13.4.
In the calculations it was assumed that in the initial state, the specimen is demag-
netized and the magnetic moments are uniformly distributed among the three easy
axes in the basal plane; the isotropic paraprocess magnetostriction was also taken
into account. In fig. 13.4 phase transitions of the first order are shown dotted (with
allowance for the maximum possible hysteresis); increase and decrease of the field
are denoted by arrows.
It is evident from a comparison of fig. 13.2 and fig. 13.4 that the theory describes
a number of peculiarities that are observed on the experimental curves, such as the
abrupt change of the field dependence of magnetostriction on transition to the non-
collinear phase and the change of sign of magnetostriction. On the theoretical curve
there are wavelike sections (curve 1, fig. 13.4 b)) coinciding qualitatively with the
experimental ones. The difference between theory and experiment takes place chiefly
due to the presence of appreciable hysteresis in the experiment relations and will be
discussed below.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 497

~ab

/
I I

f
I
iI
t,
II
I
0
1

? I I

l~ 6' 170 1 0 ---''- H, kOe


~ 3

50 0 240
H, kOe

a) b)

Fig. 13.4. Theoretical field dependence of the magnetostriction; the magnetostriction behavior shown is
that for the maximum possible hysteresis. The notation is the same as in fig. 13.2; after Berezin et al.
(1980).

13.5. H - T phase diagram


Theoretical magnetic phase diagrams for DyCos, obtained by using formulas (12.7)
and (12.8), are shown in fig. 13.3. The thin solid lines correspond to phase transitions
of the second order; the dotted lines show the critical fields during increase, the
dashed-dotted ones - during decrease of external field for phase transitions of the
first order. Transitions of the first order that occur in the non-collinear phase in a
field applied along the hard direction as well as those that occur during technical
magnetization of the specimen, are not shown.
It is seen from fig. 13.3 that there is a partial agreement of the theoretical and
experimental phase diagrams. For a field applied along the a-axis, the phase transition
of the first order predicted by theory observed at low temperatures. For a field
applied along the b-axis, both theoretical and experimental diagrams have a 'throat'
form, characteristic of hard directions. Agreement of the experimental data with the
theoretical shows up also in the fact that for each fixed temperature the critical fields
of transition those along the a-axis are larger than those along the b-axis.
498 A.K. ZVEZDIN

The greatest discrepancies between the theoretical and experimental phase dia-
grams occur at low temperatures and near the compensation temperature. Without
discussing the quantitative differences, which to some degree can be explained by
inaccuracy in the choice of the parameters in calculations, we shall consider the
qualitative difference between the theoretical and experimental results. First, appre-
ciable hysteresis is observed experimentally in a field directed along the hard axis,
whereas theoretically along this direction the transition from the collinear phase to
the non-collinear one should be of the second order, and the occurrence of hysteresis
is in principle impossible. Second, along the easy direction the collinear phase is
absent within the interval of 85-135 K; according to the theory, it is unobservable
within a considerably narrower temperature interval (110-135 K). To the qualitative
differences we can be add the fact that the experimentally observed hysteresis along
the a-axis exceeds the calculated hysteresis by an order of magnitude (curve 1 in
figs 13.2 a) and 13.4 a)).
It is possible that the presence of such hysteresis can be explained by taking into
account the dependence of the hexagonal anisotropy constant on the external field.
According to the single-ion theory at low temperatures, on increase of the field the
low-field non-collinear structure transforms directly to the high-field structure. Then
on the field-dependent curves of magnetization and magnetostriction approximately
horizontal sections can occur, imitating the phenomenon of saturation. If it is so, then
the above-stated value of the anisotropy constant K is too low (by more than an order
of magnitude), and the observed hysteresis phenomena are actually connected with
phase transitions within the uncollinear phase. This assumption would also explain
the appreciable hysteresis in a field directed along the a-axis. A final explanation of
this problem requires additional experimental data.
The presence in DyCo5 of hexagonal anisotropy considerably exceeding that mea-
sured by Berezin et al. (1980) would lead to a substantial broadening of the range
of metastability near the compensation temperature. In this case, a collinear struc-
ture may not be observed in a field directed along the a-axis, for the following
reason.
In the absence of an external field, as has it already been mentioned, several
types of domains exist in a hexagonal crystal. If it turns out that the fields for
displacement of domain boundaries exceed the critical field for stability loss of the
collinear phase. Then the transition to the high-field non-collinear phase occurs
from a multi-domain (nonuniform) state, avoiding the uniform collinear phase. Such
situation is especially probable near the compensation temperature, where, because
of the weak interaction of the magnetic moments of the domains with the external
field, the field for displacement of domain boundaries may be large.
There is also another reason explaining the absence of a collinear phase along the
a-axis. In a real crystal there may be homogeneous regions (grains) with somewhat
different values of the magnetic moment per elementary cell; that is, with different
compensation points. Such possibility exists because of the replacement, which may
not be completely statistical, of dysprosium atoms by cobalt. Then what is observed
experimentally will be a mean value of magnetostriction produced by asynchronous
rotation of the sublattices in each such grain. The phase transition may then turn out
FIELD INDUCEDPHASETRANSITIONSIN FERRIMAGNETS 499

to be so 'smeared out' that determination of the critical field becomes impossible.


Along the a-axis this will lead to an apparent absence of a collinear phase near the
compensation temperature; along the b-axis to an increase of the range of existence
of the non-collinear phase.
The presence of grains with different magnetic moments in the crystal may also
explain other anomalies. For example the fact that the spontaneous magnetization
does not vanish exactly at the compensation point (the 'residual' magnetization is
approximately 10 G) - a result already noted by Ermolenko et al. (1975), Nowik and
Wernik (1965) - and the unusual behavior of the field dependence of magnetostriction
at Tc in a field applied along the b-axis (curve in fig. 12.2 b)). The increase
of magnetostriction observed here experimentally indicates that at least some of
the magnetic moments rotate toward the field direction, whereas theoretically there
should occur a change-over of domains forming an angle of 60 ° with the field
direction to domains oriented perpendicularly to the field (curve 5 in fig. 13.4 b)).

14. Spin-flop and spin-reorientation phase transitions in the anisotropic


ferrimagnet HoCo3Ni2 with Tcomp : ~SR1

14.1. Spin reorientation in HoCo3Niz


The aim of this section is to describe the field-induced phase transition in the
ferrimagnetic compound HoCo3Ni2. Ferrimagnet HoCo3Ni2 is a unique member
of the hard magnet family of RCos-type compounds. It is characterized by the
compensation of magnetic moments of 4f and 3d sublattices occurring at T = 160 K.
This temperature marks also the start of the process of spontaneous change of easy
axis direction (spin-reorientation) from the basal plane (a-axis) to the hexagonal
axis (c-axis) in the temperature interval from TSR 1 = 160 K up to TsR2 = 200 K
(Drzazga 1981, Drzazga and Drzazga 1987). This coincidence, i.e. (Teomp ~ TSR1),
is the reason why the critical fields are very low (Zawadzki et al. 1993). This fact
leads to a rich variety of possible magnetic phase transformations induced both by
temperature changes and by magnetic fields applied along the basic crystallographic
axes. It allowed us to verify experimentally the main theoretical results.
Ferrimagnet HoCo3Ni2 crystallizes into the hexagonal CaCus-type structure. It is
well established that in compounds of this type the exchange and crystal-field inter-
action do not differ very much in magnitude and that the SR transition is governed
mainly by two (K1 and K2) anisotropy constants (Asti 1990). Therefore, to obtain
magnetic phase diagrams of HoCo3Ni2, it was necessary to take into account the ex-
change interaction as well as second and fourth-order contributions of the anisotropy
energy.
The anistropy energy has been expressed as

/ ( = t(1 sin 2 0 + / ( 2 sin 4 0

where /(I(T) changes the sign at T = Tsm (/(a > 0 for T > Tsm) and /£2 > 0.
This is a typical temperature dependence of the magnetic anisotropy energy for the
occurrences of spin-reorientation processes (Belov et al. 1979).
500 A.K. ZVEZDIN

The thermodynamic potential (3.1), after taking into account this magnetic aniso-
tropy energy has been minimized with respect to the polar 0 and azimuthal qa angles
of the 3d sublattice (Krynetski et al. 1993, Zawadzki et al. 1994). The exchange
constant, the temperature dependencies of both sublattice magnetization as well as
the anisotropy constants have been taken from Zawadzki et al. (1993).

14.2. The H - T phase diagrams


The obtained magnetic phase diagrams are presented in figs 14.1 and 14.2. The
magnetic phases are labeled as follows: FO - ferromagnetic, FI - ferrimagnetic,
C - canted. The numbers following these labels denote possible sub-phases. The
subscripts [] and _k indicate the direction of magnetic field in relation to c-axis. The
subscript 0 denotes H = 0.

14.2.1. Case 1: Magnetic field perpendicular to the c-axis


1) Low temperature range (T < TSR1): For T < TSR1 the magnetic moments of both
sublattices are lying in the basal plane (0 = 7r/2). The anisotropy energy is constant
which results in the-well known case of isotropic ferrimagnet. Three magnetic phases
exist now ferrimagnetic FII± (or FI10), canted CI± and ferrimagnetic FO±. They
are separated by phase boundaries described as follows

Hcl,2 = ,~(mf(Heff) m md).

H, 10xkOe 1 I" ] T T

100

FO±

CI±
50

lOO 200 300


T,K
Fig. 14.1. The magnetic phase diagram for the field parallel to the basal plane. The meaning of the
symbols can be found in the main text in section 14.2 (after Zawadzki et al. 1994).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 501

H, 10xk0e i J t J J

100

FOIl

50
Fllll

Cll~ C20~/] FI2I,


I B I ~ I
0 100 200 300
r,K
Fig. 14.2. The magnetic phase diagram for the field parallel to the hexagonal c-axis. The meaning of
the symbols can be found in the main text in section 14.2 (after Zawadzki et al. 1994).

The signs ' - ' and '+' correspond to the low and high-field limits of the canted
phase region, respectively.
2) High temperature range (T > TSR2): In this case three phases are stable, too. The
low field stability region of the collinear ferrimagnetic phase is reduced to the line
H = 0 (FI20). For magnetic fields lower than He, where

2(K1 + K2)
He=
Md(2 - X(Heff)A)

canting of ferrimagnetic structure is induced, and the magnetic moments are bent
away from the c-axis creating the C2± phase. The angle of deflection increases with
the field, and finally, in He the magnetic moments are oriented perpendicularly to the
c-axis. In fields higher than Hc the canting vanishes and the collinear phase appears
again. Further increase of the field reduces the net moment of the Ho-sublattice. It
reaches zero for H = AMa, and then grows along with the field. As a result, the
high-field collinear phase is divided into two sub-phases: ferrimagnetic FI3± and
ferromagnetic FOx.
3) SR transition temperature range (TsR1 < T < TSR2): The noncollinear magnetic
structure (C20) is stable in the SR range even at H = 0. This weakly canted phase
appears as a result of competition between anisotropy and exchange energies was
investigated both theoretically and experimentally (Decrop et al. 1982, Irkhin and
Rozenfeld 1974). In non-zero magnetic fields, apart from the above described C2±
502 A.K. ZVEZDIN

and FOx phases, a new canted phase C3± appears in the SR range. Such cone-canted
phase is allowed if the condition

I~1
sine 0 -
2K2
is fulfilled. This phase is characterized by constant susceptibility X(Heff) = 1/A.
The region of stability of the phase is limited at high temperature by the following
critical field
) 1/2
M2 cos 2 0
Hcl,2 = AMd sin0 :t: AMf 1 Mff(H~ff)

The low temperature limit of the C3± phase is T = TSR1. The properties of the
new cone-canted phase C3± resemble the properties of CI± in the isotropic case.
Nevertheless, due to the conical character of anisotropy, magnetic moments of the
3d and 4f sublattices during the canting progress are lying on the easy cone, not
in the basal plane. Comparing the magnetic structures of the cone-canted phases
C2± and C3± it is worth noting that in both phases, magnetic moments of the
3d and 4f sublattices are not collinear (~rf /]]Md). However, their components
perpendicular to the e-axis remain collinear (.~rf,±ll~d,±)
in C2±, whereas in C3±
they are noncollinear (~rf,± is not parallel with ~rd,±). Moreover, the susceptibility
of C2± phase, contrary to that of C3±, is constant.

14.2.2. Case 2: Magnetic field parallel to the c-axis


In this case the magnetic phase diagram is simpler. Phase boundaries are given by

Mf(H~ff))
MdH 1 - A Heft T 2K1 O.

The curve corresponding to the ' - ' sign surrounds the stability region of collinear
ferrimagnetic phase (FIIlI). In the region bounded by both these lines exists a
canted phase. This region is divided into two parts, the low-field part (ClU,), where
rotation of the weakly canted structure from the easy plane (easy cone) to the c-axis
takes place, and the high-field part (Clnll), where a change of magnetic order from
ferrimagnetic to ferromagnetic is induced. It can be seen in fig. 14.2, that the link
between low-field (ClUll) and high-field (c1H") regions of the canted phase is very
narrow.

14.3. The magnetization measurements


In order to verify the existence of the newly proposed phase $3± the magnetization
measurements have been carried out on high quality single crystals using both SQUID
(up to 5 T) and ballistic (up to 7 T) magnetometer.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 503

1.0 I I I I
. , I
I I

0.8 T= 170 K /

0.6

0.4

0.2 W~,~,~, ;, . ;, ~,2.~.~-0

I I [ I I I
l 2 3 4 5 6 7
H, 10×kOe
Fig. 14.3. Longitudinal (1) and transversal (2) components of magnetization at 170 K as function of
magnetic field applied parallel to the a-axis (after Zawadzki et al. 1994).

Longitudinal and transverse magnetization components have been measured si-


multaneously (Zawadzki et al. 1993, 1994). It allowed us to examine precisely the
magnetic structure evolution during the magnetization process.
The application of a ballistic magnetometer allowed us to observe phase transitions
occurring in higher magnetic fields.
A good agreement between calculated and experimental magnetic phase diagrams
has been found in the attainable field range. The transition from C2± to C3± can
be deduced from the field dependence of the longitudinal Ma and transverse Mc
magnetization components, in the case of the field applied along the a-axis. An
example of such a field dependence measured at 170 K in presented in fig. 14.3.
The linear relation Ma(H) = ~ - I H for fields higher than the critical field (Hc ~ 4
T) is similar to that often reported for field-induced phase transitions in easy-plane
in ferrimagnetic intermetallics (Ballou et al. 1989). Moreover, the relation Mc(H)
supplies still more information on the magnetization process. The beginning of the
phase transition is now more distinct. Besides, it is easy to notice that canting of
magnetic structure takes place on the easy cone, not on the basal plane (C2± -+ C3±
transition). Otherwise, the transverse component Mc should vanish above Hc. . . . .
• ,2-

14.4. Magnetostriction and spin-flop transitions ~;


To obtain further information on magnetic phase diagrams of HoCo3Ni2 Krynetski
et al. (1994, 1995) have measured its linear magnetostriction since it is known
(Belov et al. 1979) that magnetostriction is a more sensitive tool for magnetic
504 A.K. ZVEZDIN

phase diagram study than magnetization. Furthermore, the study of magnetoelastic


properties of intermetallic ferrimagnet HoCo3Ni2 is a very interesting problem by
itself, particularly when the role of rare-earth sublattices in magnetostriction of 3d-4f
compounds is concerned.

14.4.1. Magnetic field parallel to the a-axb


Spin-reorientation temperature range (160 < T < 200 K). In this temperature range
the non-collinear magnetic structure caused by the competition between anisotropy
and 4f-3d-exchange energies is stable even at H = 0. Therefore, the transformation
of this phase induced by the external field has a very complex character. In fig. 14.4
the longitudinal magnetostriction )~a~ of HoCo3Ni2 at T -- 162.5 K is shown as
an example. The low-field step-like anomaly at about 0.3 T is due to the domain
process. The high-field anomaly on this curve results from the spin-flop transition.
The value of the critical field agrees with magnetization data (see insert to fig. 14.4).
It is seen from fig. 14.4 that magnetostriction measurements allow to determine the
values of the critical fields of induced phase transitions really more precise. The
additional mid-field anomaly seems to be due to the fact that at this temperature the
primary direction of the weakly canted 4f and 3d magnetic moments does not coincide

Zaa(10-5)

HoCo2Ni 3 T= 162.5 K

f M(10_l~
" T= 170K

Bc5
0 1 2 3 4 5 6 Bcr
2 I I ,F I
0 1 2 3 4 5
B(T)
Fig. 14.4. The longitudinalmagnetostriction)~aa of HoCo3Ni2 at temperature T = 162.5 K. The
insert shows the magnetization of HoCo3Ni2 along a-axis versus magnetic field parallel to a-axis
at temperatureT = 170 K.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 505

with the a-axis due to the spin-reorientation process. It must be emphasized that
transformation of magnetic structure of HoCo3Ni2 caused by a spin-flop transition
is accompanied by the relative positive deformation along the a-axis that is equal to
3 × 10 -5.
High temperature range (T > TSR2). At these temperatures the stable zero-field phase
is a collinear ferrimagnet with magnetic moments directed along hexagonal c-axis.
The magnetic field along the a-axis causes both canting of ferrimagnetic structure
and continuous rotation of the magnetic moments towards the basal plane. Finally,
in Her the magnetic moments are directed perpendicularly to the c-axis. In fig. 14.5
the field dependence of longitudinal magnetostriction )~aa for T = 210.2 K is shown.
The process of gradual spin reorientation to the basal plane is finished at the critical
field equal to 10 T. The appropriate magnetostriction deformation equals 2.2 × 10 -5.
The additional high-field anomaly is due to the field-induced canting suppression
and the collinear ferrimagnetic phase recovery again. The essential reason for such
conclusion is the fact that total magnetostrictive deformation (fig. 14.5), within the
limits of experimental accuracy, agrees with the thermal expansion anomaly caused
by the reorientation of magnetic moments from hexagonal c-axis to the basal plane
Aa~ = 3 x 10 -5 (Krynetski et al. 1994). The process of canting suppression has the
significant field hysteresis, so when the external field decreases to zero the residual
strains of the crystal lattice of HoCo3Ni2 exists.

&aa(lO-s)

HoCo2Ni3
T = 210.2 K /

r1 BerI1
I I I I I I

4 8 12
~(T)
Fig. 14.5. The longitudinalmagnetostriction)~aa of HoCo3Ni2 at temperature T = 210.2 K.
506 A.K. ZVEZDIN

14.4.2. Magnetic field parallel to the c-axis


Rather unexpected results are obtained at temperatures just above TSR2. At these
temperatures the stable zero field phase is collinear ferrimagnet with magnetic mo-
ments directed along the hexagonal c-axis. The magnetostrictive anomaly (Her equal
to 11 T for T = 207 K) indicates that the field induced phase transition occurs. This
transition has the significant hysteresis, for instance, at T = 207 K it is equal to
about 5.5 T. The nature of this transition needs further investigation.
The magnetostriction measurements show that the ferrimagnet HoCo3Ni2 has not
only the magnetic moment compensation occurring at T = 160 K but the magne-
tostriction compensation, too. The magnetostriction due to the domain process goes
passes zero point at T = 120 K with rare-earth contribution being positive while
d-sublattice portion has the negative sign.

15. Surface anisotropy effects and surface phase transition

The magnetic behavior near the surface of a crystal or film may drastically differ
from that of the inside. An attempt to describe the influence of the surface inevitably
would lead to consideration of the surface shifts of energy levels and to modification
of the s-d hybridization. This is beyond the scope of this review. Here we would
like to focus our attention on only one aspect of surface magnetism, the surface
anisotropy (a more complete discussion of surface magnetism including the state of
the art of this problem can be found in a detailed review by Kaneyosi (1991a, b)).
There are many experimental data which confirm its existence in a large class of
crystals and films. This surface anisotropy is due to the various causes associated
with material, preparation techniques as well as with aging processes (oxidization of
RE ions, crystallization). Further we will consider surface anisotropy constants as
phenomenological constants.
Let us come back to the expression for total free energy with allowance made for
the homogeneous exchange energy A((grad0) 2 + sin2 0 (grad~)2), where A is the
exchange stiffness constant (see section 7)

~St = ~bulk q- ~surf. (15.1)

Here ~bulk is given by the eqs (3.1) and (7.1).


The uniaxial anisotropy follows from the most common surface energy term
~surf = - K s cos 2 0s, (15.2)
where 0s is the angle between the easy axis (z-axis) and the spontaneous magneti-
zation M, the subscript s denotes that this angle refers to the surface of the film.
Let us consider a plate of thickness 2d. For the sake of simplicity we assume the
surface anisotropy values on the both sides of the film to be equal: Ks(d) = Ks(-d).
Besides, we imply the film to be homogeneous in the xy plane. Then we can state
that

O(z) = O(-z) and --dzdO~=o = O.


FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 507

A simple calculation yields us the Euler-Lagrange equation


d20 ()~3eff
2A-----
dz 2 00 '
and the boundary conditions
d0 I 0,
d z z=0
dO z=±a Ks
d-~ = - 2---Asin20s,
where #eff is defined by eq. (3.1).
There are two approaches to this problem: a) to solve the first integral of eq.
(15.3) and do numerical calculations, and b) to use the bifurcation theory to study
the stability of eq. (15.3). For our purposes it is sufficient to investigate the stability
of this equation (Kaganov 1980, Kaganov and Chubukov 1982).
We have three solutions of eq. (15.3): two collinear phases {0 = 0, 7r} and a
canted inhomogeneous phase {0 < 0(r-) < 7r}, their stability regions depending now
on the surface anisotropy. In particular, the stability curves of collinear phases are of
the same shape as in the bulk case, the only difference being the uniaxial anisotropy
(Zvezdin et al. 1991)
Ku - + K = Ku(1 - h),
where Ku is the bulk uniaxial anisotropy and h is a dimensionless parameter defined
by the equation
,=(~)l/2arctanh(h) 1/2,

K,a
0"-- (~--
KuA' A
As appears from the first approximation above, the presence of the surface leads
to changes in the bulk anisotropy. Thus, the characteristic points and curves in the
phase diagrams depend on the surface anisotropy. It is quite easy to determine the
influence of the surface anisotropy on the coercive force of the film (Zvezdin et al.
1991).
The presence of domain wall-like solution of eq. (15.3) is another interesting
property of this equation. It differs from the domain wall equation only by the
boundary conditions. In our case such equivalence leads to the existence of surface-
induced domain walls (in the canted inhomogeneous phase) which divide the film
into surface and bulk domains. This statement agrees well with Mrssbauer data
(Kaneyosi 1991a, b). Here it means that a surface anisotropy can induce surface
phase transition, when a new field induced phase is generated only near the surface
(Kaganov 1980, Zvezdin et al. 1991).
So, the surface anisotropy affects considerably the critical curves and points of
phase diagrams, particularly in thin and ultrathin films. The presence of surface
anisotropy can lead to the existence of surface domain walls and to the surface
field-induced phase transition.
508 A.K. ZVEZDIN

16. Phase transition at the local defect. Dislocations and FIPT's in Gd3Fe~O~

The problem of elucidation of the role of point-like, linear, and other defects of the
crystal lattice in processes that determine phase transitions is of considerable interest
for solid state physics from a basic point of view (Levanyuk et al. 1979, Ginsburg
1981). Exceedingly useful may be experimental research involving a considered here
spin-flop transitions, due to the change of direction of the magnetization vectors of
magnetic sublattices with changing external magnetic field or temperature.
Dislocations are particularly appealing to the study of this problem. The effect
of strains induced by dislocations may be expressed in the free energy (3.16) by a
contribution to the magnetoelastic interaction

X_I_ XII
H 2 sin 2 0 - M ° H cos 0 - K](0, ~) + e(~, 0, ~o),
2

where 0 is the angle between the 3~rd a n d / t , e is the energy increment due to the
elastic stresses of dislocation, f" is the distance from dislocation. This contribution
can be considered as the density of magnetic anisotropy induced by dislocation
deformations. The value of this induced magnetic anisotropy in the vicinity of the
dislocation can be superior to the magnetic anisotropy of the perfect crystal. This
additional magnetoelastic anisotropy causes a considerable deflection of Md from its
orientation away from dislocation. Figure 16.1 illustrates schematically this effect.
It is of interest to study the FIPT near the dislocation. A particular feature of this
case is the dependence of critical fields on distance from the dislocation.
Vlasko-Vlasov et al. (1981, 1983), Vlasko-Vlasov and Indenbom (1984) inves-
tigated singularities in the course of spin-flop phase transition near an individual
dislocation in single crystal gadolinium iron garnets. The samples used in their ex-
periments were obtained from a single crystal of gadolinium iron garnet cut into
wafers parallel to the (110) plane, Their thickness after mechanical and chemical
polishing was 30-50 #m. Because thin wafers of this material are transparent to vis-
ible light, one can use methods based on polarized light to study simultaneously the
distribution of internal stresses (by the photoelastic effect) and the magnetic domain

----~'lllllj 81 ~ & l l l 4 ~V7 . . . . . . . --~-~-.~k, ¢ ] i r T!|2~, . . . . .


"" ~- -- '-, - ' ~ PI | ' | l l ~ l l it* i l ' ~ . ~. .-.-.-.-. .. - - ~ ~" ,,,,v,
I I It !
V v,,
v t I t It
,~,_--
--,, , t ,, .... ,,v,;l, ,,,,,,

Fig. 16.1. Domains arisen near a dislocation in a cubic ferromagnet with K l < 0 (i.e. the easy axes
are parallel to [001]) (after Ditchenko and Nikolaev 1979, Dichenko et al. 1983).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 509

structure (by the Faraday and Cotton-Mouton effects). This circumstance affords a
unique possibility for direct experimental observation of the influence of elementary
dislocations on the rotation of magnetic moments, nucleation of domains of different
phases, and motion of phase boundaries during spin-flop transitions.
The spin-flop transitions in constant field ~ll[ll~] of 177 Oe, near a single 60 °
dislocation with a glide plane (111) and axis [110], is illustrated in fig. 12.2. Ap-
proaching the compensation point Tc 282.5 K from below, in the vicinity of the
dislocation a low-temperature canted phase D, in the form of a light three-lobed
rosette appeared against the background of the collinear phase B. In the latter the
magnetization MII [1 l i] was in the plane of the plate (fig. 16.2 a)). One of the lobes
of the rosette was considerably smaller than the other two. The region occupied by

Fig. 16.2. Change of the shape of the interphase boundary near a dislocation in the course of a
spin-flop transition in an external magnetic field H I I [ l l i ] (the Nicol prisms are slightly uncrossed): a)
T = 280.6 K; b) T = 281.0 K; c) T = 281.7 K; d) T = 282.1 K; H = 177 Oe (after Vlasko-Vlasov et
al. 1983).
510 A.K. ZVEZDIN

phase D in fig. 16.2 a) is brighter than surrounding dark region of phase B, since
the magnetization of the iron sublattices in phase D does not coincide with the plane
of the light polarization.
With increasing temperature, the rosette of the canted phase near the dislocation
increased and merged with the macroscopic region of the same phase advancing
from the bulk of the sample (fig. 16.2 b). In the course of the subsequent succes-
sive (fig. 16.2 c)) redistribution of the volume fractions of phases B and D, their
sectoral arrangement in definite sections of the dislocation field of microstresses was
rigorously preserved. The collinear phase B, decreasing with temperature formed a
rosette (fig. 16.2 d) symmetric relative to the dislocation axis and to the phase-D
rosette that existed at low temperatures (fig. 16.2 a)). The phase B vanished at
284 K.
A similar process took place when approaching Tc from above. The picture of
replacement of the high-temperature collinear phase C is practically identical to that
considered above when replacing the corresponding values of T (see fig. 16.2) by
2Tc T.-

Figure 16.3 shows the H - T diagrams of the phase transitions in four points that
are symmetric with respect to the dislocation and are located at different distances
from the dislocation axis. Each plotted line corresponds to the values of T and H at
which the size of the lobe of magnetic dislocation rosette (see fig. 16.2) in the [111]
direction retained a fixed value of R. It can be seen that in local sections, the H - T
diagram does not undergo fundamental changes as the dislocation is approached,

/-/ kOe
8

4 3
i
I I I I I I
282 283 T, K
Fig. 16.3. Local phase diagrams in the vicinity of the dislocation at a distance R from its axis in the
[111] direction. 1, 1') R = 13/zm; 2, 2') R = 35/zm; 3, 3') R = - 3 5 / z m ; 4, 4 ' ) / ~ = 13/zm. Unprimed
and primed numbers refer to the low- and high-temperature transition from collinear into canted phase,
respectively (after Vlasko-Vlasov et al. 1983).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 511

whereas from the opposite side (R < 0) the diagram changes qualitatively when the
distance to the dislocation is decreased. At the point where the transition curves
change from a diverging type (curves 1-3) to an intersecting type (curve 4), the
dislocation microstresses and the stresses from other sources cancel each other, as
discussed below.
The form of the phase diagram obtained at different points of the crystal depends
substantially on whether the induced anisotropy at these points contributes to the
appearance of a canted phase, as in the case of curves 1-3 of fig. 16.3, or hinders
it, as at the point corresponding to curve 4. Both cases are shown in fig. 16.4. In
fig. 16.4 b), in the metastable existence region of the high-temperature phases at
T < Tc = 283.1 K, there is no experimental phase-transition line, because in the
corresponding magnetic field the temperature of this high-temperature transition is
lower than the temperature of nucleation of the compensation boundary in the region
with the larger value of To.
At H > 5 kOe, in all cases, the asymptotes of the experimental points are
straight lines (shown dashed in fig. 16.4) passing through Tc and having a slope
IdH/dTI ,~ 12.8 kOe/deg. The compensation-boundary motion which causes the
transition between the low- and high-temperature phases, took place through the
chosen sections of the crystal at different field-independent temperatures T = T~.
The good agreement, not only qualitative but also quantitive, between the experi-
mental and theoretical data uncovers prospects for using of the determined relations

H, kOe
20
20 I

I
I
I
/
/
p
15
15
I
I

10 10

i\'
A_

281 282 283 284 282 283 284 T, K


a) b)
Fig. 16.4. Phase diagrams plotted at crystal points where the induced anisotropy enhances a) and hinders
b) the formation of the canted phase: (zx) - transition between low-temperature and canted phase; (v) -
transition between high-temperature canted and collinear phase; (o) - transition between canted phases
(180 ° rotation of sublattice magnetization), (.) - transition between low- and high-temperature collinear
phase (after Vlasko-Vlasov et al. 1983).
512 A.K. ZVEZDIN

for the spin-flip phase transitions for the purpose of studying the defect structure of
rare-earth-iron garnets. In particular, investigations of the magnetic rosette produced
near the dislocation make it possible to determine the direction and magnitude of its
Burger's vector.
From the distribution of the striction - is nonequivalent magnetic phases in the
inequivalent crystal one can determine more reliably anisotropy and the weaker
internal stresses the induced than by the photoelasticity method. Thus, by accurately
measuring the transition-temperature shift due to the stresses, e.g., to TO. 1 K we can
record a stress level of the order of 2 kgf/cm 3.

17. Free-powder samples

Single crystals are not always available for a study of the field induced phase tran-
sitions and the intrinsic magnetic properties. Free-powder samples give a good
alternative. Verhoef et al. (1989, 1990a-c), de Boer and Buschow (1992) reported
the elegant high-field free-powder method (HFFP) which has been used to determine
the intersublattice-coupling strength in a fairly large number of different intermetallic
compounds. The particles of the free-powder sample, having a size of about 40 #m,
are assumed to be small enough to be regarded as single crystallines and are free
to rotate in the sample holder during the magnetization process, so that they will be
oriented by the applied field with their magnetic moment in the field direction.
Verhoef et al. (1989, 1990a-c) analyzed magnetization curves of Er2Fe14_~MnxC
compounds by assuming i) the powder particles to be free to rotate in the sample
holder, ii) the magnetization of the rare earth sublattice to be fixed in the easy axis
direction, iii) the magnetic anisotropy of the transition metal sublattice (Kd) to be
zero, and iv) both sublattices to be spontaneously magnetized to saturation.
In this model the magnetic anisotropy does not influence the magnetization mea-
sured on a free-powder sample. These magnetization curves are the same as in
the isotropic case (Verhoef et al. 1989, 1990a-c)). Only if both sublattices display
magnetic anisotropy, an influence on the free-powder magnetization can be observed.
de Boer and Zhao (1994) and Zhao (1994) studied this problem assuming both
sublattices to be spontaneously magnetized to saturation.
The above considered model of two-sublattice anisotropic ferromagnet (see sec-
tions 2 and 3) allows to abandon the last-mentioned assumption and to obtain the
magnetization curves and the phase diagrams of free-powder samples taking into
account the real dependence of the rare earth magnetization on temperature and
magnetic field. The thermodynamic potential (3.1) can be rewritten as

F(O, a) = - M d H cos ¢ - f0 H~ Mf(h) dh + Kd sin 2 0 + Kf sin2(0 + a),

where Kd and Kf are the magnetic anisotropy constants for d- and f-sublattices, ~b
is the angle between magnetic field and the d-sublattice magnetization, 0, Of = 0 + a
are the angles between the easy axis and the d- and f-sublattice magnetization, a is
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 513

EA

oe M

Mf

Fig. 17.1. The orientations of the sublattice magnetizations relative to the external field and the easy
axis.

the angle between the two sublattice magnetization (see fig. 17.1). The angles 0 and
determine the direction of the d- and f-sublattice magnetization.
It can be shown that in the case of freely suspended sample the angle ¢ is uniquely
determined by the angle a:

Mf
COS 't/) - - - - COS ~ ,
M(~)

where

M(oO = ( M 2 + M 2 + 2MdMfCosa)U 2.

This means that the angles 0 and a can be adopted to define the magnetization process
in this case. It is appropriate at this point to recall that the condition Kf << )~MdMf
is herein taken into account as well as above in section 3.
The angles 0 and a can be found by minimizing the thermodynamic potential

Off
O---'O= Kd sin 20 + Kf sin 2(0 + c~) = 0, (17.1)

O~ mdmf
- - = Kf sin 2(~ + 0) - AMdMf sin a + H - - sin c~ = 0 (17.2)
0c~ M
Here we make use of the following formula for Here:

H~ff = H c o s ( a - ¢ ) - Amd cos ¢ .

After some algebra, eq. (17.1) can be brought to the form (Zhao 1994)

]K I sin[2(0 + c~)] = Kd sin 2c~, (17.3)


514 A.K. ZVEZDIN

where

IK[ = [Kg + + 2 K d K f c o s 2o~11/2.

Equation (17.3) can be used to eliminate 0 from eq. (17.2), which becomes

sinc~ --M - ~ + I K B M d M f cosc~ = 0. (17.4)

In contrast with the theory elaborated by de Boer and Zhao (1994) and Zhao (1994),
the approach presented here is unbounded by the severe requirement Mf = const. The
magnetization of the rare-earth sublattice in this approach depends on the temperature
and Heff.

2.5

1.5

0.5

,,,I,,,,I,

J
a) 0.5 1 1.5 2 2.5 3 3.5

2.5

1.5

0.5 ~ l

lllllllllrllll~
0.5 1 1.5 2 2.5
b)
Fig. 17.2. Representation of the uniaxial anisotropy influence on the magnetization curves of the free
powder sample of the uniaxial ferrimagnet. These magnetization plots are calculated by formula (17.6)
at a = 3 and: a) e = -0.1, b) e = +0.1, c) e = +0.2. The straight line /~ = h on the graphs
refers to the 'isotropic' case Kd = 0. The portions of the magnetization curves having negative first
derivative d#/dh correspond to the unstable phases. The first-order transitions points of the jump-like
magnetization curves should be determined using the Maxwell rule, i.e. the thermodynamic potentials
of the corresponding phases should be equated at the point of the transitions.
FIELD INDUCED PHASE TRANSITIONSIN FERRIMAGNETS 515

Figure 17.2 schematically shows the influence of the magnetic anisotropy on the
high-field magnetization curves of free-powder samples. Critical fields Hcl and Hc2
can be written as (Zhao 1994)

Hc ,2 = IMd Mfl( + za),

where

2KdKf
A=
MaMelKa + Kfl

and the upper signs being valued for Hcl and the lower ones for Hc2. An additional
term A appears in the formula for Hcl and Hc2. This additional term will be zero
when either Ka or Kf is zero.
For simplicity assume that Kf >> Ka then

2Kd
A ~ - - =e (17.5)
MdMf

does not depend on the angle a. Notice that the same result is true when the rare-
earth sublattice is described in the Ising approximation, i.e.

where ~ is the unique vector of the Ising axis. This approximation often has a
considerable utility for the description of large magnetic anisotropy of rare-earth
ions.
Magnetization curves in the canted phase follow immediately from eq. (17.4). It
is convenient to express these curves in parametric form:

h = (a + 2cos o01/2(1 - ecosc0,

# = (a + 2 cos oL)1/2, (17.6)


where
M H
-- h -
(MdMf)l/2 ' A(MdMf)I/2'

Me ma
a= m+__
md Me'
O < a < rr, e = 2Ka/ MaMf.
Evidently, the portions of magnetization curves having negative first derivative d#/dh
correspond to the unstable phases. The first-order transitions points of the jump-
like magnetization curves should be determined using the Maxwell rule, i.e. the
516 A.K. ZVEZDIN

Fig. 17.3. The h-¢ phase diagram of the free powder sample for a uniaxial anisotropy); h =
H / A ~ , ~ = 2 K a / A M a M f . A C , B D are the second order curves, C - E - D is the first order
curve (see figs 4.1, 4.2 as well).

z--N
I
O' ~ D E
B f

lit

~4
Fig. 17.4. The h-¢4 phase diagram of free powder sample for tetragonal symmetry;A6', BC' (CD,
O 0 ~, C ' D , D E ) are the second (first) order curves (see figs 4.3, 4.8).

thermodynamic potentials of the corresponding phases should be equated at the point


of the transitions.
Figure 17.3 show the typical magnetization curves for the different values of the
parameter ~ defined by the formula (17.5) and the associated phase diagrams.
By the same procedure as in the case of uniaxial anisotropy, one obtains for the
tetragonal and hexagonal structure similar equations in the parametric form.
Figures 17.4 and 17.5 show schematically the phase diagrams of the tetragonal
and hexagonal ferrimagnets. As above, it is suggested here that IKfl << IKdl. There
is a striking similarity of topological features of these phase diagrams with phase
diagrams of monocrystals discussed in the section 4. In the same way it is possible
to construct H - T phase diagrams for these free-powder samples, but this is beyond
the scope of this chapter.
The associated magnetization curves can be determined by the following para-
metric equations taking into account the corresponding equation for the magnetic
anisotropy energy of the tetragonal and hexagonal ferromagnets:
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 517

O~~o 2 ~-

~6
Fig. 17.5. The h - e 6 phase diagram of free powder sample for hexagonal symmetry; Bt3', AA' ( O 1 0 ~ ,
020~, BD, AD, 03D, DE) are the second (first) order curves (see figs 12.2, 12.5).

20

16 DyCo12B6
...........o ..........
..o"'"
.~.........-""'
12
.,0.....,0.0
.....
8
,...C'"
~000.0.0.0-0O"

0 i I I I t l I

0 10 20 30 40
/~(r;
Fig. 17.6. Free-powder magnetization for D y C o l z B 6 at 4.2 K. The circles represent measurements in
quasi-continuous field and the dotted line represents a measurement in a field varying linearly with time
(after Zhou et al. 1992a).

Tetragonal s y m m e t r y :
# = (a + 2COS o01/2,

h=(a+2c°sc0]/2
(sine°)
l-e4 sinc~ "

Hexagonal symmetry:
# = (a + 2 c o s cO 1/2,

h = (a + 2 c o s a ) 1/2
sin 6 a
1 - e 6 sin------~ '
)
w h e r e 0 < a < 7r, e 4 = K4/mdmf for tetragonal symmetry, and ~6 = K 6 / M d M f for
h e x a g o n a l symmetry. T h e constants K4 and K6 are d e t e r m i n e d in subsection 3.2.1.
518 A.K. ZVEZDIN

Figure 17.6 shows the free-powder magnetization for D y C o 1 2 B 6 a t 4.2 K. This


is a good example of the complete bending process in a ferromagnet. The magne-
tization curve in the canted phase is nearly linear indicating that the Dy sublattice
anisotropy is much larger than that of the Co sublattice. It is seen that the effect of
the magnetic anisotropy of the Co sublattice is rather small. From the slope in the
field region where the canted phase exist, the interaction between the rare-earth and
cobalt sublattices can be derived to amount JDyco/k = --4.1 K (Zhao 1994).
Figure 17.7 show the free-powder magnetization curves of Ho2Fe17_~Al~ com-
pounds in which only the first critical field Hcl is attainable. The second critical
field Hc2 is very large here. In all measurements in figs 17.6 and 17.7 a quite good

20 ' I ' i ' I

16 ...... ~ifOD::'~:'a'~
• O ............ 0~:::::~...... <>
ooo~.ooOOO-O.."o'"'o'o''""°'"'°'c" ........ []
,"3.
12
1200 0 m O [ ] • O 0 [] rn•
o[]o §
$$ o x=l
[] x = 2
~<>oooooo<> o o•
A e~ <> X = 3
A X=4
mm vv
v X=5
V VVV
~V
0 i l i l I 1 I
0 10 20 30 40

16 ' I ' I ~ I '

A ............. ~'-
......... ~'" ......,"

12 A
o .............
..... @-'''÷'ae ............ 4-................. '~" ....... "' """ -
,+...÷-q: .... /~ ............... ,
_., +..... .....
~:..~÷'*"
8 -.,~,~'"''""~"..x~-.-'X~A
o~
O
o x=5
.~,~'V'"V'~'~AZxA • I~ Ov [] x = 6
'xA"zxzx o o 08 o x = 7
_ ^OOO O 8
4 eO<>C'v unS8 A x= 8
Du [] o o v x=9
moo• ° ° + x = 10
o
oo
0 I I i I i I i

0 10 20 30 40

Fig. 17.7. Free-powder magnetization curves of Ho2Fel7_xAlz compounds measured at 4.2 K. Circles
stand for the step-field measurements and dotted lines represent measurements in which the field decreases
linearly with time (after Jacobs et al. 1992b).
FIELD INDUCEDPHASE TRANSITIONS1N FERRIMAGNETS 519

linear magnetization in the canted phase is found with little influence of the magnetic
anisotropy.
A number of ferrimagnets have free-powder magnetization curves which are more
complicated in the canted phase.
Figure 17.8 shows the free-powder magnetization of Erz_~YxCoTB3 for z = 1.9
and 1.7. The occurrence of the first order magnetic phase transitions is clearly seen
in these curves.

12.0 i I i I i

......... O . ............... o
•. 0,," .....
..::y"

,-, 8.0
..~i?.6%:,Cr..O'..::%7.::';'% "Q~"

<,7

4.0 / Er0.sY2.sC°l 1B4


/
"" Free powder
~ 0 0 . 0 0 "0'0 T=4.2 K

0.0 I I I I I I i
0 10 20 30 40
B(r)

15.0 I ' I i

Er0.75Y2.25CoI 1B4
Free powder /o----- u ~o-

,-,10.0
T=4.2K
/
/ i'
50 i /
i
:"i / <,

rvO.00.,...~
''''U''
0.0 '°°°'~i I i I I I I
0 10 20 30 40
/~ (/3
Fig. 17.8. Free-powder magnetization at 4.2 K of Er3_xY~CollB4 compoundswith z = 2.0 (top) and
1.9 (bottom) (after Zhao 1994).
520 A.K. ZVEZDIN

16.0

_.--.---O

ErY2Co 11B4 .....::....oi"


...- ."
12.0 ,_owder
Free 13 /0 //
T=4.2 K ~
//
~ 8.0
~J /o /
- /
/ /
/ -

4.0
000.0.0.0. 0 O0 0 OD':f''"

0.0 i I = f i I i

10 20 30 40
(73

L I ~ I i I i

16.0
Erl.IY1.9COllB4 f
Free powder -i¢
~, 12.0
T=4.2 K .//

8.0 //
//
.//

4.0 ~ 0 0 0 . 0 , 0 , 0 .....0....0,..., 0,0" ........

0.0 = I f I i I I

0 10 20 30 40
/~ (73
Fig. 17.9. Free-powder magnetization at 4.2 K of Er3_~Y=CoHB4 compounds with z = 2.5 (top) and
2.25 (bouom) (after Zhao 1994).

Figure 17.9 shows the magnetization curves for (Tm,Y)Co7B3 and (Er, Y)CollB4,
(Tm,Y2)CoIIB4 compounds.
All these results illustrate that when the magnetic of the d-sublattice in the basal
plane can not be neglected in comparison with the anisotropy of the rare-earth sub-
lattice the magnetization process is rather complex.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 521

18. Spin-flop transitions in itinerant metamagnets

Spin-flop transitions in ferrimagnets have been considered above in detail both the-
oretically and experimentally. The most common materials, which have been the
setting for most of the research on these phase transitions so far, are compounds of
rare-earth elements and transition metals of the iron group: the iron garnets R3FesO12
and the intermetallic compounds R,~Tm, where R is a rare-earth element and T a
transition metal. Such systems are generally described in the approximation of a
'rigid' d-sublattice, in which the magnetization of the d-sublattice is assumed to be
independent of the external magnetic field and the exchange field exerted on it by
the rare-earth sublattice.
Strong interest has been caused by itinerant f-d metamagnets, which in a sense
are the opposite of the materials mentioned above. In them, the d-subsystem is a
weak itinerant ferromagnet and can undergo a metamagnetic phase transition with
a substantial jump in magnetization. Some typical members of this family are
YCo2, LuCo2, and RCo2, which exhibit metamagnetic transitions in the electronic
d-subsystem from a paramagnetic state to a ferromagnetic state under the influence
of a magnetic field.
Magnetic field-induced transitions of this sort were predicted by Wohlfarth and
Rhodes (1962). According to the theory of Bloch et al. (1975), Cyrot and Lavagna
(1979), Yamada et al. (1984), such transition is possible in YCo2; it was indeed
observed (Goto et al., 1989, Murata et al. 1991) in fields on the order of HM ~ 106 0 e
and is explained by the peculiarities in energy dependence of the density of states
N(F_,) of the d-electrons near the Fermi level. The critical field decreases in the
compounds RxYl-zCo2, since metamagnetic transition arises in an effective field
which is the sum of the external field and the molecular field exerted on the d-ions
by rare-earth ions (Wohlfarth and Rodes 1962, Cyrot and Lavagna 1979, Duc et al.
1988, Steiner et al. 1978, Ballou et al. 1992).
Another possibility for a decrease in the critical field of a metamagnetic transition
is realized in compounds of the type (R,Y)Co2_xAI~ (Aleksandryan et al. 1985).
Our purpose in this section is to discuss magnetic-field-induced phase transitions
and the H - T phase diagrams in itinerant metamagnets with an unstable d-sublattice.
This problem is important to the magnetism of f-d intermetallic compounds in gen-
eral, since according to the present understanding of itinerant ferromagnetism the
d-subsystem in such compounds is frequently an unsaturated (weak) ferromagnet
with all the characteristics of this situation.
We first consider the ground state of an f-d itinerant ferrimagnet at T = 0. The
thermodynamic potential of the d-subsystem of a weak itinerant ferrimagnet is cus-
tomarily expressed as an expansion of magnetization power (Wohlfarth and Rodes
1962)
1 1 1
F = ~ a~T~2 -4- ~ b~r~4 -[- g cfn 6, (18.1)

where a, b, and c are coefficients whose temperature dependence is determined


by the particular band structure and/or by spin fluctuations. Let us assume, as
522 A.K. ZVEZDIN

is customarily done in the theory of itinerant metamagnetism, that the following


relations hold: a > 0, b < 0, c > 0.
We consider the case in which the d-subsystem is itself paramagnetic down to
T = 0. It would seem at first glance that in this case the f-d system as a whole
should also be paramagnetic, since the molecular field created by the d-subsystem
at the f-ion must be zero. However, we will show that this is not the case. Actually,
the paramagnetic state (i.e. a state with a zero spontaneous magnetization) of such
a subsystem is unstable in the presence of an f ion with a degenerate ground state,
and a spontaneous magnetization arises in the f-d systems.
The onset of spontaneous magnetization below the threshold., i.e. formally, in the
paramagnetic phase, can be seen directly on the magnetization curves reported by
Ballou et al. (1992). The physical meaning of this phenomenon can be explained
as follows. We consider a spin fluctuation in a system of d ions surrounding the
f ion of interest. We assume that this fluctuation causes a splitting of the ground
state of the f ion. The meaning here is that the f ion becomes magnetized, and in
its turn causes further magnetization of the surrounding ions by virtue of the f-d
interaction. A spontaneous magnetization thus arises in the d-f system. In other
words, the symmetry under time reversal is spontaneously broken. (In the case of
an isolated ion, of course, quantum fluctuations should lead to the restoration of
symmetry under time reversal, but in the case of cooperative instability discussed
below this symmetry breaking has a completely real meaning.)
This conclusion can be drawn as a consequence of a more general position - as
a manifestation of crossover instability in magnetically ordered systems (Zvezdin et
al. 1976, 1985). The lowering of the energy due to the splitting of the ground state
of the f ion is equal to

-.XmMe,

where Me is the magnetic moment of the f ion, m is the amplitude of the magneti-
zation of the cloud of magnetized ions around the f ion, and A is the f-d exchange
constant. When the paramagnetic d-subsystem becomes magnetized, its energy in-
creases by an amount

v m Z /2xcl,

where Xd = a - 1 is the magnetic susceptibility of the d-subsystem, and V is the


volume of the magnetized region around the f ion. The instability of the uniform state
(m = 0) is obvious even in an arbitrarily weak interaction, since the improvement
in terms of energy is linear in m, while the degradation is quadratic.
If the concentration of m ions is high enough, i.e. if magnetization clouds around
the f ions overlap, there will be a cooperative spontaneous magnetization of the
f-d-subsystems in the systems. In this case the total energy per molecule is

E=-~1 m Z / x d _ t A m M , (18.2)
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 523

where t is a relative concentration of f ions.


Minimizing (18.2), we find

m = XdtAmM. (18.3)

This discussion can easily be generalized for nonzero temperatures Hell. The
thermodynamic potential # of this system is

1 a m 2 + -~1 bm 4 + -~1 cm 6
~5 = -~ - - t ~o n*~ Mf(h) dh, (18.4)

where Mr(h) is the magnetization of the f-sublattice expressed as a function of the


effective field H e f f = H - Arfi (in this case, the external field H is assumed to be
zero), of temperature T, and of the magnetic moment # of the f ion. In particular, for
the Gd +3 ion we have Mr(h) = # B 7 / 2 ( # h / T ) , where B ( x ) is the Brillouin function,
and/z = 7#B. Minimizing (18.4) with respect to m and ¢, we find

a m + b m 3 + c m 5 - t A M f ( A # m / T ) = O. (18.5)
This equation determines two phase-transition points: the magnetic-ordering tem-
perature Te, which is determined by the condition XdXfA2t = 1, and the (usually
lower) temperature of the metamagnetic transition to a 'strong ferrimagnetic' state,
TM.
The threshold concentration for the t* transition to a cooperative behavior of the
system can be estimated on the basis of the following qualitative considerations,
which are based on arguments concerning the size of the magnetized cloud around
the f ion. This is determined by the correlation radius of fluctuations in the d-
subsystem. To find it we work from the Ginzburg-Landau energy

E = f (A(gradm) 2 + m2/2Xd) dV, (18.6)

where A is a nonuniform-exchange constant (exchange stiffness) of the d-subsystem,


and the integration covers the entire volume except a small volume around the f ion
with a diameter of an interatomic distance. We assume m ( r ) = rn at the boundary
of this small volume, and at infinity we naturally assume m ( r ) ~ O. Using these
boundary conditions and Euler-Lagrange equations for functional (18.6) we then find

re(r) = (ma/47rr) e x p ( - r / p ) , (18.7)

where p = (AXd)I/2 is the correlation radius which we are seeking. Setting A =


5 x 10 -9 esu/ion and Xd ~ 2 × 10 -7 esu/ion, we find p = 3A. This value is of
course too small for the continuous approximation to be reliable, but it will do for a
qualitative argument.
The magnetization distribution described by (18.7) is obviously degenerate under
the direction of the spins of the f ion and the d-subsystem (the matrix). The actual
524 A.K. ZVEZDIN

wave functions of this composite formation is therefore a linear combination of all


possible wave functions of this system which are included in the degeneracy space,
i.e. which are characterized by different spin orientations. A wave function of this
sort describes a localized spin-wave mode which includes the joint motion of f and
d spins. A detailed description of this mode goes beyond the scope of the present
chapter, but we would like to point out that this picture of an isolated f ion in a d-
matrix may be observed at a low concentration of f ions, at which local spin modes
do not overlap. The mode overlap determines the interaction of centers, and if the
concentration of centers is sufficiently high, it will lead to cooperative effects.
The threshold concentration t* can be estimated with the help of a percolation
model. If the distance between the two nearest f ions is smaller then p, then these
ions perturb each other substantially. As a result, a correlation arises in the directions
of spins. The percolation threshold can then be found approximately from N * p 3 = q,
where q is a numerical factor of the order of one; for our purposes we can set it
equal to one. Thus the critical threshold concentration is

t* ~ ( a / p ) 3 ~ 0.1,

where a is the average interatomic distance in the compound.


We turn now to phase transitions induced by a magnetic field and to phase dia-
grams, assuming that the concentration of f ions is high enough that the behavior of
the system can be assumed to be cooperative. In other words, we assume t > t*.
In this case we can ignore the nonuniform-exchange energy and we can write the
thermodynamic potential in the form (Zvezdin and Evangelista 1995)

1 1 1
q5 = L a m 2 + - b m 4 + L c m 6 -- m ( H 2 + A 2 M 2 - 2 A M f H c o s ¢ ) I / 2 -
2 4 6

- M f H cos ¢ - T S ( M f ) ,

w h e r e ~ d is the energy of the d-subsystem found from (18.1), and ¢ is the angle
between the magnetic field H and magnetization Mf of the f-sublattice; S(Mf) is the
entropy of f-subsystem,
Let us first consider the case T -- 0, when Mf = const. Minimizing (18.8) with
respect to m and ¢, we find

hd(m) = Heff(Mf, ¢), (18.9)

M f H sin ¢(1 - A m / H e f f ) = O, (18.10)

~)S cos ¢(1 - Am mA2mf (18.11)


-T c)Mf -- "~eff ) -- ~eff -- O,

where/~eff = / t - A/~ff.
FIELD INDUCEDPHASETRANSITIONSIN FERRIMAGNETS 525

Let us approximate function m(h) in the form


( Xdh, h < Hp,
m(h) mr, h > Hp, (18.12)

where Hp is a threshold field for metamagnetic transition in the d-subsystem. From


(18.12) we can directly determine the functional dependence h(m). It is simple to
verify by direct calculation that the values of magnetization on the descending slope
of the m(h) curve (or the h(m) curve determined by eq. (18.12)), i.e. the values
corresponding to a negative differential susceptibility 8m/Sh < 0, correspond to
unstable states of the system. For this reason, the interval of magnetization values
from XdHd to ml is of no interest for our purposes, in our approximation of re(h).
For simplicity we will also ignore the hysteresis in the m(h) curve.
We should also mention that in a real metamagnet the magnetization of the d-
subsystem is not constant in ferromagnetic phase. Instead it increases slightly with
an increasing field. This point is simple to deal with mathematically, but taking it into
account makes the analysis more complicated without leading it to any qualitatively
new facts. Since it tends to obscure the overall picture, we will ignore it.
Equations (18.9), (18.10), and (18.11), along with the stability conditions (82~ > 0),
determine the following solutions (phases) and regions in which they exist:

W :¢=0, m=Xd (H-AMf),


Fsl : ¢ = 0 , 0=Tr, m=ml,
Fs2 : ¢ = 7r, O=0, m=ml,
C :0<¢<7r, m=ml,
Fe : ¢ = 0 , O=0, m=ml,
where a is the angle between the magnetic field and rS. Phases W (Fsl and Fs2)
can be weak (strong) ferrimagnetic collinear phases, Fe is a ferromagnetic phase,
while phase C is a canted (angular) phase. To plot H - T phase diagrams, we need
to equate the energies of coexisting phases and find the curves of first-order phase
transitions from these equations. As ~d(m) in our approximation we should use the
quantity

qSd = /? hd(m) dm.


Phase diagrams in fig. 18.1 give a general picture of magnetization curves and
the critical field of itinerant metamagnets. The nature of the H-Mr phase diagrams
depends on the relation between Hp and Aml. If Hp < Aml/2, the phase diagram will
have, in addition on the 'ordinary' lines of second-order phase transitions between
collinear and angular phases (which are determined by the known expressions for
critical fields Hcl = Alml -Mf[ and Hez = Alml +Mfl), lines of first-order transitions,
which are determined by the equation (under the condition AXd << 1)
Hp - AMf
HM -- (18.13)
1 -- 2Mf/m
526 A.K. ZVEZDIN

H.

N
Xml ~.m 1
H
"2~O
N #

[]
X-1Hp mI M ~-lHp m1 M
a) c)
H H

~,n1
Y H

~.-IHp m1 M x-lG M
b) d)
Fig. 18.1. Phase diagrams of an itinerant ferrimagnet with an unstable d subsystem, a) /fp < Am1/2;
b) Hp = ),ml; c) Am1~2 < Hp < k,ml; d) Hp > k,ml. The phases W, Fsl, Fs2, Fe are labeled
A1, A~, B2, A2. The inserts are schematic diagrams of the corresponding magnetization curves (after
Zvezdin 1993).

In the interval Aml/2 < Hp < Am1 the phase diagram (fig. 18.1 b)) has two
interesting features. Not only W --+ Fs2 phase transitions can occur here, but also
first-order transitions from the weak ferrimagnetic phase A1 to the canted phase C.
Transitions of this sort have not previously been seen in isotropic systems. Lines
QL and Q'L* are determined by eq. (18.13), and lines Q N and Q'N* by

H = AIMf + ETI~1 l, H = ,~[Mf - ETn 1 l,

where e = ((2Hp/Aml)- 1) 1/2. The points Q and Q* have coordinates Mf = l m l ( l +


e), H = 1Am1(1 q:e).
The second unusual feature of this diagram is that there can be an 'inverse transi-
tion' from a strong ferrimagnetic phase to a weak ferrimagnetic phase. This transition
would occur as the magnetic field is increased (Zvezdin and Utochkin 1992). The
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 527

same feature is seen, even more clearly, at Hp > )~ml (fig. 18.1 c)), where an inverse
transition occurs abruptly, without an intermediate canted phase.
Figure 18.2 shows the H - T phase diagrams of itinerant ferrimagnets deduced from
the minimization of the thermodynamic potential (18.8) (Zvezdin and Evangelista
1995).
Substituted intermetallic compounds of the Laves Phases Y1-tRt(COl-~Alx)2 (R is
a heavy rare earth element) with a negative intersublattice exchange interaction may
serve as representatives of the itinerant ferrimagnets with one unstable magnetic
sublattice. One, stable, magnetic subsystem of these two-sublattice ferrimagnets is
formed by the localized moments of the 4f-shells of rare earth ions. The second
itinerant, magnetically unstable, subsystem is formed by magnetic 3d-electrons of
Co hybridized with 5d(4f) electrons of rare earth ions. For this subsystem, the density
of d-states at the Fermi level and the magnitude of d-d exchange are such that the
Stoner criterion for the appearance of itinerant ferromagnetism is not fulfilled (Bloch
et al. 1975, Yamada et al. 1984). Because of this, YCo2 and LuCo2 compounds are
exchange-enhanced itinerant paramagnets (Lemaire and Schweizer 1966).
In RCo2 compounds with magnetic rare-earth ions, both magnetic subsystems are
magnetically ordered (Bloch and Lemaire 1970, Levitin and Marcosyan 1988). In
this case, the magnetic ordering of the itinerant subsystem is extrinsic and due to a
magnetizing molecular field acting on the d-subsystem. The investigations of the f-d
exchange on the properties of the d-subsystem have been carried out on (RY)Co2
compounds for R = Gd, Tb, Ho, and Er (Lemaire and Schweizer 1966, Levitin et al.
1984a, b, Duc et al. 1988a, b, 1989, Gratz et al. 1986, Baranov et al. 1989, 1990).
In order to study the influence of the f-d interaction on the magnetic order of the
unstable d-subsystem, Ballou et al. (1992) investigated the magnetic properties of
intermetallic compounds Yl-tGdt(Col-xAlz)2, where 0 ~< t ~< 0.2, 0 ~< :c ~< 0.105.
Partial replacement of cobalt by aluminum in YCo2 leads to a decrease in the field
of metamagnetic transition and to the appearance of itinerant ferromagnetism in
Y(COl_~AI~)2 compounds for z/> 0.12 (Aleksandryan et al. 1985).
Figure 18.3 shows the field dependence of magnetization at 4.2 K for certain
compounds belonging to the system Yl-tGdt(Co0.095A10.05)2. It is clear that for
small replacements by gadolinium (t < 0.12) there is no spontaneous magnetiza-
tion. Increasing the gadolinium content leads to increased weak-field susceptibility;
the magnetization curves of compounds with Gd become nonlinear, and exhibit a
tendency toward saturation in strong fields. Compounds with Gd content t /> 0.12
possess a spontaneous moment. The value of this spontaneous moment decreases as
the gadolinium content increases, passing through a minimum at tcomp ~ 0.17-0.18
and then increasing once more (i.e. this concentration marks a balance point with
respect to compensation). In compounds near tcomp, kinks in magnetization curves
take place, which are characteristic of a transition from a collinear ferrimagnet to
noncollinear phase.
Figure 18.4 shows the magnetization curves of several compounds of the
Y1-tGdt(Coo.915Alo.oss)2 system at 4.2 K. It is clear that the original compound
Y(Coo.915Alo.oss)2 is an itinerant metamagnet with a critical magnetic transition field
528 A.K. ZVEZD1N

a)

1.5

1
F~2
0.5

i I i ~ i i r

50 100 150 200 250 300


Temperature (IO

b)

1.5

0.5

50 100 150 200 250 300


Temperature (K)

e)
2.5

,•1.5 W

0.5

I I

50 100 150 200


Temperature (K)
Fig. 18.2. H - T phase diagrams of an itinerant ferrimagnet with unstable d-subsystem, a) Hp < ,kin1/2;
b) )~ml/2 < Hp < Aml/2; c) Hp > )~ml (after Zvezdin and Evangelista 1995).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 529

1M p~/f.units
a)
$++ 1
1 4-+
xX~X~~xx 2
xX +4-

0.8
.Z~Zi xx ++
^L~ XX +4-
xx / ++
0.6
xx ++ /

0.4
++++++
AZ~ xx ++
~ ~o~x+++ +++
0.2 ZIA x~.+ +
~,~l x ~.>f.+
,'~<". x~-+
~xX~+
"{ "''1" T i I I I t I

0 50 100 150 200 250

b)

0.5
- 4
+ + 4.4. + + .t.+ +.t. 4.-I-+ 4- -I-4.-I-'1-+ "1"t"+ + + ~

0.4

0.3

0.2
7 .~IA

0.1
x×6
Nxx

I t P I I I
0 50 100 150 200 250 H, kOe
Fig. 18.3. Magnetization curves of the compounds (Yl-tGdt)(Co0.95A10.05)2 at 4.2 K and t = 0.0 (1),
0.04 (2), 0.1 (3), 0.12 (4), 0.15 (5), 0.18 (6), 0.2 (7) (after Ballou et al. 1992).

HM = 225 kOe. As the gadolinium concentration increases, the metamagnetic tran-


sition field HM decreases, and for concentrations t/> 0.04 these compounds possess
a spontaneous magnetization. At comparatively small gadolinium concentrations
(0.04 ~< t ~< 0.06) the spontaneous magnetization is small in the magnetically or-
dered region, and the application of a field leads to a metamagnetic transition from
a weakly ferrimagnetic to a strongly ferrimagnetic state. For a larger gadolinium
content (t >~ 0.06) metamagnetic transitions are not observed: these compounds are
in a strongly ferrimagnetic state even at a zero field. The saturation of magnetiza-
530 A.K. ZVEZDIN

M /as/f.units
0.6 0.6
0.5 DoDoDo~ 0.4
xXXXXXxXxXX

0.4
D °
um~
0.3
XXX x:JS :÷
,~
n= ooooo 2
0.3 = ~o~
0.2
5 ++~,~
0.2 nmO~O~~ ' ~
, o aa~aaaaaaA1
0.1 6:
0.1
:natc~az~a
; aaaa ' , ,
I I I I I

0 50 100 150 200 0 50 100 150 200 H, kOe


a) b)
Fig. 18.4. Magnetization curves for the compounds (Yl-tGdt)(Coo.915Alo.o85)2 at 4.2 K and t = 0.0
(1), 0.02 (2), 0.04 (3), 0.10 (4), 0.15 (5), 0.18 (6), 0.20 (7) (after Ballou et al. 1992).
M, t~/f.uniis

0.8 0.8 .~

0.6 0.6 -

0.4 0.4 -

0.2 0.2

0 0.05 0.1 0.15 0.2 0


J I

0.05 0.1 0.15 0.2


a) b)

0.8 0.8

0.6 L 0.6
0.4 0.4

0.2 0.2

0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2 t


e) d)
Fig. 18.5. Dependence of the magnetization on gadolinium content t for various compounds from the
system (Yl_tGdt)(Col_xAl~)2 at 4.2 K. (o) - spontaneous magnetization of the weakly ferrimagnetic
compounds, (A) - magnetization of the weakly ferrimagnetic compounds in the field of 270 kOe, ([]) -
spontaneous magnetization of the strongly ferrimagnetic compounds: a) x = 0.105, b) 0.075, c) 0.07,
d) 0.05 (after Ballou et al. 1992).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 531

HM, kOe

300

25O

200

150

100 \~~,~. i k

50 z~

0.02 0.04 0.06 t


Fig. 18.6. Dependence of the metamagnetic transition field on gadolinium concentration t for the
system (Yl_tGdt)(Col_~Al~)2; x = 0.07 (1), 0.085 (2), 0.105 (3). The straight line is calculated
using eq. (18.13), and the points are experimental data (after Ballou et al. 1992).

tion decreases with increasing t, and this causes it to increase once more. As with
compounds with a low aluminum content, near this concentration a transition is ob-
served from the collinear ferrimagnetic phase to a noncollinear phase in the external
magnetic field.
A decrease in the metamagnetic transition field is also observed in other systems,
i.e. Y~_tGdt(COl_~Al~)2 with a high aluminium content (x + 0.07 and 0.105), as
the gadolinium content increases. This decrease is followed by the appearance of
a weakly ferrimagnetic phase, which is then replaced by a strongly ferrimagnetic
phase. The only difference is that the increase of the aluminium content causes
the metamagnetic transition field in these compounds to increase. It also causes
the concentration region, where paramagnetic and weakly ferromagnetic metamag-
netic phases can exist, to shrink. At the same time, the values of magnetization
and gadolinium concentrations, at which a compensation of the magnetic moments
of f- and d-subsystems is observed, are close for all these systems, although with
certain differences. Furthermore, the spontaneous magnetization of a low-aluminium-
content system with x = 0.05 is also close to the magnetization of those compounds
corresponding to systems with larger amounts of aluminium (x = 0.07, 0.085, and
0.105).
All of these features are easy to see in figs 18.5 and 18.6, where the basic char-
acteristics of all systems under discussion are plotted. Figure 18.5 presents data on
magnetization as a function of gadolinium concentration for systems with various x
values. Here the spontaneous magnetization of weakly ferrimagnetic and strongly
ferrimagnetic samples are shown as well as magnetization in the field of 270 kOe
for samples with metamagnetic transitions. Figure 18.6 shows measured dependence
of the metamagnetic transition field on gadolinium content for systems with various
532 A.K. ZVEZDIN

aluminium contents. A comparison of the experimental data with calculations for


f-d magnetic systems shows that they agree in most cases, at least qualitatively.

Conclusion

Field induced phase transitions connected with breaking of a collinear alignment of


sublattices in ferrimagnetics have been considered here. It is assumed that the mag-
netic anisotropy of these materials is small in the sense that the energy of a magnetic
anisotropy is smaller in comparison with the energy of the exchange interaction of
sublattices.
The magnetic anisotropy in such materials is best manifested near a compensation
point, where critical fields of transitions to a canted phase tends to zero. In this
connection H - T phase diagrams are very complicate near Tc and composed of curves
of the first and second order phase transitions and of critical, tricritical points, etc.
From this standpoint the anomalous behavior of many physical properties near Tc
can be explained.
Many rare earth ferrimagnets should be referred to as strongly anisotropic mate-
rials, i.e. crystal field interactions are much stronger than the exchange interaction
between sublattices. The field induced phase transitions in these materials are more
complicated and of interest. We shall address these questions elsewhere with a spe-
cial attention paid to mechanisms that can lead to jumps in a magnetization curve:
field induced crossing levels and the magnetic Jahn-Teller effect, crystal field related
transitions, metamagnetism of 3d-sublattice and other.
Finally, mention should be made of field induced phase transitions in an amor-
phous rare-earth-transition metal system where strong local anisotropy effects are of
importance.

Acknowledgements

I acknowledge fruitful interactions with A.S. Andreenko, K.R Belov, A.S. Borovik-
Romanov, S.L. Gnatchenko, L.R. Evangelista, N.E Kharchenko, A.M. Kadomt-
seva, I.K. Kamilov, B.E Khrustalev, I.B. Krynetskii, R.Z. Levitin, EV. Lisovskii,
V.M. Matveev, A.A. Mukhin, S.A. Nikitin, R.V. Pisarev, A.E Popkov, A.I. Popov,
R. and H. Szymczak, V.K. Vlasko-Vlasov, Y. Zawadzki.
In particular I would like to thank Prof. Dr. K.H.J. Buschow for his valuable
remarks and kind help in preparing this report.

19. Appendix. Microscopic calculation of the thermodynamic potential of the


non-equilibrium state

We shall find the thermodynamic potential of the non-equilibrium state as follows


(according to Leontovich (1944)). The non-equilibrium state of a magnetic crystal
containing interacting d and f electrons will be defined by the magnetization of the
d-subsystem, i.e. by the value of ~rd. As we have mentioned above, this means that
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 533

the f subsystem is in equilibrium with the subsystem of d ions whatever the value
of 2Qd. We shall also introduce an auxiliary (imaginary) field h which is conjugated
with l~rd and only affects d ions. The Hamiltonian of the d-f system interacting with
this field is

(A.1)

where H is the Hamiltonian of the d-f system where ft = 0 and includes interaction
with the external field, the exchange interaction and the crystal field; the summation
is carried over all d ions; #d and Sd are the magnetic moment and spin of the ion.
The free energy depending on h, T and H is

F(f0=-TlnSpexp -~- . (A.2)

For the sake of brevity we shall omit variables T a n d / t in the functions F and ~b
(see below).
We wish to obtain the potential in variables 2krd, T, H. It can be found from (A.2)
by using the Legendre transformation:

+(l~rd) = F(h) + fiM'd; (A.3)

]~rd _ OF (A.4)

The latter relations define the function ~(l~rd) in a parametric form (h is the param-
eter). It can be taken as a thermodynamic potential of a non-equilibrium state with
a given value l~rd. The auxiliary field h 'prepares' this state. It is easy to see that
• (Md) reaches the minimum in equilibrium.
So the calculation of the potential ~(Md) is reduced to a calculation of the equilib-
rium function F(h) with eq. (A.2) followed by the elimination of the auxiliary field
with the Legendre transformation. This procedure is e~ily generalized to obtain
potential q~(Md,Mr) in anon-equilibrium state with any Md and Me.
We shall calculate ~(Md) for a specific d-f system with the Hamiltonian

(A.5)
d 2 d,f

in which

Itf= E H i.
f,
i

= V° + V° + V° + V $ - 9j .Z ;
534 A.K. ZVEZDIN

V~ are the operators of the crystal field. H should be given in the form

g---~ H d - - h E # d ~ - ~ - ~1 ~ i d f ( ~ f ) ~ -'}-Hf +
d d,f

(A.6)
+ ~1 Z Iaf<~a/<Jf/+ 1 E If(Sd - (Sf))(Sf- (Sf))
d,f d,f

- ~ro + V.

Here
V= 1
d,f

Sp A exp(- FIo/T)
(A) =
Sp exp(- £ro/T)
Expanding F(ft) defined by eq. (A.2) over V, we get

F(h) = - T l n Sp exp ~ - / / d - h~--~#aSd + ~ ~ Idf(Sf)Sd --


d d,f /

- T In Sp exp - ~ Hf + ~ d,f (A.7)

1
2 ~ / d f ('°qf)('~d/ + o(g2)"
d,f

Putting F(flt) into (A.3) we calculate

~(~rd) = ~d(~rd) + Ff(Beff), (A.8)

~ d ( ~ ) = - T l n Sp exp - y Ha - a J ~ + ~ Z ~fg~<Jf> -
d d,f
(A.9)
1
2 E fdf<~'~d)('J~f) -b h]~fd,
d,f

/~rd - 8F
Jz'
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 535

Ff0qeff) = - N T In Sp exp - 1 (vo + vo + v O _ g j ~ B O j , ) } , (A.10)

where N is the number of f ions;

~qoff=/~ + ~%,
/~M -- g j- 1
- - E fdf(ffd) = --/~fd;
29j d,f
~d(Md) is the actual thermodynamic potential of an anisotropic ferromagnet in a
magnetic field. In particular, if the magnetic energy ( - M H ) , the anisotropy energy
and the energy of f-d exchange are lower than the d-d exchange energy and if we
expand the thermodynamic potential Oa(Md) defined by eq. (A.9) in these energies
then

Od(Md) = ~i~0(Md) -- MdH ~- Hd(Md). (A.11)

The small quantity of this expansion is

where a is atomic spacing, R - radius of d-d-interaction, z - number of nearest


neighbors of the d-type in the interaction sphere of a d-ion.
The function Ff(Heff) is also expanded over V6 (llV~ll << IlV°ll, ~ = 2, 4, 6)

Ff(/~eff) = Fc(/]eff) -}- (V6), (A.12)

FC(Heff) = - N T l n Sp p~,
(V6) : Sp V6p°,

p°:exp -~ (VO + VO + v60-gJ#BJHeff) .

Of course, the function FC(/~eff) does not vary with rotations of Heff around the
hexagonal axis c because it depends only on two variables:

I/4effl, 0 = a r c c o s -
0qeff~
Heft

If/~effJ-~ then

Fc(/]eff) = FC(IHeffl) = _ [ H e , Mf(x) dz, (A. 13)


J0
536 A.K. ZVEZDIN

(V~) = ~ (M[V66]M')(M']p°[M),
MM ~
where IM) are the eigenfunctions of Iz.
In a crystallographic system of coordinates in which the plane ZOX is the sym-
metry plane, the matrix elements (~r[V66]~r') are known to be real. The sign ..~
indicates that wave functions [M) are taken in the crystallogr~iphic coordinates sys-
tem. However, it is best to calculate (V66) in a system coordinate rotating around the
2 axis relative to the crystallographic system, so that the new axisJ?'l[/goff. Wave
functions in the new IM) and the original system of coordinates [M) are related by
[M) = ei~brM[J~r),
in which ~bf is the angle in the basis plane between 3~ and/~eff. In view of this we
get

(V6) = ~ exp[i(M - M')g)e](M[V661rM')(M'lp°[M ')


MM'

: e i64~r ~ ( 2 ~ f l V 6 6 1 ~ f q- 6) +
M
(A.14)
-1- e -i6~bf Z ( M t -[- 6[V6IM') (M'[p°IM' + 6)
M~

1
= -- - g f ( H e f f ) c o s 6(~bf + a ) ,
6
where
1
g Kf = ~ b 2 '
a
cos 6c~ -
v ~ + l)2
b
sin 6a -
b2'
a = 2Re ~ (~rlv661~r+ 6)(M + Sip°[M),
M
b = 2Im Z ( ~[V6[~r + 6 ) ( M + 6[p°lM).
M
Adding together (A. 1 1), (A. 13) and (A. 14), we then find the required thermodynamic
potential of the nonequilibrium state
~ [Herr
~ ( ] ~ d ) = ~iSd(J~d) -- MdH - Mf(x) d z -
J0
1
-- - Kf(Heff) cos 6(qgf+ a).
6
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 537

References

Akselrad, A. and H. Callen. 1971, Appl. Phys. Belov, K.P., A.I. Goryaga and Li Zhan Da,
Lett. 19(11), 464. 1960b, Zh. Eksp. Theor. Fiz. (Sov. Phys. -
Alben, R., 1970a, Phys. Rev. B 2, 2767. JETP) 38, 1914 (in Russian).
Alben, R., 1970b, Phys. Rev. Lett. 24(2), 68. Belov, K.E, A.I. Goryaga and Li Zhan Da, 1961,
Aleksandryan, V.V., A.S. Lagutin, R.Z. Levitin, Zh. Eksp. Theor. Fiz. (Soy. Phys. -JETP)
A.S. Markosyan and V.V. Snegirev, 1985, 40(3), 752 (in Russian).
Sov. Phys. - JETP 62, 153. Belov, K.P., E.V. Talalaeva, L.A. Chernikova
Aleonard, R., 1960, J. Phys. Chem. Solids and V.I. Ivanovskii, 1968, JETP Lett. 7(11),
15(1-2), 167. 423.
Allain, V., M. Bihara and A. Herpin, 1966, J. Belov, K.E, R.Z. Levitin, B.K. Ponomarev and
Appl. Phys. 37, 1316. Yu.E Popov, 1969, JETP Lett. 10(1), 13.
Amadesi, G., 1971, Phys. Lett. A 35, 289. Belov, K.E, L.A. Chernikova, E.V. Talalaeva,
Anderson, E.E., 1964, Phys. Rev. A 164(6), R.Z. Levitin, T.V. Kudryavzeva, S. Amadezi
1581. and B.I. Ivanovskii, 1970a, Zh. Eksp. Theor.
Antonov, A.V. and V.I. Jilin, 1976, Fiz. Tverd. Fiz. (Soy. Phys. - JETP) 58(6), 1923 (in
Tela (Soy. Phys. Solid State) 18, 1269 (in Russian).
Belov, K.P., R.Z. Levitin and Yu.E Popov,
Russian).
1970b, Zh. Eksp. Theor. Fiz. (Soy. Phys. -
Antonov, A.V., V.I. Jilin and V.I. Ribak, 1976,
JETP) 59(6), 1985 (in Russian).
Fiz. Tverd. Tela (Sov. Phys. Solid State)18,
Belov, K.E and S.A. Nikitin, 1970, Zh. Eksp.
2124 (in Russian).
Theor. Fiz. (Soy. Phys. - JETP) 58, 937 (in
Asomoza, R., I.A. Campbell, H. Jouve and R.
Russian).
Meyer, 1977, J. Appl. Phys. 48(9), 3829.
Belov, K.E, 1972a, Ferrites in Strong Magnetic
Asti, G. 1990, First-order magnetic processes,
Fields (Nauka, Moscow, in Russian).
in: Ferromagnetic Materials, Vol. 5, eds Belov, K.E, A.M. Bisliev, C.A. Nikifin and
K.H.J. Buschow and E.P. Wohlfarth (Elsevier, V.E. Kolesnitchenko, 1972b, Fiz. Met.
Amsterdam) p. 397. Metalloved. 34(3), 470.
Asti, G. and E Bolzoni, 1980, J. Magn. Magn. Belov, K.E, A.K. Zvezdin, A.M. Kadomtseva,
Mater. 20, 29. I.B. Krinetskii and T.L. Ovchinnikova, 1974,
Avaeva, G., EV. Lisovskii and V.I. Schapovalov, Fiz. Tverd. Tela (Soy. Phys. Solid State)
1975, Fiz. Tverd. Tela (Soy. Phys. Solid 16(9), 2615 (in Russian).
State) 17(8), 2488 (in Russian). Belov, K.E, A.K. Zvedin, R.Z. Levitin, A.S.
Babushkin, G.A., A.K. Zvezdin, R.Z. Levitin, Marcosyan, B.V. Mill and A.A. Mukhin,
V.N. Orlov and A.I. Popov, 1983, Sov. Phys. 1975, Zh. Eksp. Theor. Fiz. 68(3), 1189
- JETP 58, 792. (Soy. Phys. - JETP 41, 590).
Ballou, R., B. Gorges, R. Lemaire, H. Rakoto Belov, K.E, A.K. Zvezdin, A.M. Kadomtseva
and J.C. Ousset, 1989, Physica B 155, 266. and R.Z. Levitin, 1976, Usp. Fiz. Nauk (Sov.
Ballou, R., Z.M. Gamishidze, R. Lemaire, R.Z. Adv. Phys.) 119(3), 447 (in Russian).
Levitin, A.S. Markosyan and V.V. Snegirev, Belov, K.P., A.K. Zvezdin, A.M. Kadomtseva
1992, Sov. Phys. - JETP 75, 1041. and R.Z. Levitin, 1979, Spin-Reorientation
Balanda, M. and S. Niziol, 1979, Phys. Status Transitions in Rare-Earth Magnetic Materials
Solidi B: 91, 291. (Nauka, Moscow, in Russian).
Baryakhtar, V.G., A.E. Borovik and V.A. Popov, Belov, K.P., A.K. Zvezdin and A.M. Kadomtseva,
1972, Zh. Eksp. Theor. Phys. (Sov. Phys. - 1987, Phys. Rev., Sov. Sci. Rev. 9, 118.
JETP) 62(6), 2233 (in Russian). Belyaeva, A.I., Yu.N. Stelmakhov and V.A.
Baryakhtar, V.G., E.P. Stephanovskii and D.A. Potakova, 1977, Fiz. Tverd. Tela (Sov. Phys.
Yablonskii, 1976, Fiz. Met. Metalloved. Solid State) 19(10), 3124 (in Russian).
42(4), 684. Berezin, A.G., R.Z. Levitin and Yu.E Popov,
Belov, K.P. and Ped'ko, 1960a, Zh. Eksp. Theor. 1980, Sov. Phys. - JETP 52(1), 135.
Fiz. (Soy. Phys. - JETP) 39(4), 961 (in Bernasconi J. and D. Kuse, 1971, Phys. Rev. B
Russian). 3(3), 811.
538 A.K. ZVEZDIN

Bertaut, E.E and E Forrat, 1956, Compt. Rend. Decrop, B., J. Deportes, D. Givord and R.
242(3), 382. Lemaire, 1982, Appl. Phys. 53(3), 1953.
Bisliev, A.M., A.K. Zvezdin, D. Kim, S.A. Demidov, V.G. and R.Z. Levitin, 1977, Zh.
Nikitin and A.F. Popkov, 1973, Pis'ma Zh. Eksp. Theor. Phys. 72, 1111 (Sov. Phys. -
Eksp. Teor. Fiz. (JETP Lett.) 17, 484 (in JETP 45, 581).
Russian). Derkachenko, V.N., A.M. Kadomtseva and V.A.
Bloch, D. and R. Lemaire, 1970, Phys. Rev. B Timofeeva, 1974, Pis'ma Zh. Eksp. Teor.
2, 2648. Phys. (JETP Lett.) 20, 236 (in Russian).
Bloch, D., D.M. Edwards, M. Shimizu and J. Derkachenko, V.N., A . K . Zvezdin, A.M.
Voiron, 1975, J. Phys. F 5, 1217. Kadomtseva, N.M. Kovtun, M.M. Lukina and
Blume, M., L.M. Corliss, J.M. Hastings et al., A.A. Mukhin, 1984, Phys. Status Solidi A:
1974, Phys. Rev. Lett. 32, 544. 84, 215.
Borovik-Romanov, A.S., 1959, Zh. Eksp. Theor. Deryagin, A.V., 1976, Usp. Fiz. Nauk 120, 393
Fiz. (Sov. Phys. -JETP) 36(6), 1954 (in (Sov. Phys. Usp., 1977, 19, 909).
Russian). Dichenko, A.B. and V.V. Nikolaev, 1979, Fiz.
Buschow, K.H.J., 1980, in: Ferromagnetic Met. Metalloved. 48(6), 1173.
Materials, Vol. 1, eds E.P. Wohlfarth and Dichenko, A.B., V.V. Nikolaev and E.A. Turov,
K.H.J. Buschow (North-Holland, Amsterdam) 1983, J. Magn. Magn. Mater. 3!(34), 1031.
p. 298. Dikstein, I.E., EV. Lisovskii, E.G. Mansvetova
Buschow, K.H.J., 1988, in: Ferromagnetic and V.V. Tarasenko, 1980, Sov. Phys. - JETP
Materials, Vol. 4 (North-Holland, Amsterdam) 52(2), 260.
p. 1. Dikstein, I.E., EV. Lisovskii, E.G. Mansvetova
Buschow, K.H.J., D.B. de Mooil, X.P. Zhong and V.V. Tarasenko, 1983, Fiz. Tverd. Tela
and F.R. de Boer, 1990, Physica B 162, 183. (Sov. P h y s . Solid State) 25, 3591 (in
Callen, H.B., 1971, Appl. Phys. Lett. 18(7), Russian).
311. Dikstein, I.E., EV. Lisovskii, E.G. Mansvetova
Clark, A.E. and E.R. Callen, 1968, J. Appl. Phys. and V.V. Tarasenko, 1984, Zh. Eksp. Theor.
39, 5972. Phys. (Sov. Phys. - JETP) 86, 1473 (in
Clark, A.E. and E. Callen, 1969, Phys. Rev. Russian).
Lett. 23, 307. Dikstein, I.E., 1991, Thesis (Insitute of
Clark, A.E., U. Belson and N. Tamagava, 1974, Radioengineering and Electronics of RAS,
Proc. ICM-1973 Conf., Vol. 6 (Nauka, Moscow, in Russian).
Moskva) p. 335. Dillon, J.E, E. Yi Chen and W.E Wolf, 1974,
Cyrot, M. and M. Lavagna, 1979, J. Phys. (Paris) in: Proc. ICM-1973 Conf., Vol. 6 (Nauka,
40, 763. Moskva) p. 38.
Date, M., 1990, J. Magn. Magn. Mater. 90-91, Druzhinin, V.V., S.P. Zapasskii and V.M.
1. Povyshev, 1977, Fiz. Tverd. Tela 19, 159
de Boer, ER., Huang Ying-kai, D.B. de Mooij (Sov. Phys. Solid State 19, 90).
and K.H.J. Buschow, 1987, J. Less-Common Druzhinin, V.V., A.I. Pavlovskii, G.S. Krinchik,
Met. 135, 199. O.M. Tatsenko, M.I. Dolotenko, N.P. Kolokol-
de Boer, ER., Huang Ying-kai, Zhang Zhi-dong, chikov and A.I. Bikov, 1981, Fiz. Tverd.
D.B. Mooij de and K.H.J. Buschow, 1988a, Tela (Sov. Phys. Solid State) 23(11), 3470
J. Magn. Magn. Mater. 73, 167. (in Russian).
de Boer, ER., R. Verhoef, D.B. Mooij de, Zhang Druzhinin, V.V., A.I. Pavlovskii, V.V. Platonov,
Zhi-dong and K.H.J. Buschow, 1988b, J. O.M. Tatsenko, and A.S. Lagutin, 1992,
Magn. Magn. Mater. 73, 263. Physica B 177, 315.
de Boer, ER., X.P. Zhong, K.H.J. Buschow and Drzazga, Z., 1981, J. Magn. Magn. Mater. 25,
T.H. Jacobs, 1990, J. Magn. Magn. Mater. 11.
90-91, 25. Drzazga, Z. and M. Drzazga, 1987, J. Magn.
de Boer, ER. and K.H.J. Buschow, 1992, Physica Magn. Mater. 65, 21.
B 177, 199. Duc, N.H., T.D. Hien, P.E. Brommer, J.J. Franse
de Boer, ER. and Z.G. Zhao, 1994, in: and W. Gaben, 1988, J. Phys. F 18, 275.
Programme of the PHMF'94, Nijmegen, Dworschak, G. and Y. Khan, 1974, J. Phys.
PA-14 (The Netherlands, Nijmegen). Chem. Solids 35, 1021.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 539

Eremenko, V.V. and N.E Kharchenko, 1977, Fiz. Gnatchenko, S.L., N.E Kharchenko, O.M.
Tved. Tela (Soy. Phys. Solid State) 19(7), Konovalov and V.M. Pusikov, 1977, Ukr. Fiz.
2210 (in Russian). Zh. 22(4), 555 (in Russian).
Eremenko, V.V. and N.F. Kharchenko, 1979, Gnatchenko, S.L., 1989, Thesis (Physics and
Phase Transitions 1, 61; 2 (1980), 207. Technics Institute of Low Temperatures,
Ermolenko, A.S., Ye.V. Rozenfel'd, Yu.P. Irkhin, Khar'kov, in Russian).
V.V. Kelarev, A.E Rozhda, S.K. Sidorov, Goranskii, B.P. and A.K. Zvezdin, 1969a, Zh.
A.N. Pirogov and A.P. Vokhmyanin, 1975, Eksp. Theor. Fiz. (Sov. Phys. -JETP)
57(2), 547; Sov. Phys. -JETP, 1970, 30,
Zh. Eksp. Theor. Fiz. (Sov. Phys. -JETP)
299.
69, 1743 (in Russian). Goranskii, B.P. and A.K. Zvezdin, 1969b, Pis'ma
Eschenfelder, A.N., 1981, Magnetic Bubble Zh. Eksp. Teor. Phys. 10, 196. (JETP Lett.
Technology (Springer, Berlin). 10, 124 (1972)).
Feron, J.L., G. Fillion and G. Hug, 1971, Z. Goto, T., K. Fukamichi, T. Sakakibara and H.
Angew. Phys. 32(3), 219. Komatsu, 1989, Solid State Commun. 72,
Feron, J.L., G. Fillion, G. Hug, A. Berton and 945.
J. Gaussy, 1972, Solid State Commun. 10(7), Grjegorjevsldi, O.L. and R.V. Pisarev, 1973, Zh.
641. Eksp. Theor. Phys. (Sov. Phys. -JETP)
Feron, J.L., G. Fillion, G. Hug and P. Morin, 65(2), 633 (in Russian).
1974, in: Proc. ICM-1973, Vol. 1 (Nauka, Guillot, M. and H. Le Gall, 1976, Phys. Status
Moskva) p. 330. Solidi B: 77, 121.
Gurtovoy, K. G., A.S. Lagutin and V.I. Ozhogin,
Fillion, G., 1974, Thesis, Grenoble.
1980, Zh. Eksp. Theor. Phys. (Sov. Phys. -
Fisch, G. L., B.P. Khrustalev, GT Frolov and JETP) 78, 847 (in Russian).
V.Ya. Yakovchuk, 1986, Fiz. Tved. Tela Gusev, A.A., 1959, Sov. Phys. Crystallogr.
(Sov. Phys. Solid State) 28(7), 2205 (in 4(4), 695.
Russian). Gusev, A.A. and A.S. Pakhomov, 1963, Soy.
Forestier, H. and G. Guiot-Gullion, 1950, Compt. Phys. Crystallogr. 8(1), 63.
Rend. p. 1844. Gyorgy, E.M., A. Rosencwaig, E.I. Blount, W.J.
Forestier, H. and G. Guiot-Gullion, 1952, Compt. Tabor and M.E. Lines, 1971, Appl. Phys.
Rend. 235, 48. Lett. 18(11), 478.
Franse, J.J.M., R.J. Radwanski and R. Verhoef, Hansen, P. and J.P. Krumme, 1973, J. Appl.
1990, J. Magn. Magn. Mater. 84, 2999. Phys. 44, 3805.
Franse, J.J.M., 1990, J. Magn. Magn. Mater. Hanton, J.P., 1967, IEEE Trans. Magn.
MAG-3(3), 505.
90-91, 20.
Hug, G. and P. Morin, 1971, J. Phys. (Paris) 32
Geller, S., 1960, J. Appl. Phys. 31(5), 258. Suppl. C1, 15.
Geller, S. and M.A. Gilleo, 1957, J. Phys. Chem. Hug, G., 1972, Thesis, Grenoble.
Solids 3, 30. Irkhin, Yu.P. and E.V. Rozenfeld, 1974, Soy.
Geller, S. and M.A. Gilleo, 1960, Acta Cryst. Phys. Solid State 16(2), 310.
10(1), 239. Jacobs, T.H., K.H.J. Buschow, G.E Zhou and
Gilleo, M.A. and S. Geller, 1958, Phys. Rev. ER. de Boer, 1992a, Physica B 179, 177.
110(1), 73. Jacobs, T.H., K.H.J. Buschow, G.E Zhou, X. Li
Ginsburg, V.L., 1960, Fiz. Tverd. Tela (Sov. and F.R. de Boer, 1992b, J. Magn. Magn.
Phys. Solid State) 2, 2031 (in Russian). Mater. 116, 220.
Ginsburg, V.L., 1981, Soy. Phys. Usp. 24, 585. Jacobs, T.H., K.H.J. Buschow, G.E Zhou, J.P.
Liu, X. Li and ER. de Boer, 1992c, J. Magn.
Giordano, V. and W.P. Wolf, 1977, Phys Rev.
Magn. Mater. 104-107, 1275.
Lett. 39, 342.
Kadomtseva, A.M., I.B. Krynetskii, M.D.
Givord, D., H.S. Li, J.M. Cadogan, J.M.D. Coey, Kuzmin and A.K. Zvezdin, 1989, J. Magn.
J.P. Gavigan, O. Yamada, H. Maruyama, M. Magn. Mater. 81, 196.
Sagawa and S. Hirosawa, 1988, J. App. Phys. Kaganov, M.I., 1980, Sov. Phys. -JETP 52(4),
63, 3713. 779.
Gnatchenko, S.L. and N.F. Kharchenko, 1976, Kaganov, M.I. and A.V. Chubukov, 1982, Zh.
Zh. Eksp. Theor. Fiz. 70, 1379 (Soy. Phys. Eksp. Theor. Phys. (Sov. Phys.- JETP)82,
- JETP 42, 719) (in Russian). 1617 (in Russian).
540 A.K. ZVEZDIN

Kamilov, I.K. and G.M. Schachschaev, 1972, Krynetskii, I., H. Szymczak, R. Szymczak and
Pis'ma Zh. Eksp. Teor. Phys. (JETP Lett.) J. Zawadzki, 1994, J. Magn. Magn. Mater.
15(8), 480 (in Russian). 129, 322.
Kamilov, I.K., H.K. Aliev, G.M. Schachschaev, Krynetskii, I., R. Szymczak, J. Zawadzki, A.K.
G.G. Musaev and M.M. Magomedov, 1975, Zvezdin and S. Piechota, 1995, J. Magn.
Zh. Eksp. Theor. Phys. (Sov. Phys. -JETP) Magn. Mater. 140-144, 939.
68(6), 2290 (in Russian). Kurtzig, A.J. and EB. Hagedorn, 1971, IEEE
Kamdaurova, G.S., V.O. Vas'kovskii and V.V. Trans. Magn. MAG-7(3), 473.
Lesnich, 1982, Fiz. Met. Metalloved. 53(4), Kuz'ma, Y.B. and N.S. Bilonizhko, 1974, Sov.
713. Phys. Crystallogr. 18, 447.
Kandaurova, G.S., V.O. Vas'kovskii, A.A. Landau, L.D. and E.M. Lifshitz, 1964, Statistical
Kazakov and V.V. Lesnich, 1985, Fiz. Met. Physics (Gostekhizdat, Moscow).
Metalloved. 60(4), 718. Landau, D.P., B.E. Keen, B. Schneider and W.P.
Kaneyosi, T., 1991a, J. Phys. Condens. Mater. Wolf, 1971, Phys. Rev. B 3(7), 2310.
3, 4497. Lagutin, A.S., 1993, Thesis (Kurchatov Institute
Kaneyosi, T., 1991b, Introduction to Surface of Atomic Energy, Moscow, in Russian).
Magnetism (CRC Press, Boca Raton, FL). Lagutin, A.S. and A.V. Dmitriev, 1990, J. Magn.
Khalturin, V.I., 1976, Fiz. Met. Metalloved. Magn. Mater. 90-91, 83.
41(2), 271. Lagutin, A.S. and R.E Druzhinina, 1990, J.
Kharchenko, N.F., V.V. Eremenko and L.I. Belii, Magn. Magn. Mater. 90-91, 85.
1968, Sov. Phys. - JETP 26, 869. Lemaire, R. and T. Schweizer, 1966, Phys. Lett.
Kharchenko, N.E, V.V. Eremenko and S.L. 21, 366.
Gnatchenko, 1974, Pis'ma Zh. Eksp. Teor. Leontovich, M.A., 1944, Statisticheskaja Fizika
Phys. (JETP Lett.) 20, 612 (in Russian). (Statistical Physics) (Gostekhizdat, Moscow,
in Russian).
Kharchenko, N.F., V.V. Eremenko, S.L. Gnat-
Levanyuk, A.P., 1959, Zh. Eksp. Theor. Fiz.
chenko, L.I. Belyi and E.M. Kabanova,
36, 810 (in Russian).
1975a, Zh. Eksp. Teor. Fiz. 68, 1073 (Sov.
Levanyuk, A.P., V.V. Osipov, A.S. Sigov and
Phys. - JETP, 1975, 41, 531).
A.A. Sobyanin, 1979, Sov. Phys. - JETP 50,
Kharchenko, N.F., V.V. Eremenko and S.L.
512.
Gnatchenko, 1975b, Zh. Eksp. Teor. Fiz. Levina, S.S., B.N. Novogrudskii and I.G.
69(5), 1697 (in Russian). Fakidov, 1963, Zh. Eksp. Theor. Fiz. (Soy.
Kharchenko, N.F., V.V. Eremenko and L.I. Belii, Phys. - JETP) 45(1), 52 (in Russian).
1978, Zh. Eksp. Teor. Fiz. 28(7), 351 (in Levitin, R.Z., B.K. Ponomarev and Yu. F.
Russian). Popov, 1970, Zh. Eksp. Theor. Fiz. (Sov.
Khrustalev, B.P., V.G. Pozdnyakov, G.I. Frolov Phys. - JETP) 59, 1952 (in Russian).
and V.Yu. Yakovchuk, 1989, Fiz. Tverd. Levitin, R.Z. and Yu.E Popov, 1975, in:
Tela (Sov. Phys. Solid State) 31(3), 112 (in Ferromagnetism, ed. K.P. Belov (Moscow
Russian). State University Press, Moscow) p. 6 (in
Khrustalev, B.P. and V.G. Pozdnyakov, 1993a, Russian).
Solid State Commun. 85(9), 763. Levitin, R.Z. and A.S. Markosyan, 1988, Sov.
Khrustalev, B.R, V.G. Pozdnyakov and G.I. Phys. Usp. 31, 730.
Frolov, 1993b, Fiz. Tverd. Tela (Sov. Phys. Li, H.S., J.P. Gavigan, J.M. Cadogan, D. Givord
Solid State) 35(4), 921 (in Russian). and J.M.D. Coey, 1988, J. Magn. Magn.
King, A.R. and D. Paquette, 1973, Phys. Rev. Mater. 72, L241.
Lett. 30(14), 662. Lisovskii, F.V. and V.I. Shapovalov, 1974,
Kirchmayr, H.R. and C.A. Poldy, 1979, in: Pis'ma Zh. Eksp. Teor. Phys. 20, 128,
Handbook on Physics and Chemistry of Rare (JETP Lett. 20, 55).
Earths, eds K.A. Gshneidner and J. Eyring, Lisovskii, F.V., E.G. Mansvetova, V.V. Tarasenko
(North-Holland, Amsterdam) p. 55. and V.V. Schapovalov, 1976a, Fiz. Tverd.
Kittel, C., 1949, Rev. Mod. Phys. 21, 611. Tela (Soy. Phys. Solid State) 17, 1729 (in
Krinchik, G.S. and M.V. Chetkin, 1969, Usp. Russian).
Fiz. Nauk 98(1), 3 (in Russian). Lisovskii, EV., E.G. Mansvetova and V.I.
Krumme, J.P. and P. Hansen, 1973, Appl. Phys. Schapovalov, 1976b, Fiz. Tverd. Tela (Sov.
Lett. 22(7), 312. Phys. Solid State) 17(11), 3504 (in Russian).
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 541

Lisovskii, F.V., E.G. Mansvetova and V.I. N6el, L., 1936, Compt. Rend. 203, 304.
Shapovalov, 1976c, Zh. Eksp. Theor. Phys. N6el, L., 1948, Ann. Phys. (Paris) 3(3-4), 137.
(Sov. Phys. - JETP) 71(4), 1443 (in Russian). N6el, L., 1954, Compt. Rend. 239(1), 8.
Lisovskii, EB., E.G. Mansvetova and V.I. Nikitin, S.A., D. Kim, A.K. Zvezdin and A.E
Shapovalov, 1980, Zh. Tech. Fiz. 50(1), 198 Popkov, 1975, Pis'ma Zh. Eksp. Teor. Fiz.
(in Russian). 22(5), 297 (Soy. JETP Lett. 22, 137).
Lisovskii, F.V., 1980, Thesis (Institute of Novogrudskii, B.N. and I.G. Fakidov, 1964, Zh.
Radioiengineering and Electronics of RAS, Eksp. Theor. Fiz. (Soy. Phys. - J E T P )
Moscow, in Russian). 47(1), 40 (in Russian).
Liu, J.P., K. Bakker, ER. de Boer, T.H. Jacobs Nowik, I. and J.H. Wernick, 1965, Phys. Rev.
and K.H.J. Buschow, 1991a, J. Less-Common A 140, 131.
Met. 170, 109.
Okamoto, K., T. Shirakawa, S. Matsushuta et al.,
Liu, J.P., ER. de Boer and K.H.J. Buschow,
1974, IEEE Trans. Magn. MAG-10(3), 799.
1991b, J. Magn. Magn. Mater. 98, 291.
Pakhomov, A.S. and A.A. Gusev, 1964, Phys.
Liu, J.E, ER. de Boer and K.H.J. Buschow,
Met. Metalloved. 18(I), 156.
1991c, J. Less-Common Met. 175, 137.
Liu, J.P., X.P. Zhong and ER. de Boer, 1991d, Pakhomov, A.S. and V.V. Gerbutov 1971, Fiz.
J. Appl. Phys. 69, 5536. Tverd. Tela (Soy. Phys. Solid State)14(1),
Liu, J.P., ER. de Boer, P.E de Ch,~tel, R. 10 (in Russian).
Coehoorn and K.H.J. Buschow, 1994, J. Pauthenet, R., 1958a, Ann. Phys. (Paris) 3(5~5),
Magn. Magn. Mater. 132, 159. 424
Lutes, C.S., J.O. Holmen, R.L. Kooyer and Pauthenet, R., 1958b, J. Appl. Phys. 29(3), 253.
O.S. Aadland, 1977, IEEE Trans. Magn. Pavlovskii, A.I., N.P. Kolokolchikov, V.V.
MAG-3(5), 1615. Druzhinin, O.M. Tatsenko and A.I. Bikov,
Makarov, V.V., V.I. Karpovich and B.I. Lukin, M.I. Dolotenko, 1979, JETP Lett. 30, 211.
1980, Fiz. Tverd. Tela (Soy. Phys. Solid Pavlovskii, A.I., V.V. Druzhinin, A.S. Lagutin,
State) 22(8), 2495 (in Russian). V.V. Platonov and O.M. Tatsenko, 1992, Fiz.
Matsuuva, M., Y. Okuma, M. Morotomi, H. Tverd. Tela (Soy. Phys. Solid State) 35(12),
Mollymoto and M. Date, 1979, J. Phys. Soc. 3755 (in Russian).
Jpn 46(3), 1031. Pearson, R.F., 1962, J. Appl. Phys. 33(3), 1236.
McCuire, T.R., R.J. Gambino and R.C. Taylor, Pirogov, A.N. and A.P. Vokhmyanin, 1975, Zh.
1977, J. Appl. Phys. 48(7), 2965. Eksp. Teor. Fiz. 69, 1743 (Sov. Phys. -
Merkulov, B.S., E.G. Rudashevskiy, A. Le Gall JETP 42, 885).
and K. Legcuras, 1981, Zh. Eksp. Theor. Pisarev, R.V., I.G. Sinii and G.A. Smolenskii,
Fiz. (Sov. Phys. - JETP) 80(1), 161 (in 1969, Zh. Eksp. Theor. Fiz. (Soy. Phys. -
Russian). JETP) 57(3), 737.
Mitchell, I.V., J.M.D. Coey, D. Givord, I.R. Pisarev, R.V., I.G. Sinii, N.N. Kolpakova and
Harris and R. Hantisch, eds, 1989, Concerted Yu.M. Yakovlev, 1971, Zh. Eksp. Theor.
European Action on Magnets (CEAM) Fiz. (Sov. Phys. - JETP) 60, 2188 (in
(Elsevier, London). Russian).
Mitsek, A.I., N.P. Kolmakova and P.E Gaidanskiy, Popkov, A.E, 1976a, Fiz. Tverd. Tela (Sov.
1969, Fiz. Tverd. Tela 11(5), 1258 (Sov. Phys. Solid State) 18, 357 (in Russian).
Phys. Solid State 11, 1021).
Popkov, A.E, 1976b, Fiz. Tverd. Tela (Sov.
Mitsek, A.I. and P.M. Serebryanik, 1976, Fiz.
Phys. Solid State) 18, 2585 (in Russian).
Tverd. Tela (Sov. Phys. Solid State) 18(6),
Popov, G.V., V.A. Seredkin and V.Yu. Yakov-
1716.
Miura, N., G. Kido, I. Oguro, K. Kawauchi, S. chuk, 1990, Fiz. Met. Metalloved. (2), 61.
Chikazumi, J.E. Dillon and L.G. Uitert, 1977, Popov, Yu.E, 1971, Thesis (Moscow State
Physica 86--88B, 1219. University, in Russian).
MOiler, H.R., W. Kelling, R. Koscik, P. Poulis, N.J., J. van den Hendel and J. Ubbink,
Rosemann and Z. Frait, 1978, Phys. Status 1951, Phys. Rev. 82, 552.
Solidi A: 50(2), 537. Radwanski, R.J., ER. de Boer, J.J.M. Franse and
Murata, K., K. Fukamichi, H. Komatsu, T. K.H.J. Buschow, 1989a, Physica B 159, 311.
Sakakibura and T. Goto, 1991, J. Phys.: Radwanski, R.J. and J.J.M. Franse, 1989,
Condens. Matter. 3, 2515. Physica B 13(154), 181.
542 A.K. ZVEZDIN

Radwanski, R.J., R. Verhoef and J.J.M. Franse, Turov, E.A., V.G. Schavrov and Yu.P. Irchin,
1990a, J. Magn. Magn. Mater. 83, 141. 1964, Zh. Eksp. Theor. Fiz. (Sov. Phys. -
Radwanski, X.P. Zhong, ER. de Boer and K.H.J. JETP) 47(1), 296 (in Russian).
Buschow, 1990b, Physica B 164, 131. Turov, E.A., 1987, Zh. Eksp. Theor. Fiz. (Soy.
Ratajczak, H. and I. Goscianska, 1980, Phys. Phys. - JETP) 92(5), 1886 (in Russian).
Status Solidi A: 62(1), 163. Tyablikov, S.V., 1956, Phys. Met. Metalloved.
Rode, V.E. and A.V. Vedyaev, 1963, Sov. Phys. 3(1), 3.
- JETP 18, 286. Tyablikov, S.V., 1958, Phys. Met. Metalloved.
8(I), 152.
Rosencwaig, A., W.J. Tabor and R.D. Pierce,
Tyablikov, S.V., 1965, Methods of Quantum
1971, Phys. Rev. Lett. 26(10), 779.
Theory of Magnetism (Nauka, Moscow, in
Rudashevsky, E.G., C. Legcuras and H. Le Gall, Russian).
1977, Solid State Commun. 24(11), 587. Tyablikov, S.V., 1967, Methods in the Quantum
Sannikov, D.G. and T.M. Perekalina, 1969, Zh. Theory of Magnetism (Plenum, New York).
Eksp. Theor. Fiz. (Sov. Phys. -JETP)56, Valiev, U.V., G.S. Krinchik, R.Z. Levitin and
730 (in Russian). K.M. Mukimov, 1979, JETP Lett. 29, 214.
Schavrov, V.G. and E.A. Turov, 1963, Zh. Eksp. Velge, W.AJ.J. and K.H.J. Buschow, 1968, J.
Theor. Fiz. (Soy. P h y s . - JETP)45(2), 349 Appl. Phys. 39, 1717.
(in Russian). Veselago, V.G., L.P. Maksimov and A.M.
Schavrov, V.G., 1965, Zh. Eksp. Theor. Fiz. Prokhorov, 1968, Vest. Akad. Nauk USSR
(Sov. Phys. - JETP) 48(5), 1419 (in Russian). 12, 58 (in Russian).
Silant'ev, V.I., A.I. Popov, R.Z. Levitin and A.K. Vedernikov, N.E, A.K. Zvezdin, R.Z. Levitin
Zvezdin, 1980, Sov. Phys. - JETP 51, 323. and A.I. Popov, 1988, Zh. Eksp. Theor. Fiz.
ShRimann, E., 1960, in: Solid State Physics, 94(8), 358 (Sov. Phys. - JETP 66(6), 1233).
Electronics and Telecommunications, Vol. 3, Verhoef, R., 1990, Thesis, Magnetic Interactions
Part I (Academic Press, New York) p. 322. in R2Fe14B and Some Other R-T Inter-
Sinnema, S., R.J. Radwanski, J.J.M. Franse, metallics, Universiteit van Amsterdam, The
Netherlands.
D.B. de Mooij and K.H.J. Buschow, 1984,
Verhoef, R., ER. de Boer, J.J.M. Franse, CJ.M.
J. Magn. Magn. Mater. 44, 333. Denissen, T.H. Jacobs and K.H.J. Buschow,
Sinnema, S, J.J.M. Franse, R.J. Radwanski, A. 1989, J. Magn. Magn. Mater. 80, 41.
Menovski and ER. de Boer, 1987, J. Phys. Verhoef,, R., R.J. Radwanski and J.J.M. Franse,
F: 17, 233. 1990a, J. Magn. Magn. Mater. 89, 176.
Smirnova, E.L., V.I. Smirnov, Ju.I. Ukhanov and Verhoef, R., EH. Quang, J.J.M. Franse and R.J.
V.A. Petrusevich, 1970, JETP Lett. 11(9). Radwanski, 1990b, J. Magn. Magn. Mater.
Smit, H.H.A., R.C. Thiel and K.H.J. Buschow, 83, 139.
1988, J. Phys. F: 18, 295. Verhoef, R., RH. Quang, J.J.M. Franse and R.J.
Stacy, W.T. and C.J. Rooymans, 1971, Solid Radwanski, 1990c, J. Appl. Phys. 67, 4771.
State Commun. 9(10), 2005. Visnovsky, S., V. Prosser, E. Zvara and P.
Steiner, W., E. Gratz, H. Ortbauer and W. Gaben, Polivka, 1974, Phys. Status Solidi A: 26(2),
1978, J. Phys. F: 5, 1525. 513.
Szewczyk, A., A. Dabkovski, S.L. Gnatchenko, Vlasko-Vlasov, V.K., L.M. Dedukh and V.I.
N.E Kharchenko and H.M. Krsymanska, Nikitenko, 1981, Fiz. Tverd. Tela 23, 1857
1985, Acta Phys. Pol. A 67(5), 877. (Soy. Phys. Solid State 23, 1085).
Szymczak, R., 1983, J. Magn. Magn. Mater. Vlasko-Vlasov, V.K., L.M. Dedukh, M.V.
Indenbom and V.I. Nikitenko, Sov. Phys. -
35(1-3), 243.
JETP 57(1), 1983.
Szymczak, R., 1987, J. Magn. Soc. Jpn 11
Vlasko-Vlasov, V.K. and M.V. Indenbom, 1984,
Supplement S1, 39.
Zh. Exsp. Theor. Fiz. (Sov. Phys. -JETP)
Tagirov, R.I. and A.A. Glazer, 1978, Fiz. Met. 86(3), 1084 (in Russian).
Metalloved. 46(1), 75. Vlasov, K.B., B.N. Novogrudskii and E.A.
Taylor, K., 1975, Adv. Phys. 24, 681. Rosenberg, 1976, Fiz. M e t . Metalloved.
Taylor, A.S. and A. Gangulee, 1976, J. Appl. 42(3), 513.
Phys. 47, 4666. Vlasov, K.B., E.A. Rosenberg and A.M.
Toxen, A.M., T.N. Geballe and R.M. White, Burchanov, 1980, Fiz. Met. Metalloved.
1988, J. Appl. Phys. 64(10), 5431. 49(1), 1005.
FIELD INDUCED PHASE TRANSITIONS IN FERRIMAGNETS 543

Vlasov, K.B. and E.A. Rosenberg, 1982, Fiz. Zvezdin, A.K. and V.M. Matveev, 1972b, Izv.
Met. Metalloved. 53(3), 493. Akad. Nauk SSSR, Ser. Fiz. 36, 1441 (in
Vonsovsh."i, S.V., 1971, Magnetism (Nauka, Russian).
Moscow).
Zvezdin, A.K. and S.G. Kalenkov, 1972, Fiz.
Wohlfarth, E.P. and P. Rhodes, 1962, Philos.
Tverd. Tela (Sov. Phys. Solid State)14(10),
Mag. 7, 1817.
Yamada, H., J. Inoue, K. Terao and H. Komatsu, 2835 (in Russian).
1984, J. Phys. F: 14, 1943. Zvezdin, A.K. and A.E Popkov, 1974, Fiz.
Yamada, M., H. Kato, H. Yamamoto and Y. Tverd. Tela (Sov. Phys. Solid State)16(4),
Nakagawa, 1988, Phys. Rev. B 38, 620. 1082 (in Russian).
Zawadzki, J., R. Ballou, R. Gorges and R. Zvezdin, A.K., A.A. Mukhin and A.I. Popov,
Szymczak, 1993, Acta Phys. Pol. A 83, 209. 1976, Pis'ma Zh. Eksp. Teor. Fiz. (JETP
Zawadzki, J., R. Szymczak, A.K. Zvezdin and Lett.) 23(5), 267.
I.B. Krynetski, 1994, IEEE Trans. Magn. 30,
Zvezdin, A.K., A.A. Mukhin and A.I. Popov,
866.
Zhao, Z.G., 1994, Thesis, Magnetization 1977, Zh. Eksp. Theor. Fiz. 72, 1097 (Sov.
Processes of Ferrirnagnets, van der Waals- Phys. - JETP 45, 573).
Zeeman Laboratorium, Universiteit van Zvezdin, A.K. and A.E Popkov, 1977, Fiz. Met.
Amsterdam, The Netherlands. Metalloved. 44, 689.
Zhong, X.P., R.J. Radwanski, ER. de Boer, T.H. Zvezdin, A.K. and A.F. Popkov, 1980, Fiz. Met.
Jacobs and K.H.J. Buschow, 1990a, J. Magn. Metalloved. 50(2), 11.
Magn. Mater. 86, 333.
Zvezdin, A.K., Zorin I.A., Kadomtseva A.M.,
Zhong, X.P., ER. de Boer, R.J. Radwanski, T.H.
Krynetskii I.B., Moskvin A.S. and Mukhin
Jacobs and K.H.J. Buschow, 1990b, J. Magn.
Magn. Mater. 92, 46. A.A., 1985, Soy. Phys. - JETP 61(3), 645.
Zhong, X.P., F.R. de Boer, D.B. de Mooij and Zvezdin, A.K., V.M. Matveev, A.A. Mukhin
K.H.J. Buschow, 1990c, J. Less-Common and A.I. Popov, 1985, Rare Earth Ions
Met. 163, 123. in Magnetically Ordered Crystals (Nauka,
Zhou, G.E, X. Li, ER. de Boer and K.H.J. Moscow, in Russian).
Buschow, 1992a, J. Magn. Magn. Mater. Zvezdin, A.K., G.V. Sayko and S.N. Utochkin,
109, 265. 1991, Fiz. Tverd. Tela (Soy. Phys. Solid
Zhou, G.F., ER. de Boer and K.H.J. Buschow, State) 33, 3175 (in Russian).
1992b, J. Alloys Comp. 187, 299.
Zvezdin, A.K., 1993, JETP Lett. 58(9-10), 719. Zvezdin, A.K. and S.N. Utochkin, 1992, J. Magn.
Zvezdin, A.K. and V.M. Matveev, 1972a, Zh. Magn. Mater. 104-107, 1479.
Eksp. Teor. Fiz. 62, 260 (Sov. Phys. -JETP Zvezdin, A.K. and L.R. Evangelista, 1995,
62, 140). J. Magn. Magn. Mater. 140-144.
chapter 5

PHOTON BEAM STUDIES OF


MAGNETIC MATERIALS

Stephen W. Lovesey
ISIS Facility
Rutherford Appleton Laboratory
Oxfordshire OX11 OQX
U.K.

Handbook of Magnetic Materials, Vol. 9


Edited by K. H.J. Buschow
©1995 Elsevier Science B.V. All rights reserved

545
CONTENTS

1. Prologue ................................................................... 547


2. Orientation ................................................................. 550
3. Survey of experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 558
3.1. Dichroism ............................................................. 559
3.2. Diffraction ............................................................ 565
3.3. Elastic resonant scattering ................................................ 571
3.4. Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 576
4. K r a m e r s - H e i s e n b e r g amplitude . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 579
5. Scattering by orbital m a g n e t i s m ................................................ 581
6. Scattering by spin m a g n e t i s m .................................................. 583
7. Dichroisrn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 585
7.1. Circular dicbroism ...................................................... 592
7.2. Linear dichroism ....................................................... 595
8. Diffraction ................................................................. 598
8.1. Unpolarized primary b e a m ............................................... 599
8.2. Linear polarization ...................................................... 604
8.3. Circular polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 609
9. Elastic resonant scattering ..................................................... 610
10. Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 613
10.1. Scattering by free charges ................................................ 613
10.2. B o u n d electrons ........................................................ 616
10.3. C o m p t o n scattering ..................................................... 617
11. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 618
Acknowledgements .............................................................. 619
Appendix. Polarization effects and magnetic scattering amplitude . . . . . . . . . . . . . . . . . . . . . . . . 619
Polarization states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 619
Scattering amplitude . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 622
List of important s y m b o l s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 625
Numerical values of units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 626
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 626

546
1. Prologue

The past decade has seen a surge of activity in the use of photon beam techniques
to study magnetic properties of materials. By and large, in this period the experi-
mental work has been accomplished with beams produced by electron synchrotron
facilities. At the time of writing, several new major facilities are ready for routine
operation, while other facilities are being steadily improved. So, it seems, the appli-
cation to magnetic materials is likely to remain a burgeoning activity for some more
years to come. To date, it is fair to say that the surge of activity is underpinned
by improvements in instrument performance, enjoyed at synchrotron sources, rather
than outstanding intellectual advances. Improvements over conventional laboratory
X-ray generators include (Margaritondo 1988, Gerson et al. 1992): a high brightness
with the concomitant option of superior resolution; a high degree of linear polar-
ization; tuneability of the primary photon energy; and the provision of good beams
of circularly polarized photons. Observable effects due to magnetic properties of a
sample are usually relatively small compared with charge induced effects, e.g., in
diffraction experiments magnetic intensities are typically five orders of magnitude
smaller than Thomson scattering. In consequence, improvements in the intensity at
the sample, and the provision of good beams of polarized photons, which enable
polarization induced discrimination effects to be exploited as a means of increasing
the signal-to-noise ratio, are particularly significant in the use for magnetic studies
of photon beam techniques.
Broadly speaking, it is useful to consider applications of photon beam techniques
in one of two regimes of the primary photon energy. These are the limits of high
TABLE 1
Possible (El) dipole transitions and X-ray energies for some elements occuring in
materials of current interest. The corresponding wavelength (]~) = (12.40/E) with/~
in units of keV.

/i/ E
Transition Edge (keV) Edge (keV)
3d Fe p-d L2 0.72 L3 0.71
4d Rh p-d L2 3.14 L3 3.00
5d Pt p-d L2 13.27 L3 11.56
4f Gd d-f M4 1.22 M5 1.19
4f Ho p-d L2 8.92 L3 8.07
5f U d-f M4 3.73 M5 3.55

547
548 S.W. LOVESEY

photon energies, when the primary photon energy lies above the excitation energy of
an absorption edge, and near resonance, when the primary energy is in the vicinity
of an absorption edge. The tuneability of synchrotron sources is essential for the
exploitation of a resonant enhancement to magnetic signals. As a guide to the energy
scale, we refer the reader to table 1 and remark that the L absorption edges of the
rare earth elements are in the range 6-10 keV, and the M absorption edges of the
actinides are 3-5 keV. The corresponding wavelengths are convenient for Bragg
diffraction by crystals, since 1 keV = 12.40 A. At M-edges in the actinides resonant
enhancements can reach many orders of magnitude, leading to magnetic diffraction
satellite intensities of order 0.1% of the charge Bragg peaks. In the second, high
energy, regime Compton spectroscopy emerges in the spectrum of secondary photons
when the primary photon energy is in the hard X-ray region, and typically larger
than 30 keV. As a general point, we mention that for diffraction experiments there is
a gain in scattered intensity on moving to harder X-rays due to the attendant increase
in the penetration depth, e.g., for salts changing from 10 to 80 keV X-rays increases
the penetration depth by some three orders of magnitude.
As we have already noted, the magnetic contribution to the scattering amplitude
is generally weak compared to the charge contribution. This discrepancy in size can
be interpreted as due to the magnetic scattering being a (small) relativistic correction
(Grotch et al. 1983, Bhatt et al. 1983, Sakurai 1987). In practical terms, it is usually
necessary to worry about achieving adequate discrimination in the two contributions
to data sets. Discrimination in elastic scattering is possibly provided naturally by
magnetic order if it differs from the chemical order, e.g., antiferromagnetic materials
and a spiral structure displayed by some rare earth magnets. Resonant enhancements
of elastic scattering have been observed; particular attention has so far been given
to large enhancements found for primary photon energies near the M4 absorption
edges in actinides, and near the L3 absorption edges in rare earth and transition
metals. In scattering from ferromagnets it is possible to discriminate between charge
and magnetic scattering by inducing interference in the two contributions. The
interference generated by circular polarization is convenient for making difference
experiments in which the polarity of the interference is reversed either by reversing
the easy axis, by application of a magnetic field, or reversing the handedness of
the polarization. This experimental technique has been applied to ferromagnets to
obtain diffraction and spectroscopic (Compton) data. The same basic technique
has been used in measurements of the attenuation coefficient, where the quantity
isolated in the difference data is the contribution to magnetic dichroism picked out
by circular polarization. Circular dichroism, in some ways, is more intimately related
to the magnetic properties of the sample than its counterpart observed with linear
polarization.
There is no analogue of linear and circular magnetic dichroism in neutron beam
attenuation experiments, in as much that the corresponding scattering amplitude does
not usually contain a significant elastic resonant contribution (Lovesey 1987a, Bal-
car and Lovesey 1989, Byrne 1994). Other differences between neutron and photon
beam techniques applied to magnetic materials include, a possible superior spatial
resolution with photon beams, and the clean separation of spin and orbital magnetism
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 549

in photon scattering whereas only the total magnetic moment appears in the neutron
scattering amplitude. Notwithstanding these useful attributes of photon beam tech-
niques, neutron beam techniques are the preferred choice for many investigations
and, currently, provide the only means of measuring in full dispersion curves and
lifetimes for magnetic collective excitations in magnetic salts, metals and alloys.
Absorption and scattering experiments can be interpreted in terms of a Kramers-
Heisenberg formula (Hayes and Loudon 1978, Lovesey t993). It provides a descrip-
tion of the contribution from orbital magnetism to the diffraction (Bragg) pattern,
and the interpretation of dichroism and elastic resonant scattering from magnetic
materials. But, for such materials, it must be extended to include events that arise
from the spins of unpaired electrons in the sample. While the derivation of the
formula is straightforward, by either standard perturbation theory in quantum me-
chanics or, more convincingly, as an exercise in covariant perturbation theory applied
to quantum electrodynamics, bear in mind that the original formula, not including
electron spins, was derived before the advent of quantum mechanics, by use of the
correspondence principle.
For the present purpose, the framework of standard perturbation theory is adequate
to recount the basic structure of the Kramers-Heisenberg formula (Sakurai 1987,
Hayes and Loudon 1978). To this end, the photons are described by a vector potential,
A, which is a linear combination of photon annihilation and creation operators. A
one-photon scattering event in the scattering amplitude is quadratic in A, since this
function admits the absorption of a primary photon and the simultaneous emission of
a secondary photon. Thus, photon-matter interactions linear in A must be taken to
the level of second-order in perturbation theory, where on account of the intermediate
states and characteristic energy denominators they generate contributions which can
provide resonant processes. Turning now to the photon-matter interaction, it contains
the familiar (p. A) term, from the kinetic energy operator {(p - ~ A)2/2m}, and a
Zeeman interaction between an electron spin s and the photon magnetic field =curlA;
these are the only terms linear in A and, as such, their matrix elements provide the
weight of the resonant processes in the scattering amplitude. To date, the evidence
is that the (p. A) term accounts for the observed magnetic dichroism and elastic
resonant scattering, i.e. the magnetic content of these effects apparently stems from
the orbital, or geometrical, features of the atomic wave functions rather than the
direct magnetic interaction, via the Zeeman term, involving electron spins. Finally,
in the scattering amplitude there are terms quadratic in A coming from first-order
perturbation theory. One such term is the A 2 in the kinetic energy. A second, similar,
contribution arises from the spin-orbit interaction s. (A x E) where the electric field
is proportional to the time derivative of A.
At this juncture, we can make a few useful general observations about the structure
of the photon scattering amplitude. First, the explicit dependence of the amplitude
on electron spin operators arises both from terms linear and quadratic in A, which
are treated by second-, and first-order perturbation theory, respectively. After the
algebraic details of the calculation of the amplitude are put in place, therefore, it is
probably not a surprise to find that the full form of the spin-dependent contribution
to the scattering amplitude is really quite complicated. Another observation is that
550 S.W. LOVESEY

a dependence of the amplitude on orbital magnetism arises solely from the (p • A)


term, treated by second-order perturbation theory. Since spin and orbital magnetism
in the amplitude are not on an equal footing there is the possibility in an experiment
to obtain independent information for them. Clear-cut examples of this arise in
diffraction when polarization effects are exploited.
The aim of this chapter is to survey both a variety of recent experiments, and a
basic framework for their interpretation in terms of variables at the atomic level of
description. If the inclusion, in what is a relatively short chapter, of a theoretical
framework needs defence, largely, it is that in the eyes of the author its omission
engenders sciolism. However, little is given by way of the details for derivations of
results which, in many instances, are provided in their general form, rather than for
specific instrument settings, and hitherto not published by the author.

2. Orientation

Three main topics are covered in this chapter, namely, attenuation of a photon beam,
and elastic (including resonant) and inelastic scattering. In all topics, the influence
of the magnetic properties of the target sample on observed quantities - attenuation
coefficient and elastic and inelastic scattering cross-sections - is the principal focus
of attention. These quantities have a unifying factor, in so much as they are calcu-
lated from the same scattering amplitude operator. Also, the quantities all depend
on the polarization states of the photon beams used in the experiments. These few
observations lead one to anticipate that the three main topics share some essential
features. Our goal in this section is to identify the features in the underlying physics
and the formalism necessary for the interpretation of experiments. To some extent,
this exercise gathers material appearing in later sections but there, with more atten-
dant detail, common features in the topics might not be so easy to discern. At the
same time, the material given here provides the important service to newcomers to
the subject of an orientation to the main concepts and ideas.
All the observable quantities of interest are described by a common scattering
amplitude. We will denote the scattering amplitude operator by G; it is a quantum
mechanical operator, and a 2 x 2 matrix in the space spanned by the polarization
states, for which our convention is described in an appendix. Elastic processes are
determined by matrix elements of G diagonal with respect to the states of the target.
Moreover, the total attenuation coefficient is determined from a knowledge of these
matrix elements averaged over the polarization states of the primary beam, whereas
elastic (diffraction) cross-sections are proportional to products of diagonal matrix
elements averaged over the polarization states. Similarly, inelastic cross-sections
are determined by products of off-diagonal matrix elements of G averaged over the
polarization state of the primary beam. The averaging process, couched in terms of
a photon density matrix, is also covered in our appendix on polarization states.
Let us denote by fu(E) the diagonal matrix element of G, with respect to a
target state labelled #, averaged over the polarization states of a primary beam with
energy E. To describe the attenuation of a beam passing through a sample one
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 551

needs this amplitude evaluated without change in the wave vector of the photon, i.e.
the (elastic) forward scattering amplitude. If E = hcq, the total cross-section, O'tot,
which includes all possible elastic and inelastic processes for a given initial state of
the photon, and the imaginary part of the forward scattering amplitude are related in
the optical theorem by (Newton 1982),

Otot = (4~/q) Im fu(E). (2.1)

When the density, n o, of scatterers is sufficiently small their contribution to the


attenuation coefficient, "7, is simply,

"7 = n0Crtot. (2.2)

Note that, since O'tot has the dimension of (length) 2, 7 has the dimension of a wave
vector. Magnetic dichroism refers to the contribution to 7 that depends explicitly
on the polarization (circular or linear) of the beam, and the magnetic state of the
sample. A clear account of non-magnetic absorption effects is given by Templeton
and Templeton (1994).
We continue the topic of magnetic dichroism with a few more comments on circular
dichroism. Experimental data for the total absorption and magnetic circular dichroism
at the L2 and L3 edges of Ni are displayed in fig. 1 together with a theoretical
interpretation based on tight-binding theory. The diminution in the intensity of
a beam on passing through a thickness R of dilute scatterers is proportional to
exp(-TR). If all aspects of the experimental geometry in two measurements are
kept fixed apart from changing the handedness of the circular polarization the change
observed in the intensity is proportional to,

{ exp ( - (7 + 7O)R) - exp ( - ('7 - 7O)R) } ,- -27OR exp(-'TR),

where 7O, the circular dichroic component of the attenuation coefficient of immediate
interest, is assumed to satisfy I%IR << 1. The other component, '7, is generated by
multiple (geometric optics) scattering, and it is often related in a simple manner to the
imaginary part of the index of refraction for the sample. By contrast, a satisfactory
interpretation of 7o is achieved in terms of a single (Born) scattering approximation.
Indeed, in most cases, Y0 is dominated by electric dipole events in which the magnetic
state of the sample is present in matrix elements through partial, or total, removal of
magnetic quantum number degeneracies by the internal, molecular magnetic field,
and crystal field perturbations.
For the next topic, we tum to elastic scattering; by way of an illustration, form
factors for Fe determined by neutron and photon diffraction experiments are shown
in fig. 2 together with results from a spin-polarized band structure calculation. Let
{...) denote a thermal average of the enclosed quantity with respect to the states
of the unperturbed target sample. (Recall that the equivalence of thermal and time
averaging is a tenet of statistical mechanics.) For a real sample, with some impurities
and disorder, and coherent (elastic) scattering, we include in (G) an average of the
552 S.W. LOVESEY

I ~ I I I
d (o)
+ lOO NICKEL L2, 3 EDGES
b*
>-
p-
80 - - MEASUREMENT -
........ CALCULATION
zuJ 60
p-

~ 40

~o. 2o
0
m 0
J 'i ...................L..2.......
en L
I I J
I i I i

I
-__.__mL~~ .........
(b)

>" -5
l Lz

z -10
I.- MEASUREMENT -~
z
........ CALCULATION
o °15
u
~ -2o

850 870 890


PHOTON ENERGY (eV)
Fig. 1. Comparisons between soft X-ray data (full curves) and tight-binding band structure calculations
(dashed curves) of the L2 and L3 white lines in ferromagnetic Ni (Chen et al. 1991). (a) The total absorp-
tion, (b) the magnetic circular dichroism (MCD). The raw MCD spectrum shown has been multiplied by
a factor of 1.85 to account for incomplete polarization and sample magnetization.

scattering amplitude with respect to all other parameters needed to describe the
sample, e.g., Bragg diffraction occurs only when strict geometrical conditions are
satisfied in scattering by a sample with perfect translational symmetry, which in
reality is a crystal averaged with respect to all forms of disorder. The corresponding
cross-section for radiation observed in an element of solid angle dO is (Lovesey
1986),

~d./dO)-- Tr{.I<G>12}, (2.3)


where the trace operation is taken with respect to the polarization states described
by a density matrix/~. The formula applies to resonant and non-resonant scattering.
The total coherent elastic amplitude for photon scattering is the sum of pure charge
and pure non-resonant magnetic contributions, and a contribution from dispersive and
absorptive processes (de Bergevin and Brunel 1986). The latter contain both charge
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 553

°0
~
6
iI
O
k-

k.
0
0 IN
W-- 3-
E 2 t
k.
0
W.-
T
i
._o O'
a)
c-

o
:~-2
0.0 '0.2 '0.~¢ 'OJ5 '0.8 '1.0 '1.2 'l.~l.
Momentum transfer (~-1)
Fig. 2. Magnetic X-ray results on ferromagnetic iron for the atomic form factor (error bars), together
with polarized neutron data (squares) and band structure calculations (triangles) (Collins et al. 1992).

(Templeton and Templeton 1994) and magnetic interactions. Figure 3 features results
for the energy dependence of the intensity at a satellite in UAs through the M3, M4
and M5 edges, and a fit to a model based on the coherent sum of three dipole
oscillators. In the experiments reported to date the observed magnetic intensities and
polarization dependencies are accounted for by the charge (electric) contribution to
resonant processes. In this case, the magnetic character of the scattering arises from
the magnetic character of the electronic orbitals that enter the calculation of the matrix
elements of the multipole operators. For holmium in its ordered magnetic state, the
significant events are virtual dipole-allowed transitions, coupling 2p core electrons
with 5d-derived conduction band states, and quadrupole transitions, coupling 2p core
electrons to 4f atomic like states. It is found for the spiral magnetic phase that some
magnetic satellites arise solely from quadrupole transitions, while the remaining
observed satellites are mixtures of dipolar and quadmpolar transitions.
At this stage, let us enlarge on the question of the dependence of cross-sections on
the polarization states of the primary beam. There is a sense in which the question
can be answered in general form. To begin with, any 2 x 2 matrix, such as G, can
be expressed as a linear combination of the unit matrix, I, and Pauli matrices, o'.
Let us use the notation,

G =/31 + or. o', (2.4)

where/3 and ot are determined from a detailed knowledge of G, namely, the gener-
alized Kramers-Heisenberg formula to be provided later on. Next, the polarization
states of the primary beam are described by a Stokes vector P = (P~, P2,/)3)- For
554 S.W. LOVESEY

I 012 i I
3 0 x 106

15x10 6
O0
I'-"
Z
0
10~o 0 I I
::k 4.2 4.4
x
>-

Z
F-
Z

121
I---

W
F-
Z
o oo - - z . .

0
106 I I I I I I J
3.4 3.6 3.8 4 4.2 4.4 4.6 4.8 5
ENERGY (KEY)
Fig. 3. The energy dependence of the intensity at a satellite in UAs through the M3.4,5 edges, which
display a strong interference effect between the two closely spaced edges. The full curve through the
data is a fit to the coherent sum of three dipole oscillators. Inset is an expanded view of the M3 edge
(McWhan et al. 1990). See also table I.

the moment, we will not dwell on the physical significance of the three parameters
(they are discussed in the appendix) other than mention, by way of an example,
that P2 is the mean helicity in the beam, i.e. the degree of circular polarization.
Armed with this knowledge of the polarization states, the trace operation can be
performed leaving the cross-section expressed in terms of/3, ot and P. In fact, since
the cross-section, of course, is purely real, and a scalar object one can anticipate
that the expression is a linear combination of the terms,/3+/3, ~+c~,/3+(~x • P), and
P . (c~+ x c~), so the detailed work is simply to determine the coefficients of such
terms in the expansion. One finds,

Tr{izG+G} = (or + . m +/3+/3 + / 3 + ( P . o0 + ( P . o~+)/3 + i P . (o~+ x oz)). (2.5)

The elastic cross-section, mentioned above, is obtained from this expression by


replacing oL and/2 by their average values. Let us further note that, the expression
provided applies also to inelastic scattering, to which we turn later, and a similar
PHOTON BEAM STUDIES OF MAGNETICMATERIALS 555

0.4

0.2 ~ 1 0 , 0
0 6,6,0
E
~-0.2
(n
~... t .NxX~8,18,0
-0,4
14,14,0
-0.6
-0.8
0 0.5 1 1.5 2
k(A'+~
Fig. 4. The spin/orbital form factor ratios for holmiumin HoFe2 at room temperature. The solid line
represents the results of a relativisticspin-polarizedband structurecalculationwhichhas been normalized
to the Hund's rule ratio of 1/3 at k = 0 (Collins et al. 1993).

procedure can be used to generate general expressions for the Stokes vector of the
secondary beam and the forward scattering amplitude.
Moving on, in section 1 we mentioned that (a) spin and orbital magnetism in
the target sample influence scattering in different ways, and (b) the magnetic con-
tribution to scattering by a ferromagnet can be isolated through use of a circularly
polarized primary beam. Both points feature in the diffraction data for HoFe2 shown
in fig. 4, which are the ratios of the spin to orbital form factors derived from the
interference scattering induced at mixed charge and magnetic Bragg reflections by
circular polarization.
For the description of polarization states of the primary, and secondary, beams we
advocate use of a formalism based on Stokes vectors (McMaster 1961, Berestetskii et
al. 1982, Lovesey 1987b). The main argument in support of this choice is generality.
For, in reality, one is unlikely to have a perfectly polarized beam, and partially
polarized states encountered in experiments are fully described by allowing for all
three Stokes parameters {Pi}. The alternative is to provide G couched in terms of
polarization vectors, e and e' for the primary and secondary beams, respectively, and
relate these to the geometry of a particular experiment. To illustrate this aspect, let us
consider the cross-section for Thomson scattering by electrons located at positions
{Rj}. The variable measured in scattering is the spatial Fourier transform of the
charge distribution,

n(k) = E exp (i k . R j), (2.6)


J
556 S.W. LOVESEY

where k = (q - q') is the change in the wave vectors of photons in the primary and
secondary beams. The Thomson cross-section is,

(d~/dS2) = r2(e. (2.7)

in which re is the classical radius of an electron. Looking at the dependence of the


cross-section on the polarization vectors, there are several, more or less, standard
representations for (e • e,)2 in terms of angles that describe the elastic scattering
geometry. However, with the advocated formalism, in terms of Stokes vectors,
(e. e,)2 is replaced by its value averaged over the possible distributions of partial,
or total, polarization in the primary beam, viz.,

( g . ~t)2 .....> 1 { 1 + cos 2 0 q- P3 sin 2 0}, (2.8)

where 0 is the angle through which the primary beam is deflected to the detector.
Note that the result does not depend on P1 or P2. The values P3 = +1, describing
complete linear polarization perpendicular (a-polarization) to the plane of scattering,
and complete linear polarization in the plane (rr-polarization), are unlikely to be
achieved in practice. To conclude this slight digression on the formalism for handling
less than complete polarization, consider values of the Stokes parameters for the
secondary beam {P'}. Sticking with the example of Thomson scattering, one finds
(Berestetskii et al. 1982); for i = 1,2,

P" = 2Pi cos0/(1 + cos20 + P3 sin2 0), (2.9)

and,

P~ = ( sin 2 0 + P3(1 + cos z 0))/(1 + cos 2 0 + P3 sin 2 0).

It is interesting to note that, for 0 = (rr/2) the Stokes parameters of the secondary
beam, for any primary polarization, are/'/ = (0,0, 1), i.e. 90 ° charge scattering
produces complete e-polarization of the secondary beam. This well-known effect
has been exploited to discriminate between charge and magnetic components in the
secondary beam (Gibbs et al. 1985).
Inelastic magnetic scattering experiments using photon beams from synchrotron
sources have so far focused on deriving the momentum distribution, or Compton
profile, of unpaired electrons in metallic magnets (Cooper 1987, Sakai 1992). These
studies exploit the charge-magnetic interference scattering induced by circular po-
larization in the primary beam to extract the magnetic component of the Compton
scattering process, while accurate total Compton profiles can be obtained with labo-
ratory 7-ray spectrometers (Anastassopoulos et al. 1991). Figure 5 contains data and
theoretical predictions for the Compton profile of unpaired electrons in ferromag-
netic nickel. An experimental investigation has shown that Compton scattering is
not sensitive to orbital magnetism in the sample (Timms et al. 1993), i.e. the profile
extracted from data is the momentum density of the unpaired electron spins.
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 557

'o F
Ni 110
8

FLAPW

"~., 6 ............ EXPT

e
O~
0

hE4

2 -

I I I I
0 1 2 3 /-. 5 6 7 8
pz (o.u.)
Fig. 5. Magnetic Compton profile for Ni in the (110) direction. The solid line is derived from a band
structure calculation, and the experimental data are denoted by the dotted line (Kubo and Asano 1990).

A compact expression for the partial differential cross-section, which includes all
elastic and inelastic processes, is achieved with the help of correlation functions,
standard in the interpretation of a wide range of other experiments, including NMR,
#SR, electron and neutron beam scattering (Lovesey 1986, 1987a). To this end,
let G(t) be the Heisenberg operator at time t, formed from the scattering amplitude
operator; for simplicity of notation, we will write G(0) =_ G. The correlation function
required to calculate the cross-section is (G+G(t)). If No is the energy transferred
from the primary beam to the target sample, the partial differential cross-section
which gives the fraction of photons of incident energy E scattered into an element
of solid angle d£2 with an energy between E ~ = (E - No), and E' + dE ~, is,

(d2cr/d~?dE ') = (E'/E)(1/2~rh)


F dt exp(-icot)Tr{~(G+G(t))}. (2.10)

The trace operation, with respect to the polarization states of the primary beam
described by the density matrix ~, can be accomplished to the extent of creating
an expansion in terms of correlation functions formed with the operators fl and
a introduced earlier on. The strictly elastic cross-section, already encountered, is
derived from the partial differential cross-section by taking t -+ c~ in the correlation
function, since,
(G+G(t = :/G+/(6(t : : I/G/I 2,
558 s.w. LOVESEY

where the first equality follows from the law of increase in entropy, or loss of
information, which requires that, for a bulk assembly, there is no correlation between
processes well separated in time. To reach the final expression use the result, correct
for any Heisenberg operator, (G(t)) = (G(0)) which is a consequence of the condition
for a stationary system that a correlation function is independent of the origin chosen
for the time variable. Returning to the cross-section, when a time-independent value
of the correlation function is inserted in the Fourier integral the latter reduces to a
delta function with w as its argument, i.e. the cross-section vanishes except for w = 0
which corresponds to purely elastic scattering. The result for the partial differential
cross-section when integrated with respect to E ~ is identical with the previous elastic
cross-section, as required.
In view of the foregoing analysis, the partial differential cross-section formed with
the correlation function,

(<G+G(t)>-I<G>I=) <{G + - <G>*){G(t)- <G>) >, (2.11)

is exclusively inelastic in its content; this is the cross-section which described all
forms of spectroscopy, e.g., Compton and Raman processes.
Photoemission has a role to play in the exploitation of photon beams to study mag-
netic materials (Thole and van der Laan 1991, Williams et al. 1980, Halilov et al.
1993, van der Laan 1994a) but it is not included in the scope of this chapter. A sim-
ple picture of photoemission entails three independent processes; photoabsorption,
propagation of an excited electron to the surface, and the escape of the photoelec-
tron into the vacuum. To underscore the intrinsic complexity of photoemission, we
mention that linear response theory does not give rise to the photoelectric process,
which puts it in a different category of experimental methods to many others used to
study magnetic materials, including, #SR, NMR and neutron and photon scattering.
Another feature which merits comment is that, in angle-sensitive photoemission one
must be aware of macroscopic refraction and reflection effects that can be modelled
by Fresnel equations. However, this is just one of several near-surface effects in-
volved in a full analysis of photoemission data. Circular dichroism in photoemission
is treated by Thole and van der Laan (1994).

3. Survey of experiments

The recent flurry of activity with applications of photon beam techniques to address
magnetic properties of materials seems to have really got underway around 1985
(de Bergevin and Brunel 1986, Cooper 1987, Gibbs 1992). At the risk of being
invidious, we mention two pieces of work published at this time that played a
part in raising the awareness of researchers in magnetism to the potential value of
synchrotron-based techniques. Magneto-optic effects, e.g., the magneto-optical Kerr
effect in the visible region, appeared to have minimal value as an investigative tool
prior to predictions by (Thole et al. 1985) for 3d absorption edges of rare earth
materials based on atomic multiplet calculations. The relatively strong magnetic
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 559

X-ray dichroism was first observed in terbium-iron garnet (van der Laan et al.
1986). At more or less the same time, interesting science was revealed in X-ray
scattering experiments on the rare-earth metal holmium (Gibbs et al. 1985). Direct
high-resolution measurements of the nominally incommensurate magnetic satellite
reflections revealed lock-in behaviour which was successfully explained in terms of
a simple spin-discommensuration model (Bohr et al. 1989).
Looking much further back in time, theoretical work was ahead of experimental
investigations of magnetic effects in photon scattering, with the first explicit calcu-
lations published in 1938 (Tolhoek 1956, Evans 1958). While these works focused
on the basic nature of the photon-matter interaction, Platzman and Tzoar (Platzman
and Tzoar 1970) explored a potential value for fundamental investigations of mag-
netic materials, successfully demonstrated by de Bergevin and Brunel (de Bergevin
and Brunel 1981, 1986) in a series of experiments on various different materials.
Skipping forward looking for particularly significant findings, over the 1985 devel-
opments mentioned in the beginning, one lights on the successful observation of
resonant elastic scattering (Gibbs et al. 1988, Hannon et al. 1988, 1989), which is
now a small industry in part because resonant enhancement ameliorates technical
problems faced in measuring intrinsically weak scattering events. Recent reviews
of experimental investigations of magnetic systems include (de Bergevin and Brunel
1986, Cooper 1987, Gibbs 1992, Sette et al. 1991, Chen 1993). For our part, we
look at experiments which have used circular and linear dichroism, diffraction from
ordered magnetic structures (Bragg scattering), including elastic resonant scattering,
and spectroscopy. Basic concepts for the interpretation of the experimental investi-
gations are gathered in subsequent sections.

3.1. Dichroism
Here, and in section 3.3, we consider the resonant regime and discuss magnetic X-ray
dichroism and elastic magnetic resonance scattering. The relation between absorption
and scattering is the standard optical theorem, discussed in section 2 and section 7.
Within a simple one-electron picture of electronic structure illustrated in fig. 6,
in resonant scattering the incident photon promotes by a virtual transition an inner
shell electron to an unoccupied orbital above the Fermi energy, which subsequently
decays through the emission of photon. The amplitude for resonant scattering then
depends on the matrix elements which couple the initial state and the intermediate
magnetic states allowed by the Pauli exclusion principle. The scattering ampfitude
(6.4) contains charge, linear momentum and spin interaction operators. To date,
the experimental data on resonant scattering and dichroism have been successfully
interpreted in terms of the momentum interaction operation. In this instance, the
magnetic character of the observed electron-photon events in magnetic materials
stems entirely from the nature of the wave functions used to calculate the dipole,
quadrupole . . . . . matrix elements. The underlying physics is common to a range
of magneto-optic effects, including dichroism, the Faraday effect and the magneto-
optical Kerr effect. Reviews in (Kao et al. 1993, van der Laan 1990) of these effects
use the formalism outlined in sections 7 and 9.
560 S.W. LOVESEY

L///Edge

s-p
4f

EF
s-p-d
4f

s-p

E1 : 2p3/2-~5ds~

E2: 2p312--)4f7/2

2P3/2
Fig. 6. Schematic, one-electron view of resonant magnetic scattering at a n L 3 absorption edge (Gibbs
1992). The linearly polarized primary photon promotes a 2p3/2 core electron into an empty state above
the Fermi level, EF. In the rare earth elements, on which this simple example is modelled, there are
localized 5d-states available in dipole-allowedtransitions (El), and un-filled4f states available through
quadrupole transitions (E2). Scatteringresults when the virtually excited electron decays, thereby filling
a core hole and emitting a photon.

Magnetic dichroic effects make the near-edge, inner-shell absorption of polarized


photons a useful tool for investigating the magnetism of transition metal, rare earth,
and actinide elements and compounds. Table 2 is a summary of representative
examples of experimental studies. The absorption and dichroic effects at the L2,3
edges in ferromagnetic nickel are displayed in fig. 1. The relatively high precision
with which the intensity ratios can be determined provide good tests of models
of magnetism. Indeed, data for the L2,3(2p --+ 3d) and Mz,3(3p --+ 3d) magnetic
dichroism and X-ray photoemission of nickel have been subject to various theoretical
interpretations (Chen et al. 1991, Jo and Sawatzky 1991, van der Laan and Thole
1992, van der Laan 1994b), with attention to electronic correlations.
The data in fig. 1 for nickel are an example of circular dichroism observed with
soft X-rays. In contrast, the prediction (Thole et al. 1985) of strong magnetic dichro-
ism and experimental proof (van der Laan et al. 1986) was for linear dichroism in
the M4,5(3d --+ 4f) absorption edges of rare earth materials. Calculations (Thole et
al. 1985, Goedkoop et al. 1988a) are made on the basis of atomic multiplet con-
figurations, illustrated in fig. 7 for the simple case of yb3+(4f13), and outlined in
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 561

TABLE 2
Representative examples of materials investigated by magnetic linear (MLD) and circular (MCD)
diehroism, see also table 1.

Material Ref. Edges Probe Comment


Tb3FesO12 [1] M4,5 MLD applied field
Ni [2, 5, 9, 12, 14] L2,3M2,3 MCD film on Cu;
single crystal
Gd3Fe5012 [2, 4] L2,3M4,5 MCD applied field
Gd [3] L1,2,3 MCD
Co [6, 7] L2,3 MCD, Kerr effect multilayer
5d impurities in Fe [8, 11] L2,3 MCD
3d impurities in Ni [10] L MCD
rare earth intermetallics [13] L2,3 MCD

References:
[1] van der Laan et al. (1986) [8] Schtitz (1990)
[2] Tjeng et al. (1991) [9] Koide et al. (1991)
[3] Schtitz et al. (1988) [10] BOske et al. (1994)
[4] Rudolf et al. (1992) [11] Schtitz et al. (1993a)
[5] Chen et al. (1990) [12] Vogel and Sacchi (1994)
[6] Kao et al. (in press) [13] Krill et al. (1993)
[7] Kao et al. (1993) [14] O'Brien and Tonner (1994)

sections 7 and 9. Circular and linear dichroism probe, respectively, the magnetization
and mean-square magnetic fluctuations, and the effects in question are much greater
than those observed in the visible region. Circular dichroism can only be exploited
with single domain magnetically oriented samples, e.g., ferro- and ferri-magnets in
an applied field; however, with linear polarization, it is possible to examine single
crystal antiferromagnets as well (Kuiper et al. 1993) provided the magnetic moments
are aligned in preferential crystallographic directions. The usually neglected electric
quadrupole transition are predicted to be as important as the dipole terms, considered
in the foregoing discussions, for the interpretation of magnetic dichroism at rare earth
L edges and transition metal K edges (Carra and Altarelli 1990, Carra et al. 1991,
Wang et al. 1993, Jo and Imada 1993).
Various types and levels of theoretical work are applied to the interpretation of
magnetic dichroism. Atomic calculations are appropriate for 3d absorption in rare
earths (Thole et al. 1985, Goedkoop et al. 1988a) and 2p absorption in 3d transition
metals (van der Laan and Thole 1991). Group theory has been exploited to derive
a general model for spin polarization and magnetic dichroism in photoemission, and
applied to the 2p, 3s, 3p and 3d photoemission from divalent Cu, Co d 7 and F e d 6
(Thole and van der Laan 1991). A relatively simple model of circular dichroism at
the L2,3 edges of rare earth atoms has been successfully used (Jo and Imada 1993) to
interpret data on atoms from Ce through to Tm in (RE)2COl7. This theoretical work is
based on the 2p ~ 5d dipole (El) transition, and the tolerable accord with the data is
seen to vindicate the neglect in the calculations of electric quadrupole (E2) transitions.
Also included in (Jo and Imada 1993) are well discussed applications of sum rules
for dichroic signals, which is a topic taken up in more detail later in this section. For
heavy rare earth metals, circular dichroism evaluated from first-principles relativistic
562 S.W. LOVESEY

yb ~ + magnetic field
M'
5/2
/ * i i 3/2
I I , I I ,
J'=5/2
-~_ & ',; &; ', & ;,, -1/2
A' ' ' A, ' , k' [ I -3/2
I ~ I "r I i I i i I i
4 : ;; ~: ; A -5/2
' I I --i I I
I I I I I I

J
I , ,,,, ....
I { i I I
I I I I I I I
I I I I

I I I I I
j_j.~.,__~_ .&;~SJ-L. ~ j_L.
zxJ =-1 /',M= 1 ~M=O z~ M=-I
-v, ,,
] I I I i i

i i I I I
I i i I
I I I I i i I
' , |I
I I I
I
• ! t I M
• ' 7/2
, i:i i, ,i
. ,, t :i,' ii i
, ,, ;,,, 312

J =7/2
=11 I ~
,,,
t
'
I [I
] 1/2
--1/2

, I -5/2
-7/2

\x /,

?
Fig. 7. Energy level diagram of the 3d]°4f 13 --+ 3d94f 14 transition in Yb 3+ without (left) and with
(right) a magnetic field (Goedkoop et al. 1988b). The vertical arrows indicate the dipole selection rule
allowed transitions. Their relative intensities are given by the dots. (In the text, the label m in the 3j
symbol is denoted by q.) The required polarization is indicated at the bottom of the figure.

spin-polarized band-structure calculations (Wang et al. 1993), applied to the dipolar


contribution, provides an adequate account of data. Several calculations of the soft
X-ray magnetic circular dichroism at the L2,3 edges of ferromagnetic nickel have
been reported, based on the Anderson impurity model (Jo and Sawatzky 1991, van
der Laan and Thole 1992, van der Laan 1994b) and a tight-binding band structure
model (Chen et al. 1991). The latter analysis isolates features in the data which lie
beyond the physics included in one-electron band-structure models, and are attributed
to a many-body shake-up or shake-off process accompanying the creation of a core
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 563

hole. These interesting findings are supported by subsequent work (Jo and Sawatzky
1991, van der Laan and Thole 1992, van der Laan 1994b). A theory of dichroism
in iron is provided in (Smith et al. 1992).
The dichroic effects observed in the cited experiments are really not the same as
expected from the Fano effect (Fano 1969, Thole and Van der Laan 1991). Fano
predicted that polarized electrons can be produced by photoionization of unpolarized
atoms with circularly, linearly and unpolarized light. The photoelectron polarization
is caused by spin-orbit interactions in the continuum state, or in the atomic states.
These are not influenced by a polarized valence shell, leaving the core-hole - spin-
orbit coupling, for example, as the remaining interaction. But, this picture is not
appropriate when the electron is excited to a more localized state, such as the valence
d states in transition metal compounds or the f states in rare earth and activide
materials. For intra-atomic excitations there are electrostatic interactions between the
spin and orbital momenta of the valence state and core-hole, and the magneto-optical
effect, referred to as magnetic dichroism, is much larger than for spin dependent
photoabsorption.
While it was recognized many years ago that the structure of the scattering ampli-
tude afforded the possibility to separate spin and orbital magnetism in the analysis of
diffraction patterns, it was not until 1992 that a similar advantage became available
in interpretations of circular dichroism (Thole et al. 1992). In this case, a sum rule
for the integral of the signal over a given edge allows one to determine the ground
state value of the orbital angular momentum, and it has been successfully applied
to data collected on ferromagnetic nickel. Various other sum rules have since been
derived (Carra 1992, Thole and van der Laan 1993, Carra et al. 1993a, b) which
afford the possibility to extract from measurements of dichroic effects the ground
state expectation value of the magnetic field operators (orbital, spin and magnetic
dipole) of the valence electrons. While the original work used a local atomic picture,
similar results have also been deduced from a simple single-particle band structure
picture (Shtitz et al. 1993a, b).
To illustrate the form taken by so-called sum rules, consider the integrated strength
of dipole transitions. Referring to section 7, the strength of such a transition is
proportional to (AL<wlRqI~>I2),where A is the difference in energy between the
ground state labelled by /z, and an intermediate state labelled ~/, and q = 0, 5:1
denotes spherical components of the position operator. The state [#) is that of an ion
with an incomplete outer shell, angular momentum g, and h holes. The integrated
strength of dipole transitions is proportional to,

and, for instance, the ratio,

D : (I1 - I _ , ) / ( I 1 + I o + -[-1),

measures the circular dichroic signal. To proceed, one argues that for all transitions
the radial integral in the dipole matrix elements is the same, and A = (En - E~) is
564 S.W. LOVESEY

replaced by some average value and taken outside the sum. In consequence, D is
assumed to be independent of the radial integral, and energy level separation. After
algebra that entails the reduced matrix element (7.12), one finds (Altarelli 1993),
D = -(~lLzl~)/eh,
which is the sum rule first derived in (Thole et al. 1992). The derivation of this sum
rule and its application are further discussed by van der Laan (van der Laan 1994b).
By way of an apparently simple example of the use of sum rules we mention
results obtained from magnetic circular dichroism at the L2 and L3 edges in a rema-
nently magnetized Ni (110) single crystal (Vogel and Sacchi 1994). The expectation
values of orbital and spin moments (in units of #R) in the d shell are found to
be, 0.06 4- 0.01 and 0.27 :k 0.03, respectively. Mixed magnetic systems have been
studied by the authors of reference (Shtitz et al. 1993b). They disclose a tolerable
agreement between the analysis of experimental data and band structure calculations
for the average spin and orbital moments of 5d elements dissolved in iron. Equally
interesting is a comparison of two methods of analysing the data. One method is an
application of sum rules (Thole et al. 1992, Carra et al. 1993b), derived originally
for an atomic model. The other method, developed by the authors, is a two-step
model based on the Fano effect (Fano 1969) and a single-electron band structure
picture. Significant differences are found between the two sets of results, e.g., for Pt
dissolved in Fe the deduced orbital moments are of a similar magnitude but opposite
in sign. On the other hand, for Os the methods give similar results for the average
spin moment but, the near common value is very different from the value derived
from band structure theory. These and other related (Jo and Imada 1993, van der
Laan 1994b) findings, for the moment, post a cautionary tale about the application
of sum rules to analyse dichroic signals, although it seems that the sum rule for the
orbital moment, discussed in the preceding paragraph, is more robust than those for
the spin moment and magnetic dipole.
Beside the dichroic effects already mentioned, it is possible to probe local magnetic
moments from absorption of unpolarized X rays in unpolarized electron shells using
branching ratio analysis (Thole and Van der Laan 1988a, b, van der Laan and Thole
1988). The core hole spin-orbit branching ratio is extremely sensitive to the angular
momentum of the valence electrons.
The sample requirements for detection of magnetic dichroic effects (a single crystal
and magnetic alignment) are less severe in analysis of core level X-ray absorption
line shapes (Alders et al. 1994). Moreover, the information, on interatomic exchange
interactions and short range spin autocorrelation functions, is element specific and
can be used to study ferro-, ferri-, or antiferromagnetic materials in single crystals,
thin film and powder film. To date, the effect has been demonstrated in a study of
the Ni L2,3 edge in a layer by layer grown NiO film.
Measurements of dichroic effects in systems with reduced spatial dimension have
proved fruitful. Examples include, magnetic films (Idzerda et al., in press; Tjeng et
al. 1992, Heinrich and Cochran 1993, O'Brien and Tonner 1994), single crystal thin
films (Idzerda et al. 1993a), heteromagnetic multilayers (Chen et al. 1993, Idzerda
et al. 1993b), and near surface magnetism of ferromagnetic nickel (van der Laan et
al. 1992).
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 565

3.2. Diffraction
When compared with neutron magnetic diffraction, synchrotron-based photon diffrac-
tion has several intrinsic strong points. These include;

(a) High spatial resolution


In some applications, the spatial resolution obtained in photon diffraction is signif-
icantly finer than that available in a corresponding neutron diffraction experiment.
This point is illustrated for critical scattering from holmium by the data displayed
in fig. 8; the resolution obtained in the two experiments differs by an order of

Holmium (O,O,2-'f) Sample#1


2.0

3-

~-~ 1.5

X /
C)

~-~ 1.0
X_ra, Face, la XX~ ~ID
haft-width "X~ •
-i-

~- 0.5 o Int. Intensity XX~


-6 - - e--e--oJ-'-rt~ o

I I i \ ~4"01 0 0 h
0 , ,

0.025

0.020 ,~

o.o15

"r" ,~
o.olo

0.005
• ~ u u

0 i I I I

130.0 130.5 131.0 131.5 132.0


Temperature (K)
Fig. 8. Integrated intensity and half-width (HWHM) of scans taken at the (0, 0, 2 - t) peak position of
holmium metal in the vicinity of the N6el temperature ~ 131.2 K (Thurston et al. 1994). Low resolution
neutron scattering data, and X-ray data are displayed. The solid lines represent power law fits; integrated
intensity ~ (TN -- T) 2~, and HWHM ~ (T - TN) ~' where ~ and/3 are the standard critical exponents,
and v = 0.54 -4-0.04, whereas the exact meaning of/3 depends on the interpretation of the experiment.
566 S.W. LOVESEY

magnitude. With the current and planned high-brightness synchrotron sources, it has
become possible to investigate ordering phenomena with correlations extending over
micron length scales with high accuracy.

(b) Polarization analysis


There is usually more scope to benefit from polarization analysis with photon beams
than with neutron beams. This stems from both differences in the intrinsic properties
of the radiations and production methods, e.g., photon beams from a synchrotron
source have a high degree (,-~ 90%) of linear polarization while neutron beams from
reactor and spallation sources are unpolarized. The scope afforded by polarization
dependent properties is illustrated by the material gathered in fig. 4 and fig. 9 on the
spin and orbital moments in HoFe2 and Ho obtained, respectively, by use of circular
and linear polarization properties in diffraction.

(c) Extinction-free scattering


Because of the relative weakness of magnetic photon scattering, an interpretation is
appropriate within the first Born approximation, so there is no extinction correction.

(d) Static approximation


The quite broad energy resolution typical of many photon scattering observations
(5-10 eV) means that inelastic events are integrated over, to a good approximation,
i.e. the total cross-section is observed. For diffuse charge scattering this means
that the observed intensity is described by (2.7) which relates the scattered intensity
to the instantaneous value of spatial Fourier transform of the spatial distribution of
scatterers.

10~
~eL 0.5 x• L=6 "-.. L=0
, ..

• 0 ~',, ""

-0.5 ""- ..

-1.0

Fig. 9. Data obtained for holmium (Gibbs eta]. 1991). The solid square at 8 = 0 corresponds to
the degree of linear polarization of the primary beam. The open circles show the degree of linear
polarization measured for the charge scattering at chemical Bragg reflections at a temperature where the
magnetic configurationis conical T < 20 K). The full curve is the degree of linear polarizationfor charge
scattering calculatedfrom (2.9) with P3 = 0.77. The solid circles show the degree of linear polarization
measured for magnetic (satellite)reflections. The broken curves are calculations,based on results given
in section 8.2, for three different values of the total spin, S, and total orbital angular momentumL.
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 567

(e) Relatively small samples are adequate


The diffraction intensity arising from charge scattering is usually much larger than
from pure magnetic scattering. For ferromagnetic systems, in which chemical and
magnetic order coincide, one of two methods to discriminate charge and magnetic
contributions to Bragg peaks suggest themselves. Both methods rely on creating an
interference between charge and magnetic amplitudes, which has the attractive feature
of giving access to the sign and magnitude of the magnetic scattering amplitude. In
one method, interference is generated by the imaginary part of the charge scattering
amplitude (de Bergevin and Brunel 1981, Vettier et al. 1986) that is available in non-
centrosymmetric samples or through tuning the X-rays from a synchrotron source to
an adsorption edge. Alternatively, use is made of circular polarization in the primary
beam (Brunel et al. 1983, Collins et al. 1992); data for the magnetic form factor of
iron obtained from polarized X-ray and neutron diffraction are contrasted in fig. 2.
Table 3 lists some of the experiments mentioned in the previous paragraph together
with studies of rare earth metals. In the latter work, attention is given to satellite
reflections that can be purely magnetic in character.
Before mentioning more experiments, it is useful to note some simple properties
of the cross section (2.5) constructed from the four components of a and/3 listed in
(A.8). Looking at the magnetic contributions in these four components, which are
identified by the small parameter ~- = (E/mc2), it is seen that ~l and c~z are purely
magnetic while c~3 and/3 are linear combinations of charge and magnetic operators.
It is significant that, in ~3 and/3 the magnetic contributions differ from the charge
contributions by a phase factor i = x/S] .
Assume in the first instance that the diagonal matrix elements of the atomic quan-
tities in ot and 13, namely n(k), S(k) and Z(k), are purely real, and the primary
beam is unpolarized, P = 0. The cross section for diffraction is proportional to
{](a)l 2 + [(/3)12). When the matrix elements in (a) and (/3) are real it follows that
TABLE 3
Representative examples of magnetic X-ray diffraction
studies

Material Ref.
Zn0.sFe2.504 [51
Fe [4]
Ho [1, 8, 9]
Tb [8]
Er [71
Tm [61
HoFe2 [101
Gd-Y superlattice [3]
Ho15-Y12 superlattice [2]

References:
[1] Gibbs et al. (1985) [6] Bohr et al. (1990)
[2] Bohr et al. (1989) [7] Gibbs et al. (1986)
[3] Vettier et al. (1986) [8] Tang et al. (1992a)
[4] Collins et al. (1992) [9] Gibbs et al. (1991)
[5] Brunel et al. (1983) [10] Collins et al. (1993)
568 S.W. LOVESEY

the magnetic contribution to the cross section is proportional to r 2, i.e. there is no


term proportional to r. Since r << 1 (otherwise use of the amplitude (6.5) is not
valid) the magnetic content of a mixed (charge and magnetic) Bragg reflection is
dominated by the charge contribution. However, pure magnetic reflections can occur
in magnetic materials for which the chemical and magnetic structures are different,
i.e. materials other than simple ferromagnets. If the condition on the matrix elements
in (c~) and (/3) is relaxed and they possess imaginary charge components, which is
possible in non-centrosymmetric materials or from anomalous scattering, then the
diffraction cross section may contain terms proportional to ~- that are probably larger
than terms proportional to ~.2.
Polarization of the primary beam has a significant effect on the cross section. We
refer to (6.5) for the diffraction amplitude, and figs 10, 1 la, and 1 lb together with
(6.7) evaluated for elastic scattering (a = b = 1). Our choice of axes for the scattering
geometry is given in fig. 12. With regard to the influence of linear polarization,
note that the matrix B, given in (6.7), which represents the polarization dependence
of spin scattering, contains off-diagonal elements. In consequence, cr - 7r and 7r -
events occur through scattering by the spin magnetization. There are also finite
off-diagonal elements in (¢' x ¢) which represents the polarization dependence of
orbital scattering. By contrast, (e' • e), which represents the polarization dependence
of charge scattering, is diagonal. From these observations it follows that (cr-Tr) and
(Tr-cr) events revea ! purely magnetic scattering.
To appreciate the role of circular polarization in photon scattering by magnetic ma-
terials, recollect that circular polarization can be represented by complex polarization
vectors. If the matrix element in (a) and (fl) are purely real, the phase difference
between their charge and magnetic components can be negated by the imaginary
part of the complex polarization vectors leading to purely real interference terms.
So, circular polarization induces interference between charge and magnetic ampli-
tudes, and hence terms of order ~- in the cross section. Furthermore, changing from
right-to left-hand polarization, or vice versa, changes the sign of these interference
terms leading to the appealing possibility of a difference measurement to isolate the
relatively small linear magnetization contribution. This scheme has value only for
materials that have mixed (charge and magnetic) reflections, as in a ferromagnet.
Experiments that confirm many of these properties of photon scattering by mag-
netic solids have been performed by de Bergevin and Brunel; their 1986 paper
contains a review of their findings and theoretical interpretations. These researchers
observed magnetic diffraction of order T2 from bulk antiferromagnetic NiO using

_a.(a) I1(~)

GII GI2

I1(~) G21 G22


Fig. 10. The representation chosen for the polarization-dependent elements of the scattering amplitude
operator G expressed as a 2 × 2 matrix eq. (2.4).
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 569

a) (¢'. ¢)

1 0

(~) o (~ .~')

b) (¢' x s)

A_(~) I1(~)

±(cr) o

I1(~) -~' (~' × ~)

c) (~' • m)(¢ • m)

±(0) 11(70

±(~r) ,n~ m±(¢ll, m)

I1(~ I (e~l - m ) m ± (e'll.,n)(Sli.m)

Fig. 11. Matrix representations of the functions ~' • s, ¢ ' × ~, and (~' • m)(¢ • m) formed with
the (real) polarization vectors of the primary, ¢, and secondary, ¢', beams. 9 and q ' are unit vectors
in the directions of the primary and secondary beams, respectively, and 9 • q ' = cosO, c f fig. 12.
Regarding the vector m, r a ± is its projection perpendicular to the plane of scattering defined by q and
q '. Polarization vectors parallel and perpendicular to the plane are denoted by ~ll and ~ ± , respectively,
and m ± = ( m . ~ ± ) : ( m . ¢~_); cf. fig. 12. Note 9 = ~ ± × ~LI' and a similar relation for the secondary
beam.

an X-ray tube operated at 1 kW. Photon beams from synchrotron sources are pre-
dominantly linearly polarized within the plane of the electron orbit, and elliptically
polarized above and below the plane. The resolution in reciprocal space can be
superior to that currently available in neutron diffraction (Tang et al. 1992a); fig. 8
illustrates this feature for critical scattering by holmium.
Magnetic diffraction from ferromagnetic iron and HoFe2 of circularly polarized
photons has been reported by Collins et al. (1992, 1993). They used a white beam and
single crystal samples, and reversed the direction of the magnetization by an applied
field to affect isolation of the polarization induced charge-magnetic interference
contribution. The experimental technique offers several attractive features, including
570 S.W. LOVESEY

ll" I

Fig. 12. Coordinates (, r/, ~ used to define the scattering geometry. The (-axis is perpendicular to the
plane of scattering, defined by q and q ~, k = q - q ~, and the secondary beam is deflected through an
angle 0.

a fixed scattering geometry, while several harmonics of a given reflection can be


resolved with an energy dispersive detector and measured simultaneously.
The possibility in photon diffraction to separately measure spin and orbital contri-
butions to the amplitude has been realized in studies of holmium (Gibbs et al. 1991)
and HoFe2 (Collins et al. 1993); some results are shown in figs 4 and 9. In principle,
a scattering geometry can be chosen at which the magnetic scattering is either purely
orbital or spin in nature cf. section 8. With regard to magnetic neutron diffraction,
the observed intensity is always a mixture of spin and orbital contributions that are
not separately obtainable through a choice of scattering geometry (Tang et al. 1992a).
As mentioned in the previous section, analysis of dichroic effects also provides spin
and orbital moments of the ground state configuration.
Photon diffraction studies of magnetic fluctuations around the (100) Bragg reflec-
tion of the diluted antiferromagnet Mn0.ysZn0.25F2 have clarified its complex prop-
erties (Hill et al. 1991, 1993a, b). The compound has been extensively studied by
various experimental techniques because when subject to a magnetic field it models a
system with both quenched disorder and competing interactions; here, the underlying
canonical model is the random field Ising model. Neutron scattering measurements
of long range magnetic order are severely affected by extinction, while other ex-
perimental techniques are limited in value by the large disorder. Magnetic photon
scattering has been proven to be free from these shortcomings. In applications to a
sample with the rutile structure, the (100) Bragg reflection has the great advantage
of being a special point at which the charge structure factor vanishes.
Lastly, there is mention in table 3 of studies of superlattices. Advances in molec-
ular beam epitaxy deposition techniques have led to production, an atomic plane at a
time, of single crystal superlattices composed of altemating layers of magnetic rare
earth. These, and other, samples have been successfully investigated by magnetic
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 571

X-ray diffraction experiments, as often as not accompanied by corresponding neutron


scattering studies (Bohr et al. 1989, Vettier et al. 1986, Majkrzak et al. 1991).

3.3. Elastic resonant scattering


When the primary photon energy is tuned near an absorption edge, large resonant
enhancements of a Bragg intensity can be observed. Figure 3 shows the energy
dependence of the integrated intensity at the first-order (0, 0, 5/2) satellite to the
(0, 0, 2) charge reflection of UAs which occurs for the simple linear antiferromag-
netic configuration of moments established below 127 K. The intensity depends on
the specific absorption edge, photon polarization states and the magnetic state of
the sample. Of course, non-ferromagnetic materials, with purely magnetic Bragg
reflections, are most easily studied. An interpretation of elastic resonant scattering,
based on an atomic model of the magnetic state and a one-electron process, has been
shown to adequately explain the features observed in antiferromangetic UAs and
more complicated magnetic configurations. Within an intuitive one-electron picture
of the electronic structure, fig. 6 depicts the resonant enhancement process at the L3
edge of a model appropriate to rare earth materials. For this model, there are local-
ized 5d-states available in dipole-allowed transitions, and unfilled 4f-states available
through quadrupole transitions.
Some materials studied by elastic resonant scattering are listed in table 4. Reviews
of the work include (Gibbs 1992, Stirling and Lander 1993, Vettier 1994).
Elastic resonant scattering was brought to the fore of synchrotron-based scattering
studies of magnetic materials by work on holmium (Gibbs et al. 1988). Just be-
low the N6el temperature, 131.2 K, the magnetic moments order ferromagnetically
TABLE 4
Representative examples of materials studied by elastic
resonant scattering, see also table 1.

Material Ref. Edges


Er [4] L3
Ho [1, 2, 10] L3
TmSe [5] L3
UAs [3] M3,4,5
Fe [6] L2,3
Tb [7] L3
Dy [8] L3
UO 2 [9] M4,5
USb [9] M4,5
NpAs [11] M4
Hoo.sEr0.5 [12] L3

References:
[1] Gibbs et al. (1988) [7] Gehring et al. (1992)
[2] Gibbs et al. (1991) [8] Isaacs et al. (1989)
[3] McWhan et al. (1990) [9] Tang et al. (1992b)
[4] Sanyal et al. (1994) [10] Thurston et al. (1994)
[5] McWhan et al. (1993) [11] Langridre et al. (1994)
[6] Kao et al. (1990) [12] Pengra et al. (1994)
572 S.W. LOVESEY

within the hexagonal planes of the h.c.p, crystal structure to form a spiral magnetic
configuration with a turn angle per plane ~b ~ 50 ° along the c-axis. The magnetic
diffraction pattern consists of pairs of satellites offset from each of the chemical
Bragg reflections by w = (0, 0, t), where t = (2~b/c0) and c0 = 5.62 /k at room
temperature (the satellite position changes with temperature). The resonant effect
near the L3 edge at 8.07 keV, for which the corresponding wavelength = 1.54 A is
quite well suited to diffraction studies, increases the scattered intensity by a factor of
about 50. The observed intensities and polarization properties are consistent with a
model (Harmon et al. 1988, 1989) based on one-electron El, E2 processes (Gibbs et
al. 1988, 1991). Unlike a simple linear antiferromagnetic configuration, displayed by
UAs (McWhan et al. 1990), dipole (El) transitions for a spiral configuration produce
first- and second-order magnetic satellite reflections.
The model proposed in (Hannon et al. 1988, 1989) is outlined in section 9 for
the case of dipole-allowed transitions. Of course, there is a very close connection
between absorption and resonant scattering, since both processes are described by
the same part of the scattering amplitude. However, the absorption coefficient is
proportional to the imaginary part of the amplitude while the Bragg intensity is pro-
portional to the square of the absolute magnitude of the amplitude averaged over
thermal, concentration and defect fluctuations. On performing the average of the
scattering amplitude both magnetic and lattice properties are effected so, for ex-
ample, it contains Debye-Waller factors that arise from lattice vibrations. Just as
for magnetic dichroism, the thermodynamic properties of the magnetic state of the
sample influences elastic magnetic scattering through the averaging of matrix ele-
ments of the position operator with respect to magnetic quantum numbers. Regarding
the dependence of quantities measured in absorption and scattering experiments on
the polarization of the primary beam for the former experiment one averages the
amplitude, and for the latter one averages its absolute square.
The Bragg scattering amplitude is usefully developed in powers of the (unit) vector
which described the configuration of the average magnetic moments in the magnetic
structure. For example, a simple antiferromagnetic configuration is described by
re(R) = (0, 0, exp(iw. R)) where the phase factor has values -4-1 when R coincides
with moments on one or other of the two sublattices. The dipole (El) contribution to
the amplitude (9.5) is found to be a linear combination of the factors (s' .¢), m.(s' x ¢)
and ( # • m)(¢. m), i.e. terms of order 0, 1 and 2 in m. Figure 11 shows these three
factors in our 2 x 2 matrix representation of polarization states. If the moments are
contained in the plane of scattering, the term of order one is seen to have no diagonal
elements so, for example, cr-polarization is rotated to ~r-polarization. This rule, and
several others, has been confirmed by experimental studies (McWhan et al. 1990,
Tang et al. 1992b).
An incommensurate spiral configuration of magnetic moments, realized in Ho and
some other rare earth metals; is described by, re(R) = (cos(w- R), sin(w, g), 0), and
the E1 amplitude can contribute to a main Bragg reflection and satellites to it defined
by w and 2w, coming from the terms of order m and m 2, respectively. The intensity
of the first satellite in Ho approaches zero at the N6el temperature in a manner
characteristic of a continuous phase transition (Gibbs et al. 1991). This temperature
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 573

dependence of the intensity is consistent with the model under discussion. For, on
referring to section 9, the products of dipole radial integrals that multiply m . (e t x e)
in the scattering amplitude - the same combination of radial integrals which arise
in the circular magnetic dichroic contribution to the attenuation coefficient - when
evaluated within a localized model of rare earth magnetism (Thole et al. 1985) are
proportional to the ordered magnetic moment. The model predicts a temperature
dependence for the magnetic intensity at the main Bragg peak and associated 2w-
satellite which stems from thermal fluctuations in the mean-square magnetization.
The one-electron atomic model of elastic resonant scattering has been explored for
rare earth, actinide and 5d transition-metal compounds (Carra et al. 1989, Hannon

30
25 " T = 7 8 K
2o
15
~ 10
5
0 • . . . , . . , . ,

30
25 ' T = g 4 K
~2Ol
-~,

5
0 : A_
30
25 = 50K
~ 20
~ 15
-~ 10
5
0
30
25
.~ 20
;T-
g 15
-~ 10
5
0
I , I l I ' I ' I
30
25 T 34 K
.-~ 20
g 15
-~ 10
5
0
8300 8320 8340 8360 8380 8400
Energy [eV]
Fig. 13. Integrated magnetic scattering intensities (arbitrary units) of the scattering at the (0, 0, 2 + t)
satellite of erbium at various temperatures, T > 34 K, corrected for absorption. The solid lines are a fit
to a single oscillator at the L 3 = 8.36 keV absorption edge (Sanyal et al. 1994).
574 S.W. LOVESEY

et al. 1988, 1989, McWhan et al. 1990, Tang et al. 1992b), and applied to scattering
from surfaces (Fasolino et al. 1993). In Luo et al. (1993) it is extended to inelastic
resonant scattering, and sum rules are developed to separate out contributions from
spin and orbital magnetism.
Tang et al. (1992b) report ab initio atomic calculations of the scattering amplitude
at the M4 and M5 edges of uranium in UO2 (U 4+) and U S b ( U 3 + ) . With the magnetic
moments arranged to lie in the plane of scattering, the observed cr - 7r scattering is
proportional to the m. (e' x e) contribution to the amplitude (9.5). Radial integrals in
FaM(E) were obtained from a Hartree-Fock scheme, including relativistic corrections.
Fits to the experimental data are good for UO2, modelled by U 4+, indicating that an
atomic picture is useful. Some discrepancies are found between experimental data
and calculations for USb, which might indicate the possible need to go beyond an
atomic picture and include hydridization between f states and band states. In the
atomic model the magnetic order is brought about by use of a magnetic field.
For erbium at low temperatures an unusual behaviour of the magnetic scattering
as a function of energy has been observed. Looking at figs 13 and 14, the integrated
magnetic scattering intensity for the (0, 0, 2 + t ) reflection at temperatures above 34 K
is quite different from that seen at 12 K (Sanyal et al. 1994). Neutron diffraction
studies of erbium have identified three magnetic configurations. Below the Ntel
temperature of about 89 K and above 52 K, the moments are believed to be ordered
along the c-axis and longitudinally modulated. For temperatures between 52 and
18 K, an additional component of the magnetization develops within the basal plane,
forming a magnetic structure with a unique chirality. Below 18 K, there is a first-
order transformation to a commensurate magnetic structure which is believed to be
a conical phase. Looking at fig. 14 for T = 12 K, the cross-section is unlike the
single peak at the L3 absorption edge observed at the higher temperature. Instead,
what is found is a series of sharp peaks and broad humps extending as far as 2 keV
below the absorption edge.

0.20

0.15
t--

"=0.10

~0.05~
0

. e e we~,,,,, e ee
c

~,~ ~'~*" ~ .
0 t I , I , .. I

8000 8150 8300 8450


Energy [eV]
Fig. 14. As - measured integrated intensity (arbitrary units) for the (0, 0, 2 + t) satellite of erbium at
12 K, as a function of the primary photon energy. The vertical solid line indicates the position of the
L3 absorption edge (Sanyal et al. 1994).
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 575

,10 -=
I I "m I I I
t.50

IIt -
O
T=175.3K I.O0 E

0.50 "~
I.,J
10 -~ I..-
z
3

0.00
II
¢- o
~2 - % T=175K
oo
o
>- o
o
o o10-,
o o

_J 5

4~ C
0
• T=177K E
go

Beg •
>-
oo p-
,".. .,,,
Z

•. - ' : . . ' . . . . - 'z

• "".'"'.:'i g go

3.0t o

T=178K

,~ ~ o o

2o,o J/
o

1.C I , I ,
0,20 0.22 0.24 0.26 0.28

(0 K 2)
Fig. 15. Examples of the critical scattering observed for NpAs as a function of temperature at the
(0, 0.23, 2) peak (Stirling and Lander 1993, Langridge et al. 1994).
576 S.W. LOVESEY

Magnetic critical scattering has been observed in Ho (Thurston et al. 1993, 1994)
in the vicinity of the transition to a spiral antiferromagnetic phase at 131.2 K, and
NpAs (Langridge et al. 1994) in the vicinity of an incommensurate antiferromagnetic
phase at 173 K. Figure 15 contains examples of the critical scattering observed for
NpAs. In both cases, the measurements were performed with the radiation tuned in
energy close to a resonant level; for holmium a primary photon energy ~ 8.07 keV
is the dipole maximum of the resonant magnetic scattering at the L3 absorption edge,
while for the experiment on neptunium arsenide the M4 edge (~ 3.85 keV) of the
Np ion was utilized. The consensus view from the experiments is that the observed
critical behaviour is significantly different to that observed with conventional neutron
scattering techniques. There are apparently two length scales, and the new, longer
one may be related to the surface or be a result of long-range order nucleated by
impurities. Magnetic correlations with two length scales have been observed (Hill
et al. 1993a) in photon diffraction studies of the random field Ising antiferromagnet
Mn0.75Zn0.zsF2, and in this case the longer length correlations have been directly
related to surface preparation. It is early days for this type of work, and the use of
two length scales in data analysis might be a notion with a short lifetime.

3.4. Spectroscopy
It is well established experimentally that the cross-section for light scattered in-
elastically by ordered magnetic materials displays pronounced features due to the
collective spin oscillations, known as spin waves and magnons (Hayes and Loudon
1978), and in insulating materials their maximum energy is of the order of several
meV. Antiferromagnetic materials contribute two magnon effects (Cottam and Lock-
wood 1986, Rosenblum et al. 1994), in addition to one magnon events seen also
with ferromagnets. The intensities of these events are too large to be consistent
with excitation mediated by the magnetic dipole operator. Instead, a satisfactory ac-
count of the data, on one and two magnon events, is provided by the electric dipole
operator when account is taken of the spin-orbit interaction. In this instance, the
photon field couples to the orbital degrees of freedom of the unpaired electrons, and
thus through the spin-orbit interaction to the atomic spins. These, in turn, interact
through an exchange interaction, of the Heisenberg type, which supports collective
spin oscillations.
The two magnon events mentioned appear in spectra for antiferromagnetic mate-
rials and are insensitive to the application of a magnetic field. Their contribution
is in the form of a broad band of intensity, from two magnon sum and difference
events, and bound states condensed out of the two magnon continuum. The two
magnon and one magnon events have similar intensities, and the former, in fact, are
for some materials the dominant feature. These aspects of two magnon events pre-
clude second-order one magnon scattering as the underlying mechanism, since this
mechanism does not discriminate between ferro- and antiferromagnetic materials.
Instead, the key lies in considering processes that involve two neighbouring mag-
netic ions, coupled by the exchange interaction. Moreover, just as the Heisenberg
exchange arises from the electrostatic interaction, between neighbouring ions, and
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 577

the requirements of the Pauli exclusion principle so the same mechanism transfers
in opposite directions spins on neighbouring ions. For the N6el state, the transfer of
spins leads to an excited state without change in the total spin of the system which,
of course, is zero in both cases. Since there is no net magnetization the process
is not influenced by a magnetic field. In a ferromagnet, the creation of two spin
excitations requires a change (of two units) in the total spin and this is not allowed
in the proposed scheme. Finally, let us mention that there is a similarity between
the schemes proposed for two magnon events in magnetic materials and interaction
induced (sometimes called collision-induced) events in light scattering from non-
magnetic materials. For example, it is observed with a dilute fluid of noble gas ions
that scattered light is significantly depolarized, yet if the scattering is by individual
atoms the intensity is proportional to (e. e~)2 which vanishes for orthog0nal primary
and secondary polarization states. Analysis of the scattered intensity is consistent
with the idea that scattering comes from two, or more, correlated ions. At the low-
est level of approximation, the correlation is created by the electric dipole-dipole
interaction between neighbouring ions, responsible for the Van der Waals attraction
between neutral atoms.
The recent development of high resolution X-ray spectrometers, for the mea-
surement of phonon dispersions, plasmon peaks, particle-hole continuum, etc., is
reviewed by Burkel (1991). In the remaining part of this subsection, attention is
directed to measurements at synchrotron sources of Compton profiles for magnetic
materials. Background theory for the interpretation of measurements is covered in
sections 2 and 10.
At present, relatively few magnetic materials have been investigated by Compton
scattering; table 5 lists most materials for which Compton profiles are available.
This situation is likely to change in the near future as new facilities, such as the
TABLE 5
Representative magnetic materials for which Compton profiles have
been measured. In all but one example, circular polarization of the
primary beam is used to isolate the profile associated with unpaired
electron spins. Reference [2] establishes by investigation of several
materials that the Compton profile is not sensitive to orbital magnetism.

Material Ref. Comment


Ni [1] unpolarized 7 rays
Fe [6, 7] circularly polarized 7 rays
Mn ferrite [7]
Ni [3] synchrotron source
Fe [3-5]
Gd [8, 9]
HoFe2 [2]

References:
[1] Anastassopoulos et al. (1991) [6] Sakai and Ono (1977)
[2] Timms et al. (1993) [7] Sakai and Sekizawa (1987)
[3] Sakai et al. (1991a) [8] Sakai et al. (1991b)
[4] Cooper et al. (1986) [9] Kubo and Asano (1992)
[5] Tanaka et al. (1993)
578 S.W. LOVESEY

•• %
..- • •

• ° • "1
.; •. ; ;.
.." . • "%

• :- .. 0
¢' ,.-; .
%
.. "..•..-., "--.,..,_~I
;" .. ".. ". '.. ....
-'- ..
t/'l
¢- .. • • " ",,~.~.~211
%,
• . "• ,,,,,,,,
.d ~ •" ' • : a ° :. " ~ ~ 2 2
Y, -" ~e,, °.% "
• • •

ul
c- - ...... _-. "-,,,
t.- ,_. ". - . , .
J~ • ".°

,...."-'" "• : '" ". "~3 31

~" • .. - "~__ 110

-'-I":""
_ ,,_.-."..... " ;" ""

,,-" '-. - ": ~"'-,~__2~ o


i °.

°°,

"%'-,,,,,,,,,..,,~ I0
~I'IQIII~ e II

0 1 2 3 /-, 5 6 7 8 9
Pz (a.u.)
Fig. 16. Magnetic Compton profiles of ferromagnetic Fe ÷ 3 wt.% Si along fourteen crystal directions.
The evaluated momentum resolution is 0.76 a.u. (Tanaka et al. 1993).
PHOTONBEAMSTUDIESOF MAGNETICMATERIALS 579

elliptical multipole wiggler at the National Laboratory for High Energy Physics,
Tsukuba, Japan (Tanaka et al. 1992), attain full capacity. Reviews of experimental
and theoretical studies of Compton profiles of magnetic materials are provided in
Cooper (1987), Sakai (1992), while the well-established field of non-magnetic ma-
terials is reviewed in Cooper (1985). Most of the, relatively few, Compton profiles
available for magnetic materials have been subjected to detailed theoretical interpre-
tation (Cooper et al. 1986, Kubo and Asano 1992, Sakai et al. 1991a). It is likely
that, even more interest will be forthcoming with the advent of three-dimensional
Compton profiles reconstructed from many one-dimensional profiles (Tanaka et al.
1993). Figure 16 shows the magnetic Compton profiles of ferromagnetic iron along
14 crystal directions, which are the raw material for the reconstruction procedure
proposed in Suzuki and Tanigava (1989).
Recent work has used circularly polarized hard X rays, e.g., at the Tsukuba ellip-
tical multipole wiggler the energy of emitted photons is 60 keV. For materials with
a net magnetization, the Compton profile can be extracted from the polarization-
induced charge-magnetic interference scattering. This contribution to the total scat-
tering can be isolated by a differencing method; the sign of the interference scattering
is reversed by reversing the polarity of the net magnetization, by application of an
external field, or changing the handedness of the circular polarization in the primary
beam.
Turning to the data for ferromagnetic iron displayed in fig. 16, the measured
magnetic scattering intensity is typically 1% of the charge scattering and the count-
ing statistics give an accuracy of about 1% at the Compton peak. The good sta-
tistical accuracy was found to be essential for a meaningful reconstruction of the
three-dimensional Compton profile. All the main features of the data, including
the diminution of intensity at the origin (usually ascribed to a negative spin den-
sity), are reproduced by a one-electron theory based on the full-potential linearized
augmented-plane-wave method (Kubo and Asano 1990).

4. Kramers-Heisenberg amplitude

The results discussed here have applications in the interpretation of resonant and non-
resonant events. Effects due to the spin of an electron are not included, although it is
not very difficult to do so (Sakurai 1987). Instead, we delay our discussion of spin-
dependent events to section 6 where our most general expression for the scattering
amplitude is recorded. One reason for giving here a discussion of a less general
result for the amplitude is that it is possible to get an appreciation for the key ideas
and approximations while pursuing an algebraically simpler problem which, none
the less, is of use in the interpretation of experiments. In particular, the Kramers-
Heisenberg formula provides a basis for the contribution from orbital magnetism to
the diffraction cross-section, and the interpretation of dichroism and elastic resonant
scattering.
Photon scattering consists of the absorption of a primary photon and the simultane-
ous emission of a secondary photon. The wave vectors of the primary and secondary
580 S.W. LOVESEY

photons are q and q', and the scattering vector,

k=q-q'. (4.1)

The concomitant change in energy is,

hw = hc(q - q') =-- E - E'. (4.2)

The real polarization vectors are ¢ and #, and e . q = e' • q l = O.


To provide a compact expression for the scattering amplitude operator it is prudent
to define a momentum density,

J(q)=~pj exp(iq . R j ) , (4.3)


J

where the sum is over all the charged particles in the sample, and p and R are
conjugate variables. Note that the Hermitian conjugate,

J+ (q) = J ( - q ) + h q n + (q), (4.4)

so ¢. J+(q) = ~ . J ( - q ) . With this notation, the scattering amplitude operator, first


derived in 1925 by Kramers and Heisenberg, in units of re, is,

G = - ¢ . # n ( k ) - (1/m*){¢. J(q) ( E , - E ' - ~t.~)-I ~ t j + ( q , ) +


(4.5)
+ e'. J+(q')(E. + E - 7-~)-1 e . J(q)},

where ra* is the mass of the charged particles, the Hamiltonian 7-£ describes the target
sample, E u is the energy of the initial target state (7-/1#) = Eul#) ) and E = ~ q ,
E ' = hcq t. The resolvent operators (Eu - E / - 7-/)- 1 and (E u + E - 7-/)- I are defined
with an infinitesimal negative imaginary part added to 7-/ (not shown explicitly)
for pole avoidance. The avoidance rule is important when the poles of (4.5) with
respect to the energies of intermediate states are in the region of the continuous
spectrum; e.g., if the initial state is the ground state of an atom this would occur
for E exceeding the ionization threshold of the atom, whereas in a molecule the
threshold for dissociation into atoms takes the place of the ionization threshold.
With the result (4.5) for G, the differential cross-section is found to be,

d~r/df2 = ( m / m * ) 2 ( q / / q ) [ ( # 'IGI~) [2. (4.6)

This expression does not contain effects due to the spin magnetic moment of the
electrons, as discussed earlier in this section. But it does contain magnetic scattering
in the form of a contribution that can be identified as due to orbital angular mo-
mentum. The occurence of orbital angular momentum is taken up in the following
section.
PHOTONBEAM STUDIESOF MAGNETICMATERIALS 581

To conclude this section, let us consider the explicit form of one of the two
'resonance-type' contributions in G. The infinitesimal negative imaginary part added
to 7 / i s denoted by (-i~//2). The choice of notation is meant to convey the notion
that physically this contribution arises from damping, just like one has in the purely
classical theory of scattering. Selecting to examine the third term in (4.5), and
inserting a complete set of states for 7/labelled by quantum {~7} including all states
from the discrete and continuous parts of the spectrum, one finds,

6'. J+(q')[~)(~16. J(q)


(1/m*) E ~-~ ~ ~ - - ~-7 +-i7~ • (4.7)

in practice, % be neglected except when E "~ E, 7 - E u.


Most often, the resonant amplitude is much larger than the sum of non-resonant
amplitudes. This arises because the magnitude of the resonant amplitude, is of the
order k while the magnitude of the non-resonant amplitude if of the order of r~. If
the condition for resonance with a particular intermediate state is almost satisfied, so
E _~ E,7 - Et,, and the intermediate state is nondegenerate, the single-level inelastic
resonance formula is,

I ( U ' [ 6 " J + (q')177) (r/16 . J(q)l/~)12


d~r/dO = ( m r ~ / m * 2 ) Z ( q ' / q ) (4.8)
(Eu + E - E,7)2 + (%/2) z

If q and q' are small the matrix elements which appear in (4.8) can be calculated
using J(0) = P where P is the total momentum. The result,

(~'[PD) = ( i m * / h ) ( E ~ , - Ev)(dIRI~7 }, (4.9)

in which (/z'lRl~7) is an off-diagonal matrix element of the dipole moment operator,


is often utilized. The next term in an expansion of J(q) in q produces the total orbital
angular momentum operator interacting with a component of H = curiA, and the
quadrupole operator. Hence, beyond the dipole approximation, resonant scattering
contains processes that are explicitly magnetic in character.

5. Scattering by orbital magnetism

The object of the present discussion is to expose the orbital magnetic moment in
the Kramers-Heisenberg scattering amplitude, and demonstrate its simple relation
to the appropriate operator in the amplitude for magnetic neutron scattering. This
identification has an immediate practical benefit since it enables us to utilize the
knowledge of the properties of the operator in the literature on neutron scattering. In
fact, the non-resonant limit of G is shown to contain the operator Z defined in (A.7)
which is related to orbital magnetism. The experimental evidence is that orbital
magnetism is observed in Bragg diffraction but not in Compton scattering.
582 S.W. LOVESEY

Turning to the definition of the photon scattering amplitude operator (4.5), it


is evident that orbital angular momentum can stem only from the last two terms.
When E, E I are much in excess of the energies of states in the spectrum of N,
the Hamiltonian that describes the target, the resolvent operators in G can be safely
expanded in ( I / E ) and (1/E'). Keeping the first-order terms in such an expansion
leads to a non-resonant scattering amplitude operator,

G=-re . e .j(q)e,.j+(q,)}](5.1 )
[e . e ' n ( k ) + - - 1 { 1 e' . J + ( q ' ) e . J ( q ) - ~ _ E'I
m* -E
Since J(q) is the spatial Fourier transform of the linear momentum, the second two
terms are manifestly quadratic in the momenta of the particles. But, because p~ and
Rj do not commute there is, in fact, a term linear in the momenta and proportional
to k:

# . j+(q,) ~. j(q) = ~ e-iq'.R3 ¢1 .pj eiq.R~, ¢ . p y


jj'
(5.2)
-: Z e iq'R~'-iq''R~ e I. (pj + hq~jj,)¢.pj,,
j3,

in which q can be replaced by k because e'. q' = 0. In the subsequent development,


terms quadratic in p are dropped in favour of linear terms since the latter are the
most significant. Two features of the final result contribute to the licence for this
decision. First, the terms quadratic in p will generate smaller effects than those
from terms linear in p, given that, as we shall see, the latter produce a finite result.
The argument here is akin to that which leads to allowed electric dipole transitions
dominating magnetic and higher-order electric transitions. Secondly, the quadratic
terms assembled in G arise with a factor (E, - E I ) / E E ~ which vanishes for elastic
scattering (Bragg aad the static limit); terms quadratic in p are analysed in Grotch
et al. 1983. Hence, in future we shall use the approximate result,

e' . J+(q') e . J(q) = h Z exp(ik. Rj) (e' . k)(e .pj), (5.3)


J
where k = q - q l , as elsewhere in this chapter. On inserting this value in (5.1),

G = -re [e" ¢'n(k) - b-(m/ra*)(# x e). Z(k)]. (5.4)

Here, r = (hq/mc) and (5.4) is valid for ~- << 1 (Grotch et al. 1983). The operator
Z is defined in (A.7). It is closely related to an operator used in the theory of
neutron scattering to describe orbital magnetism (Balcar and Lovesey 1989). Since
the operator T is extensively studied in neutron scattering theory it is prudent to
import the knowledge to the present case of photon scattering. To this end we note
that, for elastic scattering,
Z(k) = - ( k / q ) 2 T(k), (5.5)
PHOTON BEAM STUDIESOF MAGNETICMATERIALS 583

and

(T(k)) = --(1/2#B) f dr exp(ik, r){k x ((Mz(r)) x k) }, (5.6)

in which Mt(r) is the orbital angular momentum operator, and k = k/k. Matrix
elements of T(k) for atomic 3d and 4f states, of a single multiplet, have been tabulated
by Balcar and Lovesey 1989.

6. Scattering by spin magnetism

The amplitude for spin-dependent scattering of photons by bound electrons has been
considered in great detail (Bhatt et al. 1983, Grotch et al. 1983). The expression
quoted here is derived in the cited papers by a systematic treatment of the quan-
tum electrodynamic Hamiltonian for a bound electron, and an expansion in powers
of the small parameter r = (primary energy/electron rest mass). In section 1 we
indicated how the result can be derived by addition of spin-dependent terms to the
non-relativistic interaction. While such a scheme is quite similar to the workings
reported by Grotch et al. (1983) it is more difficult to be confident about making an
internally consistent calculation. We find to order T = (hq/mc) that results from the
two approaches are identical. Hence, to this level of approximation it is not neces-
sary to employ the mathematical apparatus used by Grotch et al., but their scheme is
a safe route by which to obtain higher-order terms, containing products of spin and
momentum operators for example. If very hard X rays are used, "r is no longer a
good expansion parameter. In which case, an alternative route to follow is to expand
the amplitude in terms of the fine-structure constant.
Following on the development provided in section 4, we employ a scattering
amplitude operator G in terms of which the differential cross-section is (m = ru*),

d~r/df2 = (q'/q)l(#'lGl#)12. (6.1)

Here, # and #~ label the initial and final states of the target, and hcq and hcq~ are the
energies of the primary and secondary photons. Polarization of the primary beam is
accounted for in (2.10) in terms of a density matrix, and the result (2.5) explicitly
displays all possible polarization-induced processes.
In place of the momentum density (4.3) we need in the general case,

J(q) = Z ( p j + ihsj x q) exp(iq • Rj), (6.2)


J
for which,

{J(q) • e} + = J(-q). e, (6.3)

where e is the real polarization vector for the primary photon, ~. q = O. In (6.2), s5
is the spin operator of the electron labelled by j.
584 S.W. LOVESEY

The operator G calculated to first-order in ~- is obtained from eq. (2) in (Grotch et


al. 1983). Expressed in our notation,

G : - r ~ [e. ~' n(k) + (i/2)7-(1 + ( q ' / q ) ) S ( k ) . (~ x ¢') +

+ (l/m){~'. J(-q')(Eu + E - 7~)-1 fir • J(q) + (6.4)

+ ~. J ( q ) ( E u - E' - 7{)-1 ~,. j ( _ q , ) } ] .

Here, n(k) is the spatial Fourier transform of the charge density, and S(k) is the
spatial Fourier transform of the spin density defined in (2.6) and (A.6), respectively.
Examination of (6.4) shows that spin arises in the resonant and non-resonant events,
and the structure of (6.4) is similar to (4.5) - on setting m = ra* - apart from the
second term which is due solely to the spin density. The matrix representations of G
and the three combinations of polarization vectors are displayed in figs 10 and 11.
The evidence to date is that the result (6.4) together with (2.4), (7.1) and (6.1)
are cornerstones for the interpretation of photon absorption and scattering by mag-
netic materials. The partial differential cross section derived from (6.1), in terms of
.correlation functions, is given in (2.10).
The scattering amplitude operator G in (6.4) enables us to discuss resonant pro-
cesses and nonresonant processes. For the latter, resolvent operators are expanded in
( l / E ) and (1/E'). The expansion is justified when E, E' are much in excess of the
energies of states in the spectrum of 7~, the Hamiltonian which describes the charge
carriers in the sample. On keeping the leading term in such an expansion, we find,

G : -r~ [~. 6' n(k) - iT{S(k). B + (~' x ~). Z(k)}]. (6.5)

The operator Z(k) is defined in (A.7); it is intimately connected to the spatial Fourier
transform of the orbital magnetization, cf. (5.5) and (5.6). The quantity B in (6.5)
is a rather complicated function of polarization and photon wave vectors. Working
directly from (6.4),

8: ! x × x +
2 (6.6)
- ×

The first term in (6.6) is explicit in (6.4), whereas the three remaining terms arise
from the use of momentum and spin commutation relations in the components in
(6.4) that are quadratic in J.
The result (6.5) is the basis of our discussions of X-ray scattering by magnetic
materials. Evaluated for q = q' the amplitude (6.5) permits a discussion of Bragg
diffraction. The inelastic scattering process of main interest is Compton scattering.
For all these experiments there are benefits to be gained in utilizing the polarization of
PHOTON BEAM STUDIES OF MAGNETICMATERIALS 585

primary and secondary photons. A compact presentation of polarization phenomena,


employing a density matrix to handle states of partial polarization, is given in the
appendix.
The result (6.5) for G determines the four components c~ and #, listed in the
appendix, from which we obtain the full polarization dependence of the cross-section
and the state of polarization of the secondary photon beam. An outline of the steps
leading to our expressions for c~ and/3 begins with the definition,

G=l#+a.o,
in terms of the unit and Pauli matrices. The next move is to express the polarization
dependence that occurs in (6.5) in terms of 2 x 2 matrices using the basis defined
by figs 10 and 11. Of the three combinations of e and # appearing in G the most
complicated is B defined in (6.6), for which we find,

a(~x~') (a-b)~+(bcosO-a)~')
(6.7)
B: (a-cos0)~+(1 - a)~' (b+l - a)(~ x ~') "

with b = (q'/q) and a = (1 + b)/2.


If the matrix elements of n(k), S(k) and Z(k) that appear in the cross-section are
real then the latter, being related to the absolute square of the matrix element of G,
is quadratic in T. Hence, the magnetic contribution to the cross-section, weighted
by T2, is small compared with the charge contribution. Another factor reduces even
further the ratio of magnetic to charge contributions. Charge scattering engages all
electrons whereas magnetic scattering arises from the few which are unpaired. If, on
the other hand, the charge matrix element, say, is complex (anomalous scattering)
the cross-section contains an interference between charge and magnetic contributions,
i.e. a term weighted by T. Such a term is likely to dominate a term weighted by
~_2. Another means by which to create an interference between charge and magnetic
amplitudes is to employ primary photons with (complex) elliptic polarization. Terms
of order "r2 are most likely to be measured at purely magnetic reflections that occur
in non-ferromagnetic materials.

7. D i c h r o i s m

The diminution of intensity of a beam caused by a target foil is described by


exp(-RT), where 3' is the attenuation coefficient and R is the distance through
the foil in the direction of the incident beam. It is found that 7 depends on the
state of polarization of the primary beam and properties of the target; the polariza-
tion dependence is usually called dichroism. In many cases of practical interest, the
imaginary part of the index of refraction, n", and 7 are simply related, ~, = 2qn".
The bulk parameters 7 and n can be correlated with properties of the constituents
of the target sample. A simple result is obtained for a low density target, in which
the particle size is small compared with the average distance between particles. In
586 s.w. LOVESEY

this case, if n 0 is the density of particles in the target, each having the same total
cross-section atot, one finds,

= noO'tot.

Recall that, the total cross-section includes all possible elastic and inelastic processes
for a given initial state of the photon. Perhaps it is worth mentioning that, a target
sample can be absorptive without its individual constituent scatterers being so. Since,
if the latter only scatter, they thereby remove energy (intensity) from the forward
direction. In view of the significance of Crtot as measuring the removal of energy
from a beam, it is also sometimes referred to as the extinction cross-section.
The foregoing relation between the attenuation coefficient and O'tot stems from the
so-called optical theorem, which says that Crtot proportional to the imaginary part
of the forward scattering amplitude, denoted by f". Here, the forward scattering
amplitude means without change in wave vector and polarization. To be concrete,
we turn to our standard representation of the scattering amplitude G in terms of a
and/3, cf. eq. (2.4), where, a and/3 are now diagonal matrix elements evaluated for
= ~ . With this notation, the attenuation coefficient,

,,/= (41rno/ q) f " ( E) = (4~rno/ q) Ira{/3 + P . a}, (7.1)

where the Stokes vector P describes the state of polarization of the primary beam.
It is a matter of straightforward algebra to show that there is no contribution to "7
from the non-resonant magnetic scattering amplitude; specifically, from (6.7) and
(A.7) the matrix B = 0 and orbital operator Z = 0 for the condition of elastic
forward scattering, q = qt. In consequence, the magnetic contribution to 7 arises
from resonant processes in the scattering amplitude.
Of course, there is no explicit account in our result of the geometric optics part of
the scattering amplitude, modelled by Fresnel equations, which describes the absorp-
tion caused by the magnetically inert features of the target. The proper incorporation
of this with the magnetic scattering is not a simple task. For one thing, the magnetic
scattering is here treated by the Born approximation for single scattering events,
whereas geometric optics, and the refractive index, are intrinsically multiple scatter-
ing properties. One avenue for a consistent treatment is to consider scattering from
the sum of two systems, one defined as the geometric properties of the sample and
the second the magnetic constituents. The corresponding amplitude is the sum of the
exact treatment, using the refractive index and geometric optics, of scattering from
the sample's shape, and the amplitude for the magnetic component. The latter entails
a matrix element of the magnetic interaction operator evaluated not with plane waves
but primary and secondary waves distorted from plane wave forms by the sample's
shape. However, in our approximate treatment, we neglect all aspects of the geomet-
ric scattering to focus on the magnetic contribution calculated with the (first) Born
approximation. An improvement to our treatment, ia the form of explicit account of
the reflectance, can be accomplished in the framework of the distorted wave Born
approximation, familiar in the interpretation of particle scattering by nuclei.
PHOTON BEAM STUDIESOF MAGNETICMATERIALS 587

In the event that the target sample contains several scattering species, which in-
cludes the possibility of particles of different orientation, the attenuation coefficient
is the sum of individual attenuation coefficients.
A few more words are in order on the topic of the optical theorem. If atot(v,~) is
the total cross-section, which, in general, arises from elastic and also inelastic pro-
cesses, for a primary beam of polarization v and incident direction ~, and G~,~(~, ~)
is the corresponding scattering amplitude evaluated in the forward direction, the
optical theorem is embodied in the relation,

O'tot(u, q) =(47r/q)ImG~,~,(~,'~).

Bear in mind that IG~,~,(~,~')I 2 determines the cross-section for the scattering of the
radiation in an unchanged state, i.e. the cross-section for elastic scattering. Hence,
the optical theorem relates the imaginary part of the elastic scattering amplitude
in the forward direction to the total cross-section, which is the sum of elastic and
inelastic parts.
As we have just demonstrated, the attenuation coefficient is calculated from the
imaginary part of the forward scattering amplitude, which in turn is determined by
the so-called resonant contributions to the total amplitude (more specifically, the
contributions generated by second-order perturbation theory applied to that part of
the photon-target interaction which is linear in the vector potential). The quantity
actually observed is derived from the elastic forward scattering amplitude averaged
over the polarization states, denoted by f", as in (7.1). The influence on f " of the
magnetic state of the target appears in matrix elements of the current operator,

J(q) = ~ ( p j + ihsj × q) exp(iq • Rj), (7.2)


J

in which p, R and s are, respectively, the linear momentum, position and spin op-
erators for an electron, and the sum is over all electrons. Let the states I/z) and 1~7)
describe initial and intermediate states, respectively; the matrix element of interest
is,

Q(q) = (r/IJ(q)l/z >. (7.3)

Atomic states are appropriately labelled by total angular momentum J, and the
corresponding magnetic quantum numbers M, e.g., I/z) = IJM;t}, where £ is the
orbital angular momentum of the atomic shell, and Jzl#} = MI/Z).
The degeneracy of Q(q) with respect to magnetic quantum numbers is partially, or
perhaps wholly, lifted by the internal molecular magnetic field which results in the
magnetic splitting of multiplets of the order of 0.1 eV. In consequence, f " reflects
the target's bulk magnetic state. This result is not derived from the explicit spin
dependent term in J(q); indeed no account of the spin term is included in the subse-
quent discussion since, usually, it is a small correction to the contribution made by
the linear momentum. So, the magnetic character of f" arises from the distribution
588 S.W. LOVESEY

of magnetic quantum numbers generated by the target's internal magnetic structure,


which changes with temperature, applied magnetic field, etc. The (exchange) molec-
ular field is different from an applied field in that it acts only on the valence electron
spins and not the orbital moment or on the magnetic moment of the core hole.
Let us make the dependence of Q(q) on magnetic quantum numbers explicit. To
this end, we focus on the dipole approximation for the current operator, achieved
with q = 0, i.e. the current operator is replaced by the momentum operator (as
already mentioned, spin is neglected in the present discussion). In this limit,

Q(q = o) = (imA/h)(rllRl~), (7.4)

where the energy difference A = ( E n - Eu). Of course, the dependence on mag-


netic quantum numbers of the dipole matrix element (~IRI~) is readily derived by a
straightforward application of the Wigner-Eckart theorem (Cowan 1981, Weissbluth
1978),

J' 1 J)
(J'M';g"]RqlJM;£)=(-1)J'-M' -M' q M (J'IIRIIJ)" (7.5)

Here, q = O, + 1 labels spherical components which are related to Cartesian compo- "
nents through,

R+] = - ( 1 / v ~ ) ( R ~ + iRy),

Ro = Rz, (7.6)

R_ 1 = (1//V~)(Rx -- i R u ) .

Hence, the magnetic quantum numbers actually factor out of the matrix element, in
the guise of a 3j-symbol. Later on, we will discuss the other key factor in the matrix
element, (J' IIR[I J), which is a so-called reduced matrix element.
The 3j-symbol vanishes unless the arguments satisfy specific conditions that in
effect select some processes and forbid others. These selection rules are illustrated in
fig. 7. First, the triangular condition limits J' to the values J' = J, J + 1. Secondly,
the 3j-symbol vanishes unless the magnetic quantum numbers in the argument satisfy
M ' = q + M . From this result it follows that the matrix element of the z-component of
R has no off-diagonal elements with respect to magnetic quantum numbers, whereas
the x- and y-components contain only M' = M + 1.
A selection rule on the orbital angular momentum quantum numbers g, g' is con-
tained in the reduced matrix element (J'[[R[[J). This vanishes unless g' = g-4- 1.
While the latter result appears immediately in a full calculation of the dipole ma-
trix element, for the moment, we will justify it using an argument based on parity
considerations. Evaluation of the matrix element entails an integration over all di-
rections of the vector R. The integral vanishes if the intergrand has an odd parity,
i.e. if the integrand changes sign under the operation R --+ - R . Evidently, R has
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 589

an odd parity, so a necessary condition for a non zero integral is that the product of
the two wave functions also has an odd parity. The parity of a state with angular
momentum g depends on whether ~ is an even or odd integer. Therefore, the parity
of a product of wave functions, labelled by ~ and 6, is odd if g' differs from ~ by
an odd integer. An additional constraint imposed by the triangular condition, which
stems from coupling three spherical vector components, leads to the specific values
~' = ~4- 1.
Not surprisingly, experimental conditions also impose selection rules on events
observed in the attenuation of a beam passing through a magnetic material. For
example, if the target has a net magnetization, and the primary photon beam is not
perpendicular to the easy axis, the absorption coefficient is changed by reversing the
direction (polarity) of the easy axis, by application of a magnetic field. The effect is
absent when the beam has no circular polarization, and polarity reversal is found to
be equivalent to changing the handedness of the circular polarization. Usually called
circular dichroism, the effect is a manifestation of the interaction between the photon
spin and the net magnetization (circular dichroism is not observed in paramagnets or
antiferromagnets not subject to a magnetic field). The photon spin, also referred to
as helicity, is the angular momentum along the direction of the wave vector. It has
the values +-1, i.e. the helicity of a photon is a conserved quantity and on absorption
there is a change of + h in the component of the angular momentum of the target
in the direction of the beam. (With our notation, the component P2 of the Stokes
vector is the mean value of the helicity of the primary beam.) So, circular magnetic
dichroism is described by the transverse components of the dipole matrix element
q = +1, for these, and not the z-component, possess off-diagonal matrix elements
in which M' - M + 1. (It is amusing to observe that, if the helicity equals 4-2 then
the dipole matrix element would vanish.)
While linear polarization does not induce a coupling to the net magnetization, and
therefore provides no selection rules, nevertheless it is a useful external experimen-
tal parameter. As one might expect by physical intuition, linear dichroism yields
information on the mean square fluctuations in the magnetization, present in ordered
(ferro-, ferri-, anti-ferromagnets) and paramagnetic systems. As an example of the
absence of a selection rule present in circular dichroism we mention that, in general,
linear dichroism is present in the ordered magnetic state for all orientation of the
beam relative to the easy axis.
The foregoing treatment has been couched in terms of matrix elements evaluated
in the IJM) basis. Had we opted to use basis [gm)[sms) for the discussion we
would have noted that, the matrix element is diagonal with respect to spin quantum
numbers, and also derived the selection rules g' = g + 1 with m' = q + m.
To keep the mathematical features of the immediate discussion to a minimum, we
consider a simple physical model in which one electron in a core orbital, labelled
by #, is promoted to a single intermediate orbital, ~/, after absorption of a photon of
energy E = hcq. For an atomic model of the type used here, the various transitions
in Yb 3+ ( 4 f13
) are illustrated in fig. 7. Atomic structure and spectra calculations are
reviewed in (Cowan 1981). A more realistic model contains many intermediate states
individually weighted by the probability that they are vacant in the atomic ground
590 S.W. LOVESEY

state, cf. (7.18). However, the one electron model is a standard approximation. The
intermediate orbital has a total decay width F (of the order of a few eV for L edges
in rare earth ions) due to both radiative and nonradiative (Auger and Coster-Kronig)
processes, and it consists of the initial configuration with a hole in a core level and
an additional electron in a higher energy level (E, 7 > Eu). A nice example of the use
of the one electron model, to L2,3 absorption by ferromagnetic rare earths, is given
by Jo and Imada (Jo and Imada 1993). These authors also give an interpretation of
the working of the sum rules, and discuss the vexed question of the overall sign of
the circular dichroic signal derived from experimental data.
The starting point in the calculation of the forward scattering amplitude is (6.4).
For elastic scattering, we need the diagonal matrix element of the scattering amplitude
operator taken with respect to the state labelled #. It is convenient to employ the
notation (7.3) and A = ( E , 7 - E u ) . The relevant matrix element is,

(~IGI~) : - r e [(~l~" tin(k) + i~-S(k). (~ x d)]~ ) +


(7.7)
+ (1/ra)(z' • Q*)(e. Q ) / ( E - A + iF/2)],

and for forward scattering k = 0. It remains to average the matrix element with
respect to the polarization states. To this end, use the matrix representations of the
three combinations of polarization vectors as shown in fig. 11. The choice of axes
to describe the beam relative to the magnetic properties of the sample is displayed
in fig. 17. With the notation,

Tll = [Q~I z, T2z = IQu cos~p - Q~ sin cp[2,

Fig. 17. Geometry of a photon beam attenuation experiment. The wave vector q lies in the y - z plane,
and makes an angle ~o with the z-axis.
PHOTON BEAM STUDIESOF MAGNETICMATERIALS 591

and

Here, Z is the number of electrons per atom, S is the average spin for the state #,
and T~2 and T(~ are the real and imaginary parts of T12, respectively. The first two
terms in f(E) are purely real, and do not contribute to the attenuation coefficient,
for which we find from (7.1), on taking the limit F ~ 0,

Later, this expression is developed with the addition of two simplifying assumptions;
first, use of single atomic orbitals and, secondly, the dipole approximation to the
matrix elements Q(q) that make up Tij (the spin term in J is neglected).
Before doing so, we remark on the relation between the real and imaginary parts
of f(E) which stems from causality. The relation is the bridge between dichroism
and the Faraday effect, f(E) depends on E through the simple denominator in the
resonant contribution, and F > 0. One sees that f(E) has no poles in the upper-half
complex plane, and thus f ' and f " are one another's Hilbert transforms; in truth, it
is not f(E) which satisfies these relations but, on account of the terms in Z and S,
the subtracted function,
(I(E) - Y(Eo))/(E - Eo),
where E0 is arbitrary.
Let us continue the discussion of 7 making use of the dipole approximation (7.4)
and the Wigner-Eckart theorem (7.5). If the states # and r/in Q(q) have the same
magnetic quantum numbers the components Q~ and Qy vanish, while if they have
different magnetic quantum numbers Qz vanishes. These rules bring significant
simplifications to T~j since there are no terms involving Q~ mulitplied by either Q~
or Qy. In particular, T12 is purely imaginary, so V is independent of P1. Furthermore,
592 S.W. LOVESEY

and

r(& = -51 cos (mA/h) 2 { I< lml I.>12 - r< JR_l I >r 2 } .

Making use of the simplifications in (7.9) the attenuation coefficient reduces to,

47r2nore
7 = - mq
- (~(E - A ) t, + 1 T l l ( 1 _1_c0s2~ @ p 3
- P2T~2 sin2 ~) +
(7.11)
+ ~ (l - P3)[Q0] 2 sin 2 ~ .

Note that, for ~ = 0 (7r/2) the attenuation coefficient is independent of P3(P2)- To


complete the specification of the model we give the reduced matrix element in the
Wigner-Eckart theorem,

3 j, _1
(J'IIRI]J) = ( - 1 ) 2 (R)g,e{(2g + 1)(2g' + 1)(2J + 1)(2J' + 1)} 2 x
(7.12)
(~ 1 ~){J' 1 J}
x 0 g 1/2 gl ,

in which (R)ee is the radial dipole matrix element. The last factor in the reduced
matrix element, a 6j-symbol, contains four triangular conditions; these apply to the
coupling of (J', 1, J), (J', 1/2, g'), (J, 1/2, g) and (g, 1, g~) and the 3j-symbol vanishes
unless (g + g ' + 1) is an even integer. By way of orientation to the value of (R)e,e we
consider its value for 2p -+ 3d in nickel. A Hartree-Fock calculation (Cowan 1981,
van der Laan 1994b) gives a value (R)zp,3d = 0.125%. It is interesting to compare
this value with the value obtained by use of hydrogenic wave functions, namely,
4.748(ao/Z ) which gives 0.170% for Z = 28 (Ni). The tolerable agreement between
these two estimates of the 2p --+ 3d dipole matrix element of nickel is by no means a
typical situation. For one thing, the hydrogenic estimate is always positive, whereas
the use of non-hydrogenic wave functions, as often as not, produce negative values.
In arriving at the result (7.11), on which much of the subsequent discussion is
based, atomic states have been represented by single orbitals. The validity of this
license is open to scrutinity when the crystal field perturbation is large, since the
atomic states are most likely to be tolerably described by a linear combination of
orbitals with different magnetic quantum numbers.

7.1. Circular dichroism


As explained in section 2, circular dichroism is often detected by performing mea-
surements of beam attenuation with two opposite values of the circular polarization.
For the model under discussion, the difference in 3, for the extreme values P2 = 4-1
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 593

is proportional to Y~ which, in turn, is proportional to the difference of the radial in-


tegrals [(~[R±I I#)l2. The following model calculations of the matrix elements make
clear their dependence on the magnetic state of the sample. This entails averaging
T{t2 with respect to the magnetic quantum number M explicit in the 3j-symbol in the
dipole matrix elements. For the averaged T(~ we introduce the notation,

_ ! c o s ~ ( m A / h ) 2 ( j ' l l R I I d ) 2 F ( j , j'),
2
and calculate F(J, ji) using the molecular field approximation. We find that, F(J, jr)
is proportional to the average magnetic moment, and therefore vanishes in the para-
magnetic state. The atomic magnetic states involved for Yb3+(4f13), for example,
are illustrated in fig. 7. Our choice of axes, and the definition of the angle ~, is
provided in fig. 17.
Magnetic dichroism in the M4 and M5 absorption edge structure of rare earths
leads to consideration of the transition 3d -+ 4f. The initial state 3dl°4f'~(J) is
determined by Hund's rule. Dipole transitions are to the configuration 3d94f'~+l(J I)
with J - J' = 0, :k l; there are a plethora of states in the configuration that satisfy
these selection rules, many of which will not be resolved in energy. Even so, in a
study of linear dichroism taken up in the next subsection, Thole et al. (1985) correctly
predicted that significant structure exists in the absorption spectrum (van der Laan
et al. 1986). The dichroic effect is more pronounced when the final states with
different Jt have significantly different mean energies. This can occur in the rare
earth M4,5 edges due to the electrostatic interaction between the 3d and 4f electrons.
As a guide, in the M4.5 edges the J' = J + 1 states are 2 - 3 eV lower in energy,
and the J~ = J - 1 states 2 - 3 eV higher than the J' = J states.
Referring to the matrix elements that occur in T(~, the (2J + 1) degeneracy of the
initial state is lifted in an ordered magnet by the Weiss molecular field. The interionic
interaction responsible for this field couples pairs of ions at sites labelled by m and
n. If Z(n) is the exchange interaction, defined to have the property Z(n = 0) = 0,
the Heisenberg exchange interaction between spins is described by the Hamiltonian,

7-I = - ~Z(n)Sm " Sm+,~ = - ( g - 1)2 ~ Z ( n ) J m "Jm+n,


??rt, , T~ r P . ~n

where the second equality is valid when the exchange is smaller than the spin orbit
splitting, and g is the Land6 factor. Within the molecular field approximation, the
magnetic moment is,
J
(jz) = ( l / Z ) ~ M exp(2Mu),
M=-J

in which Z is the partition function, and

u = ( J Z ) ( g - 1)2 ~ Z ( n ) / k s T .
n
594 S.W. LOVESEY

The critical temperature Tc at which the magnet orders satisfies,

(9 - I)2 ZZ(n) = 3ksTc/2J(J + I).


n

From,

Z = E exp(2Mu): sinh {u(2J + 1)}/sinh(u), (7.13)


M

it follows that,
1 d
(J~) - Z = JBj(u), (7.14)
2Z du
where the Brillouin function satisfies,

2JBj(u) = (2J + 1)coth {u(2J + 1)} - coth (u). (7.15)

From the definition,

F(J,J') = ( 1 / Z ) Z e x p ( 2 u M ) { ( _ _ J r 1 11 j ) 2 -
M
(7.16)

1 -M -I

one finds,

F(J, J - 1) = -<JZ)/{J(2J + 1)},


(7.17)
F(J, J ) = _ < j z ) / { j ( j + 1)(2J + 1)},

and

F(J, J + 1) = ( j z ) / { ( j + 1)(2J+ 1)}.


Hence, F(J, J') is proportional to the magnetic moment. The complete expression
for T(~ which is the weighted sum of many terms proportional to F(J, JO, must
reflect the thermodynamic properties of the rare earth magnet, e.g., Y[t2 continuously
decreases as the temperature approaches Tc, and it vanishes in the paramagnetic state
T > Tc.
We conclude this section by recording expressions for the matrix elements which
are more general than those obtained so far, and also the expression used by Carra
et al. (1991) in their study of E1 and E2 magnetic dichroism in gadolinium metal.
PHOTON BEAM STUDIESOF MAGNETICMATERIALS 595

Electrons participating in the absorption process are labelled by j. The probability


that the intermediate orbital ~7is vacant for a given initial state # is denoted by pu(o),
and the (Boltzmann) probability distribution for the initial states is denoted by pu
(at absolute zero, pu = 1). Let (L = 1,2 . . . . . and - L ~< M ~< L),

WL -- q(27ce)2 ((L + 1)/L) Zp~pu(~)5(Et~ + E - Ev) x

2 (7.18)

where JL(X) is a spherical Bessel function and y L ( ~ ) is a spherical harmonic (Weiss-


bluth 1978). For small arguments,

JL(X) ~ xL/(2L + 1)!!.


The dipole (L = 1) and quadmpole (L = 2) contributions to the circular dichroic
part of the attenuation coefficient "/are,
7 (') = (67rno/q 2) cos~ { W11 - wL1 }, (7.19)

and,
3'(2)= (101rno/qZ) c o s ~ { [ w Z - w 2 _ 2 ] sin2~ + [W2 - W_21] cos2~}. (7.20)

Note that W L is dimensionless. Using the small argument expansion of the Bessel
function, it is quite easy to demonstrate that (7.19) is consistent with the previous
result obtained for a simple model with single component initial and intermediate
states with non-vanishing dipole matrix elements.
In their study of Gd, Carra et al. (1991) use relativistic Bloch functions to evaluate
E1 matrix elements with 2p core states. For comparison with experiment, they added
a core-hole lifetime (about 4 eV) and an energy-dependent intermediate-state lifetime.
The calculation of the quadrupole spectra is made with an atomic model based on
Hartree-Fock theory with relativistic corrections.

7.2. Linear dichroism


In order to exhibit the salient features of linear dichroism obtained by Thole et
al. (1985) and Goedkoop et al. (1988a), we continue to use their model of the
absorption process developed in section 7.1. Figure 18 contains their prediction
of the dichroic effect in Dy. The starting point of our calculation is (7.9) for the
attenuation coefficient.
Synchrotron radiation is predominantly linearly polarized in the plane of the elec-
tron beam orbit, and described by the Stokes vector P -- (0, 0 , - 1 ) . Inserting this
value in (7.9), the attenuation coefficient is given by,
7 = (47r2nor~/mq)6( E - A) T22. (7.21)
596 S.W. LOVESEY
| I I

Dy

,I.~ T >T c
>.-
l--
.J

ISS
0
n

Z
o b T=0

C // T =0
I I
1285 1295 1305
EXCITATION ENERGY (eV}
Fig. 18. Calculated spectrum for linear dichroism of Dy (a) for T > To, and for T = 0 with the
polarization (b) perpendicular and (c) parallel to the direction of the molecularfield (Thole et al. 1985).

If the orbitals I/z) and 1~7) are characterized by a single magnetic quantum number,
as described in the text leading up to 7.11), then some products of dipole matrix
elements vanish leaving,

T22 = (IQ~I 2 c o s 2 ~ ÷ IQ~I 2 sin 2 ~P

(77/,A/h)2[~c0s2~{I('QIRq-ll/Z)]2÷I(T]IR-II/Z)[2} ÷ (7.22)

+ sin 2 ~l<~lml/z>12],
where ~ is the angle between the photon beam and (magnetic) z-axis, as depicted in
fig. 17. We find that linear dichroism is sensitive to thermal fluctuations in the mean
square magnetization of the sample which is finite at all temperatures including, of
course, the paramagnetic state.
After averaging over the (2J + 1) degeneracy of the initial state we find,

I(O[Rol#)l a ---4 (J'llnllj)2Fo(J, J')


PHOTON BEAM STUDIESOF MAGNETICMATERIALS 597

where,
1J) 2
Fo(J, J') = M~ exp(2uM) 0 M "

Evaluating the 3j-symbol leads to,

Fo(J,J- 1)= (J2-((J~)2))/{J(4j2- 1)},


(7.23)
Fo(J, J) = ( ( j z ) z ) / { j ( j + 1)(2J + 1)},

and F0(J, J + 1) is obtained from the first result by making the substitution J --+
(J + 1). Further algebra yields,

(7.24)
> (J'[[R[[j)2{ (1/(2J + 1)) - Fo(J, J')},

for J' = J, J + 1. Hence, for the chosen model, the thermodynamics revealed in
linear dichroism appears solely in ((j~)2) which appears in Fo(J, J').
We conclude with some remarks on the value of ((jz)2). In a pure paramagnet,
the result,
1
<(jz)2> _- 3 j ( j + 1),

follows because J . J = J ( J + 1). The corresponding values of Fo(J, J') are,


Fo(J, J') = 1/[3(2J + 1)].
Inserting this in (7.22) one finds that 2"22 is independent of ~, as it should be in
the absence of a preferred magnetic axis. On the other hand, in the ordered state
(T ~< Tc) the molecular field approximation developed in section 7.1 provides the
estimate,
1 d2
- Z = J ( J + 1) - (J~} cothu. (7.25)
((j~)2) 4Z du 2

The result for <(j~)2) is consistent with the standard identity for spin 1/2 operators
((j~)2) = (jz)2 = 1/4. In the limit T -+ Tc, u -+ 0 and from (7.25),

) ~1 J ( J + l ) { 1 + ~2 u Z ( 2 J - 1)(2J+ 3) }

1 (2J+3)(2J- 1) T - T c ]
=- - J(J + 1) 1-
3
598 S.W. LOVESEY

where the second line is achieved with results provided in section 7.1. In the opposite
extreme, T --+ 0, only one initial state is accessible and ( ( j z ) 2 ) = j 2 . In this instance,
Fo(J, J - 1) = 0 while,
Fo(a, J) = J / { ( J + 1)(2J + 1)}, Fo(J, J + 1) = 1 / { ( J + 1)(2J + 3)}.

So, for T = 0 and J ' = (J - 1), the matrix element 7"22 is proportional to cos 2 ~,
while for J ' = J, J + 1 it is a weighted sum of cos 2 ~ and sin 2 ~, although this
angular dependence vanishes for J = 1/2.
In terms of the more general formalism surveyed at the end of section 7.1, the E1
attenuation coefficient is,

%(')(E)-- (6rrno/q 2)
{1
-~ [WJ + VV"I1] COS2 ~0 q" W 1 sin 2 ~
} . (7.26)

This result describes the so-called white line in the absorption spectrum, and it is
the companion to (7.19) for circular dichroism. In the appropriate limit, (7.26) is
compatible with 7.22).

8. Diffraction

From the discussion provided in section 2 it follows that strictly elastic scattering
(diffraction) is described by (2.5) and (A.5) when products of operators are taken as
products of thermal averages of the operators. So, for example, the term/3+(a • P)
in (2.5) is (/3+)(a) • P for strictly elastic scattering, achieved in Bragg diffraction.
The same interpretation holds for terms in (A.5) for the polarization of the secondary
beam.
As a first example of the use of (2.5) and (A.5) consider diffraction from a non-
magnetic material. Referring to (A.8), we have for this case,

O~1 = O~2 = O,

and

1
% = - ~ re(1 - cos O)n(k),
(8.1)
1
/3 = - - re(1 + cosO)n(k).
2

Let,

(n(k)) = N(k), (8.2)

which means that N(k) is the average particle density for the sample, and it is usually
a complex quantity, N = N' + iN".
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 599

For the cross-section we find the value,

1 2
do-/dS'2 = elN(k)12{1 + cos 2 0 + P3 sin2 0}. (8.3)

It is interesting to observe that (a) the cross-section is independent of circular po-


larization and of the Stokes parameter P1, and (b) the cross-section for photons
polarized perpendicular to the scattering plane (~r polarization, P3 = + 1), defined by
q and q~, is larger than for photons in the scattering plane Qr polarization, P3 = - 1 ) .
Stokes parameters for the secondary beam follow from (A.5) and (8.3). The results
are given in (2.9).

8.1. Unpolarized primary beam


As the second case we choose to examine diffraction of an unpolarized photon beam
by a magnetic crystal. When q = qr the operator Z(k) reduces to,

tug(k) = (i/q 2) ~ exp(i k. Rj) (k x pj), (8.4)


J

and for elastic events, Z(k) is related to the orbital magnetization through (5.5) and
(5.6). Equation (8.4) shows that k • Z = 0. The quantities a 1, a 2, a 3 and fl are
obtained from (A.8) by taking a = b = 1; in units of re they are,

O~1 =- (i/2)~-(1 - cos 0 ) ( ~ - ~ ' ) . S(k),

O~2 ~-~ - - - ~1 "r('~+ "~'). (Z(k) - (1 - cos O)S(k)), (8.5)


1
a 3 = - ~ (1 - cosO)n(k) + (i/2)'r(~ x ~ ' ) . Z(k),

and,

1
= - ~ (1 + cos O)n(k) - (i/2)~-(~ x ~ ' ) . (Z(k) - 2S(k)).

Here, the small parameter ~- = (incident energy/electron rest mass) = (E/mc2). We


choose to write,

(s(k)) - l fdr e x p ( i k . r ) (Ms(r)). (8.6)

For a perfect magnetic crystal, (Mz(r)) and (Ms(r)) are periodic functions that can
be expressed as Fourier series in reciprocal space.
600 S.W. LOVESEY

Following Lovesey (1987a), the position of a unit cell in the crystal is defined by
a lattice vector l and the cell volume is denoted by %. An atom position Rj = l + d,
where d defines the position within a unit cell. The site d = 0 coincides with the
comer of the cell, so there are (r - 1) non-null vectors d. Vectors of the reciprocal
lattice {-r} satisfy

exp(il, r ) = 1 for all 1. (8.7)

Because (Mz(r)) is periodic its spatial Fourier transform satisfies,

1 (2~) 3
f d r exp(ik, r) (Mz(r)) = E 5(k - "r)Fz('r), (8.8)
2# B at2 Uo ~.

and a similar expression for the spin density. The delta function in (8.8) expresses
the fact that diffraction from a perfect crystal does not occur unless strict geometric
conditions are satisfied, namely, Bragg's law. We refer to Ft(~-) and F~(T) as magnetic
unit-cell structure factors for orbital and spin magnetizations, respectively. Inverting
(8.8) we get,

( - 1/2#s)(Mz(r) ) = 1 ~ exp(-i r . r)Fz(r), (8.9)


T

and the conjugate relation,

FL(T) = fcell dF exp(i -i-. r) (Mz(r)}/(--2#s), (8.10)

together with analogous relations for the spin magnetization.


Consider a simple magnet with one atom per unit cell. At least for moderate
values of k, the atomic structure factors for spin and orbital contributions to the total
moment are almost the same. Let f(k) be the atomic form factor, with the standard
normalization f(0) = 1. So, to a good approximation,

FL(k)= l(L)f(k),
and (8.11)

Fs(k)=(S)f(k),

where L and S are the (total) orbital and spin angular momentum operators for the
magnetic ion. For a 3d-transition metal ion it is usual to define a gyromagnetic factor
go by,

(L) = ( g o - 2)(S), (8.12)


PHOTON BEAM STUDIES OF MAGNETICMATERIALS 601

in which case the ratio of Ft to F , is simply (g0 - 2)/2. On the other hand, a rare
earth ion is characterized by the total angular momentum J -- L + S. For a given
J-manifold, L = 9zJ and S = 9 s J with,

9t = (2 - 9), 9s = (9 - 1), (8.13)

in which 9 is the Land6 factor.


Finally we note that all the cross-sections involve products of densities of the form
(Lovesey 1987a),

( - 1/2#B ) {N(k)}* / d r exp(ik, r) (Mz(r))


(8.14)
-----No ((2rr)3/Uo) E 5(k - r ) F ; ( r ) F z ( r ) ,

where No is the number of unit cells and Fc(k) is the charge unit cell structure
factor. If the chemical and magnetic structures are different there exist r for which
F t ( r ) = 0 and/or F , ( r ) = 0 while Fc(r) ¢ 0, and vice versa, i.e. {r} generally
comprises pure charge and pure magnetic reflections, and mixed reflections.
In subsequent expressions for diffraction cross-sections the factor,

(2rr)3 E 5(k - r ) ,
No--g-
°
T

on the right-hand side of (8.14) is omitted, simply to minimize the notation. The
factor cancels in expressions for the Stokes vector for the secondary beam. Thus,
subsequent expressions for cross-sections and P' are understood to refer to a Bragg
reflection defined by {'r}. With our chosen notation, (oe} and (~3) are linear combi-
nations of the unit cell structure factors for charge (Fc), spin (F,) and orbital (Fz)
densities. Working directly from (8.5),

(OZl) ~-- r ~ ( i / 2 ) r ( 1 - cos 0)(~ - ~ ' ) . F s ,

(1)
(c@ = r~r(1 - c o s O ) ( ~ + ~ ' ) • F, + ~ F , , (8.15)

(O~3) =- -r~(1 - cos0) {'


~ Fc + ir(~ x ~ ' ) . Ft } ,

and

(/3) = - r ~ 1 (1 + cos O)F~ + r~iT(~ X ~ ' ) . {(1 -- COSO)fl + f~}.


602 S.W. LOVESEY

In general, Fc, Fz and Fs are complex quantities. For a centrosymmetric crystal


structure, Fl and F~ are purely real, as is Fo unless the photon energy is tuned to
a resonance for an atom in the unit cell. In some materials the spatial anisotropy
of the atomic environment is sufficient, for wavelengths near an absorption edge, to
give rise to dichroism, birefringence and anisotropic charge scattering. The example
of tetrahedral anisotropy, realized in germanium, is discussed by Templeton and
Templeton 1994. An example of a structure for which Fz and F~ can be complex is
the hexagonal close-packed structure adopted by rare-earth metals. It is interesting
to observe that, if the spin and orbital moments are perpendicular to the plane of
scattering defined by q and q~ then (OZl) = (OZ2) = 0, while if F~ and Ft lie in
the plane (c@ and {fl) contain only the charge density. Approaching the forward
scattering position, 0 -+ 0, all the components of (a) tend to zero, and {/3) contains
only the charge density. With the component of Fs in the plane of scattering arranged
perpendicular to k it follows that (a 1) = 0, whereas when the components of Fs and
Ft in the plane are arranged parallel to k one has (a2) = 0.
With an unpolarized primary beam the diffraction cross-section for the scattering
geometry depicted in fig. 19 is,

(d~/d£2)o = r~ { sin4(0/2)[T2[ ( ~ - ~ ' ) . F~[2+

-+- T2](q - [ - a t ) . ( 2 F t + Fs)]2 + [Fc + 2iT sin 0F~[ 2] + (8.16)

1
4- ~ [(1 + cos O)F~ - 2iT sin 0((1 - cos O)Flu + F~)12 .
}

Fig. 19. Coordinate system adopted in section 8 for the description of diffraction experiments. The x - z
plane coincides with the r/- ( plane (fig. 12), k = q - q/ and (~ x ~1) is parallel with the B-axis (not
shown) which is aligned opposite to the ~-axis. An alternative coordinate system is used by Blume and
Gibbs 1988.
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 603

In this expression, the superscript y denotes a component perpendicular to the plane


of scattering in the direction opposite to the (-axis. An interference between charge
and magnetic amplitudes is possible at mixed Bragg reflections for which there are
components of F~ and Ft perpendicular to the plane. For a pure magnetic reflection
Fc = 0 and in general both Fs and Ft are different from zero. Above the ordering
temperature Ft = F~ = 0, and the cross-section is consistent with the classical
expression for coherent scattering by a charge distribution. For small 0 in (8.16) the
leading-order magnetic contribution is F~.
It is interesting to examine the linear polarization of a purely magnetic diffraction
signal. In (8.15) take Fc = 0, and for P~,

(da/dY2)P~ = -T2(1 - cos 0) 2 x

x Re{ (2 cos(O/2))2F?(F~ + (1 - c o s O)F{)* +

+(~-~').V*(~+~'). V,+~V~ , (8.17)

where the cross-section on the left-hand side is given by (8.16) with Fc = 0. Observe
that when the moment is perpendicular to the plane, P~ is proportional to the orbital
magnetization, and for the extreme case in which there is no spin magnetization
P~ = - 1 for all qualifying reflections. Conversely, with Fc = Fz = 0 and F~
perpendicular to the plane P~ = 0. The last finding is a special case of the result,

- sin 0 sin 2T
(8. 1 8)
{ 1 - cos 0 cos 2W + 2(tan ¢ cot ½0) 2 } '

in which Fs makes an angle (~ 7r - ¢) with the y-axis, perpendicular to the plane of


scattering, and the projection of Fs in the plane makes an angle p with the scattering
vector k; see fig. 19. Note that P~ = 0 when the projection of F~ in the plane is
parallel or perpendicular to k, as is evident from the general expression (8.17) with
Fz -- 0. These various results for P~ at a purely magnetic reflection serve to illustrate
the value in analysing the linear polarization and the striking differences in photon
scattering between the spin and orbital magnetization components.
Turning now to the cross-section for mixed reflections, the appropriate expression
evaluated to leading-order in r is

1
d¢/dY2 = r 2 ~ (1 + cos 2 0)IGI 2 + T sin 0 (1 + cos 0) Im(Fc*F~u) +
(8.19)
+ T sin 20 (1 - cosO)Im(FcF{) }.
604 s.w. LOVESEY

The interference contribution is seen to involve just the components of Fz and Fs


that are perpendicular to the plane of scattering. At least with regard to experiments
performed at synchrotron sources, where the photons are predominantly polarized in
the plane of the orbit, (8.19) is actually misleading. The complete expression (8.21)
shows that, the orbital term in the cross-section vanishes for P3 = 1 and, to a good
approximation, this is realized at a synchrotron source when the plane of scattering
is perpendicular to the plane containing the electron beam orbit.
The calculation of pI reveals that P~ and P~ are proportional to the magnetic
amplitude, and hence they are small quantities. On the other hand, P~ contains both
pure charge and mixed amplitudes and the pure charge contribution is, of course,
consistent with (2.9).

8.2. Linear polarization


In this section the focus is on features due to linear polarization in the primary photon
beam, and P = (0, 0, P3). The cross-section of interest is,

der/dD = (d•/df2)0 + 2P3 {Re((/3)*(a3) ) - Im((al)*(a2) ) }, (8.20)

where (dcr/df2)0 is the cross-section for unpolarized photons given in (8.16), and
the components of (a) and (/3) are provided in (8.15). Perhaps the first thing to
observe is that, ((/3)* (a3)) involves only those components of the magnetization that
are perpendicular to the plane of scattering (the y-axis in fig. 19) while ((Cel)*(a2))
involves only those components of the magnetization that lie in the plane. Further-
more, ((al)*(a2)) vanishes if the components of Fs and Fz in the plane are either
parallel or perpendicular to the scattering vector k.
The product ((al)*{oe2)) is proportional to r 2, and consequently it does not con-
tribute to the leading-order interference between charge and magnetic amplitudes.
The results (8.19) and (8.20) produce the expression, correct to order r,

da/dY2 = r 2
{(1 )
1 + ~ ( P 3 - 1)sin20 IFc[2 +

+ ~-(1 - P3) sin 20 (1 - cos 0) Im(Fc*FzU) + (8.21)

+ Tsin0((1 + P3) + cos0(1 - P3))Im(F£ F~ ) .

Observe that, the pure charge contribution is consistent with (2.8) for the classical
limit of scattering by a free electron. The orbital contribution to (8.21) vanishes
for or-polarization when the linear polarization of the primary beam is perpendicular
to the plane of scattering, and oppositely aligned to the y-axis, cf. figs 4, 9. In
consequence, linear polarization permits a separate determination of the spin and
orbital contributions to the cross-section.
Some general features of the linear polarization of the secondary beam, described
by the Stokes parameter P~ can be deduced from the structure of the magnetic
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 605

scattering amplitude displayed in (A.2) and fig. 10. For Bragg scattering the relevant
amplitude is,

Evidently, if (o~1) = (0~2) = 0, e.g., the magnetization is perpendicular to the plane


of scattering, ~r-Tr and 7r-~r events are forbidden, and P~ = 4-1 for/:'3 = + 1. On the
other hand, if (%) = (/3) = 0 then cr-cr and 7r-Tr events are forbidden, and P~ = ±1
for/:'3 = q: 1. (N.B. P~ = 0 for both of the two cases.)
These features of the linear polarization of the secondary beam are nicely illustrated
in result for Mn0.75Zn0.25F2 (Hill et al. 1993b) shown in figs 20, 21. There is an exact
cancellation of the charge structure factor of the ruffle crystal structure at the (1 0 0)
Bragg reflection. (This is a fortunate result since the chemical and antiferromagnetic
unit cells are identical which, in general, precludes the possibility of observing pure
magnetic scattering.) For this reflection, (%) = (/3) = 0 if the magnetization is in the
plane of scattering. Turning attention to the data displayed in fig. 20, the polarization
analysis demonstrates that the polarizations of the secondary and primary beams are
opposite when the moment is in the plane of scattering, and the same when it is
perpendicular to the plane of scattering. Also included, for reference, is data taken
at the (200) Bragg reflection which is nominally purely charge. Figure 21 shows
the observed variation of the linear polarization of the secondary beam with the
projection of the moment on the plane of scattering.
The linear polarization of the secondary beam is determined by the expression,

Here, (dc~/dg2)0 is the cross-section (8.16) and the cross-section on the left-hand
side is given by (8.20). In discussing the latter, we have noted the salient features
of ((/3)*(a3)) and ((al)*(a2)). The only point that needs to be added with regard
to (8.22) is that I(%)12 and [(a2)l 2 are both of order r 2, and both vanish if the
magnetization is perpendicular to the plane of scattering. Evaluated for pure charge
scattering, (8.22) agrees with (2.9). For the case of mixed reflections,
606 S.W. LOVESEY

i J i

25
20 (200)
15

'~"

E
C2}
10
5
0
150

__j
• . , . , . , . ,
~lL~u{T{(((i,l(l~t~k&E~k~l;IfY~rCflhtft~ddtfUltltl~__

i
••
.
--

,
.
.

.

.
.
,
.
II

.
.
l
,

Q
125
@ 100 (100) @=90°
tJ
,,, 75
~, 50 I

I !

,,, 25
t-'-
=O 0 - , - ,
i . i . * . , . I

7
6
5
34 (100)~-0°!I~
02 _:_%.,J
. . . . . . . .

-0.02 0 0.02 0.04 -0.02 0 0.02 0.04


t~ [Degrees]
Fig. 20. Polarization analysis of charge (200) and magnetic (100) Bragg reflections in Mn0.YsZno.25F2
(Hill et al. 1993b). The primary polarization is largely perpendicular to the plane of scattering (~r-
polarization, P3 ~ 1) and is unchanged by charge scattering; the observed ratio (I_k/Ill) is due to the
inherent polarization of the synchrotron beam. At the magnetic Bragg reflection, (100), the same ratio
is observed with ~b = 90 ° (magnetic moment perpendicular to the plane of scattering) while for ~b ----0
it is inverted, as expected for magnetic scattering. For charge scattering (top panel) the measured value
of the secondary polarization P~ = 0.9 4-0.1, and for magnetic scattering with the moment in the plane
of scattering (bottom panel) P~ = -0.9 4- 0. I.
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 607

' o {100)'
* (200)
1.5f '
1.0 ~_ :~
f ]i I
0.5
O-

-0.5

-1.0

-1.5
I ! ! I i , I i i I

0 30 60 90
[degrees]
Fig. 21. The secondary linear polarization measured at the magnetic (open circles) and charge (full
circles) Bragg reflections in Mno.75Zno.25F2 studied by Hill et al. (1993b). The inherent polarization
of the primary beam is preserved in charge scattering. On the other hand, linear polarization at a
purely magnetic Bragg reflection (Fc = 0) is rotated as the magnetic moment is rotated into the plane of
scattering. If the primary beam contains only linear polarization, which is almost true in the present case,
and the magnetic moment is due to spin magnetization (the manganese ion spectral state is 6S5/2) the
variation of the secondary linear polarization with respect to the orientation of the moment is provided
by eq. (8.29).

with the result (8.21) for the cross-section on the left-hand side. Note that (8.21)
and (8.23) obey the requirement P~ = -4-1 for P3 = -4-1.
Expressions for the cross-section and secondary polarization are algebraically quite
complicated even for the special case of pure magnetic reflections, i.e. reflections
for which Fc(k) = 0 and the magnetic structure factors are finite. In view of this, we
elect to consider special limits of the expressions beginning with the magnetization
perpendicular to the plane of scattering, defined by q and q ' . The cross-section for
this case is,

do-/da2 = (rr~ sin 0) 2 { 2(1 - P3)(1 - cos 0) x


(8.24)
x [(1 - cos0)lFtVl a + Re(F~F,U*)] + IFsYI2},

and the corresponding result for P~ is,

(da/d£2)P~ = (rr~ sin 0)2{2(P3 - 1)(1 - cos0)[(1 - cos 0)IFtul 2 +


(8.25)
+ Re(F~FYs*)] + P3lF~]2}.
608 S.W. LOVESEY

Hence, when the magnetization is perpendicular to the plane, and the primary radi-
ation is ~r-polarized, the cross-section measures the spin magnetization. A similar
finding holds for charge-magnetic scattering. The pair of results (8.24) and (8.25)
obey P~ = + 1 for P3 = 4-1.
The last case we consider is that of pure magnetic reflections (Fc = 0) with the
magnetization confined to the plane of scattering. In this instance, (a3) = (¢3) = 0.
These results lead directly to the finding P~ = + 1 when P3 = :~ 1. The foregoing
result is manifest in the following pair of relations,

dcr/dg2 = {l(C~l)[2 q- ](c~2)l2 - 2P3 Im((c~l)*(~2) ) }, (8.26)

and,

(d,/df2)P~ = -{P3(1(%)12 ÷ [(%)1 2) - 2Im((%)*(%)) }. (8.27)

On developing the expression for the cross-section one arrives at the result,

do-/dI2 = ('rre sin2 ~ 0 ) 2 { l ( q - * q ' ) ' F s 12+ I(q+ q') • (2Ft + Fs)12 +
(8.28)
+ 2P3 R e [ ( ~ - ~ ' ) . F: (~ + ~ ' ) . (2F, + Fs)] },

from which the corresponding result for P~ quite readily derived. Inspection of these
expressions leads to the view that, the geometry for which (8.26)-(8.28) apply is not
particularly useful for a comparative study of spin and orbital magnetism in solids.
For most transition metal compounds the orbital magnetization is very small, due
to the action of the relatively large crystal field. This is motivation to continue the
study of polarization effects for Fl = 0 begun at the end of the last subsection.
At a purely magnetic reflection (Fc = 0) and P = (0, 0, P3), the secondary linear
polarization is found to be,
_{ t + (s - r)P3 }
P3 = r + 8+~73 ' (8.29)

where r, s, t are functions of the angle 0 through which the photons are deflected,
the angle q5 that defines the spin moment relative to the y-axis, (1 7r - ~b), and the
angle ~ between the direction of k = (q - q') and the projection of the moment on
the plane of scattering,
1 2
r=2(tan~bcot~O) , s = (1 - cos0 cos 2~),

and

t = sin 0 sin 2~.


PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 609

The formula (8.29) agrees with (8.18) for P3 = 0, and the corresponding result
for pure charge scattering is (2.9). Two special cases are ~b = 0 and 1 7r which
correspond, respectively, to the spin moment in the plane of scattering and the spin
moment aligned along the y-axis perpendicular to the plane. Taking ~b= 0 in (8.29),

s + tP3

from which one finds P~ = + 1 for P3 = • 1. For the second special case, 4) = ~1 7r,
P~ =/:'3. So, if the spin moment is perpendicular to the plane of scattering linear
polarization is preserved, as in pure charge scattering at fixed 0. On the other hand,
if the spin moment lies in the plane of scattering linear polarization is rotated by
magnetic scattering.

8.3. Circular polarization


Unlike linear polarization, circular polarization induces interference between the
charge and magnetic scattering amplitudes. In consequence, use of circular polariza-
tion is a valuable method for the measurement of magnetic scattering, providing as it
does both the sign and magnitude of the amplitude, good signal-to-noise ratios by a
difference technique, and potentially better flexibility than use of anomalous charge
scattering. However, it is restricted to the investigation of mixed charge-magnetic
reflections. Examples of diffraction data obtained in this way are provided in fig. 2
and fig. 4.
The existence in the cross-section of charge-magnetic interference induced by
circular polarization is apparent when we look at terms to be added to the cross-
section (8.20). From (2.5) it is found that the additional terms are,

2P2{Re((/3)* (c@) + Im((c~l)*(c@)}. (8.30)

Turning to (8.15), (c~1) and (~2) are purely magnetic while (fl) and (c@ are com-
binations of charge and magnetic terms. One general observation is that (oq) and
(c~2) vanish if Fs and Ft are perpendicular to the plane of scattering. Hence, charge-
magnetic interference induced by circular polarization does not contribute to the
cross-section unless there is a component of the magnetization in the plane, pre-
cisely as for scattering by a polarized electron.
On developing the quantities in (8.30) to leading-order in 7-, the contribution to
the cross-section induced by circular polarization is found to be,

-r~'r(1 - cos O)P2Re{Fc ((1 + cos 0)(~ + ~ ' ) . Ft + (~ cos 0 + ~ ' ) . Fs) }. (8.31)

The cross-section for a primary beam described by a Stokes vector P = (0, P2, P3) is
the sum of (8.21) and (8.31). It is interesting to observe that the spin contribution in
(8.31) is precisely the same as the spin contribution for the classical non-relativistic
cross-section for scattering by a single free electron that is at rest before the collision.
610 S.W. LOVESEY

The contribution (8.31) can be measured by forming the difference of data for
P2 = 1 and P2 = - 1 , since the contribution (8.21) will cancel in forming the
difference. Keeping P2 fixed and switching the polarity of an external magnetic
field, strong enough to saturate the magnetization, achieves the same result provided
the charge-magnetic terms in (8.21) are negligible.
The occurence of different geometric factors with the spin and orbital contribu-
tions to (8.31) offers the possibility to separately measure the two contributions.
For example, if the magnetization is aligned along the scattering vector the orbital
contribution will not participate in the diffraction event. Taking the standard case
P = (0, O, P3), P~ is small being proportional to the magnetic component of the
amplitude.

9. Elastic resonant scattering

The polarization dependence of resonant scattering, satellite selection rules, and


basic features of line shapes are quite elementary consequences of a model based
on electric multipole contributions to the resonant scattering amplitude (Carra et al.
1989, Hannon et al. 1988, 1989, Luo et al. 1993). The nontrivial calculation required
to confront experimental and theoretical findings is the calculation of radial integrals
which arise in matrix elements of multipole operators (Cowan 1981). Where this
has been attempted, ab initio calculations based on an atomic model have provided a
good account of experimental data (McWhan et al. 1990, Tang et al. 1992b). Reviews
of experimental work include (Gibbs 1992, Stirling and Lander 1993, Vettier 1994).
The description of elastic resonant scattering has much in common with dichroism,
considered in section 7. Referring to eq. (7.7), the resonant component of the elastic
scattering amplitude is,

re (#lg'J(-q')l~)(~Tl~'J(q)ll~)
f~(E) - (9.1)
re (E - A + iF/2)

Here, the intermediate state, labelled r/, is taken to be a single electronic orbital;
formulae for a more general case are provided later in this section. Neglecting the
contribution to J from the electron spin, and treating the contribution from the linear
momentum in terms of the dipole approximation (7.4), f~(E) does not depend on
the wave vector of either the primary (q) or secondary (q'). Near resonance, the
amplitude (9.1) reduces to,

(s'. Q*)(e • Q)
f,.(E) = -(eq) 2 (9.2)
(E - za + i t / 2 ) '

where,

Q = (7?lRl#),
PHOTON BEAM STUDIES OF MAGNETICMATERIALS 611

is a dipole matrix element; the present value of Q differs from (7.4) by a simple
factor proportional to A ,-~ E, near resonance. The observed cross-section is propor-
tional to the absolute square of fr averaged over thermal, concentration and defect
fluctuations, i.e. the square of the scattering length of the perfect crystal. Thermal
fluctuations will include both those arising from the magnetic and lattice vibrational
properties of the sample.
Following the development pursued in section 7, we adopt a simple representation
of the orbitals which reduces the product of matrix elements in (9.2) to an appealing
form. Let the initial and intermediate orbitals be represented by states described by
a single magnetic quantum number. In this instance, Q*Q~, (u = 0, +1) vanishes
unless the spherical component labels satisfy the condition u = u ~, and so,

(e'" Q*)(e" Q) = Z e ~ , ,. ~,IQ~,I2. = ~1 ((~,, ~)(IQ÷112 + IQ-II 2) +


1/

+ i r a . (e' x e)(lQ_ll 2 - I q + l l 2) + (9.3)

+ (~'. m)(~. m)(210012 -IQ+~I 2 - IO_a[ 2) }.

The three terms in the second equality are arranged in ascending powers of the (unit)
vector m that defines the magnetic quantization axis. It comes as no surprise that,
the combinations of dipole matrix elements in (9.3) arise also in the description of
E1 circular and linear dichroism.
The atomic model described in section 7 can be used to estimate the temperature
dependence of the products of radial integrals in (9.3). This reveals that, the first
and third terms involve thermal fluctuations of the mean square magnetization, while
the second term is proportional to the magnetization.
With regard to the dependence of the scattering amplitude on polarization states,
the representations of e ~• e, e' x ~ and (e'. m)(e. m) in fig. 11 could be useful. The
notation of cr and 7r polarization states appears in the literature; a photon polarized
perpendicular to the plane of scattering is said to exhibit ~r polarization, while a
photon polarized in the plane has 7r polarization.
From fig. 11 we see that, for the scattering amplitude formed with (9.3) there are
three forbidden events, namely, c~-Tr events in the term which is independent of m,
and cr-cr events in the term proportional to m. When the magnetic moment lies in
the plane of scattering, rn± = 0, the only event in the term of order rn 2 is 7r-Tr
scattering, while for rn± = 1 the only event allowed is a-or scattering.
To illustrate the contribution made to Bragg intensities by resonant events, consider
the term of order m in the resonant amplitude. Inspection of (9.3) shows that this
term is of the form X" re(R) where the vector R defines the position of the magnetic
atom. For coherent scattering from an array of atoms the total amplitude is,

exp(ik • R) X" ra(R),


R
612 S.W. LOVESEY

where, as usual, k = q - q'. A simple magnetic configuration that gives rise to


satellite reflections is a spiral structure in which the moments are perpendicular to an
axis, and rotation about the axis varies sinusoidally with position along the axis. Such
a spiral configuration of magnetic moments is realized in holmium in the temperature
interval 20-132 K. Let,

re(R) = (cos(w • n), sin(w • R), 0),

where the moduation wave vector, w, might change with temperature. The scattering
amplitude is then of the form,

1 ~--~exp {iR • (k +
R

from which we conclude that Bragg scattering occurs when,

k+w=r,
where r is a reciprocal vector for the magnetic lattice. Consideration of all three
terms in the resonant amplitude shows that, dipole (El) resonant scattering can
contribute to a main Bragg reflection and incommensurate satellites to it defined by
w and 2w, i.e. satellites of order 1 and 2.
The E2 amplitude reported by Hannon et al. (1988, 1989) contains terms in m
from zero up to fourth order, and there are thirteen distinct contributions. Hence, E2
resonant scattering can contribute to a main Bragg reflection and satellites to it of
order 1 through 4.
The model for E1 events used up to now in this section is based on single atomic
orbitals for the initial and intermediate states. We conclude by recording the ex-
pression for the amplitude in which more than one initial and intermediate state is
required. But, we retain the simplifying assumption that each orbital contains one M
component, e.g., consideration of crystal-field perturbations are not included. The
notation for the more general model follows that introduced in section 7. Let the
dimensionless quantity,

I@ EjJL(qRy)yL(Rj) #> :
, (9.4)
t'7/ E,, + E - E~ + i / 7 2

in terms of which the E1 amplitude is,

f~(E)- 3 {(¢,.e)(Fll+Fl_l)+im.(s,x¢)(Fll_F~l) +
4rrq (9.5)
+ ( e . m ) ( , . m)(2Uo' - U+ l -
PHOTON BEAM STUDIESOF MAGNETICMATERIALS 613

The physical significance of the three terms in this expressions, and the connection
of fr(E) to the interpretation of the magneto-optical Kerr effect, has recently been
reviewed in an application to data for Co (Kao et al. 1993). The Debye-Waller
factor is not shown explicitly in (9.5). The corresponding E2 amplitude contains
linear combinations of F ~ . For a given magnetic reflection, the amplitude is the
coherent sum of contributions from El, E2 . . . . events. Line shapes, proportional
to the absolute square of the total amplitude, as a function of photon energy can
display marked asymmetries, which result from the interference of the resonant and
non-resonant contributions.

10. Spectroscopy

Inelastic photon scattering, here referred to as spectroscopy, is described by the


partial differential cross-section (2.10) and the concomitant result, based on (A.5),
for the polarization of the secondary beam. Observed events include the excitation of
collective charge oscillations, also known as plasmons, and interband (particle-hole)
transitions (Calaway 1991). These events occur in metals for energy changes, hw, of
the order of 10 eV. At larger values of co the cross section approaches the Compton
limit, which gives access to the electron momentum distribution. For a magnetic
material there is also a contribution from the density of unpaired electrons (Platzman
and Tzoar 1970). As might be expected circular polarization can be of assistance in
efforts to discriminate magnetic from charge induced events in the scattered beam
(Gibbs 1992, Sakai 1992).
Here, we first recall exact results for scattering by free charges, after which there
is a discussion of cross-sections for bound electrons and the corresponding Compton
limit.

I0.1. Scattering by free charges


We begin with non-relativistic scattering by free charges. The corresponding cross-
section is (2.7) extended to dynamic events, as demonstrated in (2.10). The rela-
tivistic result, usually referred to as the Klein-Nishina formula (Berestetskii et al.
1982), is discussed at the end of the subsection.
In free space a free electron cannot emit or absorb a photon without violating
energy or momentum conservation. Therefore, there are no first-order processes
involving the p .A terms in the interaction provided by (p - -~A) 2. However, there are
first-order processes involving the A 2 term, which is quadratic in photon operators.
The scattering of photons, which is of interest here, is a process in which one photon
is destroyed and another created. The quantum theory of scattering applied to the
A 2 term shows that, within the first Born approximation (equivalent to the use of
Fermi's Golden Rule for transition rates),

d2cr/df2 dE' = Nr2e(q'/q)(e • e')2S(k, w), (10.1)


614 S.W. LOVESEY

with the Van Hove response function for non-relativistic charge scattering (Calaway
1991, Lovesey 1986),

S(k, w) = (1/2~rhN)
F dt exp(-iwt)(n+(k) n(k, t)).
oo
(lO.2)

Here, n(k) is the spatial Fourier transform of the microscopic particle density of N
electrons, and an explicit expression is found in (2.6).
The expressions (10.1) and (10.2) reduce (2.7) in the limit of a large primary
photon energy, since then it is appropriate to replace the correlation function in
(10.2) by its static (t = 0) value. The latter step is based on the observation that,
for E ~ ~ the duration of the scattering event is vanishingly small; this line of
reasoning is not restricted to free charges, of course. The corresponding cross-section,
(2.7) in the case of charge scattering, is an estimate of the total scattering. Indeed,
a complementary line of argument is to say that, when E --+ e¢ all allowed events
contribute to scattering and, so, the observed response approaches (10.1) integrated
over all energy transfers, hw.
Perhaps the most direct way to calculate S(k, w) is to employ a second quantized
representation for n(k) (Lovesey 1986). If the carriers obey Fermi statistics, correct
for spin- 1/2 particles,

(n +(k)n(k, t)} = 2 ~ re(1- fk+p)exp { it___~h(k 2 + 2k .p)}, (10.3)


p 2ra*

where fp is the Fermi distribution function (/3 = 1/kBT),

fp = (exp {fl(Ep - #)} + 1)-', (10.4)

and the chemical potential, #, is determined, in the usual way from the number of
carriers. Note that, the one-particle energy Ep = (hp)Z/2ra * is independent of the
spin state. From (10.2) and (10.3),

2 E S { h w + Ep - Ek+p}fp(1 - fk+p)
P
(lO.5)
={l+n(w)} 2 a{r + uj, - k+p }(/j, -
P

where, in the second form of the result, the new factor is defined by n(w) =
{exp(hwfl) - 1}- 1.
The Boltzmann limit is recovered from (10.5) when the chemical potential satisfies
exp(#/3) << 1, and in this case it has the value,
n0
exp(#p) = -~- ( 2 r r h 2 ~ / m * ) 3/2 << 1, (10.6)
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 615

which is a condition on particle density, n o, temperature and mass.


Another interesting limiting case of (10.5) is when the scattering vector k is large.
For sufficiently large k, fp >> fk+p, whence S(k,w) in (10.5) approaches the value,

S(k, w) ~- { 1 + n(w)} 2 Zp + Ep -
(10.7)

which describes the Compton limit of scattering. The structure of the right-hand side
is a delta-function, expressing conservation of energy for free particles, weighted by
fp which is the momentum distribution function for free particles. The Compton
limit of the response function for bound particles has precisely the same structure,
but the momentum distribution is the one appropriate to the energy surface of the
binding potential.
As the second topic in this subsection, let us turn to the Klein-Nishina formula for
relativistic scattering a photon by a free electron that is at rest before the collision.
The latter is polarized, and the average value of the spin is (s). The primary radi-
ation is assumed to contain linear and circular polarization, represented by Stokes
parameters P2 and P3. The values P3 = + 1 and P3 = - 1 correspond to complete
linear polarization and, respectively, labelled a- and rr-polarization states by many
authors. The parameter P2 represents the degree of circular polarization; with our
convention, the probability that the primary photon has right-hand or left-hand circu-
lar polarization is respectively (1 + P2)/2 and (1 - P2)/2. In scattering, the primary
photon is deflected through an angle 0, and the relation between relativistic photon
energy change and the scattering angle is,

1 1
q' q
- (h/me)(1 - cos 0). (10.8)

Having dispatched these necessary definitions, the exact cross-section is,

do'/d~ = ~1 r2~ q + ~ + (P3 - 1) sin 2 0 -


~7
(10.9)
- 2rP2(1 - cos0)(s) • (~ cos0 + q'/q) }.

Here, the dimensionless quantity "r = (/'N/me). The contribution from the electron
spin vanishes if (s) is perpendicular to the scattering plane, defined by q and q '. When
this condition holds, the cross-section is that for scattering by unpolarized electrons;
photons polarized perpendicular to the scattering plane (P3 = 1) have a larger cross-
section than photons in the plane (P3 = - 1 ) . To engage in scattering the spin of the
electron must have a projection on the scattering plane, as already mentioned, and
the primary photon must be circularly polarized (P2 ¢ 0). The spin-dependent term
can be thought of as an interference between charge and magnetic (spin) scattering
616 S.W. LOVESEY

induced by circular polarization. There is no orbital angular momentum for a free


electron, of course.
The classical, non-relativistic result is obtained for hq << rac, i.e. r << 1. In this
instance q' ~ q, and the cross-section (10.9) reduces to,

do"
d[? = r e
{ '
2 1-4- (t:)3-1)sin20- rP2( 1-c°sO)(s)" (q c°sO +q
2
(10.10)

For/:)2 =/:)3 = 0, (10.10) is the same as (2.7) and (2.8) evaluated for one electron
(N = 1), i.e. Thomson's formula is recovered, as it should be.

10.2. Bound electrons


The~expression developed in this section includes circular polarization in the primary
beam, P = (0, P2, 0), and magnetic contributions to leading-order in r = E/rnc 2.
The starting point is (2.10) taken together with (2.5). Examination of (A.8) for the
components of a and/3 shows that % and ~2 are proprotional to r so the terms
ce+ce,
1 l and a+% are not included in our expression. Probably the most compact
formalism is achieved with correlation functions introduced in section 2. Recall that
the frequency change in scattering is denoted by co = c(q - q') and the concomitant
wave vector change k = q - q'. For ease of notation, the k-dependence of quantities
will not always be displayed, e.g., the charge variable n(k), defined in (2.6), is written
as n in subsequent working. The partial differential cross-section is,

d2o-
d[2dE'
_ q'
q 27rh
1
i7 ~
dt exp(-icot)
{<~+a3(t) +/3+/3(0) + (10.11)

+ P2(/3 +O~2(t) 4- Ol;/3(t)÷ i [O~3+OZl(t) -- 0~+0~3(~)]) },

where the term proportional to P2 is magnetic in character in as much that it vanishes


for r - + 0.
On using the expressions listed in (A.8) and developing results for correlation
functions to leading-order in -r, the dynamic quantity in (10.11) for a ~ 1 becomes

1 (1 + cos 2 o)(,~+n(t)) +

ir
+-~ (~ × ~'). { cosO (n+z(t)- z+n(t)) -

- a(1 + cosO)(n+S(t) - S+n(t))} .4-


(lO.12)

•4"~ /92 (q(1-- a + (cos 0 -- a) cos 0) +

+ ~ ' ( 3 a c o s 0 - 2 cos 0 - a ) ) . (n+S(t) + S+n(O) +


+ (~ + ~' cos0). @+z(t)+ Z+n(t)>}.
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 617

Here, a = (1 + qt/q)/2 and the spin and momentum operators S and Z are defined in
(A.6) and (A.7). The first term in (10.12) is the pure charge contribution, evaluated
in section 10.1 for free particles.
In the subsequent work, the various correlation functions in (10.12) are evaluated
in the Compton limit, discussed in section 10.1 for the particular case of charge
scattering.

10.3. Compton scattering


The Compton limit is achieved with a high primary energy and large scattering
vectors. Scattering in this limit is dominated by incoherent events that involve
individual particles. Compton profiles of unpaired electrons in Ni and Fe are shown
in fig. 5 and fig. 16.
Experiments (Timms et al. 1993) have shown that Compton scattering from mag-
netic materials measure the spin density and supporting arguments are discussed
by Lovesey (1993). This subsection addresses the structure of the spin correlation
functions in (10.12) evaluated in the Compton limit.
We argue that the Compton limit of the mixed correlation function (n+S} is,

(n + (k) S(k, t)) ..~ h3 / dQ ~((~(BQ - pj)sj) x


J (10.13)
x exp {it(Ek + h2k • Q/m)/h},

where Ek = (hk)2/2m is the recoil energy. On taking the time Fourier transform of
this result, to obtain the appropriate response function, the exponential term produces
a delta function which expresses conservation of energy, c.f. (10.7). An argument
which produces the estimate (10.13) rests on the use of R(t) = (R(0)+ tp/m) where
p is the momentum conjugate to R = R(0); this expression for R(t) is manifestly a
short-time expansion. The first few steps leading to (10.13) are,

(n+(k) S(k, t)) = ~ ( exp(-ik • Ry) exp (ik. Rj(t))sj(t))


jj'

~ ( exp(-i k. R3) exp {i k. (R3 + tpj/~)}


J

= ~ ( exp { ( i t k . p j / m ) + i tE~/h}sj(t)),
J

where the last line is reached by using a standard identity for the product of two
exponential operators. Note that, just the self part of the correlation function is
retained, i.e. the incoherent approximation is made.
Before collecting the results for the partial differential cross-section in the Compton
limit, let us remark that in the extreme limit E -+ oe scattering is elastic, and the
618 s.w. LOVESEY

scattered intensity is proportional to the static limit of the correlation function. The
static limit of (n+S) in the incoherent approximation is simply the average value of
the spin density (S).
In subsequent working, the static correlation function in (10.13) is assumed to be
purely real, in which case the polarization-independent spin contribution to (10.12)
vanishes. Starting from (10.12), and gathering results from the foregoing discussions,
the Compton limit of the cross-section is,

d2~r/dY2 dE' -- h3(q'/q)r 2 ×


1
× { (l+cos 2 O)N
. / dQ p(Q)~(Energy) + (10.14)

+ ~-P2 (~(1 - a + (cos 0 - a) cos 0) + ~'(3a cos 0 - 2 cos 0 - a)) ×

x J dQ E (~(hQ-pj)sj)~(Energy)),
J

where the momentum density

p(Q) = I ( ~ ~(hQ-pj) I, (10.15)

and

3(Energy) = ~(hw- Ek -- h2k • Q/m).


If the inelasticity is very small, a = (1 + q'/q)/2 ,-~ 1 and the orientation factor in
the polarization dependent term is identical with that found for the spin contribution
to the diffraction cross-section given in (8.31). The orientation factor in (10.14) is
not the same as the one derived from the Klein-Nishina formula. This result is
attributed to the different models; the Klein-Nishina result is the exact cross-section
for a free electron that is at rest before the collision, whereas (10.14) is derived from
an approximate amplitude, appropriate for bound electrons, derived by perturbation
theory.

11. Concluding remarks

Frenetic activity over the past few years by researchers at synchrotron light sources
has established the value of photon beam techniques to investigate magnetic proper-
ties of materials. At the moment, the most productive area of work is with methods
using linear and circular dichroism. Like elastic resonant diffraction, the capabilities
of these methods are not shared by techniques based on neutron beams. Nonresonant
PHOTON BEAM STUDIESOF MAGNETICMATERIALS 619

scattering, including Bragg diffraction and Compton scattering, suffer in full the dis-
advantages that stem from the intrinsic weakness of the magnetic photon scattering
interaction.
Looking to the future, one can readily anticipate improvements in methods and
techniques, and the interpretation of data. In light of the rapid pace of the expansion
in experimental activities over the past decade, current instrumentation might not yet
be optimal. Increased sensitivity in scattering experiments will open the way to more
work on ferromagnetic materials, for example. Different approaches could prove
useful. For example, the use in diffraction of high energy photon beams (,-~ 80 keV)
improves penetration, which increases the scattered intensity, and decreases surface
sensitivity thereby making diffraction closer to a bulk probe as in neutron beam
diffraction (Brtickel et al. 1993). Furthermore, techniques will be refined to the
stage where some measurements are routine. Understandably, inelastic magnetic
photon scattering has yet to be exploited.
While sum rules are proving to be valuable in the interpretation of data on magnetic
dichroic effects there is more scope to relate observed signals to physical variables
built from atomic quantities. Probably, it will not be too long before the domain of
validity of currently known sum rules is firmly established through the confrontation
of data with model calculations based on one-electron pictures of atomic and band
states. Turning to another issue, the quantitative value of elastic resonant scattering
depends on the quality of absorption corrections. Several techniques are apparently
sensitive to near-surface structure and defects. An understanding of these effects is
important, as currently witnessed in the efforts to fully account for the two, or more,
length scales appearing in the analysis of critical magnetic scattering.

Acknowledgements

The work of preparing this chapter contributes to an EC funded project SCI 0467-
M(SMA).
Discussions and correspondence with the following researchers is gratefully ac-
knowledged: U. Balucani, J. Bohr, E Carra, C.T. Chen, S.E Collins, D. Gibbs, G. van
der Laan, S. Manninen, D.B. McWhan, G.D. Priftis, N. Sakai, G.A. Sawatzky, and
G.T. Trammell. Dr. G. van der Laan commented on the first draft of the text.
Preparation of the text for publication was made by Shirley Fortt (ISIS Facility).
Figures 1-9, 13-16, 18, 20, and 21 are reproduced with the consent of the authors,
and the American Physical Society, Institute of Physics, Springer-Verlag, Francis-
Taylor, and Gordon-Breach.

Appendix. Polarization effects and magnetic scattering amplitude

Polarization states
The scattering amplitude operator, and observable quantities derived from it, depends
on the polarization of the primary radiation, described by a real unit vector e. In this
620 S.W. LOVESEY

appendix we describe a formalism for handling the consequences of this dependence,


and a more general example in which the beam of secondary radiation has also a
definite polarization. Additional material is provided in Lovesey (1993), Newton
(1982), Berestetskii et al. (1982).
Let us begin by recapping the properties of a classical light wave propagating in
the z-direction, say. The electric E and magnetic H field vectors are perpendicular
to the direction of propagation of the light, whence E = (E~, Ey, 0) and for a plane-
wave H = ( - E u, E~, 0) or H = (~x E) where ~ is a unit vector along the z-direction.
It is often convenient to use a complex notation for the electric field, remembering
that in the end the real components lead to observable quantities.
The polarization state of the light wave is directly related to the E vector. Examples
of particular interest include:
(a) E v = 0, plane polarized in the x-direction,
(b) E~ = 0, plane polarized in the y-direction,
(c) E~ = Ey, polarized at 7r/4,
(d) Ey = e i ~ / 2 E x , the y-component lags the x-component by 90 °, and the wave is
right circularly polarized,
(e) Ey = -iE~, the wave is left circularly polarized.
Let us choose to employ real polarization vectors; e for the primary and e' for
the secondary radiation. By definition, 6 . q = e , . q t _- 0. The properties of partially
polarized radiation are best described in terms of a density matrix.
Although several polarization effects are adequately treated by elementary meth-
ods, slightly more formalism is required to provide a complete description of po-
larization contributions to the cross-section and secondary beam. Fortunately, it is
possible to derive master formulae for the cross-section and polarization of the sec-
ondary beam from which special cases are readily obtained. Hence, at the end of the
day there is no need to grapple with the formalism. The master fomulae to which
we refer are (2.5) and (A.5). First some material to define concepts and notation,
beginning with results that provide a physical interpretation of the density matrix.
Any polarization ~ can be represented as a superposition of two mutually orthog-
onal polarizations, chosen in some specified manner. These polarizations can be
taken to be two mutually perpendicular linear polarizations, or two circular polariza-
tions having opposite directions of rotation. For the moment, we choose to denote
complex vectors of right-hand and left-hand circular polarization by ~(+) and ¢(-),
respectively. Seen head-on right-hand circular polarization is counter-clockwise.
In coordinates (, r/, i, shown in fig. 12, with the if-axis in the direction of the
photon q, we take,
i(i ) ,

and ¢(-) = (¢(+))*. The spin operator for the photon is usually taken to be,
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 621

and,

S¢ (+) = e (+), right-hand,

S¢ (-) = - e ( - ) , left-hand.

The component of angular momentum along the direction of the momentum vector
q (the C-axis), called the helicity of the photon, is a conserved quantity. The vectors
e (+) and ~(-) correspond to the helicity values +1 and - 1 , respectively. Thus, the
component of the photon angular momentum along the direction of its motion can
have only the two values 4-1; the value zero is not possible.
We employ a density matrix to describe states of partial polarization; applications
of this method to photon scattering are reviewed in Newton (1982), McMaster (1961 ),
Berestetskii et al. (1982), Tolhoek (1956). The polarization density matrix is a tensor
/ , ~ of rank two, in a plane perpendicular to the wave vector q which is the ~--~-
plane with our notation and conventions. The tensor is required to be Hermitian
# ~ = (/,~)* and normalized such that the trace is unity.
The probability that the photon has any given polarization e is determined by the
quantity,

Writing the density matrix in the form,

(~11, /3'12/
z ~21, ~22 '

with/z21 = (1Z12)*, the components//,11 and//,22 are the probabilities of linear polar-
izations along the ~- and rj-axes. Choosing s to be the vector e (+) we find that the
probability for right-hand circular polarization,

(~(q-))*~(+) : ! {1 "+ i(~12 --/Z21)}.


2

Next let us consider the two extreme cases of unpolarized and completely polarized
photons. For unpolarized photons all directions of polarization are equally probable,
and,

#,~ = 6,~/2.

A photon beam with complete polarization is described by the tensor,

]l/, = ~ * ,

for then the probability for the state e is unity, as required.


622 s.w. LOVESEY

In general, it is convenient to describe partial polarization by means of three real


Stokes parameters P1, P2 and P3, which form a vector P. The latter is a purely
formal step and is done only for the sake of notational convenience. The density
matrix t* can be written,

1 ( 1+P3 Pl-iP2) a (A.1)


/z= k, P1 +iP2 1 -P3 =
In the second equality I is the unit matrix, and ~r are Pauli matrices.
Referring back to our comments on the physical significance of the elements of/z
we see immediately that the parameter/:'3 defines the linear polarization along the
~- and ~/-axis. For the scattering geometry depicted in fig. 12, P3 = 1 corresponds to
complete polarization perpendicular to the scattering plane and P3 = - 1 is complete
polarization in the scattering plane. Very often, P3 = 1 and P3 = - 1 are referred to
as states of ~r- and 7r-polarization, respectively.
The parameter P2 represents the degree of circular polarization; the probability of
fight-hand polarization is (1 + P2)/2, and the probability for left-hand polarization
is (1 - P2)/2. Since these two polarizations correspond to helcities :~1, it is clear
that P2 is the mean value of the helicity of the photon. With our conventions,
which depend on our use of a right-handed Cartesian co-ordinate system, the state
of complete fight-hand circular polarization (seen head-on this is counter-clockwise)
is described by the polarization vector ¢(+), and the density matrix for the state with
P2 = 1 (positive helicity) is t* = ~(+)(¢(+))*. The probability for finding the state
¢(+), by definition, is,

(~(+))./z¢(+) = 1 (1 + P2).
2

Our conventions for photon polarization states and those in Berestetskii et al. (1982)
concur. A complete discussion of helicity states can be found in section 16 of
Berestetskii et al. (1982). The reader should be constantly wary about conventions
used to describe the polarization and helicity states of a photon beam because there
appears to be a plethora of conventions and plenty of scope for inconsistent conven-
tions. Other frequently used consistent conventions are described in (Shore 1990)
and references therein.
Choosing ¢ = (1, ~ I ) / V ~ reveals that P1 describes the linear polarization along
directions at angles ±7r/4 to the ~-axis. Complete linear polarization of the photon
is described by P2 = 0, p2 + p2 = 1. In the unpolarized state P1 = P2 = P3 = 0,
while for a completely polarized photon p2 + p2 + p2 = 1.

Scattering amplitude
The scattering amplitude operator G is written as a 2 × 2 matrix by constructing
four elements for the two values of ~ and #. Our scheme for this is defined in
fig. 10, in which the diagonal elements Gll and G22 are amplitudes for both ~ and #
perpendicular or-or and parallel 7r-Tr to the scattering plane, respectively. The element
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 623

G21 is the amplitude for the event in which the initial polarization e is perpendicular
to the scattering plane and the final polarization e' is parallel to the plane a-Tr, while
G12 describes the event in which the polarizations are in the reverse order 7r-or.
It is convenient to write the matrix G in the form,

/3 -t- 0:3 0:1 - i0:2


G=/31+c~.cr= 0:1+i0:2 /3 0:3 ,/' (A.2)

where/3 is a scalar quantity, cz = (0:1, 0:2, 0:3), and I is the unit matrix. The conve-
nience arises when we calculate the cross-section,

(da/d$2) = Tr{ttG+G}, (A.3)

where the trace operation is taken with respect to the photon polarization states.
A similar situation arises in the calculation of Stokes parameters for the secondary
photon beam pI obtained from the relation,

(dcr/df2)P' = Tr{ttG+crG}. (A.4)

The trace operations in (A.3) and (A.4) entail quite a bit of algebra. The trace
operation yields (2.5) for the cross-section, while for P ' we obtain,

(da/df2)P' = [/3+a + oz+/3 - i(c~+ x a) +/3+/3P + a + ( a . P) +


(A.5)
+ (oz+- P)o~ - P ( a + . a) + i/3+(c~ x P) + i(P x a+)/3].

In deriving (2.5) and (A.5) we have assumed that/3 and c~ are quantum-mechanical
operators that do not commute. The expressions for (da/dO) and P ' are in the form
of products of operators, and to obtain results for the interpretation of experiments
(Bragg diffraction, Compton scattering, etc.) appropriate matrix elements have to be
formed, which are topics taken up in sections 8-10. For the moment, we empha-
size that (2.5) and (A.5) provide general statements with regard to the polarization
dependence of the cross-section, and the polarization state of the secondary beam.
The expressions for 0:1, 0:2' 0:3 and/3 appropriate for diffraction and spectroscopy
can be expressed in terms of three atomic quantities that have immediate physical
interpretations. One quantity is the spatial Fourier transform of the particle density,
n(k), defined in (2.6). The remaining two quantities relate to magnetic properties,
and they are the spin density,

S(k) = E s j exp(ik. Rj), (A.6)


J

where sj is the spin operator for the jth electron, and,

Z(k) = (i/hq) E exp(i k . Rj){ (~ - ~') x pj }, (A.7)


J
624 S.W. LOVESEY

A!
where p is the momentum operator conjugate to R, and ~ and q are unit vectors.
Note that, S(k) and Z(k) are dimensionless and Z(k) vanishes in the forward direction,
and that the sums in (A.6) and (A.7) are over all electrons in the target assembly.
Matrix elements of S(k) and Z(k), needed in the cross-section for a magnetic material
will, however, involve only the fraction of electrons that are unpaired, e.g., for the
localized atomic ion model for a rare earth magnet, core electrons will not contribute
to the matrix elements in the first approximation.
The expressions for c~ and/3 are found from the results (6.5) and (6.7) together
with figs 11 a) and 11 b); the following expressions are in units of r~, and the small
parameter 7- = E/mc 2,

i
~-S(k) • { ( 2 a - b - cos 0)~ + (1 + b cos 0 - 2 a ) ~ ' } +

i
+ ~ 7-(~-~'). z(k),

0 : 2 -_- _ 1 7-S(k). {(cosO - b)~+ (bcosO - 1)~'} -


2
(A.8)
7-(~ + ~'). Z(k),
2

0:3 ~ m ~ i
(1 - cos 0)~(k) - ~ 7-(~ × ~ ' ) . {(b + 1 - 2a)S(k) - Z(k)},

= _ (1 + cos o)~(k) - ~ ~-(~ x ~ ' ) . {Z(k) - (b + 1)S(k)},

where b = q~/q and a = (1 + b)/2, and a = b = 1 for elastic events. Notice that a 1
and 0:2 are purely magnetic in character (0:1 = 0:2 = 0 for T = 0), and the expressions
for c~ and/3 are correct to first-order in 7- (Grotch et al. 1983).
We have remarked that for elastic scattering Z(k) is proportional to the orbital
magnetization in the sample. Looking at a and/3 one is struck by the lack of any
symmetry in the spin and orbital contributions. The origin of this state of affairs
can be traced back to the perturbative calculation of G, (6.4). It is found that, the
orbital contribution arises from second-order processes, the same ones that generate
Raman scattering. On the other hand, the spin contribution is the sum of terms that
arise in first-and second-order processes. A consequence of the lack of symmetry in
the spin and orbital contributions is that scattering geometries exist were one or the
other contribution is absent, a feat not possible in neutron scattering (Vettier 1993).
P H O T O N B E A M STUDIES OF M A G N E T I C MATERIALS 625

List of important symbols

A photon vector potential


ao
Bohr radius
c velocity of light
E electric field
E, E' primary and secondary photon energies
e unit of charge for an electron
Fc charge unit cell structure factor
Fs, Fl unit cell structure factors for spin and orbital magnetism
f(E) scattering amplitude in the forward direction (2.1), = f ' + if"
G scattering amplitude operator (2.4)
g gyromagnetic factor
H magnetic field
Hamiltonian operator for electrons in a solid
kB Boltzmann's constant
J(q) current density
k=q-q' scattering vector
m mass of an electron
fr~* effective mass of a charge carrier
N(k) average particle density
~(k) Fourier transform of the single particle density function
P linear momentum operator conjugate to R
Pi, Pi' Stokes parameters, i = 1,2, 3, for primary and secondary beams (2.8)
q wave vector of photon; q, q' wave vectors for primary
and secondary photons
R position vector
re classical radius of an electron
S(k) spin density operator
s(k,w) response function
T(k) orbital density operator
°} components of the scattering amplitude operator, eq. (2.4)
7 attenuation coefficient (2.2)
polarization vector of a photon, ¢ • q = 0, ¢* = e
X wavelength
density matrix for polarization states (2.5)
#e Bohr magneton
P(q) momentum density distribution
o cross-section (2.3)
ratio of primary photon energy to electron
rest mass = (N//mc) = (E/mc 2)
"r" reciprocal lattice vector
03 frequency change in scattering event
626 S.W. LOVESEY

Numerical values of units

Quantity Symbol Value


Bohr radius % : h2 / m e 2 0.529 ,~
Planck's constant/27r h 0.658 meVpsec
Rydberg Roo = mc2a2/2 = e2/2% 13.606 eV = 911 ,~
Fine-structure constant a = e2/hc 1/137.04
Rest mass energy of the electron me 2 0.511 MeV
Bohr magneton #B = e h / 2 m c 0.579x 10-1 meV T -1
Classical radius of the electron re = e2/mc 2 0.282x 10-12 cm
hZ/2m 3.810 eVA 2

References

Alders, D., J. Vogel, C. Levelut, S.D. Peacov, Burkel, E., 1991, Inelastic Scattering of X-rays
M. Sacchi, G. van der Laan, B.T. Thole and with Very High Energy Resolution (Springer,
G.A. Sawatzky, 1995, Europhys. Lett. Berlin).
Altarelli, M., 1993, Phys. Rev. B 47, 597. Byrne, J., 1994, Neutrons, Nuclei and Matter
Anastassopoulos, D.L., G.D. Priftis, N.I. (Institute of Physics Publishing, Bristol).
Papanicolaou, N.C. Bacalis and D.A. Papa- Callaway, J., 1991, Quantum Theory of the Solid
constantopoulos, 1991, J. Phys.: Condens. State, 2nd edn (Academic Press, Boston).
Matter 3, 1099. Carra, E, in: 1993, Proc. 7th Int. Conf. X-ray
B6ske, T., W. Clemens, C. Carbone and W. Absorption Fine Structure, Kobe, J. Appl.
Eberhardt, 1994, Phys. Rev. B 49, 4003. Phys. 32, 279.
Balcar, E. and S.W. Lovesey, 1989, Theory Carra, P., M. Altarelli and E de Bergevin, 1989,
of Magnetic Neutron and Photon Scattering Phys. Rev. B 40, 7324.
(O.U.P., Oxford). Carra, P. and M. Altarelli, 1990, Phys. Rev.
Berestetskii, V.B., E.M. Lifshitz and L.P. Lett. 64, 1286.
Pitaevskii, 1982, Quantum Electrodynamics, Carra, P., B.N. Harmon, B.T. Thole, M. Altarelli
Course of Theoretical Physics, Vol. 4 and G.A. Sawatzky, 1991, Phys. Rev. Lett.
(Pergamon Press, Oxford). 66, 2495.
Bhatt, G., H. Grotch, E. Kazes and D.A. Owen, Carra, P., H. KOnig, B.T. Thole and M. Altarelli,
1983, Plays. Rev. A 28, 2195. 1993a, Physica B 192, 182.
Blume, M. and D. Gibbs, 1988, Phys. Rev. B Carra, P., B.T. Thole, M. Altarelli and X. Wang,
37, 1779. 1993b, Phys. Rev. Lett. 70, 694.
Bohr, J., D. Gibbs, J.D. Axe, D.E. Moncton, K.L. Chen, C.T., 1993, in: Proc. 7th Int. Conf.
D'Amico, C.F. Majkrzak, J. Kwo, M. Hong, X-ray Absorption Fine Structure, Kobe, J.
C.L. Chien and J. Jensen, 1989, Physica B Appl. Phys. 32, 155.
159, 93. Chen, C.T., E Sette, Y. Ma and S. Modesti,
Bohr, J., D. Gibbs and K. Huang, 1990, Phys. 1990, Phys. Rev. B 42, 7762.
Rev. B 42, 4322. Chen, C.T., N.V. Smith and F. Sette, 1991, Phys.
B~ckel, T., M. Lippert, R. Bouchard, T. Rev. B 43, 6785.
Schmidt, J.R. Schneider and W. Jauch, 1993, Chert, C.T., Y.U. Idzerda, H.-J. Lin, G. Meigs,
Acta Cryst. A 49, 679. A. Chaiken, G.A. Prinz and G.H. Ho, 1993,
Brunel, P.M., G. Patrat, F. de Bergevin, F. Phys. Rev. B 48, 642.
Rousseaux and M. Lemonnier, 1983, Acta Collins, S.P., D. Laundy and A.J. Rollason, 1992,
Cryst. A 39, 84. Philos. Mag. B 65, 37.
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 627

Collins, S.P., D. Laundy and G.Y. Guo, 1993, J. Hannon, J.P., G.T. Trammel, M. Blume and D.
Phys.: Condens. Matter 5, L637. Gibbs, 1989, Phys. Rev. Lett. 62, 2664(E).
Cooper, M.J., 1985, Rep. Progr. Phys. 48, 415. Hayes, W. and R. Loudon, 1978, Scattering of
Cooper, M.J., 1987, J. Phys. (Paris) 48, C9. Light by Crystals (Wiley, New York).
Cooper, M.J., D. Laundy, D.A. Cardwell, D.N. Heinrich, B. and J.E Cochran, 1993, Adv. Phys.
Timms, R.S. Holt and G. Clark, 1986, Phys. 42, 523.
Rev. B 34, 5984. Hill, J.P., T.R. Thurston, R.W. Erwin, M.J.
Cottam, M.G. and D.J. Lockwood, 1986, Light Ramstad and R.J. Birgeneau, 1991, Phys.
Scattering in Magnetic Solids (Wiley, New Rev. Lett. 66, 3281,
York) Hill, J.P., Q. Feng, R.J. Birgeneau and T.R.
Cowan, R.D., 1981, The Theory of Atomic Thurston, 1993a, Phys. Rev. Lett. 70, 3655.
Structure and Spectra (Univ. of California Hill, J.P., Q. Feng, R.J. Birgeneau and T.R.
Press, Berkeley). Thurston, 1993b, Z. Phys. B 92, 285.
de Bergevin, F. and M. Brunel, 1981, Acta Cryst. Idzerda, Y.U., C.J. Gutierrez, L.H. Tjeng, H.-J.
A 37, 314. Lin, G. Meigs and C.T. Chen, 1993a, J. Magn.
de Bergevin, F. and M. Brunel, 1986, Structure Magn. Mater. 127, 109.
and Dynamics of Molecular Systems II, eds Idzerda, Y.U., L.H. Tjeng, H.-J. Lin, C.J.
R. Daudel et al. (Reidel, New York). Gutierrez, G. Meigs and C.T. Chen, 1993b,
Evans, R.D., 1958, Handb. der Physik 34, 218. Phys. Rev. B 48, 4144.
Fano, U., 1969, Phys. Rev. 178, 131.
Idzerda, Y.U., C.T. Chen, H.-J. Lin, G. Meigs,
Fasolino, A., P. Carra and M. Altarelli, 1993,
G.H. Ho and C.-C. Kao, Rev. Sci. Instr., in
Phys. Rev. B 47, 3877.
press.
Gehring, P.M., L. Rebelsky, D. Gibbs and G.
Isaacs, E.D., D.B. McWhan, D.P. Siddons, J.B.
Shirane, 1992, Phys. Rev. B 45, 243.
Hastings and D. Gibbs, 1989, Phys. Rev. B
Gerson, A.R., P.J. Halfpenny, S. Pizzini, R.
40, 9336.
Ristic, K.J. Roberts, D.B. Sheen and J.N.
Sherwood, 1992, Materials Science and Jo, T. and G.A. Sawatzky, 1991, Phys. Rev. B
Technology, Vol. 2a, Chapter 8. 43, 8771.
Gibbs, D., 1992, Synchrotron Radiation News 5, Jo, T. and S. lmada, 1993, J. Phys. Soc. Jpn 62,
18. 3721.
Gibbs, D., D.E. Moncton, K.L. D'Amico, J. Kao C.-C., C.T. Chen, E.D. Johnson, J.-M.
Bohr and B.H. Grier, 1985, Phys. Rev. Lett. Brot, Y.U. Idzerda and J.B. Hastings, 1993,
55, 234. Presented at 38th MMM93 Annual Meeting,
Gibbs, D., J. Bohr, J.D. Axe, D.E. Moncton and Minneapolis.
K.L. D'Amico, 1986, Phys. Rev. B 34, 8182. Kao, C., J.B. Hastings, E.D. Johnson, D.P.
Gibbs, D., D.R. Harshman, E.D. Isaacs, D.B. Siddons, G.C. Smith and G.A. Prinz, 1990,
McWhan, D. Mills and C. Vettier, 1988, Phys. Phys. Rev. Lett. 65, 373.
Rev. Lett. 61, 1241. Kao, C.-C., C.T. Chen, E.D. Johnson, J.B.
Gibbs, D., G. Grtibel, D.R. Harshman, E.D. Hastings, H.J. Lin, G.H. Ho, G. Meigs,
Isaacs, D.B. McWhan, D. Mills and C. Vettier, J.-M. Brot, S.L. Hulbert, Y.U. Idzerda and
1991, Phys. Rev. B 43, 5663. C. Vettier, Phys. Rev. B (submitted).
Goedkoop, J.B., B.T. Thole, G. van der Laan, Koide, T., T. Shidara, H. Fukutani, K.
G.A. Sawatzky, EM.E de Groot and J.C. Yamaguchi, A. Fujimori and S. Kimura, 1991,
Fuggle, 1988a, Phys. Rev. B 37, 2086. Phys. Rev. B 44, 4697.
Goedkoop, J.B., J.C. Fuggle, B.T. Thole, G. van Krill, G., J.P. Schill6, P. Sainctavit, C. Brouder,
der Laan and G.A. Sawatzky, 1988b, J. Appl. C. Giorgetti, E. Dartyge and J.P. Kappler,
Phys. 64, 5595. 1993, Physica Scripta T49A, 295.
Grotch, H., E. Kazes, G. Bhatt and D.A. Owen, Kubo, Y. and S. Asano, 1990, Phys. Rev. B 42,
1983, Phys. Rev. A 27, 243. 4431.
Halilov, S.V., E. Tamura, D. Meinert, H. Gollisch Kubo, Y. and S. Asano, 1992, J. Magn. Magn.
and R. Feder, 1993, J. Phys.: Condens. Mater. 115, 177.
Matter 5, 3859. Kuiper, K., B.G. Searle, P. Rudolf, L.H. Tjeng
Hannon, J.P., G.T. Trammell, M. Blume and D. and C.T. Chen, 1993, Phys. Rev. Lett. 70,
Gibbs, 1988, Phys. Rev. Lett. 61, 1245. 1549.
628 S.W. LOVESEY

Langridge, S., W.G. Stifling, G.H. Lander, J. Sanyal, M.K., D. Gibbs, J. Bohr and M. Wulff,
Rebizant, J.C. Spirlet, D. Gibbs and O. Vogt, 1994, Phys. Rev. B 49, 1079.
1994, Euro. Phys. Lett. 25, 137. SchUtz, G., 1990, Physikalische Bl~itter46, 475.
Lovesey, S.W., 1986, Condensed Matter Schtitz, G., M. Kntille, R. Wienke, W. Wilhelm,
Physics: Dynamic Correlations, 2nd edn W. Wagner, P. Kienle and R. Frahm, 1988, Z.
(Benjamin/Cummings, Menlo Park, CA). Phys. B 73, 67.
Lovesey, S.W., 1987a, Theory of Neutron Schtitz, G., M. KnUlle and H. Ebert, 1993a,
Scattering from Condensed Matter, 3rd edn Physica Scripta T49A, 302.
Vol. 1 (OUP, Oxford). Schtitz, G., P. Fischer, S. St~ler, M. Knale and
Lovesey, S.W., 1987b, J. Phys. C 20, 5625. K. Attenkofer, 1993b, Jpn. J. Appl. Phys.
Lovesey, S.W., 1993, Rep. Progr. Phys. 56, 32, 869.
257. Sette, F., C.T. Chen, Y. Ma, S. Modesti and N.V.
Luo, J., G.T. Trammell and J.P. Hannon, 1993, Smith, 1991, X-ray Absorption Fine Structure
Phys. Rev. Lett. 71, 287. (Ellis Horwood, New York) chapter 23.
Majkrzak, C.E, J. Kwo, M. Hong, Y. Yafet, D. Shore, B.W., 1990, The Theory of Coherent
Gibbs, C.L. Chien and J. Bohr, 1991, Adv. Atomic Excitation (Wiley, New York).
Physics 40, 99. Smith, N.V., C.T. Chen, F. Sette and L.F.
Margaritondo, G., 1988, Introduction to Synchro- Matthels, 1992, Phys. Rev. B 46, 1023.
tron Radiation (O.U.P., New York). Stifling, W.G. and G.H. Lander, 1993, J. Appl.
McMaster, W.H., 1961, Rev. Mod. Phys. 33, 8. Phys. 73, 6877.
McWhan, D.B., C. Vettier, E.D. Isaacs, G.E. Ice, Suzuki, R. and S. Tanigawa, 1989, in: Positron
D.P. Siddons, J.B. Hastings, C. Peters and O. Annihilation (World Scientific, Singapore) p.
Vogt, 1990, Phys. Rev. 42, 6007. 626.
McWhan, D.B., E.D. Isaacs, P. Carra, S.M. Tanaka, Y., N. Sakai, H. Kawata and T. Iwazumi,
Shapiro, B.T. Thole and S. Hoshino, 1993, 1992, Rev. Sci. Instrum. 63, 1213.
Phys. Rev. B 47, 8630. Tanaka, Y., N. Sakai, Y. Kubo and H. Kawata,
Newton, R.G., 1982, Scattering Theory of Waves 1993, Phys. Rev. Lett. 70 1537.
and Particles (Springer, New York). Tang, C.C., W.G. Stifling, D.L. Jones, C.C.
O'Brien, W.L. and B.P. Tonner, 1994, Phys. Rev. Wilson, P.W. Haycock, A.J. Rollason, A.H.
B 49, 15370. Thomas and D. Fort, 1992a, J. Magn. Magn.
Pengra, D.B., N.B. Thoft, M. Wulff, R. Mater. 103, 86.
Feidenhans'l and J. Bohr, 1994, J. Phys.: Tang, C.C., W.G. Stifling, G.H. Lander, D.
Condens. Matter 6, 2409. Gibbs, W. Herzog, P. Carra, B.T. Thole, K.
Platzman, P.M. and N. Tzoar, 1970, Phys. Rev. Mattenberger and O. Vogt, 1992b, Phys. Rev.
B 2, 3556. B 46, 5287.
Rosenblum, S., A.H. Francis and R. Merlin, Ternpleton, D.H. and L.K. Templeton, 1994,
1994, Phys. Rev. B 49, 4352. Phys. Rev. B 49, 14850.
Rudolf, P., E Sette, L.H. Tjeng, G. Meigs and Thole, B.T., G. van der Laan and G.A. Sawatzky,
C.T. Chen, 1992, J. Magn. Magn. Mater. 1985, Phys. Rev. Lett. 55, 2086.
109, 109. Thole, B.T. and G. van der Laan, 1988a, Phys.
Sakai, N. and K. Ono, 1977, J. Phys. Soc. Jpn Rev. A 38, 1943.
42, 770. Thole, B.T. and G. van der Laan, 1988b, Phys.
Sakai, N. and H. Sekizawa, 1987, Phys. Rev. Rev. B 38, 3158.
36, 2164. Thole, B.T. and G. van der Laan, 1991, Phys.
Sakai, N., M. Ito, H. Kawata, T. Iwazumi, M. Rev. B 44, 12424.
Ando, N. Shiotani, E Itoh, Y. Sakurai and S. Thole, B.T., P. Carra, E Sette and G. van der
Nanao, 1991a, Nucl. Instrum. Meth. Phys. Laan, 1992, Phys. Rev. Lett. 68, 1943.
Res. A 303, 488. Thole, B.T. and G. van der Laan, 1993, Phys.
Sakai, N., Y. Tanaka, E Itoh, Y. Sakurai, H. Rev. Lett. 70, 2499.
Kawata and T. Iwazumi, 1991b, J. Phys. Soc. Thole, B.T. and G. van der Laan, 1994, Phys.
Jpn 60, 1201. Rev. B 49, 9613.
Sakai, N., 1992, Mat. Sci. Forum 105-110, 431. Thurston, T.R., G. Helgesen, D. Gibbs, J.P. Hill,
Sakurai, J.J., 1987, Advanced Quantum Mecha- B.D. Gaulin and G. Shirane, 1993, Phys. Rev.
nics (Addison-Wesley, Redwood City). Lett. 70, 3151.
PHOTON BEAM STUDIES OF MAGNETIC MATERIALS 629

Thurston, T.R., G. Helgesen, J.P. Hill, D. Gibbs, van der Laan, G., M.A. Hoyland, M. Surman,
B.D. Gaulin and P.J. Simpson, 1994, Phys. C.EJ. Flipse and B.T. Thole, 1992, Phys.
Rev. B 49, 15730. Rev. Lett. 69, 3827.
Timms, D.N., E. Zukowski, M.J. Cooper, D. van der Laan, G., 1994a, J. Elect. Spect. Related
Laundy, S.P. Collins, E Itoh, H. Sakurai, T. Phen. 68, 489.
Iwazumi, H. Kawata, M. Ito, N. Sakai and Y.
van der Laan, G., 1994b, Int. J. Mod. Phys. B
Tanaka, 1993, J. Phys. Soc. Jpn 62, 1716.
8,641.
Tjeng, L.H., P. Rudolf, G. Meigs, E Sette and
C.T. Chen, 1991, SPIE 1548, 160. Vettier, C., 1993, Physica B 192, 1.
Tjeng, L.H., Y.U. Idzerda, P. Rudolf, E Sette Vettier, C., D.B. McWhan, E.M. Gyorgy, J.
and C.T. Chen, 1992, J. Magn. Magn. Mater. Kwo, B.M. Buntschuh and B.W. Battennan,
109, 288. 1986, Phys. Rev. Lett. 56, 757.
Tolhoek, H.A., 1956, Rev. Mod. Phys. 28, 277. Vettier, C., 1994, J. Magn. Magn. Mater. 129,
van der Laan, G., B.T. Thole, G.A. Sawatzky, 59.
J.B. Goedkoop, J.C. Fuggle, J.M. Esterva, R. Vogel, J. and M. Sacchi, 1994, Phys. Rev. B
Karnatak, J.P. Remeika and H.A. Dabkowska, 49, 3230.
1986, Phys. Rev. B 34, 6529. Wang, X., T.C. Leung, B.N. Harmon and P.
van der Laan, G. and B.T. Thole, 1988, Phys.
Carra, 1993, Phys. Rev. B 47, 9087.
Rev. Lett. 60, 1977.
van der Laan, G., 1990, Physica Scripta 41, 574. Weissbluth, M., 1978, Atoms and Molecules
van der Laan, G. and B.T. Thole, 1991, Phys. (Academic Press, New York).
Rev. B 43, 13401. Williams, R.H., G.P. Srivastava and I.T.
van der Laan, G. and B.T. Thole, 1992, J. Phys.: McGovern, 1980, Rep. Progr. Phys. 43,
Condens. Matter 4, 4181. 1357.
Subject Index

A°(lat) 380 bound electrons 616


A°(val) 380 boundary conditions 507
absorption edge 571 Bragg diffraction 552, 598, 619
adiabatic magnetization 465 Bragg intensity 572, 611
alloys 3d-element based 101 Bragg law 600
anion ordering 274 Bragg scattering 572, 605, 612
anisotropic ferromagnet magnetic field 535 break of symmetry 476
anisotropic magnetostriction 470 Brillouin function 139, 412, 420, 483, 523, 594
anisotropy constants 336, 500 Brillouin zone 77
anisotropy energy 420, 441,483, 499 bulk anisotropy 507
anisotropy field 419, 420 bulk diffusion 311
anomalous scattering 568 bulk magnetostriction 470
antibonding states 379 Burger's vector 512
antiferromagnetic correlations 124
antiferromagnetic vector 464 CaCu5-type structure 499
antiferroquadrupolar ordering 155 calculated magnetization 340
Arrhenius behaviour 240 canted phase 413,414, 419, 425,429, 430, 436-
asperomagnetic state 454 438, 453, 454, 518
atomic form factor 553,600 canting 505
atomic model 589 carbonitrides 313
atomic picture 574 causality 591
attenuation coefficient 592, 595 CEF 168, 340
Auger process 590 - effects 13, 118, 139, 146, 170
average particle density 598 - Hamiltonian 141
- levels 168
ballistic magnetometer 503 - parameter 380
band structure calculations 375 - picture 18
basal plane 500 - splittings 15
Bessel function 73 centrosymmetric crystal 602
bifurcation theory 507 charge gaps 35
birefringence 446 charge ordering 165
- effects 444 charge screening 61
Bohr orbital 230 chemical radical 280
Bohr radius 626 chemical reduction via mechanical alloying 389
Boltzmann limit 614 Chevrel phase systems 166
bonded magnets 383 circular dichroism 561,589, 592
Born approximation 586, 613 circular polarization 568, 609, 620

679
680 SUBJECT INDEX

classical radius of the electron 626 crystallographic parameters 357


Co sublattice 491,495, 518 crystallographic structure data 317, 322, 350
coercive force 463, 507 Cu-O hybridization 228
coercivity mechanism 394 cubic anisotropy 422, 430
coherence length 260 cubic Laves phase 148
collinear phases 412, 438 - phase compound 148
compensation-boundary motion 511 Curie point 456
compensation point 409, 415, 416, 446, 450 Curie temperature 328, 330, 354, 356, 370, 377
compensation temperature 409, 419, 440, 442,
443,447, 453, 462, 475 d-d-interaction 535
compensational domain walls 450 d-f interaction 482
competing exchange interactions 130 d-sublattice 482, 506, 512
compounds Daniel-Friedel model 94
- 2:17-type compounds 459 Debye temperature 3
Compton limit 613, 617 Debye-Waller factors 572, 613
- of scattering 615 defects 508
Compton profile 556, 577, 579 degeneracy 425, 475
Compton scattering 617, 619 degeneracy space 524
cone-canted phase 502 degeneration 454
contact hyperfine coupling 77 demagnetization field 67, 72, 96
contact hyperfine field 73 density matrix 552, 620
contact hyperfine interaction 72, 78 dichroism 559, 579, 585
contact hyperfine Knight shift constant 75 differential cross-section 580
continuous symmetry 454 differential susceptibility 456
copper-oxide compounds 212 diffraction 565, 598
Coqblin-Schrieffer model 7 - cross-section 579, 601,602
correlation functions 557 experiments 602
-

correlation length 78, 79, 160, 264 diffusion constant 312


correlation radius 455, 456, 523 dilute magnetic systems 71
Coster-Kronig process 590 dipolar interaction 78
Cotton-Mouton effects 509 dipolar Knight shift constant 75
Coulomb correlation energy 6 dipole-allowed transitions 571
covalency effect 203 dipole approximation 581,588, 610
critical behaviour 79, 182 dipole-dipole coupling 264
critical concentration 40 dipole-dipole interaction 72, 77
critical exponent 78 dipole field calculation 87
critical field 416, 447, 493 dipole transitions 547
critical fluctuations 79, 455 dislocation 508, 509
critical phenomena 96 dodecahedral 440
critical point 430, 431,445, 460, 476, 485 domain magnetization 93
critical scattering 575 domain of 180° 477
critical temperature 78, 312, 313 domain structure 453, 475, 478
critical transition fields 494 domain wall 476, 507
cross-section 552, 554, 623 - energy 395
crossover instability 522 - width 395
crossover level transitions 411 DOS 370
crossover point 414 Dy sublattices 491,495
crystal electric field parameter 373 dynamic Kubo--Toyabe function 70, 107, 113
crystal field 7, 440 dynamic Lorentz-Kubo--Toyabe behaviour 181
- parameters 333, 338, 339, 349, 361, 372, 380 dynamic Lorentz-Kubo-Toyabe function 183
- perturbations 612 dynamic scaling theory 78
crystalline electric field (CEF) 6, 13, 135
- effects 21, 117 E1 attenuation coefficient 598
splitting 61, 75
-
E1 circular dichroism 611
SUBJECT INDEX 681

E1 linear dichroism 611 Faraday effect 441,450, 462, 478, 591


E1 magnetic dichroism 594 - measurements 417
E2 magnetic dichroism 594 Faraday rotation 441--443, 480
easy axis 419, 424, 425, 427, 482, 485 Fermi Golden Rule 613
easy magnetization direction 336 Fermi liquid picture 7
easy plane 423, 482 Fermi surface 193
elastic cross-section 557 ferrimagnetic 419
elastic resonant diffraction 618 - compounds 408
elastic resonant scattering 579, 610 - phase 414
elastic scattering amplitude 610 ferrimagnets 408
electrical field gradient (EFG) 108, 142, 347 ferromagnet: 1D-ferromagnet 281
- created by the #+ 95 ferromagnetic iron 553, 569
electrical resistance of heavy fermions 4 ferromagnetic Ni 552
electron density distribution 374, 381 ferromagnetic phase 414, 419
electron doping 237 field hysteresis 505
electronic band structure calculation 367 field induced phase transition 408
electronic phase separation 248 fine-structure constant 626
elemental metals 80 first order magnetic processes (FOMP) 411
ellipticity 443 first order phase transitions 419, 430, 434, 438,
energy product 382 439, 445, 448
enthalpy change the decomposition 311 first order transitions 431,488, 526
entropy 465 fluctuations 455
epoxy bonded magnets 386 forward scattering amplitude 587, 591
equilibrium conditions 482 four-fold degeneracy 430
Euler-Lagrange equations 477, 507, 523 Fourier analysis 66
Ewald method 73 free energy 424, 482, 506
EXAFS experiments 326 free-powder magnetization 518
exchange free-powder samples 512
- a - d 440 freely suspended sample 513
- f - d 527 Fresnel equations 558, 586
exchange constant 500 Friedel oscillations 61
exchange-coupled two-phase magnet 382 frustrated Kagom6-1attice system 215
exchange energies: 4f-3d exchange energies 504 frustration 223, 236, 238, 264
exchange-enhanced itinerant paramagnets 527 f-sublattice 482, 512, 523
exchange field 420, 466
exchange field and anisotropy constant 495 galvanomagnetic effects 464
exchange interaction 75, 328, 331, 379, 416, gapping of Fermi surface 164
440, 482 gas-phase interstitial modification (GIM) 307,
- parameter 412 310
exchange stiffness 477, 523 gas-solid solution phase 312
constant 506
-
Gaussian distribution 69
expansion of the domain walls 476 Ganssian Kubo-Toyabe function 195
explosion sintered magnet 393 Gaussian Kubo-Toyabe signal 70
extinction 566, 570 Ganssian relaxation 68, 263
Gd sublattice circular birefringence 442
F - :/~+ : F - 'hydrogen'-bonded center 216 Gd-Fe exchange energy 441
f-d exchange 527 Gd-Y superlattice 567
f--d interaction 522, 527 giant moment alloys 97
f~l intermetallic compounds 521 GIM process 313
f-d system 522 Ginzburg-Landau energy 523
f-f interaction 416, 482 ground state 414
Fano effect 563, 564 ground state multiplet 168, 264
far-infrared 24 gyromagnetic factor 600
682 SUBJECT INDEX

Hall effects 462, 464 hydride phase 257


Haldane system 279 hyperfine anomaly 103
Hamiltonian 533 hyperfine field 84, 85, 93, 103, 203, 204, 342,
Hamiltonian of the d-f system 533 363
handedness of the circular polarization 589 - 57Fe M6sbauer 363
hard permanent magnets 170 hysteresis 462--464, 497, 498, 525
HDDR method 392 - loops 463
HDDR process 307, 391 - phenomena 494, 498
heat capacity 455
heavy electron compounds 148 impurity induced strains 101
heavy electron superconductors 199 incoherent approximation 617
heavy electron systems 185 incommensurate 134
heavy electrons without free carriers 165 - magnetic structures 67
heavy fermion 3 - modulated spin structure 163
heavy fermion compounds 148 - spin density wave 199
heavy fermion superconductor 184, 186, 197
- sinusoidally modulated 277
Heisenberg antiferromagnet 240 - structures 73
Heisenberg exchange interaction 593 index of refraction 585
Heisenberg exchange parameter 222 induced anisotropy 511
Heisenberg ferrimagnet 416
induced contact hyperfine field at the/z + 96
Heisenberg ferromagnet: 2D-Heisenberg ferro-
inelastic neutron scattering 142
magnet 281
influence of an elastic strain 89
Heisenberg Hamiltonian 76
inhomogeneous line broadening 68, 69
Heisenberg model 278
inhomogeneous magnetic ordering 274
- 2D-Heisenberg model 222
integral curve 477, 478
Heisenberg operator 557
interaction
helicity 589, 621
- f~l 522, 527
helimagnet 106
- f-f 416, 482
hexagonal anisotropy 422, 482, 491,498
inter-chain coupling 279
hexagonal axis 482
intermediate nitrogen content 323
hexagonal c-axis 501
interstitial atoms: 18g-site interstitial atoms 326
hexagonal ferrimagnets 482
interstitial site: 2b interstitial site 355
hexagonal structure 422, 516
intersublatice-coupling strength 512
hexagonal structure of CaCu 5 491
intersublattice exchange field 419
hexagonal symmetry 517
lntra sublattice interactions 440
hexagonal Th2NilT-type structure 315
high coercivity 388 intrinsic magnetic domain parameters 394
high-field domains 481 inverse transition 526, 527
high-field magnetization 336 irradiated 101, 102
high-field free-powder method (HFFP) 512 Ising axis 515
Hilbert transforms 591 isoentropy slope 465
Ho-sublattice 501 isomer shifts 346
Hois-Ylz superlattice 567 lsotropic ferrimagnet 411,416
hole doping 237, 239 lsotropic magnetostriction paraprocess 494
hole dynamics 259 itinerant electron model 278
holmium 553, 566, 569, 570, 612 itinerant f-d metamagnets 521
holmium metal 565 itinerant ferrimagnet 526, 528
homogeneous line broadening 68 itinerant ferromagnet 521
H-T diagrams 418, 510 itinerant ferromagnetic system 106
Hund's rule 593 itinerant ferromagnetism 527
hybridization between Fe 3d and R 5d electron itinerant metamagnets 410, 418, 521
states 330
hybridization effects 368 Jaccarino-Peter effect 167
hybridization matrix element 6 Jahn-Teller effect 411,415
SUBJECT INDEX 683

jump-like magnetization 514, 515 local phase diagrams 510


jump-like transition 419 localized model of rare earth magnetism 573
location of interstitial atoms 324
Kagom6 lattice 113 location of N atoms 354
Kerr effect 613 longitudinal field (LF)-/~SR 65, 67
kinks in the magnetization curves 527 longitudinal magnetostriction 472, 494, 504, 505
Klein-Nishina formula 613,615, 618 - measurements 417
Knight shift 67, 114, 132, 141, 198, 199, 264, longitudinal spin fluctuations 127
265, 269 longitudinal susceptibility 424
Kondo effect 7, 13 Lorentz field 67, 96, 110
Kondo insulator 36, 37 Lorentz sphere 72, 73
Kondo lattice 8, 161, 162 Lorentz-Kubo--Toyabe signal 134
Kondo mechanism 75 Lorentzian distribution 69, 71
Lorentzian Kubo-Toyabe behaviour 181
Kondo metals 162
LTO phase 235
Kondo-necklace phase diagram 9
LTO-LTT transition 233
Kondo resonance 193
LTT phase 235
Kondo temperatures 13
Korringa constant 168
Korringa mechanism 79, 116, 141, 168, 170 m* 3
Kramers-Heisenberg amplitude 579 magnetic anisotropy 146, 333, 335, 356, 416,
Kramers-Heisenberg formula 549 421,512
Kubo-Toyabe function 69, 107 - constants 512
Kubo-Toyabe signal 111, 142, 158, 179, 230 magnetic correlations 268
magnetic critical scattering 576
magnetic domain patterns 313
Landau theory 455, 456 magnetic energy 483
Land6 factor 593, 601 magnetic fluorides 215
Larmor precession frequency 66 magnetic frustrations 229
lattice dilatation 61 magnetic insulators 199
lattice gas model 312 magnetic Jahn-Teller 415
lattice parameters 316, 321 magnetic moments 367, 376, 377, 379
- of nitrides and carbides 316 magnetic phase diagrams 445, 503
lattice relaxation 75 magnetic properties of elemental metals 80
lattice sum 72, 73 magnetic susceptibility 522
lattice vector 600 magnetism of insulators 200
Laves phases 527 magnetization 456, 457, 459, 462, 493,498, 502
layered oxides 212 - process 503, 512
left-hand polarization 622 - of sublattices 495
Legendre polynomials 194 magnetocaloric effect 455, 465
Legendre transformation 533 magnetocrystalline anisotropy 139
Levanyuk-Ginsburg criterion 456 magnetoelastic anomalies 469
line shapes 613 magnetoelastic energy 447
linear birefringence 444, 464 magnetoelastic interaction 508
linear dichroism 561,589, 595 magnetometer 502
linear magnetoelectric effect 464 magnetooptical phenomena 455
linear magnetostriction 464 magnetooptical Kerr effect 558
linear polarization of secondary beam 605 magnetoresistance 464
linear term 4 magnetostriction 455, 469, 493, 497, 498, 503
of specific heat 166
-
- deformation 505
linear-muffin-tin-orbitalmethod (LMTO) 367 - of paraprocess 470
liquid phase epitaxy 450 magnon energies 76
local DOS 378 magnon excitations 77
local field 203 martensitic transition 133
- a t / z + 84,93 matrix element 588
684 SUBJECT INDEX

maximum energy product 387, 396 - #+SR technique 62


Maxwell rule 514, 515 - / ~ + trapping 117, 118
mean field approximation 411 - #+ tunnelling 87
mean field theory 455 - / ~ + zero-point motion 87
mean field treatment 277 muonic atom 103
mean square magnetization 596, 611 muonium 61, 199, 216, 217
mechanical alloying (MA) 388, 389
mechanical grinding 388 nano-structured two-phase magnetic system 393
mechanical grounding (MG) 389 N6el antiferromagnet 408
melt spinning 389 negative differential susceptibility 525
melt-spinning method 309 nesting model 279
melt-spun interstitial carbides 309 neutron diffraction 227, 236, 238, 246, 264, 268
metamagnetic phase transition 521 neutron diffraction studies 375
metamagnetic transition 160, 411,525, 527, 531 neutron magnetic diffraction 565
metamagnetic transition field 529 neutron polarization analysis 126
metastable canted phase stability 427 neutron scattering 110, 124, 128, 155, 181, 189,
metastable phase 435, 437, 443 239, 248, 250, 262
rnicrostresses 510, 511 nitrided materials of l:12-type 393
missing fraction 66 NMR 62, 66, 76, 106, 107, 111, 155, 158, 190,
mixed reflections 603, 605 193, 219, 230, 256
mixed valence 23 NMR/NQR 251
mode-coupling theory 78 non-centrosymmetric materials 568
modulated structure 134 non-collinear phase 494
molecular crystal 220 non-collinear structure 446
- o~-O 2 199 - magnetic 409
molecular field 420, 483, 522, 527, 593 non-Fermi liquid behaviour 40
- approximation 593, 597 non-magnetic absorption 551
- coefficients 331 non-resonant scattering 618
molecular magnetic field 587 - amplitude operator 582
momentum density 580 nonuniform-exchange constant 523
momentum distribution 556 NQR 151, 215
Morin temperature 203 nuclear hyperfine interaction 264, 265
M6ssbauer 140, 259 nuclear Schottky anomaly 164
- data 260 nuclear spin lattice relaxation 107
- 57Fe 341, 363 nucleation 477
- spectroscopy 62, 76 - of domains 509
motion of phase boundaries 509 - of new phase 475
motional averaging 69
muon Knight shift 96 O-H bond 244
muon-nuclear level-crossing resonance 218 observed domain width 395
muon-quadrupolar level crossing 108 octahedral 440
muon-spin relaxation in 3d-elements 97 one-electron atomic model of elastic resonant
muon-spin relaxation in lanthanides 98 scattering 573
muon-spin rotation one-electron model 590
- #+-beams 64 operators of the crystal field 534
- # + diffusion 117, 118, 124 optical theorem 551,586, 587
- #+-Knight shift 106 orbital angular momentum operator 583
- / z - - K n i g h t shift 230 orbital contributions to the cross-section 604
- #+ motion 87 orbital magnetism 581, 582
- / z + site 61 orbital magnetization 584, 599, 624
- / z + S R 103 ordered moment 15
- # - S R 62, 103 orientation transition 455
- #SR frequency shift 96 overcritical exchange 265
- / z S R signal 64 oxygen-muon bonds 199
SUBJECT INDEX 685

PAC: 3'~' PAC 62 quadrupolar anisotropy 132


PAD 62 quadrupolar interactions 133
paramagnetic fluctuations 79 quantum fluctuations 160, 522
paramagnetic state 522 quantum Hall effect 274
paramagnetic sublattice 426 quasi-elastic neutron scattering 76
paraprocess 470
parity 588
R-Co exchange interaction 491
- violation 62
radial dipole matrix 592
partial differential cross-section 557, 616
radiation damage 61
partially polarized radiation 620
Raman process 76
particle-hole continuum 577
Raman scattering 624
Pauli paramagnetism 126
random order 228
percolation threshold 524
percolative network 260 range in matter 65
permanent magnet 396 rare-earth M6ssbauer 347
- applications 382 rare-earth orthoferrites 199, 209
phase rare-earth sublattice 416, 423, 466
- HTT 232 reaction enthalpy of interstitial nitride formation
- LTO 232 311
- LTT 230, 232 reaction enthalpy 311
phase diagram 419, 428, 430-432, 437, 439, reciprocal lattice 600
443,444, 485, 495, 526 recoil energy 617
- H-T 411,420, 424, 497, 500, 521, 525, 528 reduced matrix element 592
phase transitions reentrant ferromagnetism 166
- of 1st kind 427 reentrant-like phase transition 419
- of 2nd kind 427 related alloys 81
- of 2nd order 445 remanent magnetization 391,463
phonon dispersions 577 renormalization-group calculations 78
phonon induced relaxation 140 reorientation 505
photoelastic effect. 508 rest mass energy of the electron 626
photoemission 558 rhombohedral structure 322
piezomagnetism 464 rhombohedral Th2ZnlT-type structure 315
plasma nitriding 314 rhombohedral crystals 422
plasmon peaks 577 right-hand circular polarization 621
plasmons 613 RKKY 8
plots of the magnetization 447 - exchange mechanism 169
point charge model 142 - exchange rate 168
polarization - interaction 79, 114, 116
- analysis 566 - mechanism 72, 73, 124
- 7r 556, 622 rotational process 419
- c~ 556, 622
Rydberg 626
- states 619
- vectors 580
polarized/~+-beams 62 s--d hybridization 506
positron distribution 64 satellite reflections 612
powder broadening 114, 127 saturation magnetization 529
power law 219, 220 scaling laws 78
pre-hydrogenation treatment 385 scattering amplitude 622
precursor phenomenon 264 scattering amplitude operator 568, 580, 619
pressure 16 scattering by free charges 613
pressure dependence 12 scattering of light 431
- of Bt~ 88 scattering vector 580
- of hyperfine field 95 SCR theory 107, 108, 110, 111
proximity effect 248 SCR model 109
686 SUBJECT INDEX

SDW condensation 279 spin polarization 65


second-order phase transition 428, 436, 438, spin reorientation 337, 347, 361, 362
480, 481,488 - temperature 339
secondary linear polarization 608 - transition 237, 239, 242
selection rules 588 spin wave excitations 76, 278
self-consistent renormalization (SCR) theory 106, spin wave spectrum 36
124 spin wave stiffness 137
semiconducting 24, 35, 36 - constant 278
semiconductors 34 spin-1 linear-chain Heisenberg antiferromagnet 279
shake-off process 562 spin-orbit branching ratio 564
shake-up process 562 spin-spin or transverse relaxation rate 68
short circuit diffusion 311 spin-canted state 416
short range correlations 118 spin-flop 419, 499
short range magnetic order 223 - state 277
short range order 128, 228, 236 - transitions 410, 424, 503, 504, 508, 521
signal amplitudes 66 spin-lattice relaxation 76, 218
single crystals 502 spin--orbit interaction 441
single particle excitations 278 spin-polarized partial densities of states (DOS)
singlet ground state ions 142 370
singular point detection 337 spin-reorientation 496, 499, 504
site percolation threshold 228 - phase transitions 465
small moment magnetic order 197 - processes 499
soft mode 455 spontaneous linear birefringence 446
solid oxygen 220 spontaneous magnetization 496, 506, 522, 531
Sommerfeld constant 154, 163 spontaneous spin-reorientation 410
sound absorption 445, 473 square root stretched exponential 71
sound velocities 455 SQUID 502
sound velocity change 473 stability conditions 430, 436
specific heat 3, 445, 455, 465, 466 stability lines 462
- -y-values 3 stability regions 507
specific heat anomaly 188 stacking faults 264
spectroscopy 613 static approximation 566
spherical components 588 static correlations 263
spin canting 170, 174 static limit 618
spin contributions to the cross-section 604 static Lorentzian Kubo-Toyabe behaviour 182
spin density 584, 600, 623 stereo graphic projection 433
spin density enhancement 74 Stevens equivalent operators 333
spin density wave (SDW) 152, 164, 274 Stokes vector 553
spin dynamics 76, 77, 114, 280 Stoner criterion 527
- 4f 260, 261, 263 stretched-square root-exponential 125
spin flip transitions 67 strongly ferrimagnetic phase 531
spin fluctuations 175, 182, 330, 370, 521,522 structural phase transition 166
spin gap 35 structural transitions 175, 236
spin glass 62, 125, 127, 153, 165, 229, 230, 259, structure modification 315
274 structure of Th2Ni17 320
- Kubo-Toyabe function 259 structure transformation 320
- order 251 structure transition from hexagonal Th2NilT-type
- type 11 structure to rhombohedral Th2Znl7-type struc-
spin glass-like order 154 ture 309
spin hole 227 sublattice magnetization 411, 511
spin lattice relaxation 79, 106, 111, 230 substituted intermetallic compounds 527
spin lattice relaxation rate 67 sum rules 563, 619
spin magnetization 600 superexchange coupling 264
spin operator 620 superconducting coherence length 248
SUBJECT INDEX 687

superconducting order parameter 189 transverse field (TF)-/~SR 65, 66


superconductivity 5, 36 transverse magnetization 503
superconductors 4 transverse spin fluctuations 127
superlattices 570 transverse susceptibility 424
supertransfer of hyperfine fields 203, 210 tricritical points 418, 427, 428, 435, 436, 485
surface anisotropy 506 tunneling spectra 37
surface beam 65 twins 475
surface field-induced phase transition 507 twinned canted domains 478
surface phase transition 506, 507 twinned domains 475
susceptibility 423, 429, 456, 457, 496 two-channel relaxation 107
symmetry breaking 522 two-fold degeneracy 476
symmetry breaking field 189 two-magnon effects 576
symmetry of crystal lattice 36 two-magnon scattering 137
synchrotron sources 547, 618 two-sublattice anisotropic ferromagnet 512
two-subnet amorphous systems 410
T*-phase 239
technical magnetization 497 ultrasmall ordered moments 190
temperature dependence of the hyperfine field 89 ultrathin films 507
terbium-iron garnet 559 uniaxial anisotropy 421, 506, 507
tetragonal structure 421,516 - field 480
tetragonal symmetry 517 - magnetic 421,424
tetragonal ThMnl2-type body-centered structure uniaxial ferrimagnet 425
354 unit cell 600
tetrahedral 440 - structure factors 600
TF-/~-SR-spectroscopy 230, 250 unstable d-subsystem 527, 528
thermal expansion 472
thermal properties 465 vacuum tunneling experiments 23
thermal stability 382 van Hove response function 76, 614
thermodynamic potential 419, 420, 423, 424, Verwey temperature 206
482, 512, 523 weak itinerant antiferromagnet 113
- of the non-equilibrium state 421, 532 weak itinerant ferrimagnet 521
thermopiezic analyzer 310 weakly ferrimagnetic phase 531
Thomson cross-section 556 white line 598
Thomson formula 616 Wigner-Eckart theorem 588
three sublattice model 440
threshold concentration 523, 524 X-ray absorption (XANES) 250
threshold field 419, 424 X-ray dichroism 559
time reversal invariance 196 X-ray scattering 198
time reversal invariance violating state 189
topological frustration 113
- magnetic 127 Young modulus 455, 473
total scattering 614 - AE-effect 473
transferred hyperfine field 346
transition metals: 3d-transition metals 81 Zeeman energy 420
transversal magnetostriction 494 zero field (ZF)-/~SR 65
measurements 417
-
zero point vibrations 74
Materials Index

amorphous Bi2Sr2CaCu20 v 271 CeCo2Ge2 14


amorphous Dy-Co films 454, 465 CeCo2Si2 14
amorphous DyAg 135 CeCu2 5, 11-16, 22, 24
CeCu5 41, 144, 149, 154, 155
BaCuO2 201,212, 245, 269 CeCu6 5, 7, 40, 144, 149, 155, 160
BaY2CuO 5 201,212, 245, 269 CeCuA13 17
a-Bi203 201, 215 CeCu4A1 41
Bi2SrCaYCu208 269 CeCu6_xAuz 40
Bi2SrCaYCu2Os+ 6 273 Ce(Cul_xAl=)5 144, 149, 154
Bi2SrYCaCu2Oy 270 Ce(CUl_=Ga=)5 144, 149, 154
Ce(Cul_z(Ga, A1)=)5 159
Bi2Sr2Yo.6Cao.4Cu208 273
Bi2Sr2Yl_xCaxCu208 273 Ce(CUl_zNi=)2Ge2 16
Bi2Sr2Yl_zCazCu2Ou 271,272, 274 CeCuGa3 17
CeCu4Ga 41
Bi2Sr2YCu208 269
CeCuGe 22
Bi2Sr2YCu2Ov 270, 272
CeCul.3Sb 2 14, 15
Bi2Sr2_~YzCu208 273
CeCuSi 22
Bi2Sr2.sYo.sCu20~8 273
CeCu2Si2 144, 151, 157, 158
BizSr3_=YzCu208 270
CeCu2+=Si2 149, 159
CeCuSn 22, 24
Cao.86Sro,14CuO2 201, 214 CeCuX3 17
Ce: ternary equiatomic compounds 22 CeFe2Ge2 14
CeAg 129, 132, 133 CeFe2Si2 14
CeAgl_=In= 129, 132, 133 CeFelo.sVl.sNy 360
CeAg2Ge2 14, 15 CeFellTi 357
CeAg2Si 2 14 CeFel 1TIN1.5 357
CeA12 114, 115, 144, 148, 149, 151, 159 Ce2Fel4 B 173
CeA13 5, 7, 144, 149, 151-153, 159, 228 Ce2Fel7 317
CeA1Ga 22, 24 Ce2Fe17C 322
CeA12Ga2 14 Ce2FeI7C2.5 317
CeAs 129, 130, 178, 180 Ce2Fe17Cu 343
CeAu2Ge 2 14 Ce2Fel7N2.5 317
CeAuIn 22 Ce2Fel7N3.6 317
CeAu2Si 2 14, 16 Ce2Fe]7Ny 343
Ce3Au3Sb4 35, 36 CelrGe 22
CeB 6 144, 149, 155, 156, 157 Celr2Ge 2 14
CeCo3B 2 28, 29 CeIr2Si2 13, 14
CeCo4B 28, 29, 31 Celr2Sn2 14

689
690 MATERIALS INDEX

CeM2Si2 16 CoF2 202, 215


CeNiA1 22, 23 Chromium 82, 102
CeNiGe 22 Chromium alloys 82, 102
CeNi2Ge2 14 compounds
CeNiln 22 - 3d-4f 504
CeNio.8Pdo.2Sn 23 - ternary equiatomic 22
CeNiSb 22 Cr203 200, 204
CeNi2Sb2 14, 15 CsAs 128
CeNiSi 22 CuF2 202, 217
CeNizSi2 14 CuO 201, 212
CeNiSn 22, 23, 35, 144, 150, 162
CeNizSn2 13-15 Dy 83, 93, 96, 571,595, 596
CeOs2Si2 14 DyAg 129, 133, 134
CePb3 144, 154 DyAI2 114-117, 124
CePdA1 22 DyA1G 418
CePdzA13 15, 33 Dy3AIsO12 428
CePdGa 22 DyBa2Cu307 266
CePdGe 22 DyBa2Cu3OT_~ 263
CePd2Ge2 14 DyCo5 482, 491,496, 498
CePdIn 22, 23 DyCos+~ 410
CePdSb 22, 23 DyCos.3 492, 494, 495
CePd2Si2 14 DY2Co7 491
CePdSn 22, 23, 144, 150, 162 DyCo12B6 518
CePd2Sn2 14 DY3Fe5012 458, 467
CePtAuSi2 16 iron garnet 473
-

Ce3Pt3Bi4 35, 36 Dy2FeI4B 173


CePtGa 22 Dy2Fe17 318
CePtGe 22 Dy2FeI7C2.5 318
CePt2Ge2 13, 14 DyzFel7Cu 343
CePtln 22, 23, 24 DyFelo.sMol.sNu 359
CePtSb 22, 24 Dy2FelvN2.5 318
Ce3Pt3Sb4 35, 36 DyzFel7Ny 343
CePtSi 22, 24 DyFeO3 200, 209, 477
CePtaSi2 13, 14, 16 DyFeloSizCo.3 360
CePtSn 22, 144, 150, 162, 163 DyFell Ti 357
CePt2Sn2 14, 15, 33 DyFe11TiCu 358
CeRh3B2 29 DyFell TiNo.5 358
CeRhGe 22 DyFell TiNy 358
CeRh2Ge2 14 Dy-ferrite garnets 471,472, 478
CeRhln 22, 24 DyMn2 119, 127
CeRhSb 22, 23 DyNi2 119, 122, 124
CeRh2Si2 14, 143 DyNi5 136, 139
CeRh2Sn2 14 DySb 129, 132
CeRu2Ge2 14
Ce(Ruo.85Rh0.15)2Si2 144, 150 epitaxial ferrite-garnet films 409, 479
CeRu2Si2 14, 142, 144, 150, 160-162, 190, 196 epitaxial film Yz.6Gdo.4Fe3.9GaHO12 451,452
CeSb 129-132 erbium 573
CeT2Ge2 15 Er 83, 96, 100, 567, 571
CeT2Si2 15 ERA12 115-118
CeTX 21 ErBa2Cu3O6.2 268
CeT2X2 13, 14 ErBa2Cu3Ox 260, 265, 267-269
Co 81, 93, 101, 561 Ero.8Dyo.2FeO3 464
Co alloys 81 ErFe2 118, 119, 123, 127, 463
COC12.2H20 219, 202 ErFe3 460, 471,473,474
MATERIALS INDEX 691

Er2Fel7 319 GdBa2Cu307_~ 261


Er2FeI4B 173 GdCo2 118, 119, 121-123, 127
Er2Fe17C 322 GdFe2 118-120, 122, 123, 127
Er2Fel7C2.5 319 Gd2Fel7 318
Er2Fel7fy 337, 338, 343, 346, 347 Gd2Fel4B 173
ErzFelTCI.oNy 319 Gd2Fel7C2.5 318
Er2Fel7CxNv 338 GdzFel7Cy 343
Er2Fel4_xMn~C 512 GdzFe17N2.5 318
ErFelo.sMOl.sNy 359, 362 Gd2Fel7Ny 343
ErzFelvN2.5 319 Gd3FesOl2 411, 441,447, 464, 467, 469, 474,
Er2FelvN2.7 338 479, 508, 561
Er2Fe17Ny 343 (Gd,Y)3Fe5OI2 461
ErFeO3 201, 209, 464 - garnets 411,416, 417, 448
ErFel 1Ti 358 Gd-ferrite garnets 443, 444, 466, 471,479
ErFel 1TiCy 358 GdFelo.sMol.sNu 359
ErFel 1TiNo.5 358 GdFeloSizCo.3 360
ErFej1TiNu 358, 362 GdFeloVzNv 360
ErFel0.sV1.sNy 360 GdFe11Ti 357
ErNi5 136, 139 GdFell TiCu 357
ErRh4B4 166, 167, 172 GdFel 1TiNo.5 357
Er2_zYzCoTB3 519 Gd-Ga garnet 450
Er3_zYxCOllB 4 519, 520 Gd3GasO12 479
(Er,Y)COllB4 520 GdNi2 119, 122
Eu 83 GdNi5 136, 137
EuBa2Cu307 266 GdRh4B4 168, 172
EuFeO3 200, 209 germanium 602
EuMo6S8 166
EuMo6S7.sSe0.5 169, 172 hexagonal 421
EuNiO3 201, 211 - compounds of d-f type 482
EuO 77 - ferrimagnets 482
Euo.75Sno.25Mo6S7.6Seo.4 169, 172 HyYBa2Cu306.6 254
EUl_zSnxMo6S 8 166 HuYBa2Cu306.7 254
HuYBa2Cu307 254
Fe 81, 89, 93, 95, 96, 101, 104, 567, 571,577, HzYBa2Cu307_~ 245, 256
Ho 83, 96, 567, 571,576
617
HoA12 115-117
Fe alloys 81
FeCo alloys 96 HoBa2Cu306.2 265, 266
HoBa2Cu307 228
FeF2 418, 428
HoBa2Cu3OT_~ 264, 266
Fe3Ge 367
HoBa2Cu3Oz 261, 263
Fe3GeNy 366
HoCo3Ni2 499, 503-506
c~-Fe203 199, 200
Hoo.5Ero.5 571
Fe304 200, 206 Ho-ferrite garnets 471
FeTiO3 200, 205
HoFe2 555, 567, 569, 570, 577
FeZr alloys 96 HoFe3 460, 461
Ho2Fel7 319
gadolinium-iron garnets 409, 508 Ho2Fel7_xAl~ 518
gadolinium-ferrite garnet 429, 478 Ho2Fe14B 173, 174
Gd 83, 84, 88, 93, 97, 99, 561,577, 595 Ho2Fe17Cy 343
GdA12 114-117, 124 HoFemsMol.5Ny 359
Gd3A12(SiO4)3 439 Ho2Fe17N2.5 319
GdBa2Cu307 266 Ho2Fe17Nv 343
GdBa2Cu3Oz 261 HoFeO3 201, 209
GdBa2Cu306_ z 262 Ho3FesO12 467
692 MATERIALS INDEX

- garnet 445, 473, 474 Lal.2Tbo.8CuO4_6 226


HoFell Ti 358 La2_xTb~CuO4 239
HoFell TiCy 358 (La,Tb)2CuO4 240
HOFelo.5V1.5Nv 360 Lal.9Yo.lCaCuO6+, 222
Hoo.oosLuo.995Rh4B4 172 LuCo2 521,527
Hoo.02Luo.98Rh4B4 172 LuFe2 118, 119
HoxLUl_xRh4B4 168, 172 Lu2Fe17 319
HoRh4B4 166, 172 Lu2Fe17C 322
Hoo.sYo.sBa2Cu3OT_, 266 Lu2FelTCv 343
HOo.41Y2.59FesO12 garnet 446 Lu2Fe17N2.7 319
Ho1.05YL95Fe5012 garnet 447 Lu2Fe17Nv 343
(Ho,Y)3FesO12 449 LuFell Ti 358
(Ho,Y)IG 448, 449 LuFe11TiCu 358
LuRh4B4 172
iron 579
Mn(CH3COO)2.4H20 416
Mn ferrite 577
lantanides 83 MnF2 61, 79, 202, 216
LaAgl_~Inx 133 MnF3 202, 217
LaA12 115 MnFe2 77
La1.875BaoA25CuO4 222, 233 MnGe2 465
La2_xBa~CuO4 232 MnO 202, 215
La2_xBazCuO4_ ~ 225 MnSi 105, 106, 108-110
La2_~Ba~CuOx 232 Mno.75Zno.25F2 216, 570, 576, 605--607
Lal.875Bao.o75Sro.osoCuO4 222, 225, 233, 234
La2CaCu206+ ~ 226, 243
NbFe2 113
La(Col_~Fex)13 365, 366
Nbo.9Zro.1Fe2 113
La1.85CuO4_, 224
Nbl_xZrxFe2 105, 113
Lal.95CuO4_~ 224 Nd 83
La2_~CuO4_, 222 NdAI2 115, 116
La2CuO4 221, 223, 236, 239, 258
Ndi+yBa2_vCu207+~ 245
La2CuO4_, 214, 224
Nd1+vBa2_yCu3OT_a 266
La2.olCuO4_~ 224 Ndl.88Ceo.12CuO4_~ 226
La2Cuo.25Coo.7504 230 Nd2_~CezCuO4 238
La2(CUl_xCox)O4+6 222 Nd2_zCexCuO4_, 222, 226, 237
La2(Cul_~Zn~)O 4 222, 224, 227 NdzCuO4 221,236, 237
LaFel3_~Si~ 365, 366 Nd2CuO4_, 222, 226, 237
La2MCu206+ ~ 222, 243 Nd~_~Dy~FeI1TiNu 364
LaNi5 135, 136 Nd2Fe17 317
La2NiO4 235 NdsFe17 364
La2NiO4+a 222, 225 Nd2FeI4B 171, 173, 174
LaPd2P2 142, 143, 161 NdzFel4BN~ 365
La2_xSrxCaCuO6+ , 243 Nd2FelTC2.5 317
Lal.85Sro.olCuO4_ ~ 224 Nd2Fe17Cy 343
La1.89Sro.ÂlCuO4 229 Nd2Fel7N2.5 317
Lal.95Sro.osCuO4 231 Nd2Fel7N4.5 325
La2SrCu206_ 6 226 Nd2FelTNu 343
La2SrCu206+ ~ 243 NdFelo.sMol.sNu 359
La2_~(Sr, Ba)xCuO4 232 NdFelo.75Mol.25Nu 358
La2_z_uSrxNduCuO 4 222, 233 NdFeloMo2Nu 355, 359
La1.775Sro.125Ndo.lCuO4 225, 233 NdFellTi 357
La2_xSrxCuO 4 228, 230, 231,233, 238 Nd2(Fe,Ti)19 366
La2_xSrxCuO4_ 6 222, 225 Nd3(Fe,Ti)29 366
Lal.2Tbo.sCuO4 222, 239, 241 NdFell TiCu 357
MATERIALS INDEX 693

NdFe11TiNo.5 357 RCo12B6 416


NdFeH TiNu 357 R2Co7B3 416
NdFeloV2N 360 R2Co17Ny 350
NdFelo.5V1.sNy 360 RE-TM intermetallic compounds 410, 477
NdNiO3 201, 211 {R3+3}(Fe+3)[Fe2+3]O12 440
NdRh4B4 168 RF~ 364, 410, 416
NdRhzSi2 143, 147, 183, 185 RFel2 378
Nd2_zSrxCuO4_a 222, 226, 239 R2Fel7 315, 316, 330, 342, 410, 482
NENP 279 R6Fe23 365, 416
Ni 82, 89, 93, 95, 96, 101, 103, 561, 577, 617 R2Fe14B 170, 410, 416, 482
Ni alloys 82 R2Fe14C 416
Ni(CzHsN2)2NO2(C104) (NENP) 275 RzFeI7C 316, 330, 416
Ni2(C2HsN2)2NO2(C104) (NENP) 279 R2FelvCy 316, 330, 342
NiO 568 R2(Fel_xCox)17Ny 350
NpA12 175, 176, 178 RFel l_~Co~TiNy 364
NpAs 571, 575 R2(Fel_xMx)lTNy 352
R2(FeMn)17 482
R2(FeMn)14C 416
p-NPNN 275, 280
R2Fe17Nu 316, 330, 342
Pd alloys 82
R3Fe5012 439
Pd-based alloys 101
RFeloSiCu 366
PdFe 97 RFe12_~TxNu 375
Pdo.97Feo.o3 98 RFel 1TiNy 362
PdFeo.oo35Mno.o5 102 RFe12Z (Z -- N or C) 367, 378
PdMn 102 R2Fel7Z3 (Z = N or C) 367
PdNi 97 R2M17 459
Pr 83 RNiO3 201,211
PrA12 115, 116 - perovskites 199, 209
PrBa2CuO6 254 RT3 416
PrBa2Cu307 250 RT12 416
PrBa2Cu3Oz 266 R2T7 416
Pr2_zCexCuO 4 226 RnTm intermetallic compounds 521
PrCo2Si2 143, 146 (RY)CO2 527
PrzCuO4 237, 238 (RY)CO2_x Al~ 521
Pr2CuO4_~ 222, 226
Pr2Fe17 317 single crystal ferrite garnet films 450
PrzFeI4B 171, 173
Sm 83
Pr2Fel7C2.5 317 Sm2CuO4_~ 222, 226
PrFelo.5MOl.5N~ 358 SmFe2 364
PrzFe17N2.5 317 SmFe3 364
PrFell Ti 357 SmzFel7 317
PrFellTiN1.5 357 SmzFe14B 173
PrFellTiNy 363 Sm2Fe17C 322
PrNi5 61, 135, 136, 141, 142 Sm2Fe17C2.5 318
PrNiO3 201,211 Sm2Fe17Cy 335, 343
PrRuzSi2 143, 146 SmzFe17Cz.6H1.1 318
SmzFelvCNu 318
rare-earth-Co(Fe) alloys 464 Sm2Fel7CxNy 336, 383
rare-earth-ferrite garnets 409, 414, 416, 422, SmzFel4Ga3C1.5 390
439, 440, 460 Sm2Fe17N2.5 317
RBa2Cu30~ 245 SmzFel7N3.0 317
RCo2 411,418, 521, 527 Sm2Fe17N5.2 317
RCo5 410, 416, 482 SmzFe17Nv 323, 335, 343, 385, 388, 394
R2Co17 410, 482 Sm2Fe17 nitride 313
694 MATERIALS INDEX

Sm7Fe93 nitride 393 TmRh4B4 168


Sm2Fel7N3.0Ho.8 318 TmSe 571
Sm2(Fel_zCox)17N2.7 350 (Tm,Y)Co7B 3 520
Sm2(Fel_zMx)17Ny 382 (Tm,Y2)COll B4 520
SmFeO3 200, 209 (TMTSF)2C104 275, 278
SmloFe79SillNy 366 (TMTSF)2NO3 275, 278
SmFeH Ti 357 (TMTSF)2PF6 275, 276, 278
SmFe11TiCu 357 (TMTSF)2X 274
SmFel 1TiN 357 (TMTSR)2-X, X = C104, NO3, PF6 277
Sm2 (Feo.982Tio.ol8) 17N2.3 352
Sm2(Fel_xVz)17Ny 352 UA12 5, 175, 176
SmFeloV2Nu 360 UAs 176, 178-180, 571
(Sml-zNdz)2(Fel_zCOz)N2.7 353 UAu2Si2 19
SmNiO3 201, 211 UAuSn 25
(Sm1_xR:c)2Fe17Nu 352 U3Au3 Sn4 35
SmRh4B4 166, 172 UBel3 4, 5, 39, 42, 177, 195
Sm3Se4 145, 150, 164 UCdll 5, 177, 193, 194
Sr2CuO3 201, 214 UCo3B2 28, 29
Sr2CuO2C12 214, 222, 226, 236, 242 UConB 28
SrCr8Ga4019 201,215 U(Col_zCuz)2Ge2 21
U(Col_zFe~)2Ge2 21, 37, 38
UCo2Ge2 10, 18, 19
Tb 83, 567, 571
U2Co2In 28
TbCo5 482, 491
U(Coo.25Nio.75)nB 28
Tb-ferrite garnets 471 U(Co0.sNi0.5)4B 28
TbFe3 365 U(Co0.75Nio.25)4B 28
Tb2Fe17 318 U(Col_zNiz)2Ge2 21, 39
Tb2Fe14B 173 UCo2P2 19, 20
TbFel0.sMoi.sNv 359 U3Co3Sb4 35
Tb2Fel7N2.5 318 UCo2Si2 19, 176, 186
TbFeO3 464 U2Co2Sn 28
Tb3 Fe5012 561 UCr2Si2 19
TbFeli Ti 357 UCu5 37, 41, 42, 162, 177, 187, 190, 192
TbFetl TiNo.5 357 UCu4A18 44
TbFelo.sV1.5Ny 360 UCu4+~A18_x 44
TbMn2 119, 127 UCu2Ge2 18, 19
TbNi5 136, 138 UCu2P2 20
(TbY)3Fe5OI2 411,417, 448, UCus_xPd~ 42
- garnets 416 U3Cu3Sb4 34, 35
Th2Fe17Cv 322 UCuzSi 2 19
Tm 567 U3Cu3Sn4 35
TmA12 114-116 UFe2 178
TmFe2 118, 119, 122, 123, 127 UFe3B2 28
Tm2Fe17 319 UFe2Ge2 19
Tm2Fe14B 173 UFezSi2 19
UIn3 176, 182
Tm2Fe17Cy 338, 340, 343
Tm2Fe17C1.oNy 319 U(Ino.sSno.5)3 182
UIr3Br2 28
Tm2Fel7CxNy 338
UIrzGe2 17, 18, 19
Tm2FeI7N2.5 319 UIrzSi2 19
Tm2Fel7N2.7 338 UzIr2Sn 28
Tm2Fel7Ny 343 UMn 2 175, 176
TrnFell Ti 358 UN 176, 182
TmFellTiC~ 358 UNiA1 25
TrnNi5 136, 140 UNi2A13 5, 28, 29, 32, 177, 187, 198
MATERIALS INDEX 695

UNi4B 28, 30, 31 U3Si2 27


UNi2B2C 20 USn3 176, 182
U(Nil_zCuz)2Gez 21 U(Sno.sIno.5)3 176
U3(Nil_xCuw)3Sb4 35 UT3B2 29
UNiGa 26 UT4B 29
UNiGe 25, 38, 39 UT2Ge2 16, 18, 20
UNi~Ge2 19 U3T3Sb4 34, 35
U2Ni2In 28 UT2Si2 16, 19, 20
U3Ni3Sb4 35, 36 UTX 24, 26
U3N!3(Sbl_zSnx) 4 35 UT2X2 17, 19
UNi2Si2 10, 18, 19, 37 U2T2X2 26--28
UNiSn 24, 25 UTe 176, 182
U2Ni2Sn 28 Uo.965Tho.o35Be13 195, 196
U3Ni3Sn4 35 U1_~Th~Be13 162, 177, 187, 189, 195, 196
UO2 571, 574 Ul_~ThzBel3_yBu 177, 195
UOs3B2 28 Ul_xTh~Pt3 177
UOs4B4 28 UxYI_~Pd3 40, 42
UOs2Si2 19 U2ZnI7 5, 37, 177, 187, 194, 195
UP 176, 180, 181
UPd2AI3 4, 5, 28, 31-33, 37, 39, 44, 177, 187, V203 200, 206
199
UPd2Ga3 28
YBa2Cu306 248, 251
UPd2Ge2 19
YBa2Cu306+,~ 247, 252, 262, 263, 269
U2Pd2In 28
YBa2Cu30~ 244-246, 249, 259
U3Pd3Sb4 35 YBa2(Cul_u Fey)30~ 245, 254, 259, 260
UPd2Si2 19 YBa2(CUl_yTu)3Ox 258
UPdSn 25
YBa2(CUl_~Zny)2Ox 245, 254
U2Pd2Sn 28 YBa2(Cul_yZny)307 254
UPt3 5, 10, 12, 35, 39, 42, 162, 177, 186-189, YBa2(Cuo.96Zno.o4)30~ 258, 259
196 YCo2 411, 521, 527
UPt2Ge.2 19 YCo5 493
U2Pt2In 28 Y9Co7 105, 111, 112
U(Ptl_~Pdz)3 188 Y(CoI_zAI:~)2 527
U3Pt3Sb4 35 YCo2Si2 142, 143, 161
UPt2Si2 19, 176, 186 Y2Cu205 201, 215
U2Pt2Sn 28 YFe2 118, 119, 121, 123, 127
uranium 574 YFe3 364
URe2Si2 19 Y2Fe17 319
URhA1 25 Y6Fe23 364
URh2B2C 20, 21 Y2FeI4B 171, 173
URh2Ge2 19 Y2Fe14BNv 365
URhln 25 Y2FeI7C 322
U2Rh2In 28 Y2Fel7C2.5 320
U(Rho.35Ruo.65)2Si2 176, 184, 185 Y2FelTCy 321, 323, 343
URh3Sb4 35 Y2Fe17Cl.2Nu 320
URh2Si2 19, 176, 183, 198 Y2(Fel_~Coz)lTNy 350
U2Rh2Sn 28 Y2FeI7H2.7 343
URu3B2 28 YFelo.sMol.sNu 359
URunB4 28 YFelo.2MOl.sNy 355
U4RuTGe6 18 YFell.sMoo.5N 359
U(RUo.TRho.3)4B4 28 YFell MoNu 355
URu2Si2 4, 5, 10-13, 17-18, 19, 33, 37, 39, 42, YFe9Mo3N 359
142, 177, 187, 197 YFe9Mo3Nv 355
USb 176, 181,571, 574 Y2Fe17N3.1 320
696 MATERIALS INDEX

Y2Fel7Ny 343 Y(Mnl_zFe~)2 119


YFeO3 200, 209 Yl_uPr~Ba2Cu307 253
Y3Fe5012 garnet 417, 418 Yl_yPryBa2Cu3Ox 245, 250
Y3(Fe,Ga)5 O12 417 Yl_uPryCu307 256
(YTb)3FesO12 477 (Yl_zPrx)Ba2Cu307 251
YFellTi 358 Yl_tRt(COl_xAlx)2 527
YFe11TiCy 358 YRh4B4 172
YFellTiNo,5 358 Yo.9Tbo.lMn2 119, 125
YFell TiNv 355 Yo.95Tbo.osMn2 119, 125
YFellTiZy (Z = N and C) 363 Yb--ferrite garnet 450, 466
YFeloVaNv 355 YbBe13 42
YFelo.sV1.sNu 360 YbBiPt 43, 44, 145, 150, 151, 163, 165
Y2Fel7Z3 with Z = H, C or N 368 YbCu5 42, 43
(YI _ t Gdt )(Coo.915Alo.o85)2 530 YbCunAg 5, 43
(YI_ t Gdt)(Coo.95Alo.o5)2 529 YbCu4Au 42, 43
(Yl_tGdt)(Col_~Alz)2 530, 531 YbCu4In 43
- film 451 YbCu4Pd 42
Yl_tGdt(COl_~Alx)2 527, 531 YbCu2Si2 43
Y2.6Gdo.4Fe3.9Gal.lO12 450, 452 Yb2Fe17N2.8 319
Y3_zGd~Fe5O12 garnet 460 YbNi5 136, 141
(Y,Gd,Yb,Bi)3(Fe, A1)5012 453,454, 479 YbNi4In 43
(YR)Co2 418 YbPtSb 44
(YR)(CoA1)2 418 Ybo.sYo.sBiPt 145, 150, 163
YMn2 119, 124-126 ytterbium-ferrite garnet 431
Y(Mnl_zAI~)2 119, 127
Y(Mno.97Feo.o3)2 126 Zno.sFe2.504 567

You might also like