Download as pdf or txt
Download as pdf or txt
You are on page 1of 182

First Principles Study of Electronic and Magnetic Structures in

Double Perovskites

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy
in the Graduate School of The Ohio State University

By

Molly Ball

Graduate Program in Materials Science and Engineering

The Ohio State University

2017

Dissertation Committee:

Professor Wolfgang Windl, Advisor

Professor Maryam Ghazisaeidi

Professor Patrick Woodward


Copyrighted by

Molly Ball

2017
Abstract

At present, electronic devices are reaching their storage and processing limit causing a

major push to find materials that can be used in the next generation of devices. Double

perovskites with A2BB′O6 stoichiometry form one of the leading classes of materials

currently being studied as a potential candidate because of their extremely wide range

and tunability of functional properties, along with economic and highly scalable synthesis

routes. Having a thorough understanding of their electronic and magnetic structure and

their dependence on composition and local structure is the basis for targeted development

of novel and optimized double perovskites. While the body of knowledge and rules

within the field of materials chemistry has enabled many previous discoveries, recent

developments within density functional theory (DFT) allow by now a rather realistic

description of the electronic and magnetic properties of materials and especially

identification of their origin from geometry and orbital structure.

This thesis details computational work based on DFT within several collaborative studies

to better understand the electronic and magnetic properties of double perovskites and

related materials that show promise for future use in multifunctional devices. First, we

will begin with a general introduction to the double perovskite structure, their properties,

and the computational methods used to study them. In the next section, we will look at

the case of the antiferromagnetic, insulating double perovskite Sr2CoOsO6, where

ii
measurements showed that the transition metal ions in the two sublattices undergo

magnetic ordering independently of each other, indicating weak magnetic short-range

coupling and a dominance of longer-range interactions, which has previously not been

observed. Here, we performed DFT calculations to extract the exchange strengths

between the ions and explain this unique dominance of the long-range interactions.

Then, we will look at studies done on thin films of Sr2CrReO6, where our experimental

collaborators found extraordinarily large anisotropy fields and record-breaking strain-

tunable magnetocrystalline anisotropy (MCA). We employed first principles calculations

that examine the dependence of MCA on strain and could identify orbital magnetism on

the Re atoms as the origin of this unique phenomenon.

In the last section, we introduce double perovskites as novel lead-free halide solar cell

materials, with current focus on Cs2AgBiBr6 and Cs2AgBiCl6. While organic Pb based

halides that can be synthesized without expensive clean rooms have achieved within

record time efficiencies that rival that of traditional semiconductor based materials,

creating quite a buzz within the field of photovoltaics, their Pb content and lacking air

stability represented severe roadblocks towards market introduction. Here, we show with

band structure calculations that spin-orbit coupling is a much more dominant interaction

than in traditional semiconductors and thus needs to be considered when designing novel

materials for maximum efficiency. The results of this study have given momentum to

investigate additional halides double perovskites.

Finally, we will summarize and discuss the importance of computational modeling in

order to explore the wide and to date little explored composition space of double

iii
perovskites, one of the currently most promising materials classes for novel devices with

unique and extremely tunable properties.

iv
Dedication

To Tom Moe.

v
Acknowledgments

I would like to first acknowledge my advisor, Professor Wolfgang Windl, for the

guidance he has given me throughout my graduate school career. His patience and

encouragement have been vital to my success. I thank him for introducing me to the

incredible research opportunities I have been able to be a part of through the Center for

Emergent Materials (CEM).

I am grateful for all members in CEM for their interest in my theoretical work and

providing so many chances to participate in unique and interesting collaborations, which

led to many more successful publications than I would have ever been able to achieve

alone. Thank you to Dr. Oscar Restrepo for his willingness to answer the overwhelming

amount of questions that I asked on a weekly basis and always being accommodating

when I asked him to explain portions of density functional theory yet again. I am grateful

for the time you spent passing your knowledge onto me.

I would like to thank my research group members, both past and present, not only for

their insightful academic discussions, but also for wonderful company both in and out of

the office. I am truly lucky to have been able to spend so much time with such incredible

people. I am grateful for the friendships that I have made while in Columbus both in and

out of the Department of Materials Science.

vi
I would finally like to thank my family, who taught me to be my best and encouraged me

to pursue my higher education goals. Their support has been unwavering for my entire

life and I am beyond grateful for that.

vii
Vita

May 2008 .......................................................Charleston High School

May 2012 .......................................................B.A. Physics, Monmouth College

May 2012 .......................................................B.A. Mathematics, Monmouth College

May 2016 .......................................................M.S. Materials Science and Engineering,


The Ohio State University

2012-2013 .....................................................Graduate Fellow, Department of Materials


Science and Engineering, The Ohio State
University

2013 to present ..............................................CEM Scholar, Department of Materials


Science and Engineering, The Ohio State
University

Publications

1. R. Morrow, R. Mishra, O. D. Restrepo, M. R. Ball, W. Windl, S. Wurmehl, U. Stockert, B.


Büchner, and P. M. Woodward. Independent Ordering of two interpenetrating magnetic
sublattices in the double perovskite Sr2CoOsO6, J. Am. Chem. Soc. 135(50), 18824-18830
(2013).

2. A. J. Hauser, J. M. Lucy, M. W. Gaultois, M. R. Ball, J. R. Soliz, Y. Choi, O. D. Restrepo, W.


Windl, J. W. Freeland, D. Haskel, P. M. Woodward, F. Yang, Magnetic structure in
epitaxially strained Sr2CrReO6 thin films by element-specific XAS and XMCD, Phys. Rev.
B. 89, 180402(R) (2014).

3. J. M. Lucy, M. R. Ball, O. D. Restrepo, A. J. Hauser, J. R. Soliz, J. W. Freeland, P. M.


Woodward, W. Windl, F. Y. Yang, Strain-tunable, extraordinary magnetocrystalline
anisotropy in Sr2CrReO6 Epitaxial films, Phys. Rev. B 90, 180401(R) (2014).

viii
4. E. T. McClure, M. R. Ball, W. Windl, P. M. Woodward, Cs2AgBiX6 (X= Br, Cl): New Visible
Light Absorbing, Lead-Free Halide Perovskite Semiconductors, Chemistry of Materials 28 (5),
1348-1354 (2016).

Fields of Study

Major Field: Materials Science and Engineering

ix
Table of Contents

Abstract…………………………………………………………………………………... ii

Dedication………………………………………………………………………………... v

Acknowledgements……………………………………………………………………… vi

Vita………………………………………………………………...…………….…….. viii

Table of Contents………………………………………………………………………… x

List of Tables……………………..……………………………………………………. xiv

List of Figures……………………..…………………………………………………… xvi

Chapter 1. Introductory Note………..…………………………………………………… 1

Chapter 2. Background.……………..…………………………………………………… 3

2.1 Introduction………………………………………………………………....... 3

2.2 Perovskite Structure………………………………………………………….. 4

2.2.1 Octahedral Tilting………………………………………………….. 6

2.2.2 Jahn-Teller Distortion…………………………………………….. 11

2.3 Electronic and Magnetic Structure…………………………………………. 12

2.3.1 Spin-orbit Coupling………………………………………………. 13

2.3.2 Exchange Interactions……………………………………………. 13

Chapter 3. First Principles Method…….……………………………………………….. 17

x
3.1 Introduction…………………………………………………………………. 17

3.2 From Many Electron to Electron Density…………………………………... 17

3.3 Exchange-Correlation in Perovskites……………………………………...... 19

3.3.1 Strongly Correlated Materials…………………………………...... 20

3.3.2 Spin-polarized Calculations………………………………………. 21

3.4 Summary……………………………………………………………………. 21

Chapter 4. Independent Ordering of Two Interpenetrating Magnetic Sublattices in the

Double Perovskite Sr2CoOsO6…………………………………………………………. 23

4.1 Abstract……………………………………………………………………... 23

4.2 Introduction…………………………………………………………………. 24

4.3 Experimental Section……………………………………………………….. 26

4.4 Results and Discussion……………………………………………………... 29

4.5 Conclusions…………………………………………………………………. 44

Chapter 5. Strain-dependent, Extraordinary Magnetocrystalline Anisotropy in Sr2CrReO6

Epitaxial Films..……………………………………………………………………….... 48

Chapter 6. Magnetic Structure in Epitaxially Strained Sr2CrReO6 Thin Films by

Element-specific XAS and XMCD ……………………..……………………………… 66

Chapter 7. Cs2AgBiX6 (X = Br, Cl) – New Visible Light Absorbing, Lead-free Halide

Perovskite Semiconductors ……………………..……………………………………… 84

7.1 Abstract……………………………………………………………………... 84

7.2 Introduction…………………………………………………………………. 85

7.3 Experimental Section……………………………………………………….. 88

7.4 Results………………………………………………………………………. 90
xi
7.4.1 Synthesis……………………………..…………………………… 90

7.4.2 Crystal Structure………………………………………………….. 91

7.4.3 Optical Properties………………………………………………… 96

7.4.4 Band Structure Calculations……………………………………… 98

7.4.5 Chemical and Light Stability…………………………………..... 103

7.5 Conclusions………………………………………………………………... 106

Chapter 8. Summary…………………………………………………………………... 111

Appendix A. Supplementary Information for Chapter 4……………………………… 115

A.1 X-ray Powder Diffraction………………………………………………… 115

A.2 Neutron Powder Diffraction………………………………………………. 116

A.3 Electrical Conductivity Measurements........................................................ 121

A.4 Density Functional Theory Calculations………………………………….. 123

A.4.1 Collinear Magnetic Calculations………………………………... 123

A.4.2 Non-Collinear Magnetic Calculations………………………….. 128

A.4.3 Exchange Interactions…………………………………………... 131

Appendix B. Supplementary Information from Chapter 5……………………………. 138

B.1 Experimental Details of Sr2CrReO6 Film Deposition……………………. 138

B.2 Diamagnetic Background Subtraction in Squid Magnetometry

Measurements….…………………………………………………………..139

B.3 DFT-calculated Structural Parameters for Sr2CrReO6 Films on LSAT, STO,

& SCNO/LSAT……………………………………………………………142

B.4 Atomically Resolved Spin and Orbital Magnetic Moments for SCRO Films

on LSAT, STO, and SCNO/LSAT……………………………………….. 143


xii
B.5 Calculated DOS of Rhenium Suborbitals Varying with Strain…………… 143

Appendix C. Supplementary Information for Chapter 7……………………………… 145

C.1 Electronic Structure Calculations…………………………………………. 145

C.2 Kubelka Munk Plots………………………………………………………. 145

C.3 Light and Moisture Stablility……………………………………………... 146

References……………………………………………………………………………... 148

xiii
List of Tables

Table 4.1 gives information concerning the unit cell and goodness of fit parameters for

each refinement on the longer neutron data sets. Figures and tables containing further

details on the structure of Sr2CoOsO6 at different temperatures can be found in the

supporting information in Appendix A…………………………………………………. 36

Table 5.1 Lattice constants (a, c), tetragonal distortion (c/a), coercivity (Hc) and

magnetocrystalline anisotropy field (Hu) at T = 20 K and 300 K of the Sr2CrReO6 films

grown on LSAT, SrTiO3 and Sr2CrNbO6/LSAT; as is the in-plane substrate lattice

constant (doubled for LSAT and STO for easier comparison)…………………………. 58

Table 7.1 details of the X-ray powder diffraction experiments and Rietveld

refinements……………………………………………………………………………… 93

Table 7.2 Refined atomic positions and atomic displacement parameters for Cs2AgBiCl6

(upper) and Cs2AgBiBr6 (lower, in italics)……………………………………………... 94

Table 7.3 Bond lengths and bond-valence sums………………………………………... 95

Table 7.4 Observed and calculated optical band gaps………………………………… 102

Table A.1 Refined atomic positions and displacement parameters for longer measuring

time data sets where applicable. Os atomic positions for both space groups remain at

xiv
Wyckoff site 2a (0,0,0) while Co atomic positions remain at Wyckoff site 2b (0,0,½) for

both space groups. Ueq is defined as one third of the trace of the Uij tensor, and it is

given for the oxygen atoms for which anisotropic displacement parameters were

refined…………………………………………………………………………………. 117

Table A.2 Select bond lengths and angles for Sr2CoOsO6 as obtained from Rietveld

refinements of the neutron diffraction data……………………………………………. 118

Table A.3 Relative energy of different magnetic configurations examined in this work,

along with the spin and orbital-moments of Co atoms and OsO6 units obtained for each

configuration…………………………………………………………………………... 126

Table A.4 Spin exchange parameters J ijeff = Si S j J ij for Sr2CoOsO6 (in meV)………... 134

Table A.5 Comparison of energy per f.u. of the different magnetic configurations

calculated DFT and with the Heisenberg model………………………………………. 135

Table B.1 DFT-calculated rotation angles of the Cr and Re oxygen octahedra as well as

equatorial (deq) and axial (dax) lengths of the Cr–O and Re–O bonds for the Sr2CrReO6

films grown on LSAT, SrTiO3 and Sr2CrNbO6/LSAT………………………………... 142

Table B.2 First principles calculated in-plane (average of [100] and [010] and out-of-

plane ([001]) spin and orbital magnetic moments for Cr, Re, and O atoms in SCRO films

of different strain states………………………………………………………………... 143

Table C.1 An HSE hybrid functional with spin orbit coupling was used for the

computations…………………………………………………………………………... 145

xv
List of Figures

Figure 2.1 Corner sharing BO6 octahedra create a cage in which the A cation sits;

forming the perfect cubic structure for a) a ternary perovskite and b) a double

perovskite………………………………………………………………………………… 6

Figure 2.2 Low values of the tolerance factor t will lower the symmetry of the crystal

structure. GdFeO3 have tilted octahedra and crystallize in the orthorhombic system. If the

tolerance factor t > 1 as for BaNiO3 due to a large A or a small B ion then hexagonal

variants of the perovskite structure form. In LnMnO3 the octahedra around Mn are Jahn–

Teller distorted…………………………………………………………………………… 8

Figure 2.3 Two possible distortions of an octahedron that remove the degeneracy of

unequally occupied eg orbitals for a high spin d4 ion; (right) elongation of the bonds

along z, and (left) elongation of the bonds in the x-y plane. In both instances the

remaining bonds contract to maintain the same degree of metal–ligand bonding……… 12

Figure 4.1 The A2MM′O6 double perovskite structure. The large A cations are shown in

gray, while the M and M′ centered octahedra are shown in red and blue respectively… 25

Figure 4.2 Field cooled (filled circles) and zero field cooled (open circles) magnetic

susceptibility with inverse zero field cooled data plotted against the right axis showing

the linear Curie-Weiss fit……………………………………………………………….. 31

xvi
Figure 4.3 Specific heat of Sr2CoOsO6 as a function of temperature………………….. 32

Figure 4.4 Log resistivity vs temperature……………………………………………… 33

Figure 4.5 Evolution of the (a) lattice parameters, (b) Os–O bond lengths, and (c) Co–O

bond lengths as a function of temperature, as extracted from variable temperature neutron

diffraction data. The dashed line separates the tetragonal I4/m phase from the monoclinic

I2/m phase………………………………………………………………………………. 34

Figure 4.6 Total- and atom- resolved density of states of Sr2CoOsO6 from a non-collinear

DFT calculation with UCo = 4.1 eV and UOs = 2.1 eV. The contributions of the Os 5d and

Co 3d orbitals are shown in red and blue, respectively………………………………… 37

Figure 4.7 Rietveld fits to the high d-spacing regions of the neutron diffraction patterns

showing the presence/absence of magnetic reflections at three temperatures: above TN1

(130 K), between TN1 and TN2 (80 K), and below TN2 (12 K). Black symbols represent

observed data while the red curves represent the calculated patterns. The hkl values given

in the figure are indexed on the nuclear cell……………………………………………. 38

Figure 4.8 (a) Experimentally determined magnetic structure of Sr2CoOsO6, with the Os

shown in red and the Co in blue. (b) A schematic of the superexchange pathways between

Co and Os ions with the exchange constants extracted from the DFT calculations. A

positive (negative) value of J indicates ferromagnetic (antiferromagnetic) coupling….. 41

Figure 5.1 XRD scans around the Sr2CrReO6 (004) (left panels) and Sr2CrReO6 (022)

(right panels) peaks for 90-nm Sr2CrReO6 films grown on (a) LSAT, (b) SrTiO3 and (c)

Sr2CrNbO6 buffer layer on LSAT. The off-normal Sr2CrReO6 (022) peaks are measured

at a tilt angle Ψ =45° for in-plane characterization……………………………………... 52

xvii
Figure 5.2 Schematics of (a) Cr−O (counterclockwise) and Re−O (clockwise) octahedral

rotations in the Sr2CrReO6 lattice and (b) an oxygen (red) octahedron surrounding Cr or

Re with two different bond lengths, deq and dax. (c) Octahedral rotation angle θ and (d)

bond lengths of Cr−O and Re−O octahedra as a function of the tetragonal distortion c/a of

the Sr2CrReO6 lattice…………………………………………………………………… 54

Figure 5.3 In-plane and out-of-plane magnetic hysteresis loops for the Sr2CrReO6 films

grown on (a) LSAT, (b) SrTiO3 and (c) Sr2CrNbO6/LSAT in magnetic fields up to 7 T at

T = 20, 100 and 300 K. Magnetic anisotropy fields can be obtained from the intercept of

linear extrapolation of the hard-axis hysteresis loop and the saturation magnetization as

shown in the left panels…………………………………………………………………. 56

Figure 5.4 Differences (∆m) between out-of-plane ([001] direction and in plane (average

of [100] and [010] directions (a) total magnetic moments, (b) total orbital and spin

magnetic moments, and (c) spin and (d) orbital magnetic moments for Re (squares), Cr

(diamonds), and O (circles). The lines are guides for the eye.…………........…………. 60

Figure 6.1 In-plane magnetic hysteresis loop at T = 200 K for films on (001)-oriented

LSAT (green circles), SrTiO3 (black squares), and ~200 nm Sr2CrNbO6 buffer layer on

LSAT (blue triangles)…………………………………………………………………... 70

Figure 6.2 Normalized XANES and XMCD spectra at the (a-c) Cr L, (d-f) Re L, and (g-i)

O K edges. XANES/XMCD spectra are on top/bottom within each figure. Spin and

orbital moments are given for the Cr and Re spectra…………………………………… 73

Figure 6.3 Normalized x-ray absorption spectra for different strains at the (a) Cr L2,3

edges and (b) Re L2 edge……………………………………………………………….. 75

xviii
Figure 7.1 Rietveld refinements of the XRPD patterns for Cs2AgBiCl6 (upper) and

Cs2AgBiBr6 (lower). The black dots, red and green lines are the experimental pattern,

calculated fit, and difference curve, respectively. Red tick marks at the bottom give the

expected peak positions. The inset shows the fit to the (111) and (200) reflections…… 92

Figure 7.2 Refined crystal structure of Cs2AgBiCl6. The Cs+ ions are shown as gray

spheres, the chloride ions as small green spheres, while the Ag and Bi centered octahedra

are shown as blue and green polyhedra, respectively…………………………………... 95

Figure 7.3 Diffuse reflectance spectra for Cs2AgBiCl6 and CH3NH3PbCl3 (top), and

Cs2AgBiBr6 and CH3NH3PbBr3 (bottom)……………………………………………… 97

Figure 7.4 Band Structure diagrams for Cs2AgBiCl6 (top) and cubic CsPbCl3 (bottom).

The Fermi energy is set to E = 0 and denoted with a dashed line……………………… 99

Figure 7.5 Band Structure diagrams for Cs2AgBiBr6 (top) and cubic CsPbBr3 (bottom).

The Fermi energy is set to E = 0 and denoted with a dashed line…………………...…101

Figure 7.6 Atomic partial density of states plots for Cs2AgBiCl6 (top) and Cs2AgBiBr6

(bottom)………………………………………………………………………………... 103

Figure 7.7 UV-Vis diffuse reflectance spectra showing the light stability of Cs2AgBiCl6

after two and four weeks of light exposure……………………………………………. 105

Figure 7.8 UV-Vis Diffuse spectra showing the light instability of Cs2AgBiBr6 after two

and four weeks of light exposure……………………………………………………… 106

Figure A.1 Refined powder X-ray diffraction pattern of Sr2CoOsO6. Data was collected

using a Bruker D8 Advance equipped with a copper source and Ge (111)

monochromator. Black symbols observed data, while red and blue curves indicate the

calculated pattern and difference respectively………………………………………… 115


xix
Figure A.2 Rietveld refinement histograms at (a) 300 K and (b) 12K where black

symbols represent observed data points, red curves for the refined fit and blue for the

difference curve. Pink hash marks indicate reflections from the vanadium sample can,

black tick marks indicate reflections from the nuclear structure, and green tick marks

correspond to magnetic reflections……………………………………………………. 119

Figure A.3 Colorfill plot of observed variable temperature NPD data revealing the

splitting of the (220) peak which is associated with the structural phase transition

lowering the symmetry of the structure from tetragonal I4/m to monoclinic I2/m at

approximately 110K…………………………………………………………………… 120

Figure A.4 The structure of Sr2CoOsO6 at 12 K in space group I2/m. The view looking

down [001] on the left shows out-of-phase octahedral rotations about the c-axis, while the

view looking down [110] on the right shows the lack of octahedral tilting about the a- or

b-axes (right). The higher temperature I4/m structure has an identical tilting

pattern............................................................................................................................. 120

Figure A.5 Electrical transport behavior of Sr2CoOsO6 as shown by log resistivity vs

temperature……………………………………………………………………………. 122

Figure A.6 16 formula-unit cell of Sr2CoOsO6 showing (a) the experimentally observed

magnetic configuration which is non-collinear and (b) the collinear version of the

experimentally observed magnetic configuration where all the moments have been

aligned along the [001] direction. Only Os (red) and Co (blue) atoms are shown for

clarity. The arrows show the direction of the magnetic moment on every Co and Os

ion……………………………………………………………………………………... 125

xx
Figure A.7 Spin- and atom- resolved density of states with collinear arrangement of

independently ordered Co and Os sublattices obtained using (a) GGA without Hubbard-U

and (b) with UCo = 4.1 eV and UOs = 2.1 eV………………………………………….. 125

Figure A.8 Spin- and orbital-resolved density of states of (a) Co and (b) Os in the 16

formula-unit cell of Sr2CoOsO6 with collinear arrangement of independently ordered Co

and Os sublattices obtained using UCo = 4.1 eV and UOs = 2.1 eV…………………… 127

Figure A.9 The total DOS for the 16 f.u. cell with (a) Exp(non_col), (b) Exp(col)+SOC

and (c) Exp(col) magnetic configurations……………………………………………... 131

Figure A.10 Schematic showing the different exchange interactions between Co and Os

ions…………………………………………………………………………………….. 133

Figure A.11 Schematic showing the spin arrangements in two neighboring ab planes in

Sr2CoOsO6…………………………………………………………………………….. 136

Figure B.1 Raw magnetic data for the 90 nm SCRO/LSAT film taken with a Quantum

Design SQUID for both in-plane and out-of-plane geometries at T = 20 K. The hard axis

loop (field applied perpendicular to the film) shows almost no signal differentiating the

SCRO film from the background of the LSAT substrate, so a reliable anisotropy field

analysis could not be performed. We estimate Hu to be 10's of tesla…………………. 140

Figure B.2 Raw magnetic data for 90-nm SCRO/STO film taken with a Quantum Design

SQUID magnetometer for both in-plane and out-of-plane geometries at T = 20 K. The

(linear) diamagnetic background of the substrate is shown…………………………… 141

Figure B.3 Calculated DOS of Re t2g (a) xy and (b) yz and xz suborbitals near the Fermi

energy in SCRO for tetragonal distortions c/a = 1.025 and c/a = 0.99………………... 144

xxi
Figure C.1 Kubleka-Munk plots for the silver-bismuth (blue) and lead compounds (red),

with chloride compounds on the left and bromide compounds on the right…………... 146

Figure C.2 XRPD overlays of Cs2AgBiCl6 (upper) and Cs2AgBiBr6 (lower) showing the

light stability of the compounds after a month of exposure to sunlight. The change in the

powder diffractogram for the bromine compound indicates that structural changes are

accompanying the visual darkening of the pigment…………………………………... 147

xxii
Chapter 1. Introductory Note

This dissertation is organized in the form of chapters, which have been written in varying

styles depending on the peer-reviewed journal they were initially published. Since all

work was done in collaboration with other groups involving graduate students, I will

report my work outside of the general introductory chapters and summary in their

original published form, with exact description of my contributions, in order to avoid

problems with overlap with the theses of my experimental collaborators while still being

able to discuss the interplay between experimental and theoretical results that was at the

heart of my research and that enabled the advances in materials understanding and design

described below.

In chapter 2, I will provide a general overview on the perovskite structure, as well as

properties associated with the electronic and magnetic structure. In chapter 3, I will

discuss the methods used to study the electronic structures of these materials. I will

give a basic overview of density functional theory (DFT) and review some of the

assumptions that are vital for this work. Chapter 4 is a journal article published in

the Journal of the American Chemical Society and will discuss the unique magnetic

structure of the two independently ordering transition-metal lattices in Sr2CoOsO6.

The focus of chapters 5 and 6 is magnetocrystalline anisotropy and magnetism

1
observed in Sr2CrReO6 both in verbatim form as published in Physical Review B. In

chapter 7, I present work done on the lead-free halides, Cs2AgBiBr6 and Cs2AgBiCl6, as

published in Chemistry of Materials. Finally in chapter 8, I will summarize the

conclusion from the previous chapters and provide an overall synthesis of my work.

2
Chapter 2. Background

2.1 Introduction

The limitations of conventional electronics, which are based on the charge of electrons,

are being reached more and more as devices become smaller. In order to explore devices

that go beyond the dogma of Moore’s law and incorporate materials and mechanisms

outside of charge-switching in silicon, more and more device classes have been studied in

the past decade or two. This includes enhanced research in III-V compound

semiconductors, where the wurtzite-based nitrides have already created a revolution in

lighting, which is now dominantly based on LED-based devices. Beyond that, more

complex materials classes further away from the diamond structure are explored, which

not only allow enhanced tunability, but also may allow exploitation of spin-based

operation in addition to or as replacement of charge-based functionality, which is

typically called spintronics. This requires materials with non-trivial magnetic properties,

such as for example half-magnetism, where one spin channel in a material has a band

gap, which the other one is metallic, resulting in a material where 100% of the

conduction electrons have the same spin – something that is not possible in traditional

simple metals or semiconductors. Among the materials classes that allow half metallicity

while at the same time provide large tunability for the sum of all functional properties,

3
the double perovskites is arguably one of the most prominent ones, if not the front runner.

Many composition regimes within double perovskites are to date unexplored, and most

every new regime that is investigated for the first time provides new, exciting and

unexpected findings, making this a highly rewarding field to study. In this dissertation, I

have studied the electronic and magnetic structure and their effects on the properties of

double perovskites.

2.2 Perovskite Structure

Ternary perovskites with stoichiometry ABO3, where A is usually a group two metal and

B is a transition-metal, can ideally be described as corner sharing BO6 octahedra created

by oxygen atoms on the face centers of a cubic unit cell with a B atom at its center and

the A atoms sitting outside the octahedra on the corners of the cell [1]. Figure 2.1a shows

a ternary perovskite supercell. Perovskites have been thoroughly investigated due to their

vast range of physical properties, including ferroelectricity, piezoelectricity,

superconductivity, magnetoresistance, and ionic conductivity. A commonly used example

is SrTiO3, which demonstrates good insulating properties [2] and is often employed as a

substrate for film growth. It retains its cubic structure at room temperature.

Able to accommodate a large number of different metallic elements on the A and B sites,

the perovskite structure is very versatile [3]. However, its idealized cubic form is only

stable for certain size ratios of the A and B cations, i.e. when twice the B−O bond length

equals the cell edge and twice the A−O bond distance the face-diagonal of the cube. Thus,

unlike the previously mentioned cubic perovskite, SrTiO3, the vast majority of

perovskites are not able to stably remain cubic in structure. An example for this is

CaTiO3, where the Ca cation is smaller than Sr. The resulting size mismatch can be
4
accommodated by a rotation of the octahedral, which results in the structure becoming

orthorhombic rather than cubic [4]. This flexibility of the structure is not only interesting

from a crystallographic standpoint, but can also have significant implications for the

electronic and magnetic properties. This type of distortion along with the Jahn-Teller

distortion, which occurs within the octahedra and lowers the energy of the structure by

reducing its symmetry, will be discussed below in Sec.2.2.1 and 2.2.2.

While these distortions can lead to significant changes in the functional properties of the

perovskites, even larger tunability can be achieved by expanding the structure to include

two different transition metal cations, resulting in a stoichiometry of A2BB′O6. This

works since the majority of the functional properties associated with perovskites stem

from the BO6 octahedra. The resulting quaternary compounds are referred to as double

perovskites and have ideally alternating B and B′ octahedra in rocksalt ordering along all

crystallographic directions, as seen in Figure 2.1b. A few examples of these properties

that double perovskites can possess beyond what is possible in simple perovskites include

magnetic ordering with high Curie temperatures (well above room temperature), high

spin polarization, and large anisotropy fields [5].

5
Figure 2.1 Corner sharing BO6 octahedra create a cage in which the A cation sits;
forming the perfect cubic structure for a) a ternary perovskite and b) a double perovskite

Double perovskites also undergo the same types of distortions as ternary perovskites to

accommodate certain ion substitutions. Nearly 70% of them are known to have structures

where octahedral tilting is present [6]. Even the slightest decrease in symmetry in and

around the octahedra can have significant impact on the electronic and magnetic

structure, as our results in the subsequent chapters show. Overall, these compounds have

remained relatively unexplored in comparison to ternary perovskites.

2.2.1 Octahedral Tilting

As previously stated, perovskites are able to rotate their octahedra from the perfect cubic

structure in order to optimize the local environment around the A site cation. When a

perovskite undergoes this type of tilting, there is a change in the overlapping of orbitals

seen between that of the transition metal d-states and the oxygen 2p-states. Often times,

6
the unique properties associated with this family of materials originate from this and will

be discussed later on in this thesis.

Goldschmidt Criterion

One way to gauge how “perfect” a perovskite will be is the Goldschmidt tolerance factor

[7]. In an ideal cubic perovskite, the lattice constant can be geometrically linked to the

size of the atomic radii (rA, rB, and rO). From this, the tolerance factor (t) can be derived,

seen in equation 2.1.

(rA + rO )
t= [2.1]
2(rB + rO )
Figure 2.2 shows examples of possible structures with different amounts of distortion.

For a cubic perovskite, such as SrTiO3, the tolerance factor will be approximately 1 and

at large deviations from this, the cubic structure becomes more likely to be unstable. In

the case of a tolerance factor being t < 1, as it is with CaTiO3 and GdFeO3, the crystal

will adopt an orthorhombic structure [3,4]. When the structure experiences a tolerance

factor t > 1, a hexagonal variation of the perovskite structure can form which leads to the

sharing of octahedral faces as seen the compound BaNiO3.

7
Figure 2.2 Effect of the Goldschmidt tolerance factor t in the resulting perovskite
structure. Low values t < 1 such as in GdFeO3 will lower the symmetry of the crystal
structure by octahedral tilting, resulting in an orthorhombic crystal structure. Large
values t > 1 as in BaNiO3 result in hexagonal variants of the perovskite structure. On top
of the size effect, electronic states can shift and lower the total energy by a symmetry-
breaking Jahn-Teller distortions as shown for LnMnO3 [3].

The Goldschmidt tolerance factor is not only applicable to ternary perovskites. It can also

be modified and applied to double perovskites. In the case of fully ordered stoichiometric

double perovskites the Goldschmidt tolerance factor can be calculated by the following

equation 2.2:

(rA + rO )
t=
r r [2.2]
2( B + B!! + rO )
2 2

8
Now the equation is adapted to account for the effect that both of the B site atoms have

on the structure. Calculating the tolerance factor helps to evaluate the presence and

magnitude of octahedral tilting that a structure experiences, which is vital in

understanding the effect that distortions can have on materials.

The use of a tolerance factor has also been used on mixed A-site perovskite compounds.

It was shown that Sr2FeMoO6 has almost a perfect structure with a calculated tolerance

factor of 0.990 when using experimentally found ionic radii [8]. Kim et al. proposed that

by substituting at the A-site, the tolerance level could be brought closer to 1, which could

enhance the low-field room-temperature magnetoresistance. They were successful in their

pursuits with optimal doping yielding the compound Sr0.2Ba0.8Fe0.5Mo0.5O3. Popov and

Greenblatt also studied how optimizing the tolerance factor could be applied to double

perovskites with mixed A-site cations. [7]. Their work with Ba2-xSrxMnReO6 showed that

optimizing the structure to have a tolerance factor of 1 would maximize the saturation

moment. They found that this was indeed true when the compound was about 36% Sr and

64% Ba. The lowest saturation moment was found when the compound had no Ba

present. This was attributed to the lowering of symmetry from the cubic structure. It is

interesting to see that although the tolerance factor was originally used to predict the

stable crystal structure of a perovskite, it can also help to optimize properties. The

simplicity of the Goldschmidt tolerance factor makes it a popular tool used for

perovskites, but of course, there are many other aspects of the structure that come into

play when investigating properties, which in the end make DFT calculations a very

valuable tool for full-scale exploration as discussed below.

9
Glazer Notation

While the Goldschmidt tolerance factor is able to detect the presence of octahedral tilting,

it is too simple to also describe the exact nature of the tilting. In order to do that, one can

use the Glazer notation, which was first developed in 1972 [9]. Glazer notation is a

scheme that relates the crystallographic lattice to the distortion of the octahedra and has

become the standard by which these types of distortions are described - simplifying the

usual symmetry analysis. There are three tilting characteristics that Glazer classification

describes:

• Which axis is experiencing the tilt, which is denoted by the corresponding

crystallographic direction

• The degree of tilting in the octahedra that occurs specific to one axis compared to

the other axes

• Whether subsequent layers of octahedra are tilted in-phase or out-of-phase with

one another

A group of three letters that are combinations of a, b, and c are assigned for each tilt

system, the order of which corresponds to the x, y, and z axes, respectively. Different or

equal letters correspond to equal or different degrees of tilt. For example, a+a+b+ would

represent a crystal with equal rotation along the x and y axes, but a different magnitude of

tilt along the z axis. Glazer notation also expresses whether adjacent layers are in-phase

(rotated in the same direction), out-of-phase (rotated in opposite directions), or if they are

not rotated at all. These are represented by a superscript; (+) in-phase, (-) out of phase, or

(0) not rotated.


10
Glazer notation can also be utilized when describing double perovskites. As stated before,

around 70% of all double perovskites undergo tilting. Nearly all of these distorted

structures (~97%), can be group within 5 tilt systems, Fm3m (a0a0a0), I4/m (a0a0c-), R3

(a-a-a-), I2/m (a0b-b-) and P21/n (a-a-b+) [11].

2.2.2 Jahn-Teller Distortion

Another common type of distortion present in perovskites is Jahn-Teller distortions.

These distortions are dependent on the electronic structure of the system. The Jahn-Teller

theorem predicts that “any non-linear molecular system in a degenerate electronic state

will be unstable and will undergo distortion to form a system of lower symmetry and

lower energy, thereby removing the degeneracy.” [12] Because of that, the transition-

metal octahedra will experience either elongation or compression (Figure 2.3), which in

turn have an effect on the amount that the orbitals overlap. In an undistorted octahedron,

the d-states of a transition-metal ion split into three t2g states, which are lower in energy

since they do not point toward the oxygen, and two eg states, which are higher in energy

since they do point towards the oxygen. This splitting is due to the repulsive nature of

like charges between the ligand and the d-orbitals. When these states are partially filled,

the states will move either higher or lower in energy depending on where the electrons

sit. An example of Jahn-Teller distortion can be seen in LaMnO3, which is pictured in

Figure 2.2

11
Figure 2.3 Two possible distortions of an octahedron that remove the degeneracy of
unequally occupied eg orbitals for a high spin d4 ion; (right) elongation of the bonds along
z, and (left) elongation of the bonds in the x-y plane. In both instances the remaining
bonds contract to maintain the same degree of metal–ligand bonding [13].

2.3 Electronic and Magnetic Structure

This section of the dissertation outlines some of the important aspects of electronic and

magnetic structure of double perovskites. The electronic structure of double perovskites

is largely influenced by the B-site cations. Due to the chemical flexibility, or ability to

interchange different ions in the system, and variations in oxidation states of these

cations, there is a wide range of possible properties that double perovskites can take on,

including semiconducting, metallic, half-metallic, dielectric, ferroelectric, thermoelectric

and even superconducting [16]. Double perovskites also exhibit an extensive amount of

magnetic structures including antiferromagnetic, ferromagnetic, or ferrimagnetic from the

12
various combinations of elements. Many are also known to exhibit high Tc values and

experience long range magnetic ordering.

The d-states of transition metal oxides generally overlap that is small with O 2p states,

resulting in localized d-electrons causing strong Coulomb repulsion between them and

the narrowing of bands. These are known as strongly correlated materials and such

materials have inter-atomic interactions (or bandwidth (W)) that are equivalently

balanced with electron-electron interactions (or Coulombic correlations (U)). Corrections

for the Coulombic interactions will be discussed in the next chapter.

2.3.1 Spin-orbit Coupling

Also competing to be a dominant force in electronic structure is the spin-orbit coupling.

This interaction occurs between an electron’s spin and the magnetic field generated by

the orbital rotation of the electron through a varying electrostatic potential around the

nucleus. This coupling between the spin and orbital movement causes shifts in an

electron’s energy levels. These energy shifts can cause dramatic changes in the electronic

structure. In 4d and 5d transition metals, SOC is quite strong. Its magnitude is

proportional to the fourth power of the effective nuclear charge Z. With heavier ions, the

need to consider the effects of spin-orbit coupling becomes more important. This is true

in the case of many double perovskites where an extra spin-orbit coupling term needs to

be included into calculations, which will be discussed in future chapters.

2.3.2 Exchange Interactions

As previously stated, the role that distortion of a perovskite plays on the electronic

structure and magnetic structure is an important one as it helps to govern the degree of

interaction between each atom. In perovskites, there is some covalency present between
13
the transition-metals and the oxygen ligand, which results in a sharing of valence

electrons. Consequently, this sharing can result in strong “exchange” interactions. The

magnetic ordering of transition-metal ions is facilitated by the exchange across the

ligand.

There are several types of dominant exchange interactions that can be seen in the

transition metals of double perovskites. Direct exchange, which occurs between

neighboring magnetic ions, is ferromagnetic and extremely weak in double perovskites

due to the distance between transition metals, which results in no overlap between their

electronic wavefunctions. Unlike direct exchange, double exchange requires a direct

transfer of electrons and is an indirect coupling through the oxygen ligand [17]. It is also

ferromagnetic in nature. A third type of exchange interaction is called superexchange.

The electrons present in superexchange are localized unlike the delocalized electrons

seen in double exchange. Also in contrast to the hopping of an electron seen in double

exchange, superexchange occurs when valence electrons are shared between the

transition metals and oxygen ligand [18]. This covalency from shared electrons between

the transition metal and oxygen results in magnetic ordering between the neighboring

transition metals. This ordering can be ferromagnetic or antiferromagnetic and is

determined by the Goodenough-Kanamori rules, a set of rules that takes into account

symmetry and orbital occupancy of the transition metals involved to determine the

coupling between two ions [18]. The strongest interactions are usually between the

nearest- and second-nearest-neighbors and the strength decreases rapidly as the distance

increases. Several of these rules are important to the ordering of insulating perovskites:

14
• The coupling of ions with half-filled orbitals will have antiferromagnetic

moments with respect to each other when the B-O-B bonds are ~180 ͦ

• The coupling of ions with half-filled or empty, or half-filled and fully occupied

orbitals will have ferromagnetic moments with respect to each other when the B-

O-B’ bonds are ~180 ͦ

• The coupling of ions will prefer to have ferromagnetic moments with respect to

each other when B-O-B’ bonds are ~90 ͦ

For the majority of insulating transition metal oxides, the Goodenough-Kanamori rules

provide an accurate prediction when determining their magnetic ordering, which is

extremely helpful to scientists in their search for new magnetic materials. However, there

are exceptions to the rules, and sometimes variations in orbital overlap from the idealized

cases used for derivation lead to what looks like a breaking of the rule, as in the case of

Sr2CoOsO6, which will be discussed in Chapter 4.

References

1. Piskunov, S., Heifets, E., Eglitis, R. . & Borstel, G. Comput. Mater. Sci. 29, 165–
178 (2004).

2. Van Benthem, K., Elsa sser, C. & French, R. H. J. Appl. Phys. 90, 6156 (2001).

3. Johnsson, M. & Lemmens, P. J. Phys. Condens. Matter 20, 264001 (2008).

4. Woodward, P. M. Acta Crystallogr. Sect. B Struct. Sci. 53, 32–43 (1997).

5. Lucy, J. M. et al. 000400, 1–6 (2014).

15
6. Lufaso, M. W. & Woodward, P. M. Acta Crystallogr. Sect. B Struct. Sci. 60, 10–
20 (2004).

7. Popov, G., Greenblatt, M. & Croft, M. Phys. Rev. B 67, 024406 (2003).

8. Kim, B. G., Hor, Y. S. & Cheong, S. W. Appl. Phys. Lett. 79, (2001).

9. Glazer, A. Acta Crystallogr. Sect. B Struct. Sci. B 28, (1972).

10. Mishra, R., Soliz, J. R., Woodward, P. M. & Windl, W. Chem. Mater. 24, 2757–
2763 (2012).

11. Lufaso, M. W., Barnes, P. W. & Woodward, P. M. Acta Crystallogr. Sect. B


Struct. Sci. 62, 397–410 (2006).

12. Housecroft, C. & Sharpe, A. G. Inorganic Chemistry. (Prentice Hall, 2008).

13. Woodward, P. M., Karen, P., Evans, J. S. . & Vogt, T. Functional Materials.
(Unpublished).

14. Mishra, R., Restrepo, O. D., Woodward, P. M. & Windl, W. Chem. Mater. 22,
6092–6102 (2010).

15. Popov, G., Greenblatt, M. & Croft, M. Phys. Rev. B 67, 024406 (2003).

16. Vasala, S. & Karppinen, M. Prog. Solid State Chem. 43 1-36 (2015).

17. Zener, C. Phys. Rev. 82, (1951).

18. Kanamori, J. J. Phys. Chem. Solids 10, 87–98 (1959).

16
Chapter 3. First Principles Methods

3.1 Introduction

In this chapter, I will discuss the basic fundamental elements of density functional theory

(DFT), which was the method used to do all electronic structure calculations within this

thesis. I will also discuss specific approximations and material-specific modifications

used.

3.2 From Many Electrons to Electron Density

To better understand the influence of distortions on the electronic structure and the

interactions that occur, I have employed computational modeling based on DFT. It is

impractical to solve a many-body problem, which depends on all of the electron

wavefunctions in the system. DFT is able to map a many-body system into a more

manageable electron density approach that is based on a system of single electrons

interacting with a system of ions and an effective field created by the electron density,

where the neglected many-body (correlation) and exchange effects are included by

suitable approximations, which are constantly improved by ongoing research. By now,

DFT has proven a vital tool in understanding properties of complex materials.

In order to do electronic structure calculations for groups of charged particles, the

Schrödinger equation, which is show in equation 1, needs to be solved.

17

− !! ∇! 𝜓 + 𝑉𝜓 = 𝐸𝜓 (1)

Where the first term is the kinetic energy of the electron of mass, m, 𝜓 is the

wavefunction, and E is the energy. As this is numerically impractical for the full many-

body system, approximations need to be made as mentioned above. Firstly, since the

mass of nuclei is much heavier than that of the surrounding electrons, it can be assumed

that the electrons adjust instantaneously to the motion of the nuclei, and thus the motion

of nuclei and electrons in a system can be separated into independent equations. Within

this so-called Born-Oppenheimer approximation, the time-independent Schrödinger

equation is solved for only the electrons while it is assumed that the nuclei are fixed at a

given location. Also, periodic boundary conditions are frequently applied to considerably

reduce the amount of atoms present in the computational unit cell in order to calculate

bulk properties. Even with that, especially for heavier atoms, there are still a significantly

large number of electrons present in the system. In order to reduce this number, the

explicit calculation is often restricted to valence electrons, while the core electrons are

either treated within an effect “pseudopotential” or within the projector-augmented wave

method [1].

In 1964, Hohenberg and Kohn proved that, for a given external potential, the total energy

of a system is a functional of the ground state electron density [2]. Although this was an

important simplification, the problem of obtaining the full electron density still remained.

In 1965, Kohn and Sham were able to remap the many-body interacting system into

terms of non-interacting electrons to produce the same electron density as the interacting

18
system [3]. While working functionals could be found for most interaction terms, the

functional dependence of the kinetic energy on the electron density has been elusive. In

order to work around this problem, single-electron wavefunctions are introduced, whose

summed up square modulus is set equal to the system’s electron density. For

wavefunctions, the kinetic energy can be easily calculated within the standard

Schrödinger equation.

In practice, the wave functions within a DFT run are determined from an initial

assumption for the electron density (e.g. the superposition of atomic electron densities

which are known), and then a new, updated electron density is calculated from the

resulting wave functions. This procedure is iterated until input and output electron

density agrees within a chosen convergence criterion. Despite the many approximations

that are made, DFT is a widely used and trusted computational method due to its

versatility and accuracy of results.

3.3 Exchange-Correlation in Perovskites

Since the exact nature of the exchange-correlation functional is unknown in DFT

calculations, researchers are still working on continuously developing improved

approaches for it. A simple approximation for the exchange-correlation energy is to

consider the behavior of an electron to be similar to that of a homogeneous electron gas

and use it locally for the system under consideration. This is known as the local density

approximation (LDA) [4]. Although LDA works well for systems with slowly varying

densities, it does not always perform well, especially when it comes to functional

properties such as band structure or magnetic moment calculations, which is often a

problem when doing calculations for double perovskites. In addition to that, the
19
simplified uniform electron density approximation leads to an overestimation of the

cohesive energy, leading in general to underestimating the lattice parameters in materials.

This over-binding behavior can be corrected with different exchange-correlation

functionals that do not just assume dependence on the local electron density, but also on

the local gradient of density. This approach is known as the Generalized Gradient

Approximation (GGA) and is the method used for the majority of this work. GGA tends

to be an improvement from that of LDA, but it sometimes overcorrects for the binding,

leading to larger lattice parameters than experimentally measured.

3.3.1 Strongly Correlated Materials

When transition metal oxides come into play, the system will have usually strongly

localized d-electrons. These localized electrons tend to have very strong electron-electron

many body interactions. These can be problematic when using LDA and GGA, which

significantly underestimate electron localization due to correlation effects and result

many times in extreme discrepancies between experimental and computational results.

In order to adjust for these strong correlation effects an additional term is needed in the

calculations. This is known as the on-site Hubbard U [5]. In a simple picture, it can be

thought of is an energy penalty for the localized d and f- state electrons of the strongly

correlated transition metals when they delocalize. It helps to improve the description of

the ground state with very little extra computational cost. This +U correction can be

included in the previously mentioned LDA or GGA. Incorporating the U shifts energy

levels of occupied orbitals by -U/2 and unoccupied orbitals by +U/2. This provides an

adjustable parameter for strongly correlated materials to better model the physical

behavior of the system, while still keeping computational time reasonable.


20
3.3.2 Spin-polarized Calculations

Still needing to be considered for calculations is the spin of electrons, which is essential

to address the properties of magnetic materials. This is necessary in order to gain an

understanding of the basic mechanisms, which cause magnetism to occur. In order to

incorporate spin dependence into LDA, two electron densities can be introduced, one

describing spin-up, the other spin-down electrons. This approach is known as Local Spin

Density Approximation (LSDA) [6] and works well for the vast majority of systems,

including double perovskites, provided that the spin interactions are well described by a

suitable approximation for the exchange-interaction term.

3.4 Summary

The flexibility of the perovskite structure opens itself up to the incorporation of numerous

different elements in different combinations. Many of the structures that are possible

have been unexplored and could lead to even more interesting properties. As techniques

in synthesis and growth improve, the list of available structures to investigate increases.

Since the functional properties of perovskites and double perovskites depend sensitively

on the exact geometry of the structure, which in turn is a function of the oxidation states

and size of the elements involved, examining structural distortions and how they affect

the underlying electronic and magnetic structure is key in gaining a better understanding

of these novel materials.

In order to be able to guide synthesis and explore novel composition systems, we have

shown that computational work can be an indispensable asset towards faster and more

targeted synthesis efforts. As the currently dominant approach, we have introduced the

framework of DFT and briefly discussed its physical framework. Finally, I would like to
21
mention that although DFT is a very useful tool in understanding the phenomena seen in

experiment, there are still limitations, and it is frequently difficult to know a priori which

modeling choices need to be made (such as what value the Hubbard U should have),

which sometimes makes predictive modeling work challenging. To that end,

collaboration with experiment is still extremely important, which is thus the nature of the

work discussed in the reminder of this thesis.

References

1. Blöchl, P. E., Phys. Rev. B 50, 17953−17979 (1994).

2. Hohenberg, P. & Kohn, W. Physical Review 136, B864−B871 (1964).

3. Kohn, W. & Sham, L. J. Physical Review 140, A1133−A1138 (1965).

4. Sholl, D.S. & Steckel, J.A. Density Functional Theory: A Practical Introduction
(John Wiley and Sons, Hoboken, NJ, 2009).

5. Himmetoglu, B., Floris, A., de Gironcoli, Stefano & Cococcioni, M. Int. J. Quant.
Chem. 114, 14-49 (2014).

6. Oliver, G. L. & Perdew, J. P. Phys. Rev. A 20, 397 (1979).

22
Chapter 4. Independent Ordering of Two Interpenetrating
Magnetic Sublattices in the Double Perovskite Sr2CoOsO6

Note: This chapter is being presented in the format of a journal article. It was published

as Morrow, R., Mishra, R. Restrepo, O. D., Ball, M. R., Windl, W., Wurmehl, S.,

Stockert, U., Büchner, B. & Woodward, P. M. Independent ordering of two

interpenetrating magnetic sublattices in the double perovskite Sr2CoOsO6. Journal of the

American Chemical Society. 135 18824-18830 (2013). Rohan Mishra performed the DFT

calculations of the non-collinear structures in this paper. I performed the collinear

calculations and performed the fitting for the results of this paper with the assistance of

Oscar Restrepo.

4.1 Abstract

The insulating, fully ordered, double perovskite Sr2CoOsO6 undergoes two magnetic

phase transitions. The Os(VI) ions order antiferromagnetically with a propagation vector

k =(½,½,0) below TN1 = 108 K, while the high spin Co(II) ions order

antiferromagnetically with a propagation vector k =(½,0,½) below TN2 = 70 K. Ordering

of the Os(VI) spins is accompanied by a structural distortion from tetragonal I4/m

symmetry to monoclinic I2/m symmetry, which reduces the frustration of the face

centered cubic lattice of Os(VI) ions. Density functional theory calculations show that

the long range Os−O−Co−O−Os and Co−O−Os−O−Co superexchange interactions are

23
considerably stronger than the shorter Os−O−Co interactions. The poor energetic overlap

between the 3d orbitals of Co and the 5d orbitals of Os appears to be responsible for this

unusual inversion in the strength of short and long range superexchange interactions.

4.2 Introduction

Superexchange interactions govern cooperative magnetism in insulators. The strength of

a given superexchange interaction depends upon the orbitals involved and generally

decreases rapidly as the distance between the magnetic ions increases. The Goodenough-

Kanamori rules1,2 provide a simple, yet accurate, means of determining the sign of the

superexchange interactions These simple concepts are used by scientists to guide their

search for new magnetic materials. In this article we report the crystal and magnetic

structures of Sr2CoOsO6, a quaternary transition metal oxide with the double perovskite

structure. The A2MM′O6 double perovskites are among the most interesting classes of

transition metal oxides.3 Compared to simpler AMO3 ternary perovskites, the presence of

chemical order of M and M′ cations profoundly impacts both the electronic structure and the

magnetic coupling, in part because there are a greater variety of magnetic exchange interactions

at work in the double perovskite structure. The results presented here show that long range

superexchange interactions between ions of the same type (i.e. Co–O–Os–O–Co and Os–

O–Co–O–Os) are much stronger than the shorter nearest neighbor Co−O−Os

superexchange interactions. This result has broad implications for the way we think

about magnetism in mixed 3d−5d transition metal oxides.

The double perovskite structure, with general formula A2MM′O6, consists of a cubic,

corner connected network of MO6 and M′O6 octahedra that alternate in all three

24
directions, so that each MO6 octahedron is connected only to M′O6 octahedra and vice

versa, as shown in Figure 4.1. An illustrative example is Sr2CrOsO6 where high


4
temperature ferrimagnetism (TC = 720 K) results from strong, antiferromagnetic

Cr(III)−O−Os(V) superexchange coupling, as predicted by the Goodenough-Kanamori

rules. According to those same rules adding electrons to the eg orbitals of the 3d ion, for

example by replacing Cr with Fe(III), Co(II) or Ni(II), should change the sign of the

superexchange interactions and stabilize an insulating ferromagnet with a high TC—a rare

combination.5 To test this hypothesis as well as to better understand the rules that dictate

3d-5d superexchange we have prepared and characterized Sr2CoOsO6, a compound first

prepared by Sleight et al.,6 but whose magnetic and electrical properties were previously

unknown.

Figure 4.1 The A2MM′O6 double perovskite structure. The large A cations are shown in
gray, while the M and M′ centered octahedra are shown in red and blue respectively.

25
4.3 Experimental Section

Batches of at most 1.6 g (3.075 mmol) Sr2CoOsO6 powder were prepared by a solid state

method utilizing stoichiometric quantities of SrO2 (Sigma Aldrich, 98% pure), Co3O4

(Fischer Scientific, 99.8%pure) and Os metal (Alfa Aesar, 99.98% pure). The mixture

was loaded into a capped alumina tube and placed in a quartz tube, of approximately 40

mL volume with 3mm thick walls, along with an additional capped alumina vessel

containing PbO2. The quartz tube was then evacuated and sealed prior to firing in a

furnace to 1000 °C for 48 hours within a fume hood. The PbO2 was reduced to PbO at

elevated temperatures, acting as an in-situ source of oxygen gas to fully oxidize Os

without allowing appreciable loss of volatile OsO4.7,8 The best results were obtained

when the amount of PbO2 was enough to produce an excess of one-quarter mole O2 per

mole of Sr2CoOsO6 produced. The reaction can be represented as follows:

12 SrO2 + 2 Co3O4 + 6 Os + (7 PbO2) → 6 Sr2CoOsO6 + (7 PbO) + 3/2 O2

Phase purity and chemical ordering was established via XRD utilizing a Bruker D8

Advance Diffractometer equipped with a Cu source and an incident beam Ge(111)

monochromator. Refinement indicated negligible loss of osmium. An X-ray powder

26
diffraction pattern for this sample can be found in the supporting information in

Appendix A (Figure A.1).

Note that when heating osmium metal or binary osmium oxides it is important to confine

the reactants to a sealed vessel and take precautions in case a leak of the vessel occurs,

due to the formation of the highly toxic OsO4(g) upon heating.

Variable temperature neutron powder diffraction measurements were conducted on the

POWGEN9 beamline at Oak Ridge National Laboratory’s Spallation Neutron Source. A

sample size of approximately 1.6 g contained within an 8 mm vanadium can was used for

these experiments. Two wavelength ranges were measured at each temperature

producing histograms with d-spacing ranges of 0.2760–3.0906 Å and 2.2076–10.3019 Å,

referred to as Frame 1.5 and Frame 5 respectively. Long scan times of 1 hour for Frame

1.5 and 2 hours for Frame 5 were collected at 12, 80, 130, and 300 K while brief 5 minute

scans were collected in each frame at 10 K intervals.

The k-vectors were determined by k search implemented in Fullprof10, symmetry analysis

was performed and magnetic cells were generated utilizing SARAh11, and refinements

were conducted with the Rietveld method as implemented in GSAS EXPGUI.12,13 The

magnetic form factor for osmium was input manually using the parameters given by

Kobayashi et al.14

The magnetic susceptibility of Sr2CoOsO6 was collected within the temperature range of

5–400 K with both zero-field cooled and field cooled conditions under the effect of an

applied field of 1 kG. These measurements were accomplished utilizing a Quantum

27
Design MPMS SQUID magnetometer. A 46.6 mg sample was contained in a gelatin

capsule and mounted in a straw for insertion into the instrument. No diamagnetic

corrections for this sample holder were taken into account due to the large contribution of

the sample.

Sr2CoOsO6 powder was pressed into a pellet prior to being sintered at 1100 °C and being

cut into an approximate bar shape. The density was calculated to be 66.9% of the

theoretical density of the material. Four-point contacts were secured to the pellet with

silver paint before DC electrical measurements of the sample were conducted within the

temperature range 5–350 K utilizing a Quantum Design PPMS. Below 47 K the

resistance of the sample became too high to measure accurately.

Specific heat capacity was measured with a Physical Property Measurement System

(PPMS, Quantum Design), using a relaxation method. Sr2CoOsO6 was measured in zero

magnetic field between 2 K and 298 K using small, disc-like samples of roughly 11 mg.

In order to thermally couple the sample to the calorimeter, Apiezon grease was used. The

specific heat of the Apiezon grease was measured alone first over the whole temperature

range, and then subtracted from the total value of the subsequent measurement which

included the sample, to derive the specific heat contribution from the sample only.

Density functional theory calculations were performed using the Vienna ab-initio

Simulation Package (VASP).15,16 The influence of the core electrons was incorporated

using projector augmented wave (PAW) potentials17 within the spin-polarized

generalized gradient approximation (GGA) exchange-correlation functional.18 The plane-

wave cutoff energy was set at 525 eV. The Brillouin zone was sampled using dense
28
Monkhorst-Pack k-point meshes,19 with mesh divisions Ni such that the product of Ni with

the corresponding lattice vectors was as close as possible to 50 Å. The low temperature

(12 K) experimental lattice constants and ionic coordinates obtained from neutron

diffraction studies were used in the calculations without performing any structural or

ionic relaxations. The collinear magnetic calculations were performed on 20 to 160 atoms

supercells. The experimentally observed magnetic structure was studied using a 160 atom

cell and non-collinear calculations were performed by including spin−orbit interactions as

implemented in VASP. 20

To include the strong-correlations in the transition metal elements, we have used the

rotationally invariant Dudarev approach21 to GGA+U,22 in which only one effective

Hubbard parameter Ueff is used. Given that Ueff is not precisely determined by experiment

or theory, we have examined a range of Ueff values: 0 and 4.1 eV for Co and 0, 2.1 and

4.1 eV for Os. While the order of stability of different magnetic configurations remains

the same for all combinations of UCo and UOs, we find that UCo = 4.1 eV and UOs = 2.1

eV, also reproduce the experimentally observed insulating nature. Hence, we use this

particular combination of UCo and UOs to determine the exchange energies.

4.4 Results and Discussion

The magnetic susceptibility of Sr2CoOsO6 as a function of temperature is shown in

Figure 4.2. As signaled by the cusps in the data, Sr2CoOsO6 appears to undergo two

antiferromagnetic transitions at temperatures of approximately TN1 = 108 K and TN2 = 70

K. The inverse susceptibility in the paramagnetic regime 200 – 400 K follows the Curie-

29
Weiss law. The Weiss temperature, θ, extracted from the Curie-Weiss fit was found to be

−51 K. The atypical observation that θ is lower than either of the Neél temperatures

suggests the presence of ferromagnetic interactions that compete with the

antiferromagnetic interactions. As we will see later density functional theory (DFT)

calculations confirm the presence of competing ferromagnetic and antiferromagnetic

superexchange interactions. The effective magnetic moment per formula unit obtained

from the Curie-Weiss fit is µeff = 4.46 µB, which is in reasonably good agreement with the

expected spin only value calculated assuming the presence of Os(VI) and high spin (HS)

Co(II), µspin = [µspin(Co)2 + µspin(Os)2]1/2 = 4.80 µB. However, it should be noted that HS

Co(II) has been reported in related double perovskites to have an effective moment which

is higher than the spin-only value of 3.87 µB due to a constructive contribution from

unquenched orbital moment,23-26 and the orbital contribution if present for Os(VI) would

tend to reduce the moment of that ion from its spin only value of 2.83 µB.27 Thus it is

quite plausible that orbital angular momentum contributes to the moment of both ions,

but the effects average out to give an overall µeff close to the spin-only value.

30
240
1.4e-2
Molar Susceptibility (emu/mol)
200

Inverse Susceptibillity
1.2e-2
160

1.0e-2
120

8.0e-3
80

6.0e-3 40

4.0e-3 0
0 100 200 300 400
Temperature (K)

Figure 4.2 Field cooled (filled circles) and zero field cooled (open circles) magnetic
susceptibility with inverse zero field cooled data plotted against the right axis showing
the linear Curie-Weiss fit.

Specific heat measurements on Sr2CoOsO6 reveal two pronounced anomalies on cooling,

at 108 K and 70 K (Figure 4.3). These temperatures correspond closely to the cusps seen

in the magnetic susceptibility measurements. The 70 K transition peak has a l-shape,

normally seen for a second-order phase transition. This observation, taken together with

the susceptibility measurements, suggests a purely magnetic phase transition. On the

other hand, the rather symmetric shape of the peak at 108 K suggests a first-order phase

transition, which would be expected if a structural phase transition accompanies the

magnetic transition, as confirmed by the diffraction measurements discussed below.

31
Figure 4.3 Specific heat of Sr2CoOsO6 as a function of temperature

Electrical transport measurements reveal insulating behavior as resistivity rises rapidly

with decreasing temperature, as shown in Figure 4.4. The log of the conductivity shows a

linear T−1/4 temperature dependence (see supporting information) which is consistent with

variable range hopping transport, whereby electrons hop between localized states. There

is a continuous deviation which occurs between the phase transitions, leading to two

separate linear fits, above and below the transitions. The mathematical form of variable
!/! )
range hopping conductivity is given as σ=σ0𝑒 (!!/! . The fitted A factors found above

and below this transition temperature are 43.12(2) and 46.14(9), respectively.

32
Figure 4.4 Log resistivity vs temperature.

Variable temperature time-of-flight neutron powder diffraction measurements were

collected to obtain accurate structural parameters and to probe the magnetic structure. At

room temperature both X-ray and neutron powder diffraction patterns can be fit to a

tetragonally distorted, ordered double perovskite structure with I4/m symmetry and

complete ordering of Co and Os. The tetragonal distortion occurs due to rotations of the

octahedra about the c-axis (a0a0c– tilting)3 and a relatively subtle axial elongation of the

octahedra. This relatively common distortion of the structure is also observed at room

temperature in double perovskites such as Sr2NiOsO6 7, Sr2FeMoO628 and Sr2CoReO6.22

The evolution of the structure was studied with variable temperature neutron powder

diffraction, and the results are shown in Figure 4.5. On cooling below 108 K, where

33
magnetic ordering sets in, an abrupt structural distortion occurs lowering the symmetry to

the monoclinic space group I2/m. The splitting of the (220) reflection signaling the

transition from tetragonal to monoclinic symmetry is shown in Figure A.3, in the

supporting information in Appendix A.

Figure 4.5 Evolution of the (a) lattice parameters, (b) Os–O bond lengths, and (c) Co–O
bond lengths as a function of temperature, as extracted from variable temperature neutron
diffraction data. The dashed line separates the tetragonal I4/m phase from the monoclinic
I2/m phase.
34
The Co–O and Os–O distances are plotted as a function of temperature in Figure 4.5. In

the high temperature tetragonal structure the axial Co–O(1) and Os–O(1) bonds are

slightly longer than the Co–O(2) and Os–O(2) bonds that lie in the xy-plane. This type of

distortion could be associated with a Jahn-Teller distortion arising from the orbital

degeneracy of either the HS d7 Co(II) ion or the d2 Os(VI) ion. The magnitude of the

distortion is relatively small, but this is not unexpected for a Jahn-Teller distortion arising

from partial occupation of the t2g orbitals.

In both the tetragonal and monoclinic phases the bond lengths are consistent with the

assignment of high spin Co(II) and Os(VI) oxidation states,29,30 so we can rule out a

change in spin-state or some type of Co–Os charge transfer as the driving force behind

the transition. Furthermore, the a0a0c– pattern of octahedral tilting is maintained, so we

can also rule out octahedral tilting as the driving force for the phase transition. Instead

the transition appears to be most closely linked to distortions of the CoO6 octahedra, as

shown in Figure 4.5.

In the low temperature monoclinic structure the symmetry is lowered in such a way that

there are three unique bond distances within each octahedron. Despite the reduction in

symmetry, the 2 long/4 short distortion of the osmium octahedron is essentially retained

with bond distances of 2×1.946(1) Å, 2×1.902(2) Å and 2×1.906(2) Å. In contrast the

distortion of the cobalt octahedron changes at the phase transition and this change is

responsible for lowering the symmetry to monoclinic. In the tetragonal structure all four

of the equatorial Co–O(2) bonds are the same length, 2.036(1) Å at 130 K. This square
35
base distorts in the monoclinic structure to give two bonds of 2.013(2) Å and two bonds

of 2.056(2) Å. The longer of these two sets of bonds is comparable to the axial Co−O(1)

distance, 2.067(1) Å, so that the octahedron now has 4 long and 2 short bonds. This type

of distortion is unusual for Co(II). A thorough literature search revealed only one other

Co2+ containing perovskite with a similar distortion, La2CoIrO6.31

Table 4.1 gives information concerning the unit cell and goodness of fit parameters for
each refinement on the longer neutron data sets. Figures and tables containing further
details on the structure of Sr2CoOsO6 at different temperatures can be found in the
supporting information in Appendix A.

12K 80K 130K 300K

Space Group I112/m I112/m I4/m I4/m

a (Å) 5.5061(1) 5.5079(1) 5.51867(7) 5.54786(8)

b (Å) 5.5079(1) 5.5097(1) - -

c (Å) 8.0258(1) 8.0235(1) 8.0049(1) 7.9565(1)

V (Å)3 243.386(3) 243.48(1) 243.796(9) 244.891(9)

γ (°) 90.623(1) 90.574(1) - -

χ2 3.38 3.27 3.89 3.86

Rwp 1.88 1.86 2.03 2.02

Rp 3.02 2.9 3.06 3.05

Os moment 1.81(4) µB 1.57(5) - -


µB

Co moment 2.90(5) µB - - -

36
Figure 4.6 shows the electronic density of states calculated using density functional

theory (DFT). The 5d t2g orbitals of osmium make the major contribution at the Fermi

level, but correlations open up a small gap in what would otherwise be a 1/3 filled band.

The prediction of insulating behavior is consistent with our resistivity measurements.

Figure 4.6 Total- and atom- resolved density of states of Sr2CoOsO6 from a non-collinear
DFT calculation with UCo = 4.1 eV and UOs = 2.1 eV. The contributions of the Os 5d and
Co 3d orbitals are shown in red and blue, respectively.

Upon cooling below TN1 = 108 K, an additional set of reflections arise in the neutron

diffraction pattern that cannot be attributed to the structural phase transition, one such

reflection is found at d = 7.7 Å as shown in Figure 4.7. These reflections signal the onset

of antiferromagnetic order, and can be accounted for with a magnetic propagation vector

k = ½, ½, 0. Two irreducible representations are consistent with this propagation vector


37
and I2/m symmetry, but only the Γ(3) representation, with basis vectors in the a and b

directions, is able to correctly reproduce the observed intensities of the magnetic

reflections. Refinements of a long data set collected at 80 K yield a moment of 1.57(5)

µB if one assumes ordering of Os(IV), or 1.68(5) µB if one assumes ordering of Co(II). It

was also possible to fit the data reasonably well with non-negligible moments on both Co

and Os by fixing the moment on one atom to an intermediate value, such as 0.8 µB, and

refining the moment on the other. However, those refinements always gave slightly

higher Rwp values, and poorly fit one of the magnetic reflections. This observation lends

support to the hypothesis that between 70−110 K only one magnetic sublattice orders.

Figure 4.7 Rietveld fits to the high d-spacing regions of the neutron diffraction patterns
showing the presence/absence of magnetic reflections at three temperatures: above TN1
(130 K), between TN1 and TN2 (80 K), and below TN2 (12 K). Black symbols represent
observed data while the red curves represent the calculated patterns. The hkl values given
in the figure are indexed on the nuclear cell.

38
On cooling below TN2 a second set of magnetic reflections emerge that can be indexed

with a propagation vector k = ½, 0, ½, the strongest of which is the peak at d = 9.1 Å seen

in Figure 4.7. Due to the similarity of the a and b lattice parameters, it is possible to

produce a nearly identical quality solution using a propagation vector of vector k = 0, ½,

½, due to powder averaging. The following analysis and conclusions are not affected by

this choice of propagation vector. A symmetry analysis yields only one irreducible

representation, Γ(2), for this second set of magnetic peaks. Although the Γ(2)

representation has three basis vectors, only the two oriented in the a and c directions were

needed to model the intensities of the magnetic reflections. It should also be noted that

the emergence of a second set of magnetic reflections below TN2 does not lead to any

noticeable change in the intensities of the first set of magnetic reflections. A low

temperature magnetic structure model was derived by combining the k= ½, ½, 0 ordering

scheme on one magnetic sublattice, and the k= ½, 0, ½ scheme on the other sublattice,

leading to a 2 × 2 × 2 magnetic unit cell, shown in Figure 4.8(a). In this structure the

cobalt and osmium spins are non-collinear, with both sublattices adopting

antiferromagnetic order.

While the neutron diffraction experiments provide strong support for independent

ordering of cobalt and osmium ions, it is not immediately apparent which ion orders at

the higher temperature. If we assume that the cobalt ions order at 108 K and the osmium

ions order at 70 K, refinements of the 12 K data set yield atomic moments of 1.99(4) µB

and 2.71(4) µB for these two ions, respectively. These values are not physically

reasonable because a moment of 2.71 µB for the d2 Os(VI) ion is larger than the spin-only

39
value of 2 µB, and any orbital contribution would be expected to reduce, not increase, the

moment. Using the opposite assumption, osmium ordering at 108 K and cobalt ordering

at 70 K, gives moments of 1.81(4) µB and 2.90(5) µB that are quite close to the high field,

spin-only values of 2 µB and 3 µB expected for the Os(VI) and HS d7 Co(II) ions,

respectively. DFT calculations give moments of 1.69 µB and 2.71 µB for osmium and

cobalt, respectively. These values are in close agreement with the experimentally

obtained values, and the agreement strongly supports the conclusion that the osmium ions

order at 108 K and the cobalt ions at 70 K.

40
Figure 4.8 (a) Experimentally determined magnetic structure of Sr2CoOsO6, with the Os
shown in red and the Co in blue. (b) A schematic of the superexchange pathways between
Co and Os ions with the exchange constants extracted from the DFT calculations. A
positive (negative) value of J indicates ferromagnetic (antiferromagnetic) coupling.

The neutron refinements reveal a complex magnetic structure, but they do not provide

direct information as to why this magnetic structure is more stable than simpler

alternatives normally seen in perovskites. DFT calculations were performed in order to

extract the magnetic coupling constants that are responsible for the unusual magnetism

observed in this compound.32 There is one prior report of DFT calculations on

41
Sr2CoOsO6, but in that study they only looked at the ferromagnetic and ferrimagnetic

configurations, concluding that Sr2CoOsO6 should be a ferrimagnetic half-metal.33

However, as shown here the actual ground state is an antiferromagnetic insulator, which

lessens the validity of the conclusions made in the earlier computational study.

The relative energies of ten magnetic configurations were calculated and the

experimentally observed non-collinear, antiferromagnetic ground state with a 2 × 2 × 2

unit cell was found to be the lowest energy configuration (see supporting information).

The ferromagnetic configuration was 809 meV per f.u. higher in energy, while the

ferrimagnetic configuration was 776 meV higher in energy. The exchange constants, for

seven exchange pathways of potential importance were extracted from these calculations

and are given in Figure 4.8. To minimize correlations in the fitting procedure the through

bond superexchange constants J1eff, J2eff, and J3eff were constrained to be the same in all

three directions. Because the direct exchange coupling between ions of the same type

across a face of the unit cell can be quite sensitive to octahedral tilting, we have used two

different J values for Co−Co coupling, J4eff, and J5eff, and two for Os−Os coupling, J6eff,

and J7eff.

!""
The calculations reveal very weak coupling, 𝐽! = −1.3 meV for the nearest neighbor

Co–O–Os superexchange pathways that are the shortest interactions and would normally

be expected to be the strongest. The largest superexchange coupling is found for the four
!""
bond Co(↑)–O–Os–O–Co(↓) interaction, which is antiferromagnetic with 𝐽! = – 47.2

!""
meV The four bond Os(↑)–O–Co–O–Os(↑) interaction is also quite strong, with 𝐽! =

20.2 meV, but ferromagnetic. The in-plane Co−Co coupling, J4eff was also relatively
42
strong, 29 meV, but this value was somewhat dependent on the constraints used with

fitting J values to the calculated energies of the various magnetic structures.

To our knowledge it is unprecedented that these long superexchange pathways could

emerge as being much stronger than the shorter nearest neighbor interactions along the

same path, but a closer look at the experimental structure shows that all Os–O–Co–O–Os

interactions are indeed ferromagnetic, while all of the Co–O–Os–O–Co interactions are

antiferromagnetic. Examination of the partial density of states of the Co and Os ions

(Figure 4.6) shows that there is only a small overlap of the Co 3d and Os 5d states

(particularly for the eg orbitals), which helps to explain the weak nearest neighbor

interaction between the two ions. In other words the poor energetic overlap of the Co 3d

and Os 5d orbitals (once hybridized with oxygen) leaves the two sublattices relatively

uncoupled.

Before finishing let’s return to the unusual 4 long/2 short distortion of the cobalt

octahedron that occurs on cooling below TN1. While there is no way to rationalize this

type of distortion based solely on the electron configuration of the HS Co(II) ions, the

fact that the distortion accompanies ordering of the osmium spins provides an important

clue. In the double perovskite structure the osmium ions sit on a face centered cubic (fcc)

lattice, which is a geometrically frustrated lattice. While all Os–O–Co–O–Os

superexchange interactions are ferromagnetic, there are four independent but

interpenetrating sublattices of Os ions that do not interact with each other via this

exchange route. These sublattices interact through nearest neighbor Os–Os exchange

given as 𝐽! and 𝐽! in Figure 4.8. In the high temperature tetragonal structure, the Os–Os

43
nearest neighbor distances are 5.51 Å (×4) and 5.59 Å (×8). This arrangement is slightly

distorted from a perfect fcc lattice, but still highly frustrated. In the low temperature

monoclinic structure the distances become 5.51 Å (×4), 5.57 Å (×4), and 5.61 Å (×4).

These distances represent the nearest neighbor distances from one Os sublattice to the

other three independent Os sublattices. The two sublattices that are aligned

ferromagnetically with each other are separated by the longest of these three distances,

5.61 Å. Thus the structural distortion that accompanies TN2 reduces the frustration of the

fcc arrangement of osmium ions by shortening the Os–Os distances between the eight

nearest neighbor antiferromagnetically coupled Os ions and increasing the distances

between the four nearest neighbor ferromagnetically coupled Os ions. Based on this

observation we hypothesize that the structural transition at 108 K, and the resulting

unusual distortion of the CoO6 octahedron, is driven by magnetic ordering of the Os

spins.

4.5 Conclusions

The double perovskite Sr2CoOsO6 undergoes two magnetic phase transitions on cooling

from room temperature. The two interpenetrating sublattices of Os(VI) and Co(II) each

order antiferromagnetically, but independent of each other. The Os(VI) spins order

antiferromagnetically below 108 K, while the Co(II) spins order antiferromagnetically

below 70 K. The observed magnetic structure suggests and the computational results

confirm that four bond superexchange interactions, Os–O–Co–O–Os and Co–O–Os–O–

Co, are stronger than the nearest neighbor Os–O–Co superexchange interactions. This

highlights a fundamental difference between double perovskites containing 3d and 5d


44
ions, with more familiar perovskites containing only 3d ions. These results provide a

fascinating glimpse of superexchange coupling in mixed 3d−5d transition metal oxides,

revealing a complexity that was not previously appreciated. Further studies on a variety

of A2MM′O6 double perovskites with varying electron count, orbital energies, and

structural distortions are needed to better understand the factors that control the strength

and sign of superexchange coupling in mixed 3d−5d transition metal oxides.

Furthermore, the presence of competing long range antiferromagnetic and ferromagnetic

exchange interactions provides a route to materials where dramatic changes in properties

can result in response to relatively subtle changes in structure.

References

[1] Goodenough, J. B. Phys. Rev. 1955, 100, 564-573.

[2] Kanamori, J. J. Phys. Chem. Solids 1959, 10, 87-98.

[3] Lufaso, M.W.; BarnesP. W.;Woodward, P.M. Acta Cryst. B 2006, 62, 397-
410.

[4] Krockenberger, Y; Mogare, K.; Reehuis, M.; Tovar, M.; Jansen, M.;
Vaitheeswaran, G.; Kanchana, V.; Bultmark, F.; Delin, A.; Wilhelm, F.; Rogalev,
A.; Winkler, A.; Alff, L. Phys. Rev. B. 2007, 75, 020404(R).

[5] Rogado, N. S.; Li, J.; Sleight, A. W.; Subramanian, M.A. Adv. Mater. 2005, 17
(18), 2225-2227.

[6] Sleight, A. W.; Longo, J.; Ward, R. Inorg. Chem. 1962, 1 (2), 245–250.

[7] Macquart, R; Kim, S. J.; Gemmill, W. R.; Stalick, J. K.; Lee, Y.; Vogt, T.; Zur
Loye, H. C . Inorg. Chem. 2005, 44, 9676-9683.

[8] Lufaso, M. W.; Gemmill, W. R.; Mugavero, S. J.; Kim, S. J.; Lee, Y.; Vogt, T.;
zur Loye, H. C. J. Solid State Chem. 2008, 181 (3), 623-627.

[9] Huq A.; Hodges J. P.; Gourdon, O.; Heroux, L. Zeitschrift für Kristallographie
Proceedings 2011, 1, 127-135.
45
[10] Rodriguez-Carvajal, J. Physica B. 1993, 192, 55-69.

[11] Wills, A. Physica B. 1991, 276, 680-681.

[12] Larson, A. C.; Von Dreele, R. B. Los Alamos National Laboratory Report LAUR
86-748 2000.

[13] Toby, B. H. EXPGUI, a graphical user interface for GSAS. J. Appl. Cryst. 1991,
34, 210-213.

[14] Kobayashi, K.; Nagao, T.; Ito, M. Acta Crys. A 1991, 67, 473-480.

[15] Kresse, G; Hafner, J. Phys. Rev. B 1991, 47, 558−561.

[16] Kresse, G.; Hafner, J. Phys. Rev. B 1991, 49, 14251−14269.

[17] Blöchl, P. E. Phys. Rev. B 1991, 50, 17953−17979.

[18] Wang, Y.; Perdew, J. P. Phys. Rev. B 1991, 44, 13298−13307.

[19] Monkhorst, H. J.; Pack, J. D. Phys. Rev. B 1991, 13, 5188−5192.

[20] Hobbs, D.; Kresse, G.; Hafner, J. Phys. Rev. B 1991, 62, 11556–11570.

[21] Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.; Sutton, A. J.
Phys. Rev. B 1991, 57, 1505–1509.

[22] Anisimov, V. I.; Zaanen, J.; Andersen, O. K. Phys. Rev. B 1991, 44, 943-954.

[23] Viola, M. C; Martinez-Lope, M. J.; Alonso, J. A.; De Paoli, J. M.; Pagola, S.;
Pedregosa, J. C.; Fernandez-Diaz, M. T.; Carbonio, R. E. Chem. Mater. 2003, 15
(8), 1655-1663.

[24] Ivanov, S. A.; Eriksson, S. G.; Tellgren, R.; Rundlöf, H.; Tseggai, M. Mater. Res.
Bull. 2005, 40, 840-849.

[25] Lopez, C. A.; Saleta, M. E.; Curiale, J.; Sanchez, R. D. Mater. Res. Bull. 2012,
47, 1158-1163.

[26] Retuerto, M.; Martínez-Lope, M. J.; García-Hernández M.; Fernández-Díaz, M.


T.; Alonso, J. A. Eur. J. Inorg. Chem. 2008, 4, 588-595.

[27] Choy, J. H.; Kim, D. K.; Kim., J. Y. Solid State Ionics. 1998, 108, 159–163.

46
[28] Chmaissem, O.; Kruk, R.; Dabrowski, B.; Brown, D. E.; Xiong, X.; Kolesnik, S.;
Jorgensen, J. D.; Kimball C. W. Phys. Rev. B 2000, 62, 14197-14206.

[29] Shannon, R. D. Acta Cryst. A 1976, 32, 751-767.

[30] Brese, N. E.; O’Keefe, M. Acta Cryst. B 1991, 47, 192-197.

[31] Currie, R. C.; Vente, J. F.; Frikkee, E.; IJdo, D. J. W. J. Solid State Chem. 1995,
116 (1), 199-204.

[32] Tian,C.; Wibowo, A. C.; Zur Loye, H.-C.; Whangbo, M-H. Inorg. Chem. 2011,
50, 4142-4148.

[33] Wang, J.; Meng, J.; Wu, Z. Chem. Phys. Lett. 2011, 501, 324-329.

47
Chapter 5. Strain-dependent, Extraordinary
Magnetocrystalline Anisotropy in Sr2CrReO6 Epitaxial Films

Note: This chapter is being presented in the format of a journal article. It was published

as Lucy, J. M., Ball M. R., Restrepo, O. D., Hauser, A. J., Soliz, J. R., Freeland, J. W.,

Woodward, P. M., Windl, W. & Yang, F. Y. Strain-dependent, Extraordinary

Magnetocrystalline Anisotropy in Sr2CrReO6 Epitaxial Films. Physical Review B. 90

180401(R) (2014). I performed all the DFT calculations including relaxing the structures,

ascertaining the density of states, and determining the magnetic moments. The

subsequent analysis for this work was also done by me with assistance from Oscar

Restrepo.

Abstract

We report the discovery of extraordinarily large anisotropy fields and strain-tunable

magnetocrystalline anisotropy in Sr2CrReO6 epitaxial films. We determine the strain-

induced tetragonal distortions and octahedral rotations in Sr2CrReO6 epitaxial films

grown on (LaAlO3)0.3(Sr2AlTaO6)0.7 (LSAT), SrTiO3 (STO), and a relaxed

SrCr0.5Nb0.5O3/LSAT substrates using x-ray diffraction and density functional theory.

The structural distortions drive dramatic changes in magnetocrystalline anisotropy. We

use magnetometry measurements and first principles calculations to determine the atomic

origins of the large anisotropy observed. These techniques elucidate the interplay

48
between structural deformations and magnetic behavior and lay the groundwork for the

study of other strongly correlated systems in this class of ferromagnetic oxides.

Magnetocrystalline anisotropy (MCA) has significant implications in a range of

applications such as power generation and magnetic data storage. The search for and

study of materials for such applications is of both scientific and technological interest.

There is much focus on ferromagnets such as SmCo5, Nd2Fe14B and FePt which exhibit

high anisotropy due to crystal symmetries and strong spin-orbit coupling [1,2]. Here we

report the discovery of exceptionally large anisotropy fields and strain-tunable MCA in

Sr2CrReO6 epitaxial films. We determine the strain-induced tetragonal distortions and

octahedral rotations of the Sr2CrReO6 lattice which lead to dramatic changes in MCA and

the capability to switch the magnetic easy axis from in-plane to out of plane via strain.

Furthermore, we perform first principles calculations in order to determine the atomic

origins of the large anisotropy observed. This Rapid Communication provides a

combination of experimental and theoretical work to elucidate the atomic magnetic

behavior of a complex material.

The advances in fabrication techniques of crystalline materials, particularly epitaxial

films in recent years, have enabled engineering of materials with desired, and sometimes

exotic, electronic and magnetic properties such as new ferroelectric materials [3,4] and

ferromagnets (FMs) with large MCA [1,2,5]. Most importantly, epitaxial strain offers the

capability to significantly alter the electronic and magnetic properties of the films and

even creates new phenomena [3,4] that do not exist in bulk. Crystal structures that are not

isotropic, such as tetragonal (e.g. CrO2 and FePt) and hexagonal lattices (e.g. Co and Dy),

49
typically result in large MCA due to magnetization-lattice coupling [5-7]. Since spin-

orbit coupling (SOC) generally scales with Z4, where Z is the atomic number, high-

anisotropy FMs typically contain 4f or 5d elements. In cubic systems, MCA is usually

small due to the high symmetry, such as in 3d FM metals and Heusler compounds [8,9].

The ABO3 perovskites are a large family of complex materials that exhibit many

fascinating phenomena. However, magnetic perovskites typically have low TC and

modest MCA. Meanwhile, the A2BB’O6 ferrimagnetic double perovskites have been

shown to possess versatile magnetic properties such as high spin polarization, high TC (up

to 725 K), strong and tunable SOC, and electrical conductivity ranging from insulating to

conducting [10-13]. We have demonstrated growth of fully ordered, high quality

Sr2CrReO6 epitaxial films [14-16] using off-axis sputtering, which exhibit SOC-enhanced

magnetization. Previous studies of Sr2CrReO6 in bulk and epitaxial film form can be

found in Refs. [10,11,14-16]. It has been recently reported that Sr2CrReO6 films exhibit

an abrupt change of magnetic coercivity of 1.2 T when subject to structural

transformations of an underlying BaTiO3 substrate [17]. Theoretical studies also

elucidated the contributing factors leading to the large MCA in Sr2CrReO6 [18,19]. These

results point toward the possibility of tuning the large MCA in S Sr2CrReO6 films via

strain.

We use off-axis sputtering [9,14-16,20-23] to deposit 90 nm thick Sr2CrReO6 (001) films

on (LaAlO3)0.3(Sr2AlTaO6)0.7 (LSAT), SrTiO3 (STO), and a relaxed SrCr0.5Nb0.5O3

(SCNO) buffer layer on LSAT with lattice constants a = 3.868 Å, 3.905 Å and 3.946 Å,

respectively (see Appendix B for growth parameters). As a comparison, the pseudocubic

50
lattice constants of bulk Sr2CrReO6 are ap = 3.907 Å and cp = 3.905 Å [10]. It should be

noted that in application it would be favorable to apply continuous strain to the

Sr2CrReO6 films through the use of a piezoelectric substrate. However, at this time no

piezoelectric substrate is available that can apply epitaxial strain within the

approximately ± 1% range used here, particularly around the lattice constant of

Sr2CrReO6.

The 90 nm Sr2CrReO6 films are thin enough to be fully strained to the underlying

substrates and the similarities in film thickness reflect our precise control of deposition

rates. The x-ray diffraction (XRD) scans in Fig. 5.1 show that the films are pure phase

with substrate-limited rocking curve full-width-at-half-maximums (FWHMs) as small as

0.0063°. The left panels in Figs. 5.1(a)-5.1(c) show the θ−2θ scans near the Sr2CrReO6

(004) peak for films on LSAT, SrTiO3 and SCNO/LSAT, from which the out-of-plane

(perpendicular to he film-substrate interface) lattice constants c = 7.926 Å, 7.860 Å and

7.804 Å, respectively, are obtained. From the off-axis θ−2θ scans for the Sr2CrReO6

(022) peak (right panels in Fig. 5.1), we calculate the in-plane lattice constants a = 7.732

Å, 7.806 Å and 7.876 Å, for films on LSAT, SrTiO3 and SCNO/LSAT, resulting in

tetragonal distortions c/a = 1.025, 1.007 and 0.991, respectively. All three films are fully

strained, compressive (c/a > 1) or tensile (c/a < 1), to the substrates or buffer layer (Table

5.1). Finally, Laue oscillations are observed in all three samples, indicating high

uniformity through the films. From the spacing of Laue oscillations, we obtain

thicknesses of 91.1, 90.2 and 89.6 nm for the Sr2CrReO6 films grown on LSAT, SrTiO3

and SCNO/LSAT, respectively.

51
(a) SCRO/LSAT

LSAT(002)

LSAT(011)
SCRO/LSAT
105

Intensity (c/s)

SCRO(004)

SCRO(022)
103

101 = 45o
(b) SCRO/STO SCRO/STO

STO(002)

STO(011)
105
Intensity (c/s)

103

SCRO(022)
SCRO(004)
101 = 45o
(c) 45 45.5 46
SCRO/SCNO/LSAT 46.5 47 SCRO/SCNO/LSAT

LSAT(011)
105 LSAT(002)
Intensity (c/s)

SCNO(004)

103
SCRO(004)

SCNO(022)

SCRO(022)
101 = 45o
45 45.5 46 46.5 47 32 32.5 33
2 (deg) 2 (deg)

FIG. 1

Figure 5.1 XRD scans around the Sr2CrReO6 (004) (left panels) and Sr2CrReO6 (022)
(right panels) peaks for 90-nm Sr2CrReO6 films grown on (a) LSAT, (b) SrTiO3 and (c)
Sr2CrNbO6 buffer layer on LSAT. The off-normal Sr2CrReO6 (022) peaks are measured
at a tilt angle Ψ =45° for in-plane characterization.

In the presence of epitaxial strain, it is energetically


1 favorable for the oxygen octahedra in

perovskites to rotate in order to accommodate the tetragonal distortion [25]. Rotation of

the octahedra alters both the bond lengths and bond angles between transition metals and

oxygen, potentially affecting both their electronic and magnetic properties. We utilize the

measured lattice parameters of the Sr2CrReO6 films to determine the rotation of the Cr
52
and Re oxygen octahedra by performing density functional theory (DFT) calculations

within the generalized gradient approximation (GGA) [26] using the Vienna ab initio

simulation package (VASP) [27,28] with projector augmented wave (PAW)

pseudopotentials [29]. Correlation effects were treated within the (GGA+U) approach

[30] with a value of U = 3 eV and an exchange parameter J = 0.87 eV for the Cr d

orbitals [31]. The resulting geometries and structural parameters are shown in Fig. 5.2

and Supplemental Material Table B.1 in Appendix B. As expected, the changes in

equatorial (in-plane) and axial (out-of-plane) bond lengths, deq and dax, respectively, in

the octahedra are considerably smaller than the changes in the lattice constants,

accompanied by rotations of the octahedra, as the films undergo progression from

compressive to tensile strain. Given the linear relationship in Figs. 5.2(c) and 5.2(d), we

find that the increase in the in-plane bond lengths is 23% (17%) of the changes in the in-

plane lattice constant a for Cr (Re) octahedra, while the decrease of the perpendicular

bond length to be 25% of the value of the c lattice constant for both octahedra. The

rotation of the octahedra is similar for both cations and changes by 2.3° across our strain

range (Supplemental Material Table B.1 in Appendix B). These structural changes, as

shown below, drastically affect the magnetic behaviors of the Sr2CrReO6 films.

53
(a) 6
(c) Cr
5.5

LSAT
Re

(deg)
5

SCNO/LSAT
4.5

STO
4

3.5
1.99 (d) d eq(Re-O)
d ax(Re-O)

Bond length (A)


1.97
(b)

1.95
d eq(Cr-O)
1.93 d ax(Cr-O)
0.99 1 1.01 1.02
compressive c/a tensile

FIG. 2

Figure 5.2 Schematics of (a) Cr−O (counterclockwise) and Re−O (clockwise) octahedral
rotations in the Sr2CrReO6 lattice and (b) an oxygen (red) octahedron surrounding Cr or
Re with two different bond lengths, deq and dax. (c) Octahedral rotation angle θ and (d)
bond lengths of Cr−O and Re−O octahedra as a function of the tetragonal distortion c/a
of the Sr2CrReO6 lattice.

To characterize the MCA in our Sr2CrReO6 films, we measure the magnetic hysteresis

loops for the three samples in magnetic fields


1
up to 7 T. Fig. 5.3 shows the in-plane and

out-of-plane hysteresis loops at temperatures T = 20, 100 and 300 K for the three films

after the subtraction of diamagnetic background (see Supplemental Material for

diamagnetic background subtraction details in Appendix B). All three substrates

(including the buffer layer) exhibit diamagnetic responses at the temperatures discussed

in this Rapid Communication.

54
We note three distinct features from these hysteresis loops. First, the magnetic easy axis

changes from in plane for the films on LSAT and STO with compressive strain (c/a > 1)

to out of plane for the film on SCNO/LSAT with tensile strain (c/a < 1). This is similar to

the observation in strained Sr2FeMoO6 films [21] since MCA favors the magnetic easy

axis along a shorter axis of the tetragonally distorted crystal lattice. A recent theoretical

calculation predicts the change in sign of the MCA energy in Sr2CrReO6 films at a tensile

strain of 0.7% using the local-spin-density approximation (LSDA) or 0.3% using the

GGA method [18]. For our Sr2CrReO6 film on SCNO/LSAT with a tensile strain of

0.9%, the magnetic easy axis is indeed out of plane and complements the theoretical

predictions.

55
(a)
1 T = 20 K T = 100 K T = 300 K
H || film
M ( B/f.u.)
0.5
0
H film
-0.5
-1 SCRO/LSAT
-6 -3 0 3 6 9 12 15 18 -6 -3 0 3 6 -6 -3 0 3 6
H (T) H (T) H (T)
(b) H || film
1 T = 20 K T = 100 K T = 300 K
M ( B/f.u.)

0.5
H film Hu
0
-0.5
-1 SCRO/STO
-6 -3 0 3 6 9 12 15 18 -6 -3 0 3 6 -6 -3 0 3 6
H (T) H (T) H (T)
(c)
1 T = 20 K H film T = 100 K T = 300 K
M ( B/f.u.)

0.5
H || film Hu
0
-0.5
-1 SCRO/SCNO/LSAT
-6 -3 0 3 6 9 12 15 18 -6 -3 0 3 6 -6 -3 0 3 6
H (T) H (T) H (T)

FIG. 3
Figure 5.3 In-plane and out-of-plane magnetic hysteresis loops for the Sr2CrReO6 films
grown on (a) LSAT, (b) SrTiO3 and (c) Sr2CrNbO6/LSAT in magnetic fields up to 7 T at
T = 20, 100 and 300 K. Magnetic anisotropy fields can be obtained from the intercept of
linear extrapolation of the hard-axis hysteresis loop and the saturation magnetization as
shown in the left panels.

Second, the MCA indicated by the hysteresis loops along the hard axis is very large for

all three films. In the left panel of Fig. 5.3(b) for Sr2CrReO6/SrTiO3, we extrapolate the

high-field region of the out-of-plane hysteresis loop to find the intercept with the

saturation magnetization (Ms) of the in-plane hysteresis loop. This intercept is an


56
1
approximate representation of the out-of-plane, uniaxial anisotropy field Hu = 18.1 T

(Table 5.1). Since the magnetization is almost, but not fully saturated at 7 T [14,32], the

obtained Hu from the intercept is an underestimate of the anisotropy field. The

demagnetization field 4πMs of 1260 G [14] is much smaller than the anisotropy field and

can be neglected. We calculate the anisotropy energy by finding the area between the

easy- and hard-axis loops. We find Ku, the uniaxial MCA energy density to be Ku = 4.77

× 106 erg/cm3 (4.77 × 105 J/m3), which is very high for Sr2CrReO6 with a small Ms = 0.85

µB/f.u. (as a reminder, this is for SCRO/STO). For comparison, SmCo5 films exhibit one

of the highest Ku = 7.6 × 107 erg/cm3, while the Ms of SmCo5 is much higher than that of

Sr2CrReO6 [1]. For Sr2CrReO6/LSAT, the MCA is considerably larger than that in

Sr2CrReO6/SrTiO3, as can be seen from the much larger difference between the in-plane

and out-of-plane hysteresis loops in Fig. 5.3(a). The magnetometry data for

Sr2CrReO6/LSAT does not allow a reliable determination of Hu from the hysteresis loops

since there is very little magnetic signal from the Sr2CrReO6 film for the hard-axis loop

(see Supplemental Material Figure B.1 in Appendix B for the raw magnetic data for

Sr2CrReO6/LSAT). However, we estimate that it should be at least tens of T. For the

Sr2CrReO6 film on SCNO/LSAT, the hard axis is in plane and the anisotropy field Hu = -

11.0 T is obtained from Fig. 5.3(c), where the negative sign indicates out of plane

anisotropy. For this film, we obtain Ku = -1.46 × 106 erg/cm3. Clearly, there is a strong

dependence of the anisotropy fields on the strain, in particular the c/a ratios of the

Sr2CrReO6 films, indicating a strong magnetization-lattice coupling. Using a simple

linear extrapolation of the anisotropy fields for the Sr2CrReO6 films on SrTiO3 (Hu = 18.1

57
T) and SCNO/LSAT (Hu = -11.0 T), we find that the MCA should change sign at 0.30%

tensile strain, which agrees with the value predicted by the GGA calculations [18].

Table 5.1 Lattice constants (a, c), tetragonal distortion (c/a), coercivity (Hc) and
magnetocrystalline anisotropy field (Hu) at T = 20 K and 300 K of the Sr2CrReO6 films
grown on LSAT, SrTiO3 and Sr2CrNbO6/LSAT; as is the in-plane substrate lattice
constant (doubled for LSAT and STO for easier comparison).

Substrate Hc (T) Hc (T) Hu (T) Hu (T)


or Buffer as (Å) a (Å) c (Å) c/a
Layer 20 K 300 K 20 K 300 K

LSAT 7.736 7.732 7.926 1.025 0.856 0.287 >10’s >10’s

STO 7.810 7.806 7.860 1.007 0.819 0.113 18.1 16.6

SCNO/LS
7.892 7.876 7.804 0.991 1.876 1.128 11.0 11.0
AT

Finally, the anisotropy fields remain essentially unchanged from T = 20 to 300 K, which

is promising for applications. It should be noted, however, that coercivity (Hc) decreases

with increasing temperature (Table 5.1). This can be understood as follows. The MCA is

mainly determined by the magnetic interaction, particularly SOC, and not defects in the

films. Since the magnetization of Sr2CrReO6 changes only slightly from 20 to 300 K

because of the high Tc [14,32], the MCA is essentially unchaged below 300 K.

Meanwhile, the coercivity is sensitive to defects in the films, thus Hc decreases at higher

temperatures due to increasing thermal energy.

58
To further understand the origin of the change in easy axis under strain seen in

experiment, we need to examine the difference in total energies for magnetic orientations

along different crystalline axes. This is known as magnetic anisotropy energy (MAE).

The origin of MAE resides in the spin-orbit coupling between the magnetization and the

lattice. The crystalline axis associated with the lowest energy determines the most

favorable direction of spontaneous magnetization (easy axis). A simple relation between

MAE and the difference between hard and easy axis orbital magnetic moments, which we

call moment anisotropy ΔmL, was proposed by Bruno [33],

𝜉
MAE = − 𝛼 Δ𝑚!
4𝜇!

where ΔmL is the difference between hard- and easy-axis orbital magnetic moments, µB is

the Bohr magneton, ξ is the spin-orbit coupling constant, and the prefactor α depends on

the electronic structure and is on the order of 0.05-1 [34]. This relation was shown to be

valid when the majority spin is fully occupied [33], which is the case for Sr2CrReO6, and

subsequent studies have confirmed it [35-37]. Eq. (1) suggests that when the moment

anisotropy is calculated for two directions, a sign change in the moment anisotropy

indicates a change in easy axis. Performing GGA+U calculations including spin orbit

coupling, for the different substrate-induced strain states (Supplemental Material Table

B.1 in Appendix B), we indeed find that Δmtot (defined as the difference between the

magnetic moment in the [001] direction and the average magnetic moment in the [100]

and [010] directions) changes sign with increasing c/a ratio [Fig. 5.4(a)], in agreement

with our experiments [Fig. 5.3]. This agreement confirms the validity of our theory

59
within numerical uncertainties. The calculated Δmtot values are much larger, by 1-2 orders

of magnitude, than those found in Ref. [35] for Ni and Fe, demonstrating the robust

nature of the magnetic anisotropy in Sr2CrReO6. However, the Δmtot values are small

enough that the precision required to measure them experimentally is not accessible by

current magnetometry techniques.

Figure 5.4 Differences (∆m) between out-of-plane ([001] direction and in plane (average
of [100] and [010] directions (a) total magnetic moments, (b) total orbital and spin
magnetic moments, and (c) spin and (d) orbital magnetic moments for Re (squares), Cr
(diamonds), and O (circles). The lines are guides for the eye.

Since the easy axis corresponds to the most energetically favorable axis (lowest energy),

Eq. (1) tells us that the easy axis aligns with the axis where the orbital magnetic moment

60
is the largest [38]. This overall agreement allows us now to examine the origin of the

easy-axis change. For that, we first separate the total moment anisotropy in spin (ΔmS)

and orbital (ΔmL) components [Fig. 5.4(b)] and find that they have approximately linear

strain dependence with opposite slopes, with the orbital moment anisotropy dominating.

Separating these contributions further according to their atomic origins, we find that for

the spin moment anisotropy [Fig. 5.4(c)], Cr and Re contribute at approximately the same

magnitude, but with opposite sign, thus canceling each other and leaving the small

contribution from the hybridizing O atoms as the dominant spin component. We have

previously observed experimentally an x-ray magnetic circular dichroism signal at the O

K edge (work performed at the Advanced Photon Source, Argonne National Laboratory),

complementing our theoretical findings that the O site carries a substantial magnetic

moment [39]. The overall dominant orbital moment anisotropy, however, is nearly

exclusively caused by the Re atoms, which contribute 93% to the combined ΔmL [Fig.

5.4(d)]. The calculated spin and orbital moments for the Cr, Re, and O atoms in

Sr2CrReO6 for each strain state (and both in plane and out of plane) can be found in

Supplemental Material Table B.2 in Appendix B.

Having identified the Re atoms as the origin of the magnetic anisotropy, we now can

further look into the orbital origin of the observed magnetic anisotropy. For that, we

examine the changes in calculated Re t2g (xy, yz, and xz) density of states (DOS) near the

Fermi level for tensile (c/a = 0.99) and compressive (c/a = 1.025) strain (Supplemental

Material Fig. B.3 in Appendix B). We observe a downward shift in energy for the xy

DOS (Supplemental Material Fig. B.3(a) in Appendix B) and an upward shift in energy

61
for both yz and xz DOS (Supplemental Material Fig. B.3(b) in Appendix B) when

transitioning from compressive to tensile strain, leading to a change in electron

distribution eventually resulting in the observed moment anisotropy.

In conclusion, we reveal large magnetic anisotropy and achieve dramatic changes in the

MCA of Sr2CrReO6 films via epitaxial strain. We use DFT calculations to understand the

structural distortions and to elucidate the dependence of MCA on the structure of the

films. A switching of the magnetic easy axis from in plane for compressive strain to out

of plane for tensile strain is observed via superconducting quantum interference device

(SQUID) magnetometry, and Sr2CrReO6 films exhibit some very high anisotropy fields.

Finally, we use first principles calculations to probe the atomic origins of the large and

tunable MCA in Sr2CrReO6, for which the anisotropy is driven primarily by Re orbitals.

A thorough understanding of the atomic magnetic behavior of this complex system can

guide the design of other versatile and applicable materials. Our results suggest that

substitution of other heavy transition metals, such as W or Os, for Re an drastically affect

the magnetic properties of FM oxides via tuning of the spin-orbit coupling.

References

[1] M. Seifert, V. Neu, and L. Schultz, Appl. Phys. Lett. 94, 022501 (2009).

[2] T. Shima, K. Takanashi, Y. K. Takahashi, and K. Hono, Appl. Phys. Lett. 85,
2571 (2004).
[3] J. H. Haeni, P. Irvin, W. Chang, R. Uecker, P. Reiche, Y. L. Li, S. Choudhury, W.
Tian, M. E. Hawley, B. Craigo, A. K. Tagantsev, X. Q. Pan, S. K. Streiffer, L. Q.
Chen, S. W. Kirchoefer, J. Levy, and D. G. Schlom, Nature 430, 758-761 (2004).

62
[4] J. Wang, J. B. Neaton, H. Zheng, V. Nagarajan, S. B. Ogale, B. Liu, D. Viehland,
V. Vaithyanathan, D. G. Schlom, U. V. Waghmare, N. A. Spaldin, K. M. Rabe,
M. Wuttig, and R. Ramesh, Science 299, 1719 (2003).

[5] P. V. Lukashev, N. Horrell, and R. F. Sabirianov, J. Appl. Phys. 111, 07A318

(2012).

[6] D. R. Behrendt, S. Legvold, and F. H. Spedding, Phys. Rev. 109, 1544 (1958).

[7] F. Y. Yang, C. L. Chien, E. F. Ferrari, X. W. Li, G. Xiao, and A. Gupta, Appl.


Phys. Lett. 77, 286 (2000).

[8] G. H. O. Daalderop, P. J. Kelly, and M. F. H. Schuurmans, Phys. Rev. B 44,


12054 (1991).

[9] B. Peters, A. Alfonsov, C. G. F. Blum, S. J. Hageman, P. M. Woodward, S.


Wurmehl, B. Büchner, and F. Y. Yang, Appl. Phys. Lett. 103, 162404 (2013).

[10] H. Kato, T. Okuda, Y. Okimoto, Y. Tomioka, K. Oikawa, T. Kamiyama, and Y.


Tokura, Phys. Rev. B 69, 184412 (2004).

[11] D. Serrate, J. M. De Teresa, and M. R. Ibarra, J. Phys.: Condens. Matter 19,


023201 (2007).

[12] K. I. Kobayashi, T. Kimura, H. Sawada, K. Terakura, and Y. Tokura, Nature 395,


677-680 (1998).

[13] Y. Krockenberger, K. Mogare, M. Reehuis, M. Tovar, M. Jansen, G.


Vaitheeswaran, V. Kanchana, F. Bultmark, A. Delin, F. Wilhelm, A. Rogalev, A.
Winkler, and L. Alff, Phys. Rev. B 75, 020404(R) (2007).

[14] A. J. Hauser, J. R. Soliz, M. Dixit, R. E. A. Williams, M. A. Susner, B. Peters, L.


M. Mier, T. L. Gustafson, M. D. Sumption, H. L. Fraser, P. M. Woodward, and F.
Y. Yang, Phys. Rev. B 85, 161201(R) (2012).

[15] J. M. Lucy, A. J. Hauser, H. L. Wang, J. R. Soliz, M. Dixit, R. E. A. Williams, A.


Holcombe, P. Morris, H. L. Fraser, D. W. McComb, P. M. Woodward, and F. Y.
Yang, Appl. Phys. Lett. 103, 042414 (2013).

[16] A. J. Hauser, J. M. Lucy, H. L. Wang, J. R. Soliz, A. Holcomb, P. Morris, P. M.


Woodward, and F. Y. Yang, Appl. Phys. Lett. 102, 032403 (2013).

[17] F. D. Czeschka, S. Geprägs, M. Opel, S. T. B. Goennenwein, and R. Gross, Appl.


Phys. Lett. 95, 062508 (2009).
63
[18] M. Komelj, Phys. Rev. B 82, 012410 (2010).

[19] X. Chen, D. Parker, K. P. Ong, M. H. Du, and D. J. Singh, Appl. Phys. Lett. 102,
102403 (2013).

[20] A. J. Hauser, R. E. A. Williams, R. A. Ricciardo, A. Genc, M. Dixit, J. M. Lucy,


P. M. Woodward, H. L. Fraser, and F. Y. Yang, Phys. Rev. B 83, 014407 (2011).

[21] C. H. Du, R. Adur, H. L. Wang, A. J. Hauser, F. Y. Yang, and P. C. Hammel,


Phys. Rev. Lett. 110, 147204 (2013).

[22] H. L. Wang, C. H. Du, Y. Pu, R. Adur, P. C. Hammel, and F. Y. Yang, Phys. Rev.
B 88, 100406(R) (2013).

[23] C. H. Du, H. L. Wang, Y. Pu, T. L. Meyer, P. M. Woodward, F. Y. Yang, and P.


C. Hammel, Phys. Rev. Lett. 111, 247202 (2013).

[24] See Supplemental Material at http://link.aps.org/supplemental/


10.1103/PhysRevB.90.180401 or Appendix B of this document for: (1) film
growth parameters, (2) methods used for SQUID magnetometry data analysis, (3)
tabularized structure, spin moment, and orbital moment parameters obtained from
DFT calculations, and (4) density of states obtained from DFT calculations.

[25] P. M. Woodward, Acta. Cryst. B 53, 44 (1997).

[26] J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J.


Singh, and C. Fiolhais, Phys Rev B 46, 6671–87 (1992).

[27] G. Kresse and J. Hafner, Phys. Rev. B 49, 14251-14269 (1994).

[28] G. Kresse and J. Hafner, Phys. Rev. B 47, 558 (1993).

[29] P. E. Blöchl, Phys. Rev. B 50, 17953–17979 (1994).

[30] V. I. Anisimov, J. Zaanen, and O. K. Andersen, Phys Rev B 44, 943–954 (1991).

[31] H.-T. Jeng and G. Y. Guo, Phys. Rev. B 67, 094438 (2003).

[32] J. M. De Teresa, J. M. Michalik, J. Blasco, P. A. Algarabel, M. R. Ibarra, C.


Kapusta, and U. Zeitler, Appl. Phys. Lett. 90, 252514 (2007).

[33] P. Bruno, Phys. Rev. B 39, 865 (1989).

64
[34] O. Hjortstam, K. Baberschke, J. M. Wills, B. Johansson, and O. Eriksson, Phys.
Rev. B 55, 15026 (1997).

[35] I. Yang, S. Y. Savrasov, and G. Kotliar, Phys. Rev. Lett. 87, 216405 (2001).

[36] F. Wilhelm, P. Poulopoulos, P. Srivastava, H. Wende, M. Farle, K. Baberschke,


M. Angelakeris, N. K. Flevaris, W. Grange, J.-P. Kappler, G. Ghiringhelli, and N.
B. Brookes, Phys. Rev. B 61, 8647 (2000).

[37] S. Gold, E. Goering, C. König, U. Rüdiger, G. Güntherodt, and G. Schütz, Phys.


Rev. B 71, 220404 (2005).

[38] H.-T. Jeng and G. Y. Guo, Phys. Rev. B 65, 094429 (2002).

[39] A. J. Hauser, J. M. Lucy, M. W. Gaultois, M. R. Ball, J. R. Soliz, Y. Choi, O. D.


Restrepo, W. Windl, J. W. Freeland, D. Haskel, P. M. Woodward, and F. Yang,
Phys. Rev. B 89, 180402 (2014)

65
Chapter 6. Magnetic Structure in Epitaxially Strained

Sr2CrReO6 Thin Films by Element-specific XAS and XMCD

Note: This chapter is being presented in the format of a journal article. It was published

as Hauser, A. J., Lucy, J. M., Gaultois, M. W., Ball M. R., Soliz, J. R., Choi, Y.,

Restrepo, O. D., Windl, W., Freeland, J. W., Haskel, D., Woodward, P. M. & Yang, F.

Magnetic structure in epitaxially strained Sr2CrReO6 thin films by element-specific XAS

and XMCD. Physical Review B. 89 180402(R) (2014). All of the DFT calculations

including the structural relaxations and partial density of states to find the number of

holes for each system were performed by Molly Ball with assistance from Oscar

Restrepo.

Abstract

We have analyzed the magnetic configuration for highly-ordered Sr2CrReO6 films as a

function of epitaxial strain using magnetometry and x-ray magnetic circular dichroism

(XMCD) measurements of Cr, Re, and O sites. The in-plane magnetic moments change

significantly when tensile strain is applied. O K edge XMCD indicates the O sites carry

at least a portion of the bulk magnetization. Spin moment values measure for Cr match

calculations incorporating spin-orbit effects, while both spin and orbital moments
66
measure for Re sites are slightly higher than previously predicted. Finally, we discuss

large changes in the x-ray absorption near-edge structure that are observed at the Cr and

Re L edges due to epitaxial strain.

Sr2CrReO6 has proven to be a highly interesting double perovskite material due to the

convergence of many attractive properties: a Curie temperature well above 300 K[1-4]

enabling room temperature applications, a scientifically interesting double exchange

phenomenon that suggests high spin-polarization for spintronic applications[5-7],

significant spin-orbit interactions and correlated electron behavior due to the presence of

Re 5d orbitals, and the recently discovered semiconducting behavior in highly ordered

epitaxial films which, combined with high temperature ferrimagnetism, may be useful in

nonvolatile logic devices[3,8]. Progress toward device applications, however, will require

an in-depth understanding of the electronic and magnetic configurations of the system.

Theoretical modeling has suggested a double perovskite double exchange model, wherein

the Cr 3d and Re 5d orbitals are hybridized via the oxygen 2p orbitals [5,6,9,10]. In these

models, the rocksalt ordered Cr and Re atoms have oppositely aligned spin moments and

almost exclusively account for the magnetization in the system. Previously, element-

specific x-ray magnetic circular dichroism (XMCD) measurements of the Re site in

polycrystalline bulk samples have been studied [11], but to date such work on highly

ordered crystalline materials has not been undertaken, and no measurements to

characterize the Cr or O sites in bulk or thin film have been reported. The L edge x-ray

absorption near-edge structure (XANES) of Cr and Re can be analyzed to provide


67
significant information regarding orbital configuration and bonding environments. By

applying varying epitaxial strain to Sr2CrReO6 thin films, we can draw conclusions from

the system’s reaction to the induced distortions. These results are needed to properly test

our understanding of this material system.

Epitaxial Sr2CrReO6 films were grown by off-axis magnetron sputtering as previously

described and characterized [3] on (LaAlO3)0.3(Sr2AlTaO6)0.7 (LSAT) and SrTiO3 (STO)

substrates, as well as on LSAT substrate with a fully relaxed Sr2CrNbO6 (SCNO) buffer

layer. These substrates impart 0.78% compressive (LSAT), 0.17% tensile/nominally

relaxed (STO), and 1.09% tensile (SCNO/LSAT) strains upon their respective Sr2CrReO6

films. All films were grown to a thickness of 90 nm, as confirmed by modeling of Laue

oscillations seen in x-ray diffraction spectra. A Quantum Design Superconducting

Quantum Interference Device (SQUID) magnetometer was used to characterize the in-

plane magnetization.

Element-specific x-ray absorption spectroscopy (XAS) and XMCD measurements of the

Cr L, O K and Re L edges were then taken at the soft (Cr, O, 4-ID-C) and hard (Re, 4-ID-

D) x-ray beam lines at the Advanced Photon Source, Argonne National Laboratory. The

soft x-ray measurements were taken in the surface sensitive total electron yield mode,

with the applied magnetic field offset ten degrees from the plane of the film surface. The

hard x-ray measurements were taken in glancing angle fluorescence mode to optimize

probed volume at the higher energy of Re L edges, with the plane of the film tilted 3.8(1)

degrees with respect to the applied magnetic field. Edge energies were calibrated with

standards at known energies (e.g., Cr2O3 for the Cr L edge, and W foil for the Re L edge).

68
Systematic error introduced by the placement of the absorption edge step is the dominant

source of error in our XMCD measurements, so we have analyzed the data for the range

of all reasonable step edge centroid energies, and report the average moment and error

values consummate with the range of possible moments calculated. Within the ranges of

error we report, the choice of placing the step edge at the white line maximum intensity

will generally correspond with the lower magnitude moment values. Placing the step

edge at the maximum derivative of the XAS (center of rising edge) with respect to energy

will yield magnitude moment values in the upper end of the error range. The hole

number values used in this work are the result of theoretical calculations and cannot be

assigned a reasonable uncertainty value, but should be noted as a possible cause of

discrepancies with future work.

69
0.8

Magnetization (µB /f.u.)


0.4
T = 200 K
0.0 H = 3.5 T

-0.4
LSAT
STO
-0.8 SCNO/LSAT

-8 -6 -4 -2 0 2 4 6 8
Magnetic Field (T)

Figure 6.1 In-plane magnetic hysteresis loop at T = 200 K for films on (001)-oriented
LSAT (green circles), SrTiO3 (black squares), and ~200 nm Sr2CrNbO6 buffer layer on
LSAT (blue triangles).

In-plane magnetic hysteresis loops for the three Sr2CrReO6 films at temperature T = 200

K are shown in Figure 6.1. Experimental limitations of the 4-ID-D beamline at the time

of the experiment imposed a maximum measurement field of 3.5 T, which suffices to

largely saturate each film at 200 K. At 3.5 T (shown in Figure 6.1 as a vertical dashed

line), the magnetization of films grown on LSAT, STO and SCNO/LSAT are 0.76, 0.79,

and 0.36 µB per formula unit (f.u.), respectively. These numbers represent the

“macroscopic” total moment of the Sr2CrReO6 unit cell and will be compared to the

individual Cr and Re moments. It should be noted that the film under large tensile strain

(Sr2CrReO6/SCNO/LSAT) has a relatively small in-plane moment compared to

compressive (Sr2CrReO6/LSAT) or small tensile (Sr2CrReO6/STO) strains. Such a

70
significant drop in moment is not due to loss of crystallinity or decreased Cr/Re cation

order as indicated by our previous work [3,12,13], which evidences high crystallinity in

all strained films by x-ray diffraction.

Figure 6.2 shows a full summary of the Cr L edge, Re L edge, and O K edge XANES and

XMCD spectra for all three films. All L edge data is normalized to a 2:1 L3/L2 edge jump

ratio. Spin (mS) and orbital (mL) moments were then extracted by sum rule analysis

[14,15]. Due to the narrow 2p spin-orbit splitting in Cr and the resultant quantum

mechanical mixing of j1/2 and j3/2 excitations, a spin correction factor of 2 was applied for

all three Cr spectra [16]. To analyze these data, we use density functional theory (DFT)

calculations to calculate the number of holes for each system. We use projector

augmented wave (PAW) pseudopotentials [17] as implemented in the Vienna ab initio

simulation package (VASP) [18,19]. Correlation effects were treated within the

generalized gradient approximations (GGA) +U approach [20] with a value of U = 3 eV

and an exchange parameter J = 0.87 eV for the Cr d-orbitals [21]. We have relaxed the

atomic positions within the Sr2CrReO6 unit cells for the different strains using the

experimental lattice parameters. The number of valence electrons occupying an orbital

can be obtained by integrating its partial density of states up to the Fermi level, which

subtracted from 10, gives the number of holes. We obtain 5.69 (5.17), 5.72 (5.28), and

5.68 (5.28) holes for Cr (Re) in Sr2CrReO6/LSAT, Sr2CrReO6/SCNO/LSAT, and

Sr2CrReO6/STO, which lead to the calculated averaged spin moments of 1.48 (-0.68),

1.50 (-0.69), and 1.49 (-0.68), respectively. The measured spin and orbital moments for

the Cr and Re sites are shown in their respective spectra.

71
We again note the stark difference in the results of the Sr2CrReO6 films on LSAT and

STO as compared to the film on SCNO/LSAT. On the Re site, both the spin and orbital

moments are reduced by roughly a factor of 3, and the Cr spin moment sharply decreases

in magnitude. Given this information and the shape of the hysteresis loop, it is likely that

a strong anisotropy turns the magnetic easy axis out-of-plane and, consequently, an in-

plane field of 3.5 T cannot align the Sr2CrReO6 magnetization. Determination of

magnetocrystalline anisotropy requires careful comparison of the in-plane and out-of-

plane magnetizations, coercive fields, and anisotropy fields, and is the subject of a more

detailed study [22].

72
10 (a) SCRO/LSAT(001) (b) SCRO/STO(001) (c) SCRO/SCNO/LSAT(001)
Cr L3
Cr L3
Cr L3
Normalized Intensity
L2 L2
L2
5

0
ms = 1.10(1) µB/Cr ms = 1.26(1) µB/Cr ms = 0.58(1) µB/Cr
m = 0.039(2) µ /Cr m = -0.041(2) µ /Cr mL = -0.012(1) µB/Cr
L B L B
-5
575 580 585 590 575 580 585 590 575 580 585 590
Energy (eV) Energy (eV) Energy (eV)

(d) SCRO/LSAT(001) (e) SCRO/STO(001) (f) SCRO/SCNO/LSAT(001)


3
Re L3 L2
L2 L2
Normalized Intensity

2
Re L3 Re L3

0
ms = -0.76(4) µB/Re x4 m = -0.77(3) µ /Re ms = -0.24(1) µB/Re x4
s B x4
-1 mL = 0.24(1) µB/Re mL = 0.26(1) µB/Re mL = 0.079(4) µB/Re

10.5 10.6 11.9 12.0 10.5 10.6 11.9 12.0 10.5 10.6 11.9 12.0
Energy (keV) Energy (keV) Energy (keV)
2.5
SCRO/LSAT(001) SCRO/STO(001) SCRO/SCNO/LSAT(001)
(g) (h) (i)
2.0
Normalized Intensity

O K edge O K edge O K edge


1.5

1.0

0.5

0.0

-0.5
525 530 535 540 525 530 535 540 525 530 535 540
Energy (eV) Energy (eV) Energy (eV)

Figure 6.2 Normalized XANES and XMCD spectra at the (a-c) Cr L, (d-f) Re L, and (g-
i) O K edges. XANES/XMCD spectra are on top/bottom within each figure. Spin and
orbital moments are given for the Cr and Re spectra.

Figures 6.2(a) and (b) show the Cr XMCD spectra for films grown on LSAT and STO,

respectively, which indicate strongly quenched orbital moments and spin moments
73
antiparallel to that of Re, as expected in ferrimagnetic Sr2CrReO6. The respective spin

moments of 1.10(1) µB/f.u. (LSAT) and 1.26(1) µB/f.u. (STO) for Cr are smaller than our

calculated values (1.48 and 1.49, µB/f.u., respectively). However, it is important to note

that the calculated values represent a material system at T = 0 K and full magnetic

saturation. It is clear from Figure 6.1 that at T = 200 K, an applied field of 3.5 T does not

fully saturate any film in this study. From previous work, we can conservatively expect a

10-15% loss as compared to the calculated magnetization due to thermal energy, and

another 10-15% loss due to incomplete magnetization [3]. This accounts for the

difference between experiment and theory, even without considering the small decrease

in magnetization expected from Cr/Re anti-site disorder.

Figures 6.2(d) and (e) show the Re XMCD spectra for films grown on LSAT and STO,

which yield mS and mL values matching closely with previous theoretical predictions

[10,23] as well as our calculations and bulk powder XMCD results [11]. The opposite

signs of the spin and orbital moments are as expected for a less than half-filled shell with

spin-orbit coupling. However, by applying the same considerations in the Cr spectra with

incomplete magnetic saturation at finite measurement temperature, the projected “fully

saturated” spin and orbital moment values may be slightly higher than previously

predicted.

When the combined spin and orbital moments for Cr and Re are added together, the

XMCD spectra imply total moments of 0.62(6) and 0.71(5) µB/f.u. for films on LSAT

and STO, respectively. Comparing these values to the macroscopic magnetizations found

74
by SQUID magnetometry, we find that 0.14(6) µB/f.u. (LSAT) and 0.08(5) µB/f.u. (STO)

are unaccounted for in the “B-site only” model.

There are two primary sources for the missing moment. It could either simply reside on

the O 2p orbitals, and/or the neglect of a magnetic dipole moment Tz in the spin sum rule

for Re may contribute to the unaccounted moment [15]. The latter is possible since

sizable spin-orbit interactions are known to occur in 5d elements such as Re [24], and can

contribute to Tz to some degree. Deviations in the octahedral environment may also

contribute to Tz. Although strain is well known to produce effects such as octahedral

rotations, tilting, and Jahn-Teller distortions [25,26], we see a similar large moment

discrepancy for strained and unstrained films alike.

Between these two, our data indicate that magnetic moments on the O 2p orbitals should

be the dominant contribution, since we found evidence of magnetic moment in the

oxygen 2p orbitals by measuring the O K edge spectra, as shown in Figures 6.2(g)- 6.2(i)

for all three films. Quantitatively, our DFT calculations yield a total moment of 0.02

µB/oxygen for all three epitaxial strains, which matches well with the equivalent missing

moment (~0.01-0.03 µB/oxygen) in our experiment. The existence of moments on the

oxygen sites can be understood by the fact that when the double exchange model is

applied to double perovskite systems, there is a high degree of orbital hybridization of B-

site cations (in this case, Cr and Re) mediated by each intervening oxygen [6]. Our

findings are also consistent with previous work that found that CrO2 had significant spin

moments on the oxygen site due to Cr 3d-O 2p hybridization [10,27,28]. It is not

uncommon for double perovskites to have oxygen with magnetic moment above this

75
range. For instance, our results are in line with previous work that found that Sr2FeMoO6

has a magnetic moment on oxygen between 0.06 µB/oxygen and 0.09 µB/oxygen [29].

(a) eg* LSAT


10 STO
SCNO/LSAT
Normalized Intensity
8 Cr L3
t2g eg*
6
t2g L2
4

0
575 580 585 590
Energy (eV)
3
(b) t2g

eg*
Normalized Intensity

Re L2

LSAT
STO
SCNO/LSAT
1
11.94 11.96 11.98
Energy (keV)

Figure 6.3 Normalized x-ray absorption spectra for different strains at the (a) Cr L2,3
edges and (b) Re L2 edge.

76
Figures 6.3(a) and (b) show the overlaid Cr L and Re L2 x-ray absorption spectra for all

three films, which result from excitation of 2p electrons to unoccupied d states. The line

shapes are sensitive to changes in coordination geometry, and the edge absorption energy

given by the inflection point is sensitive to changes in bonding and charge density

[30,31].

Figure 6.3(a) shows the Cr L3 edge, where the absorption energy increases by 0.25 eV

under both ~1% compressive and tensile strain when compared to the nominally

unstrained Sr2CrReO6 film grown on STO. The absorption peak maximum exhibits an

identical shift. The change in the Cr L3 edge energy is small compared to shifts seen in

Cr compounds due to changes in Cr formal charge [32], suggesting epitaxial strain has

only a weak influence on the Cr ground state electron density. In general, a decrease in

Cr L edge absorption energy is caused by higher charge density on Cr, which leads to a

lower effective nuclear charge. This destabilizes the electronic levels and decreases the

energy required to promote core electrons to unoccupied states.

The Cr L edge also undergoes minor changes in spectral line shape with strain, but the

line shapes of 3d L edge spectra are complicated by multiplet effects. These multiplet

effects are caused by the overlap of partially filled core and valence wave functions in the

final state, which makes interpretation difficult by altering the line shape and intensity

ratios of the resulting spectra [33]. (The large 2p spin-orbit coupling in 5d systems

reduces the influence of multiplet effects in Re 5d L edge spectra discussed later, which

makes interpretation of Re L edge spectra more feasible.) The t2g/eg* intensity ratio in the

Cr L2 edge is ~1.03 in epitaxial Sr2CrReO6 films on LSAT and STO, but 1.11 in films on

77
SCNO/LSAT, a 7.7% change. Similarly, the L3 t2g/eg* intensity ratio increases by 5.2%

on SCNO/LSAT. Although there is a significant change in the line shape of the films

grown under 1.09% tensile strain on SCNO/LSAT, these films are complicated by the

presence of the strong magnetic anisotropy discussed earlier. No significant change in

line shape is seen in films grown on LSAT under 0.78% compressive strain. The L3/L2

integrated area branching ratios are 1.47 and 1.48 for films grown on STO and LSAT,

and 1.64 for films grown on SCNO/LSAT. Overall, there are only small changes between

films grown on LSAT and those grown on STO, and although there are slightly larger

changes in films grown on SCNO/LSAT, drawing conclusions from the Cr L edge is not

straightforward given the many aforementioned considerations.

The Re L edge spectra of Sr2CrReO6 films show more significant changes. The Re L3

edge absorption energy (given by the inflection point) decreases by ~1 eV under ~1%

applied tensile or compressive strain (0.78% compressive strain: ΔE(STO – LSAT) = –1.2 eV,

1.09% tensile strain: ΔE(STO – SCNO/LSAT) = –1.0 eV). The L2 edge line shape experiences a

corresponding change; the intensity of the primary transition decreases substantially in

strained films, shown in Figure 6.3(b), and the L3 edge white line intensity also decreases

by ~45%. The L3/L2 integrated area branching ratios are 1.99 and 2.02 for films grown

on STO and LSAT, and 1.62 for films grown on SCNO/LSAT.

In general, the probability of a transition is proportional to the number of unoccupied

states, so the intensity of the primary absorption feature (i.e., the L edge white line) is

sensitive to changes in the electronic structure. Under most circumstances, the decrease in

absorption energy and white line intensity would suggest an increase in the ground state

78
charge density of Re, though the magnitudes of the changes are unexpectedly large given

that there is no change in the formal charge of Re. The stability of the Re valence states

is supported by DFT calculations performed here, which suggest the number of electrons

per Re site increases by 0.1 from films grown on STO to films grown on LSAT.

Furthermore, any change in the Re charge would be difficult to reconcile with other

sample considerations. Maintaining charge neutrality with considerably more valence

electrons in Re would require either oxidation of the Cr – there is no appreciable change

in Cr valence states seen by Cr L edge XANES – or formation of oxygen vacancies

commensurate with the change in the Re valence state. However, based on the work by

Clancy et al., the change in valence state correlating to the observed absorption energy

shift would require nearly two oxygen vacancies per unit cell [34]. Such a large

concentration of vacancies would likely have resulted in significant changes to the

material properties of the system, but we see no such effects. For instance, oxygen

vacancies have been previously reported to lead to rapid relaxation of epitaxial strain in

Sr2CrReO6 films [13]. However, off-axis x-ray diffractometry reveals the films in the

present work are coherently strained to the substrate, even at film thicknesses of 90 nm.

In the case of Sr2CrReO6, it is more likely that these changes are the result of the

tetragonal distortion of Re centers induced by epitaxial strain. The tetragonal distortion

breaks the degeneracy of unoccupied Re states and leads to broadening of the resulting

transitions. In a simple crystal field interpretation, tetragonal distortion splits the partially

occupied Re octahedral t2g states, leading to non-degenerate unoccupied states. Instead of

79
a single transition with high intensity in unstrained Sr2CrReO6, there are several

transitions with less intensity in strained samples.

These changes can occur without any change in the Re formal charge. Analogous results

have been observed in Al K-edge investigations of regular and distorted Al3+ octahedral

centers in oxides, where distortion decreases the Al K-edge absorption energy by ~2 eV

and splits the primary absorption feature [35]. In fact, we estimate that the integrated

white line intensity shown in that paper decreased by approximately 50% due to

octahedral distortion, roughly similar to our observations in Sr2CrReO6. The same

phenomenon is also seen in Ti L-edge spectra of octahedral Ti4+ centers in SrTiO3

(regular) and TiO2 (distorted) [36].

In summary, we have analyzed a strain-dependent series of Sr2CrReO6 films via in-plane

SQUID magnetometry and element-specific XANES/XMCD analysis of the Cr and Re

L2,3 edges. We find a significant change in the in-plane magnetic moment by both

techniques when tensile strain is epitaxially applied. For nominally relaxed and

compressively strained films, Cr-site XMCD shows a slightly lower spin moment than

previously predicted, but matches well with calculations incorporating spin-orbit effects.

Re-specific XMCD yields spin and orbital moments that agree well with previous theory

and bulk powder results, but considerations of incomplete magnetic saturation suggest

calculations may underestimate the ideal spin and orbital moment values. Interestingly,

we find a small moment (0.01-0.03 µB/oxygen) at the oxygen site, which is confirmed by

both theory and experiment. Finally, comparative analysis of Cr and Re L edge XANES

80
shows that epitaxial strain leads to large changes in the Re ground state, while Cr is

largely unaffected.

References

[1] D. Serrate, J.M. De Teresa, and M.R. Ibarra, J. Phys. Condens. Matter 19, 023201
(2007).

[2] H. Kato, T. Okuda, Y. Okimoto, Y. Tomioka, Y. Takenoya, A. Ohkubo, M.


Kawasaki, and Y. Tokura, Appl. Phys. Lett. 81, 328 (2002).

[3] A.J. Hauser, J.R. Soliz, M. Dixit, R.E.A. Williams, M.A. Susner, B. Peters, L.M.
Mier, T.L. Gustafson, M.D. Sumption, H.L. Fraser, P.M. Woodward, and F.Y.
Yang, Phys. Rev. B 85, 161201 (2012).

[4] H. Asano, N. Kozuka, A. Tsuzuki, and M. Matsui, Appl. Phys. Lett. 85, 263
(2004).

[5] G. Vaitheeswaran, V. Kanchana, and A. Delin, Appl. Phys. Lett. 86, 032513
(2005).

[6] O.N. Meetei, O. Erten, A. Mukherjee, M. Randeria, N. Trivedi, and P.


Woodward, Phys. Rev. B 87, 165104 (2013).

[7] S.A. Wolf, D. Awschalom, R. Buhrman, J. Daughton, S. von Molnár, M. Roukes,


A. Chtchelkanova, and D. Treger, Science (80-. ). 294, 1488 (2001).

[8] K. Ando, Science (80-. ). 312, 1883 (2006).

[9] H. Das, P. Sanyal, T. Saha-Dasgupta, and D.D. Sarma, Phys. Rev. B 83, 104418
(2011).

[10] G. Vaitheeswaran, V. Kanchana, M. Alouani, and A. Delin, EPL (Europhysics


Lett. 84, 47005 (2008).

[11] P. Majewski, S. Gepra gs, O. Sanganas, M. Opel, R. Gross, F. Wilhelm, A.


Rogalev, and L. Alff, Appl. Phys. Lett. 87, 202503 (2005).

[12] J.M. Lucy, A.J. Hauser, H.L. Wang, J.R. Soliz, M. Dixit, R.E.A. Williams, A.
Holcombe, P. Morris, H.L. Fraser, D.W. McComb, P.M. Woodward, and F.Y.
Yang, Appl. Phys. Lett. 103, 042414 (2013).

81
[13] A.J. Hauser, J.M. Lucy, H.L. Wang, J.R. Soliz, A. Holcomb, P. Morris, P.M.
Woodward, and F.Y. Yang, Appl. Phys. Lett. 102, 032403 (2013).

[14] B.T. Thole, P. Carra, F. Sette, and G. van der Laan, Phys. Rev. Lett. 68, 1943
(1992).

[15] C. Chen, Y. Idzerda, H. Lin, and N. Smith, Phys. Rev. Lett. 75, 152 (1995).

[16] E. Goering, Philos. Mag. 85, 2895 (2005).

[17] P.E. Blochl, Phys. Rev. B 50, (1994).

[18] G. Kresse and J. Hafner, Phys. Rev. B 47, (1993).

[19] G. Kresse and J. Hafner, Phys. Rev. B 49, (1994).

[20] V.I. Anisimov, J. Zaanen, and O.K. Andersen, Phys. Rev. B 44, (1991).

[21] H.-T. Jeng and G. Guo, Phys. Rev. B 67, 094438 (2003).

[22] J. M. Lucy, M. R. Ball, O. D. Restrepo, A. J. Hauser, J. R. Soliz, J. W. Freeland,


P. M. Woodward, W. Windl, and F. Y. Yang, Phys. Rev. B 90, (2014).

[23] V. Kanchana, G. Vaitheeswaran, and M. Alouani, J. Phys. Condens. Matter 18,


5155 (2006).

[24] L. Mattheiss, Phys. Rev. 151, 450 (1966).

[25] P.M. Woodward, Acta Crystallogr. Sect. B Struct. Sci. 53, 32 (1997).

[26] P.M. Woodward, Acta Crystallogr. Sect. B Struct. Sci. 53, 44 (1997).

[27] D. Huang, H.-T. Jeng, C. Chang, G. Guo, J. Chen, W. Wu, S. Chung, S. Shyu, C.
Wu, H.-J. Lin, and C. Chen, Phys. Rev. B 66, 174440 (2002).

[28] E. Goering, A. Bayer, S. Gold, G. Schutz, M. Rabe, U. Rudiger, and G.


Guntherodt, 58, 906 (2002).

[29] R. Mishra, O.D. Restrepo, P.M. Woodward, and W. Windl, Chem. Mater. 22,
6092 (2010).

[30] R. Leapman, L. Grunes, and P. Fejes, Phys. Rev. B 26, 614 (1982).

[31] J.G. Chen, Surf. Sci. Rep. 30, 1 (1997).

82
[32] T.L. Daulton and B.J. Little, Ultramicroscopy 106, 561 (2006).

[33] F. de Groot, Coord. Chem. Rev. 249, 31 (2005).

[34] J.P. Clancy, N. Chen, C.Y. Kim, W.F. Chen, K.W. Plumb, B.C. Jeon, T.W. Noh,
and Y.-J. Kim, Phys. Rev. B 86, 195131 (2012).

[35] J. van Bokhoven A., H. Sambe, D. Koningsberger C., and D. Ramaker E., J. Phys.
IV Fr. 7, 835 (1997).

[36] R. Laskowski and P. Blaha, Phys. Rev. B 82, 205104 (2010).

83
Chapter 7. Cs2AgBiX6 (X = Br, Cl) – New Visible Light

Absorbing, Lead-free Halide Perovskite Semiconductors

Note: This chapter is being presented in the format of a journal article. It was published

as McClure, E. T., Ball, M. R., Windl, W. & Woodward, P. M. Cs2AgBiX6 (X = Br, Cl) –

New visible light absorbing, lead-free halide perovskite semiconductors. Chemistry of

Materials. 28 1348-1354 (2016). I performed all electronic band structure and density of

states calculations. The carrier mobility analysis was also done by me.

7.1 Abstract

The double perovskites Cs2AgBiBr6 and Cs2AgBiCl6 have been synthesized from both

solid state and solution routes. X-ray diffraction measurements show that both

compounds adopt the cubic double perovskite structure, space group Fm 3 m , with lattice

parameters of 11.2711(1) Å (X = Br) and 10.7774(2) Å (X = Cl). Diffuse reflectance

measurements reveal band gaps of 2.19 eV (X = Br) and 2.77 eV (X = Cl) that are

slightly smaller than the band gaps of the analogous lead halide perovskites, 2.26 eV for

CH3NH3PbBr3 and 3.00 eV for CH3NH3PbCl3. Band structure calculations indicate that

the interaction between the Ag 4d-orbitals and the 3p/4p-orbitals of the halide ion

84
modifies the valence band leading to an indirect band gap. Both compounds are stable

when exposed to air, but Cs2AgBiBr6 degrades over a period of weeks when exposed to

both ambient air and light. These results show that halide double perovskite

semiconductors are potentially an environmentally friendly alternative to the lead halide

perovskite semiconductors.

7.2 Introduction

In a remarkably short period of time, metal-halide perovskites have gone from relative

obscurity to an intensely studied class of materials. The efficiencies of solar cells

constructed from metal-halide perovskites have risen from 3.8% in a dye-sensitized solar

cell configuration [1], to NREL certified 20.1% in a planar heterojunction cells [2]. The

most studied materials by far are the AMX3 perovskites where A is an alkyl ammonium

cation, such as CH3NH3+, M is Pb2+, and X is a halide ion (I−, Br−, Cl−). The allure of the

organohalide perovskites for photovoltaic and LED applications stems from the fact that

they contain cheap, earth-abundant elements and are amenable to a variety of processing

techniques, including solution-based deposition methods [3].

Metal-halide perovskite solar cells have already reached efficiency levels that are

commercially viable. The remaining obstacles to commercialization are threefold: (i)

develop cost effective, large scale manufacturing routes, (ii) improve the moisture

stability so that modules can operate in an outdoor environment without resorting to

expensive encapsulation methods, and (iii) discover metal-halide alternatives that don’t

contain environmentally harmful elements like lead [3]. The first obstacle is an industrial

85
challenge that lies outside of the scope of fundamental research, but the latter two

challenges fall squarely in the realm of materials chemistry. While there are many reports

of CH3NH3PbX3 perovskites being stable for hundreds of hours when left exposed to the

air in the dark [4,5], their stability plummets when simultaneous exposed to moisture and

sunlight [6]. Regarding the presence of lead there are varying opinions in the scientific

community about the viability of working with lead. Lead is present in everyday items

like car batteries and solder, and equally toxic elements like Cd can be found in

commercial photovoltaic modules [7,8]. At the same time much of the world is trying to

move away from materials containing lead. While there may be a place for applications

of the lead halide perovskites in solar energy conversion devices, there can be little doubt

that discovery of a lead-free material whose properties matched perovskites like

CH3NH3PbI3 or CH3NH3PbBr3 would be a boon. In short that is the goal of this work, to

find metal-halide perovskites whose properties approximate those of the extensively

studied CH3NH3PbX3 perovskites without incorporating toxic elements like Pb, Cd, Tl, or

Hg.

The Sn2+ ion is most obvious substitute for Pb2+, and tin-halide perovskites with attractive

optical and electrical transport properties can be made [9-11]. Unfortunately, the tin-

halide perovskites are extremely unstable in air, degrading over time even when handled

in an inert atmosphere glovebox. It seems unlikely that the Sn-based perovskites will

ever have the stability to be commercially viable alternatives to the Pb-based perovskites.

The role of the organic CH3NH3+ cation seems to be less critical. The optical band gaps

of CsPbBr3 (2.25 eV) [12] and CsPbCl3 (2.97 eV) [13] are quite similar to their

86
organohalide analogs CH3NH3PbBr3 (2.26 eV) and CH3NH3PbCl3 (3.00 eV), vide infra.

Furthermore, Kulbak, et al. have shown that the efficiencies of solar cells made from

CsPbBr3 are comparable to equivalent cells made from CH3NH3PbBr3, with the

advantage that the all-inorganic CsPbBr3 cells have better thermal stabilities and can

operate longer without degrading [14,15]. Unfortunately the perovskite polymorph of

CsPbI3, which has the most suitable band gap of the CsPbX3 family for solar cell

applications, is very difficult to stabilize at room temperature [9]. Hence, the role of the

CH3NH3+ ion is largely to stabilize the perovskite form of APbI3 at room temperature.

Simple combinatorics of ionic charges in the A+M2+X3 formula does not lend much hope

that alternatives to the APbX3 perovskites with comparable band gaps, carrier mobilities,

and good moisture stability can be found. There are few if any suitable 2+ cations with a

d10s2 or d10s0 configuration that are both nontoxic and stable against oxidation.

Fortunately, the situation improves if we expand the search from ternary A+M2+X3

perovskites to quaternary A2M+M3+X6 double perovskites. There are numerous

combinations of 1+ and 3+ ions with suitable electron configurations that are quite stable

when exposed to the air, such as Cu+, Ag+, Bi3+, Sb3+, and In3+. In this paper we

document our initial investigations into the structural and optical properties of metal-

halide double perovskites by reporting on the synthesis and properties of Cs2AgBiX6 (X

= Br, Cl). These compounds have band gaps and moisture stability that are comparable to

their CH3NH3PbX3 analogs, demonstrating that double perovskites are a promising

alternative to lead halide perovskites.

87
7.3 Experimental Section

Starting materials CsCl and CsBr were prepared by reacting Cs2CO3 (99+%, Strem

Chemicals) with HCl (Sigma-Aldrich, 37%) and HBr (Fluka, ≥ 48%), respectively. The

solutions were evaporated, and the resulting solids were filtered and washed with ethanol.

AgCl and AgBr were precipitated from as prepared aqueous solutions of AgNO3

(99.9+%, Alfa Aesar) and NaCl (ACS Reagent, GFS Chemicals) or KBr (99+%, Alfa

Aesar). BiBr3 was prepared by reacting Bi2O3 (≥99.0%, J.T. Baker) with HBr (Fluka, ≥

48%). The mixture was heated until fully dissolved, evaporated to dryness, and then

filtered and washed with ethanol. Pure BiCl3 could not be prepared from this method,

thus a commercial reagent was used (≥98%, Aldrich).

Polycrystalline Cs2AgBiX6 samples were prepared by precipitation from a solution of the

hydrohalic acid and hypophosphorous acid. A mixture of 8 mL of 12.1 M HCl (or 8.84 M

HBr) and 2 mL of a 50% solution of H3PO2 was added to a round bottom flask and

heated to 120 °C. Then 1.89 mmol of AgCl and an equal amount of BiCl3 (1.41 mmol

AgBr and BiBr3) were dissolved in the hot solution. Once dissolved and well mixed, 3.78

mmol CsCl (2.82 mmol CsBr) was added to the flask, immediately triggering a

precipitate reaction. The precipitate was collected on filter paper, washed with ethanol,

and then dried overnight.

Polycrystalline Cs2AgBiX6 samples were prepared via a solid state route as well by

mixing cesium, silver, and bismuth halide salts in a 2:1:1 molar ratio. Reagents were

88
ground together for 20 minutes and placed in an alumina crucible as a loose powder. The

samples were heated in a box furnace in air at 210 °C for 10 hours. It was found that at

least two heating cycles with grinding in-between was needed to obtain nearly phase pure

samples.

X-ray powder diffraction (XRPD) data were collected on a Bruker D8 powder

diffractometer (40 kV, 50 mA, sealed Cu X-ray tube) equipped with an incident beam Ge

111 monochromator and Lynx Eye position sensitive detector. Rietveld refinements of

laboratory XRPD data were carried out using TOPAS Academic software package to

determine the crystal structure [16]. A Rigaku MiniFlex II bench top X-ray powder

diffractometer (30 kV, 15 mA, sealed Cu X-ray tube) with a NaI scintillation detector

was also used for phase identification, pattern indexing, and for probing air and light

stability of the synthesized compounds.

UV-Visible diffuse reflectance data were collected over the spectral range of 200-1100

nm with an Ocean Optics USB4000 spectrometer equipped with a Toshiba TCD1304AP

(3648-element linear silicon CCD array). The spectrometer was used in conjunction with

an Ocean Optics DH-2000-BAL deuterium and halogen UV-Vis-NIR light source and a

400µm R400-7-ANGLE-VIS reflectance probe.

Electronic band structure and density of states (DOS) calculations were obtained with

density functional theory implemented in the Vienna Ab-initio Simulation Package

[17,18]. These calculations were performed using projector augmented wave (PAW)

potentials [19] based on the PBE exchange-correlation functional [20] within the hybrid

HSE hybrid approach [21-23] with an exact exchange fraction of 0.26. Cutoff energies of
89
250 and 280 eV were used for Cs2AgBiBr6 and Cs2AgBiCl6, respectively. The

calculations include spin-orbit coupling.

7.4 Results

7.4.1 Synthesis

When synthesizing the bromine compound in the solid state, evidence for the perovskite

phase was not seen in the XRPD pattern until the sample was heated to 150 °C. At a

heating temperature of 185 °C the target compound was the dominant phase in the

sample mixture. After heating to 210 °C the yellow-orange sample was nearly phase

pure.

For the solid state synthesis of the chlorine compound, evidence of the perovskite phase

was seen upon room temperature grinding of the reactants. Since bismuth (III) chloride

is hygroscopic, the sample mixture took on moisture from the air while being ground,

causing the sample to clump together in a gray mass that had the appearance of wet clay.

The sample was left to sit on the lab bench for 10-15 minutes, and during that time the

gray mass dried out. Additional grinding led to the formation of a fine powder with a pale

yellow color, similar to the final product. Evidently the moisture taken on by the

reagents promotes reactivity and is then displaced once the double perovskite phase

forms. However, two cycles of heating to 210 °C are still needed to obtain phase pure

samples.

90
7.4.2 Crystal Structure

The refined fits to the XRPD patterns of Cs2AgBiBr6 and Cs2AgBiCl6 are shown in

Figure 7.1. Both compounds adopt the cubic double perovskite structure with Fm 3 m

space group symmetry. Reflections with odd-odd-odd Miller Indices signaling a rock salt

ordering of Ag+ and Bi3+ ions are readily apparent. Close inspection of the XRPD

patterns reveals no signs of secondary phases.

91
Figure 7.1 Rietveld refinements of the XRPD patterns for Cs2AgBiCl6 (upper) and
Cs2AgBiBr6 (lower). The black dots, red and green lines are the experimental pattern,
calculated fit, and difference curve, respectively. Red tick marks at the bottom give the
expected peak positions. The inset shows the fit to the (111) and (200) reflections.

92
Table 7.1 details of the X-ray powder diffraction experiments and Rietveld refinements.
Cs2AgBiCl6 Cs2AgBiBr6

Crystal system Cubic

Space group Fm-3m (#225)

a (Å) 10.7768(1) 11.2712(1)

V (Å3) 1251.64 1431.904


−3
Density (g cm ) 4.22 (1) 4.926(8)

Radiation Cu Kα (Å) 1.5418

Collection Limits (° 2θ) 8.00 – 90.00

Step Size (° 2θ) 0.01460

Rwp 20.379 20.963

χ2 1.148 1.163

Structural parameters extracted from the refinements of Cs2AgBiCl6 and Cs2AgBiBr6 are

given in Table 7.1, while atomic coordinates and displacement parameters are given in

Table 7.2. A visualization of the double perovskite structure is shown in Figure 7.2.

Like the basic perovskite structure there is a three-dimensional framework of corner

connected octahedra, with Cs+ ions occupying the cuboctahedral cavities in the

framework. The double perovskite structure is then obtained by alternating Ag+ and Bi3+

centered octahedra in all three directions building up a superstructure that is typically

referred to as rock salt ordering.

Bond distances and bond valence sums are tabulated in Table 7.3. In both compounds

the Ag+ and Bi3+ ions show complete ordering into a rock-salt supercell of the simple

perovskite structure. The halide ion undergoes small displacements toward the bismuth

93
site, leading to Bi−X distances that are slightly shorter than the Ag−X distances. The

possible presence of Ag/Bi antisite disorder was investigated in the Rietveld refinements,

but the occupancies did not refine to values that indicated anything other than complete

ordering. Given the similarity of the Bi−X and Ag−X distances it is somewhat surprising

that antisite mixing between the two ions is not observed. The relatively low bond

valence sum for the Cs+ ion and rather large values of the displacement parameters (Beq)

suggest that fairly large dynamic rotations of the octahedra are taking place at room

temperature. It would not be surprising to see a phase transition triggered by cooperative

octahedral tilting take place upon cooling below room temperature.

Table 7.2 Refined atomic positions and atomic displacement parameters for Cs2AgBiCl6
(upper) and Cs2AgBiBr6 (lower, in italics).
Wyckoff
Site x y z Beq
site

2.0(2)
Ag 4a 0 0 0
1.8(2)

0.98(7)
Bi 4b 0.5 0 0
1.09(8)

2.68(9)
Cs 8c 0.25 0.25 0.25
3.4(1)

Cl 0.2513(7) 3.0(2)
24e 0 0
Br 0.2503(4) 3.66(8)

94
Figure 7.2 Refined crystal structure of Cs2AgBiCl6. The Cs+ ions are shown as gray
spheres, the chloride ions as small green spheres, while the Ag and Bi centered octahedra
are shown as blue and green polyhedra, respectively.

Table 7.3 Bond lengths and bond-valence sums.


X = Cl X = Br

Bond lengths (Å)

Cs – X (× 12) 3.810(5) 3.985(3)

Ag – X (× 6) 2.708(8) 2.821(4)

Bi – X (× 6) 2.680(8) 2.814(4)

Bond valence sums

Cs 0.76 0.73

Ag 1.13 1.18

Bi 3.49 3.55

X 1.02 1.03

95
7.4.3 Optical Properties

In order to determine the optical band gaps UV-Vis diffuse reflectance spectra of the

samples were measured (Figure 7.3). Spectra for the methylammonium lead halide

samples are also shown (in red) for comparison. Aside from some minor differences seen

at photon energies above the absorption onset there are striking similarities between the

Cs2AgBiX6 double perovskites and their CH3NH3PbX3 analogs. However, the onset of

absorption is not quite as sharp in the double perovskites. We believe this change can be

attributed to the indirect band gap of the double perovskites. This hypothesis is

supported by band structure calculations discussed below.

96
Figure 7.3 Diffuse reflectance spectra for Cs2AgBiCl6 and CH3NH3PbCl3 (top), and
Cs2AgBiBr6 and CH3NH3PbBr3 (bottom).

In order to extract an optical band gap the reflectance data were transformed to pseudo-

absorption data using the Kubelka-Munk equation, which expresses the absorbance as a

function of reflectance: F(R) = α = (1-R)2/(2R), where R is the reflectance and α is the

optical absorption coefficient. Kubleka-Munk plots for the silver-bismuth and lead

compounds are shown in the Supplemental Information in Appendix C. The Tauc

Method is then applied to the Kubelka Munk transformed data. It utilizes the equation

[F(R)hν]1/ϒ vs. hν, where ϒ is equal to ½ for direct transitions (applicable to


97
CH3NH3PbX3 compositions) and 2 for indirect transitions (applicable to Cs2AgBiX6

compositions). Using this approach band gaps of 2.19 eV and 2.77 eV are extracted for

Cs2AgBiBr6 and Cs2AgBiCl6, respectively. Meanwhile, those for CH3NH3PbBr3 and

CH3NH3PbCl3 are found to be 2.26 eV and 3.00 eV.

7.4.4 Band Structure Calculations

Band structure calculations were performed to further investigate the electronic structure

of the two compounds. Calculations were also performed on lead halide systems for

comparison using Cs+ as the A-site cation for simplicity. The approach is justified by the

establishment fact that neither the CH3NH3+ nor the Cs+ ions make significant

contributions to the band structure near the Fermi level [12,24-26]. The space group

symmetry for the lead compounds was taken to be cubic with Pm-3m symmetry, both for

ease of comparison with the cubic double perovskites and because the average structures

of CH3NH3PbBr3 and CH3NH3PbCl3 are both cubic. The band structure diagrams for the

chlorides are shown in Figure 7.4, and the bromides in Figure 7.5.

98
Figure 7.4 Band Structure diagrams for Cs2AgBiCl6 (top) and cubic CsPbCl3 (bottom).
The Fermi energy is set to E = 0 and denoted with a dashed line.

99
The calculated band gaps are in good agreement with the experimental values (Table

7.4). While the band structures for the Cs2AgBiX6 and lead-halide systems contain many

similarities, there is one significant difference: the double perovskites possess an indirect

band gap as opposed to the direct band gap seen for the lead halide perovskites. This

comes from the fact that the valence band maximum has moved away from the (111)

Brillouin zone boundary (the R point in a primitive cubic cell and the L point in a face

centered cubic cell, respectively) to the X-point. Calculations performed with and without

spin-oribit coupling show that the band degeneracy present at the Γ-point of the

conduction band is lifted when spin-orbit coupling is included. This leads to the

emergence of a narrow heavy electron band that is separated from the two next higher

energy bands by 1 to 1.5 eV.

In order to estimate the carrier mobilities, we determined effective charge carrier masses

from the curvature of the band extrema (assuming parabolic bands). Along the R to X

direction for CsPbCl3 (CsPbBr3) the electron effective mass is 0.41me (0.34me) and the

hole effective mass along the same direction is 0.35me (0.37me). By comparison

Cs2AgBiCl6 (Cs2AgBiBr6) has an approximate electron effective mass of 0.53me

(0.37me) along the L to W direction, and a hole effective mass of 0.15me (0.14me) along

the X to Γ direction. These estimates are only a qualitative comparison since they

neglects possible differences in carrier scattering rates. Although these double

perovskites have a less disperse conduction band, it is encouraging that estimates of the

hole effective masses for the double perovskites are lighter than their lead analogs.

100
Figure 7.5 Band Structure diagrams for Cs2AgBiBr6 (top) and cubic CsPbBr3 (bottom).
The Fermi energy is set to E = 0 and denoted with a dashed line.

The valence to conduction band transition is primary from filled halogen 3p/4p states to

antibonding Ag 5s and Bi 6p states. This can be seen by looking at the partial density of

101
state diagrams (PDOS) in Figure 7.6. The participation of both Ag and Bi orbitals to the

lower energy conduction bands is a key to maintaining some dispersion of the conduction

band. Although the valence band is largely halogen 3p/4p in character there is extensive

admixture of Ag 4d states, which results in the presence of several relatively flat bands at

energies between −1.5 and −4 eV. The presence of Ag 4d states plays a role in reducing

the band gap and is partially responsible for the indirect band gap of these phases.

Table 7.4 Observed and calculated optical band gaps.

Compound Experimental Calculated

CH3NH3PbCl3 3.00 eV 2.97 eV

Cs2AgBiCl6 2.77 eV 2.62 eV

CH3NH3PbBr3 2.26 eV 2.25 eV

Cs2AgBiBr6 2.19 eV 2.06 eV

102
Figure 7.6 Atomic partial density of states plots for Cs2AgBiCl6 (top) and Cs2AgBiBr6
(bottom).

7.4.5 Chemical and Light Stability

To probe the stability of the compounds, samples were exposed to an ambient atmosphere

in both light and dark conditions. Each compound was loaded into a flat sample holder

103
used for XRPD measurements and then left exposed to the ambient air. Visual inspection

of those samples that were kept in the dark for two weeks, showed no apparent change.

Subsequent XRPD measurements, confirmed that no change had occurred. When the

samples were left sitting next to a large glass window, thereby exposing them to both

atmosphere and visible light, darkening of the exposed surface was observed, particularly

for Cs2AgBiBr6. It should be noted that, during the month of exposure, the samples only

received about six hours of direct exposure each day. The samples were analyzed by

XRPD and UV-Vis diffuse reflectance spectroscopy after two and four weeks exposure to

sunlight. As can be seen in Figure 7.7, the overall features of the reflectance spectrum

for Cs2AgBiCl6 remain intact, but the total reflectance decreases ever so slightly, in

agreement with the visual darkening of the surface of the sample that was observed. No

apparent change was detected in the XRPD patterns of Cs2AgBiCl6 that were exposed to

both light and air.

104
Figure 7.7 UV-Vis diffuse reflectance spectra showing the light stability of Cs2AgBiCl6
after two and four weeks of light exposure.

The reflectance spectrum for Cs2AgBiBr6, shown in Figure 7.8, changes drastically after

two weeks exposure to light and ambient air, accompanied by the appearance of

unidentified phases in the XRPD (see Supporting Information in Appendix C). These

observations clearly show degradation of the double perovskite phase over time when

simultaneously exposed to light and moisture.

105
Figure 7.8 UV-Vis Diffuse spectra showing the light instability of Cs2AgBiBr6 after two
and four weeks of light exposure.

7.5 Conclusions

Given the challenges of finding air-stable, non-toxic alternatives to the lead halide

perovskites, it’s highly encouraging to find double perovskites with band gaps that are

comparable to their CH3NH3PbX3 analogs. Given the intense level of interest and

research activity being devoted to the lead halide perovskites the emergence of double

perovskites that possess comparable optical properties is significant.

While the structure and properties of Cs2AgBiX6 double perovskites are exciting as a

proof of concept there are still challenges to be overcome. The mixing of silver 4d

orbitals with the halogen 3p/4p orbitals changes the valence bands sufficiently that an

indirect band gap is observed. While not ideal for solar cell applications it should be

noted that the most widely used material in photovoltaic cells, silicon, has an indirect

106
gap, so this does not completely rule out their use in solar cells. The air and moisture

stability are promising, but the instability of Cs2AgBiBr6 to simultaneous exposure would

necessitate encapsulation for long term use. It should be noted that its stability is similar

to that of CH3NH3PbBr3 and CH3NH3PbI3. Finally, for single junction cells there is a

need to reduce the band gap. However, metal-halide perovskites with band gaps of ~2.2

eV, such as CH3NH3PbBr3 and Cs2AgBiBr6, are of interest for use with silicon in tandem

solar cells [3,14].

It is worthwhile to compare the properties of Cs2AgBiBr6 with those of the defect ordered

perovskites A3Bi2I9 and A3Sb2I9 (A = Cs, Rb, K), which have also been investigated as

lead-free alternatives to CH3NH3PbI3 and CH3NH3PbBr3. The band gaps of Cs3Bi2I9 (1.9

eV), Rb3Bi2I9 (2.1 eV), K3Bi2I9 (2.1 eV) and Cs3Sb2I9 (2.05 eV) are slightly smaller but

comparable to Cs2AgBiBr6 [27-29]. While the A3Bi2I9 and A3Sb2I9 compounds have

direct or near direct band gaps, their band dispersions are quite small which will translate

to heavier carrier effective masses than estimated here for Cs2AgBiBr6 [28,29]. Saparov

et al. estimated electron and hole effective masses in the [100] direction where they are

smallest, to be 0.44me and 0.60me, respectively [29]. Both values are heavier than those

estimated for Cs2AgBiBr6, particularly the hole effective mass. Although they have not

been estimated from calculations, the effective carrier masses for phases appear to be

even heavier.

Finally, it is important to remember that Cs2AgBiCl6 and Cs2AgBiBr6 are just two

members of what is likely to be a large family of halide double perovskites. The

replacement of Bi3+ with ions like Sb3+ and In3+, or replacing Ag+ with ions like Cu+ and

107
Au+, is an obvious strategy for tuning the electronic structure to narrow the band gap and

move back toward a direct band gap. Double perovskites such as Cs2NaSbCl6 and

Cs2NaInCl6 have been reported; however, little is known about their optical properties

[30], and chemical substitutions such as Ag+ for Na+ and Br− for Cl− have not been

explored. The results reported here provide a compelling motivation to prepare and

characterize additional halide double perovskites.

References

[1] Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal halide
perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc.
2009, 131 , 6050–6051.
[2] National Renewable Energy Labs (NREL) efficiency chart (2015);
http://www.nrel.gov/ncpv/images/efficiency_chart.jpg (accessed 30 October
2015).
[3] Stranks, S. D.; Snaith, H. J. Metal-halide perovskites for photovoltaic and light-
emitting devices. Nat. Nanotechnol. 2015, 10, 391–402.
[4] Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon,
S.-J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J. E.; Grätzel, M.; Park, N.-G.
Lead iodide perovskite sensitized all solid state submicron thin film mesoscopic
solar cell with efficiency exceeding 9%. Sci. Rep. 2012, 2, 591.
[5] Noh, J. H.; Im, S. H.; Heo, J. H.; Mandal, T. N.; Seok, S. Il. Chemical
management for colorful, efficient and stable inorganic-organic nanostructured
solar cells. Nano Lett. 2013, 13, 1764–1769.
[6] Leijtens, T.; Eperon, G. E.; Pathak, S.; Abate, A.; Lee, M. M.; Snaith, H. J.
Overcoming ultraviolet light instability of sensitized TiO2 with meso-
superstructured organometal trihalide perovskite solar cells. Nat. Commun. 2013,
4, 2885.
[7] Fthenakis, V. M.; Kim, H. C.; Alsema, E. Emissions from photovoltaic life
cycles. Environ. Sci. Technol. 2008, 42, 2168–2174.
[8] Held, M.; Ilg, R. Update of environmental idicators and energy payback time of
CdTe PV systems in Europe. Prog. Photovoltaics Res. Appl. 2011, 19, 614–626.

108
[9] Stoumpos, C. C.; Malliakas, C. D.; Kanatzidis, M. G. Semiconducting tin and
lead perovskites with organic cations: Phase transitions, high mobilities, and near-
infrared photoluminescent properties. Inorg. Chem. 2013, 52, 9019–9038.
[10] Noel, N. K.; Stranks, S. D.; Abate, A.; Wehrenfennig, C.; Guarnera, S.;
Haghighirad, A.-A.; Sadhanala, A.; Eperon, G. E.; Pathak, S. K.; Johnston, M. B.;
Petrozza, A.; Herz, L. M.; Snaith, H. J. Lead-free organic–inorganic tin halide
perovskites for photovoltaic applications. Energy Environ. Sci. 2014, 7, 3061.
[11] Hao, F.; Stoumpos, C. C.; Cao, D. H.; Chang, R. P. H.; Kanatzidis, M. G. Lead-
free solid-state organic–inorganic halide perovskite solar cells. Nat. Photonics
2014, 8, 489–494.
[12] Stoumpos, C. C.; Malliakas, C. D.; Peters, J. A.; Liu, Z.; Sebastian, M.; Im, J.;
Chasapis, T. C.; Wibowo, A. C.; Chung, D. Y.; Freeman, A. J.; Wessels, B. W.;
Kanatzidis, M. G. Crystal growth of the perovskite semiconductor CsPbBr3: A
new material for high-energy radiation detection. Cryst. Growth Des. 2013, 13,
2722–2727.
[13] Myagkota, S.; Voloshinovskii, A.; Stefanskii, I.; Mikhailik, M.; Pashuk, I.
Reflection and emission properties of lead-based perovskite-like crystals. Radiat.
Meas. 1998, 29, 273–277.
[14] Kulbak, M.; Cahen, D.; Hodes, G. Planar How important is the organic part of
lead halide perovskite solar cells? Efficient CsPbBr3 solar cells. J. Phys. Chem.
Lett. 2015, 6, 2452–2456.
[15] Kulbak, M.; Gupta, S.; Kedem, N.; Levine, I.; Bendikov, T.; Hodes, G.; Cahen,
D. Cesium enhances long-term stability of lead bromide perovskite-based solar
cells J. Phys. Chem. Lett. 2016, 7, 167−172.
[16] Topas Academic, General Profile and Structural Analysis for Powder Diffraction
Data; Bruker AXS: Karlsruhe, Germany 2004.
[17] Kresse, G.; Hafner, J. Ab initio Molecular-Dynamics for Liquid- Metals. Phys.
Rev. B 1993, 47, 558–561.
[18] Kresse, G.; Hafner, J. Ab Initio Molecular-Dynamics Simulation of the Liquid-
Metal Amorphous-Semiconductor Transition in Germanium. Phys. Rev. B 1994,
49, 14251–14269.
[19] Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B. 1994. 50, 17953-
17979.
[20] Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made
simple. Phys. Rev. Lett. 1996, 77, 3865–3868.

109
[21] Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid Functionals Based on a Screened
Coulomb Potential. J. Chem. Phys. 2003, 118, 8207–8215.
[22] Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid Functionals Based on a Screened
Coulomb Potential. J. Chem. Phys. 2006, 124, 219906.
[23] Paier, J.; Marsman, M.; Hummer, K.; Kresse, G.; Gerber, I. C.; Angyan, J. G.
Screened Hybrid Density Functionals Applied to Solids. J. Chem. Phys. 2006,
125, 249901.
[24] Giorgi, G.; Fujisawa, J. I.; Segawa, H.; Yamashita, K. Small photocarrier
effective masses featuring ambipolar transport in methylammonium lead iodide
perovskite: A density functional analysis. J. Phys. Chem. Lett. 2013, 4, 4213–
4216.
[25] Melissen, S. T. A. G.; Labat, F.; Sautet, P.; Le Bahers, T. Electronic properties of
PbX3CH3NH3 (X = Cl, Br, I) compounds for photovoltaic and photocatalytic
applications. Phys. Chem. Chem. Phys. 2015, 17, 2199-2209.
[26] Yin, W.-J.; Yang, J.-H.; Kang, J.; Yan, Y.; Wei, S.-H. Halide perovskite materials
for solar cells: a theoretical review J. Mater. Chem. A 2015, 3, 8926-8942.
[27] Park, B. W.; Philippe, B.; Zhang, X.; Rensmo, H.; Boshchloo, G.; Johansson, E.
M. J. Bismuth Based Hybrid Perovskites A3Bi2I9 (A: Methylammonium or
Cesium) for Solar Cell Application, Adv. Mater. 2015, 27, 6806–6813.
[28] Lehner, A. J.; Fabini, D. H.; Evans, H. A.; Hebert, C. A.; Smock, S. R.; Hu, J.;
Wang, H.; Zwanziger, J. W.; Chabinyc, M. L.; Seshadri, R. Crystal and Electronic
Structures of Complex Bismuth Iodides A3Bi2I9 (A = K, Rb, Cs) Related to
Perovskite: Aiding the Rational Design of Photovoltaics, Chem. Mater. 2015, 27,
7137−7148.
[29] Saparov, B.; Hong, F.; Sun, J.-P.; Duan, H.-S.; Meng, W.; Cameron, S.; Hill, I.
G.; Yan, Y., Mitzi, D. B. Thin-Film Preparation and Characterization of Cs3Sb2I9:
A Lead-Free Layered Perovskite Semiconductor, Chem. Mater. 2015, 27,
5622−5632.
[30] Morss, L. R.; Siegall, M.; Stenger, L.; Edelstein, N. Preparation of cubic chloro
complex compounds of trivalent metals: Cs2NaMCl6 Inorg. Chem.,1970, 9, 1771-
1775.

110
Chapter 8. Summary

In this dissertation, I have described computational work that has guided and explained

experimentation for functional materials within the perovskites and double perovskites

families. I have tried to convey how a large number of crucial questions can only be

examined on an atomic level with the help of electronic-structure density functional

theory (DFT) calculations, and how the calculation results have led to new levels of

understanding and improved routes to discover even better materials. I have also

discussed what level of DFT calculations is necessary in order to produce reliable and

realistic properties, which often require special treatment of spin interactions, spin-orbit

coupling, and electron correlations for highly localized d-electrons.

For discovery of novel phenomena and great property tunability, I have then shown that

double perovskites are an ideal family of materials, since a huge number of transition

metal ions can be incorporated in their structure, frequently also alloyed with other ions,

which results in a wide range of electronic and magnetic as well as multiferroic

properties, making this one of the most comprehensive materials classes existing. Within

this thesis, I have examined double perovskites with

• previously unseen exchange interactions, leading to non-interaction sublattices

never observed before;

111
• record-breaking magnetocrystalline anisotropy and its strain tunability; and

• lead-free halide composition, suggesting an air-stable and less toxic alternative to

one of the most exciting new material discoveries in the field of photovoltaics.

For all these materials and effects, we have developed a detailed understanding of the

phenomena observed and values measured based on my computational results.

To discuss these results in more detail, we found novel magnetic behavior in Sr2CoOsO6,

which experiences antiferromagnetic ordering of its two transition metal ions completely

independently from each other. The Co and Os sublattices were found to be almost

entirely uncoupled from each other, which results in a magnetic structure that has a

ground state with non-collinear spins. DFT+U was used to extract the exchange

interactions quantitatively within a Heisenberg-type model to better understand the

competing forces in this structure. It was found that the nearest neighbor Co-O-Os are

significantly weaker than the long-range Co-O-Os-O-Co and Os-O-Co-O-Os interactions,

leading to weak inter-sublattice interactions. These weak interactions in turn were

examined by density of states calculations with DFT, which showed that there is little to

no overlap between the electronic states on Co and Os atoms, leading to weak interaction.

Within the second body of work described within this thesis, I have described how thin

films of Sr2CrReO6 were found to have extraordinarily large anisotropy fields and

magnetocrystalline anisotropy that is strain-tunable. These films were grown on three

substrates, LSAT, STO, and SCNO/LSAT, causing strain ranging from compressive to

tensile causing tetragonal distortions in the structure and the transition metal and oxygen

octahedral to rotate optimizing the local environment. Across this strain range, the
112
magnetic easy axis was found to change from in-plane for LSAT and STO to out-of-plane

for SCNO/LSAT. While it is thus clear that the applied strain and the resulting internal

distortions that the structure experiences drive the switching in magnetocrystalline

anisotropy, the exact mechanism could not be understood by experiment alone. Thus,

DFT was used better understand the exact dependencies. The origin of the change in easy

axis was identified by looking at the total energies for magnetic orientations along

different crystalline axes. We have then proposed that a simple relationship between the

magnetic anisotropy energy and the atomic moments, originally developed for metallic

materials, can be also applied to these double perovskites, and that the sign change in

moment anisotropy can be used to identify when the easy axis switches. The difference in

moment values was calculated and found to be an order of magnitude larger than any

materials previously reported, making this a record-breaking case. Orbital magnetism on

the Re atoms were identified to be the origin of the magnetic anisotropy, since all other

contributions canceled out in magnitude. As a deeper-lying cause, we found in the

calculated DOS that there was an opposite shift in energy between the in-plane and out-

of-plane d-orbital peaks, indicating a shift in electron distribution, which then is at the

heart of the anisotropic behavior.

The magnetic configuration and its changes with strain in Sr2CrReO6 analyzed were then

further investigated with element-specific XANES/XMCD, which confirmed the

theoretical results, especially the large strain dependence of the Re ground state. As an

interesting side result, it was confirmed that a small moment (0.01-0.03 µB/oxygen)

should reside at the oxygen site, which has been also found in our calculations. This was

113
an important result with relevance to many other calculations for magnetic oxide

materials, since it wasn’t clear previously if this was real or an artifact from partitioning

the overall moment between the different atoms.

Lastly, I explored the lead-free halides, Cs2AgBiBr6 and Cs2AgBiCl6. These double

perovskites are a part of a very promising class of materials. They were investigated to

help solve several obstacles still remaining in the commercialization of perovskite solar

cells, including high fabrication costs, moisture instabilities, and environmentally harmful

element inclusion. Finding a material that doesn’t have these issue and also remains as

efficient as existing solar cells has been a challenge. DFT calculations including spin-

orbit coupling were performed and optimized to give results very similar to the

experimentally measured electronic properties. Band structures and density of states were

presented and compared to the lead-containing analogs of these materials. While it was

found that the band gaps were comparable in size to that of the lead containing halides,

the materials possessed an indirect band gap, which is not ideal for efficient solar cells.

Since we could show that these materials can be described well within optimized DFTs,

in a follow up project, we explore the composition space in order to find optimal

materials with direct band gaps that one day may revolutionize solar cell materials in a

way similar to nitrides revolutionizing the LED field.

114
Appendix A. Supplementary Information for Chapter 4

A.1 X-ray Powder Diffraction

Figure A.1 Refined powder X-ray diffraction pattern of Sr2CoOsO6. Data was collected
using a Bruker D8 Advance equipped with a copper source and Ge (111)
monochromator. Black symbols observed data, while red and blue curves indicate the
calculated pattern and difference respectively.

115
A.2 Neutron Powder Diffraction

Neutron diffraction data was used to refine the crystal structure of Sr2CoOsO6 at 300,

130, 80 and 12 K. The magnetic structure was also refined at the latter two temperatures.

Using peak splitting and systematic absences the space group symmetry at 300 and 130 K

was determined to be tetragonal I4/m, while at 80 and 12 K it was monoclinic I112/m.

The results of the Rietveld refinements are given in Table A.1. Isotropic displacement

parameters were used for Sr, Co, and Os atoms, and anisotropic displacement parameters

for O atoms.

116
Table A.1 Refined atomic positions and displacement parameters for longer measuring
time data sets where applicable. Os atomic positions for both space groups remain at
Wyckoff site 2a (0,0,0) while Co atomic positions remain at Wyckoff site 2b (0,0,½) for
both space groups. Ueq is defined as one third of the trace of the Uij tensor, and it is given
for the oxygen atoms for which anisotropic displacement parameters were refined.

12K 80K 130K 300K


Space Group I112/m I112/m I4/m I4/m
Sr x 0.5 0.5 0.5 0.5
Sr y 0 0 0 0
Sr z 0.2502(2) 0.2499(3) 0.25 0.25
O(1) x 0 0 0 0
O(1) y 0 0 0 0
O(1) z 0.2424(2) 0.2423(2) 0.2416(2) 0.2421(2)
O(2) x 0.2049(3) 0.2056(3) 0.2053(2) 0.2111(2)
O(2) y 0.2810(3) 0.2806(3) 0.2780(2) 0.2732(2)
O(2) z 0 0 0 0
O(3) x 0.2777(3) 0.2775(3) - -
O(3) y 0.7976(3) 0.7969(3) - -
O(3) z 0 0 - -
Sr Uiso 0.0025(1) 0.0034(2) 0.0042(1) 0.0101(2)
Co Uiso 0.0014(6) 0.0020(7) 0.0025(6) 0.0064(7)
Os Uiso 0.0010(1) 0.0014(2) 0.0018(2) 0.0044(2)
O(1) Ueq 0.0044(5) 0.0054(6) 0.0066(2) 0.0148(3)
O(2) Ueq 0.0040(3) 0.0048(4) 0.0060(2) 0.0128(3)
O(3) Ueq 0.0044(3) 0.0052(4) - -

Constraints were used to avoid correlations among the variables that describe the

direction of the magnetic moment. The components of the moments in the z-direction for

both osmium and cobalt were fixed to be zero based on the fact that any moment oriented

117
in this direction produced intensity for Bragg peaks that are not present in the observed

diffraction patterns. In addition, mx1 = my2 = −mx1 = −my2 where subscripts x and y

indicate the component of the magnetic moment along those respective directions, and

the subscripts 1 and 2 refer to the two different osmium atoms in the magnetic unit cell.

Similar constraints were used for the cobalt moments, mx3 = mz4 = −mz3 = −mx4 and my3

= my4 = 0, where subscripts 3 and 4 refer to the two different cobalt atoms in the unit cell.

Table A.2 Select bond lengths and angles for Sr2CoOsO6 as obtained from Rietveld
refinements of the neutron diffraction data.

12K 80K 130K 300K


Co−O1 (×2) 2.0672(13) 2.0677(14) 2.0684(16) 2.052(2)
Co−O2 (×4) --- --- 2.0361(13) 2.0379(15)
Co−O2(a) (×2) 2.0564(16) 2.0539(18) --- ---
Co−O2(b) (×2) 2.0128(19) 2.013(2) --- ---
Os−O1 (×2) 1.9457(13) 1.9440(14) 1.9340(16) 1.927(2)
Os−O2 (×4) --- --- 1.9072(13) 1.9151(15)
Os−O2(a) (×2) 1.902(16) 1.9032(14) --- ---
Os−O2(b) (×2) 1.9056(19) 1.907(2) --- ---

∠Co−O1−Os 180 180 180 180


∠Co−O2−Os --- --- 163.45(4) 165.83(6)
∠Co−O2(a)−Os 163.02(7) 163.22(7) --- ---
∠Co−O2(b)−Os 162.49(6) 162.76(6) --- ---

118
Figure A.2 Rietveld refinement histograms at (a) 300 K and (b) 12K where black
symbols represent observed data points, red curves for the refined fit and blue for the
difference curve. Pink hash marks indicate reflections from the vanadium sample can,
black tick marks indicate reflections from the nuclear structure, and green tick marks
correspond to magnetic reflections.

119
200

150

Temperature (K)
100

50

1.90 1.92 1.94 1.96 1.98 2.00 2.02 2.04 2.06

d spacing (Å)

Figure A.3 Colorfill plot of observed variable temperature NPD data revealing the
splitting of the (220) peak which is associated with the structural phase transition
lowering the symmetry of the structure from tetragonal I4/m to monoclinic I2/m at
approximately 110K.

Figure A.4 The structure of Sr2CoOsO6 at 12 K in space group I2/m. The view looking
down [001] on the left shows out-of-phase octahedral rotations about the c-axis, while the
view looking down [110] on the right shows the lack of octahedral tilting about the a- or
b-axes (right). The higher temperature I4/m structure has an identical tilting pattern.

120
A.3 Electrical Conductivity Measurements

Transport measurements clearly indicate insulating behavior as resistivity rises rapidly

with decreasing temperature. This data was found not to follow a linear relationship with

T-1 as one would find with activated transport in a conventional semiconductor, but rather

to be linear with T-1/4 consistent with variable range hopping transport, whereby electrons

hop between localized states. These linear relationships can be fitted via equation A.1

below to extract the A coefficient. There is a deviation from the linear fitting of the

model which occurs near the structural phase transition, leading to two separate linear

fits, above and below this deviation. The fitted A factors found above and below this

transition temperature are 43.12(2) and 46.14(9) respectively.

!/! )
σ=σ0𝑒 (!!/! (A.1)

121
Figure A.5 Electrical transport behavior of Sr2CoOsO6 as shown by log resistivity vs

temperature.

122
A.4 Density Functional Theory Calculations

Additional details about the DFT calculations and the method used to fit the exchange

constants are given below.

A.4.1 Collinear Magnetic Calculations

Figure A.6(a) shows the experimentally observed low temperature magnetic

configuration, where the Co and Os spins order independently with the Co (Os) moments

aligned parallel or antiparallel along a particular direction in the ac- (ab-) plane. Based

upon their independent ordering, as observed experimentally, if the nearest neighbor Co

and Os spins are hypothesized to be non-interacting or only weakly interacting with each

other, it would be a reasonable argument to make their moments collinear, i.e. make them

align parallel or antiparallel while lying on the same plane. In order to test this

hypothesis, we have performed collinear spin-polarized DFT calculations (excluding

spin−orbit coupling effects) with ten different magnetic configurations: (a) ferromagnetic

(FM), where the spins on all the Co and Os ions are aligned parallel to each other, (b)

ferrimagnetic (FiM) (AF1), where the spins on Co ions are aligned parallel to each other

and antiparallel to the spins on Os ions (consequently the spins on Os ions are aligned

parallel to those on other Os ions), (c) C-type antiferromagnetic, with chains of parallel

aligned Co and Os ions along the c-axis and the neighboring chains being aligned

antiparallel to each other, (d) A-type antiferromagnetic (AF2), where the spins on Co and

Os ions lying on any equatorial plane (ab-plane) are parallel to each other while the

ordering between any two neighboring ab-planes is anti-parallel, (e) the collinear version

of the experimentally observed low temperature magnetic configuration (Exp(col)), as


123
shown in Figure A.6(b) and (f) five more antiferromagnetic structures (AF3-AF7) shown

in Figure A.11.The calculations were performed with the following three combinations of

(UCo, UOs): (0, 0), (4.1, 2.1) and (4.1, 4.1) eV. For all the three combinations, the energy

of Exp(col) was found to be the lowest making it the ground state. This also lends support

to our conjecture that the Co and Os spins might be non- or only weakly interacting. The

relative energies of the other four configurations with respect to the ground state

configurations calculated using UCo = 4.1 eV and UOs = 2.1 eV are shown in Table A.3.

Figure A.7 shows the spin- and atom-resolved densities of states (DOS) for the ground

state magnetic configuration for calculations (a) with GGA, i.e excluding strong

correlation effects and (b) with UCo and UOs = 4.1 and 2.1 eV, respectively. Although the

two calculations predict the same ground state magnetic configuration energetically, from

Figure A.7 it can been seen that inclusion of strong-correlation effects on both Co and Os

are necessary to reproduce the insulating nature of the compound observed

experimentally. Since the three combinations of UCo and UOs give similar results when

comparing the energies of different magnetic configurations and as UCo = 4.1 eV and UOs

= 2.1 eV also reproduce the experimentally observed insulating nature, we use this

particular combination of UCo and UOs for the remainder of the work.

124
Figure A.6 16 formula-unit cell of Sr2CoOsO6 showing (a) the experimentally observed
magnetic configuration which is non-collinear and (b) the collinear version of the
experimentally observed magnetic configuration where all the moments have been
aligned along the [001] direction. Only Os (red) and Co (blue) atoms are shown for
clarity. The arrows show the direction of the magnetic moment on every Co and Os ion.

Figure A.7 Spin- and atom- resolved density of states with collinear arrangement of
independently ordered Co and Os sublattices obtained using (a) GGA without Hubbard-U
and (b) with UCo = 4.1 eV and UOs = 2.1 eV.

125
Table A.3 Relative energy of different magnetic configurations examined in this work,
along with the spin and orbital-moments of Co atoms and OsO6 units obtained for each
configuration.

Magnetic Energy/f.u. Ion Total Orbital


configuration (meV) moment (µB) moment (µB)
FM +809 Co 2.76 -
OsO6 2.17 -
C-type AFM +789 Co 2.74 -
OsO6 1.80 -
Exp(col) +712 Co 2.72 -
OsO6 1.82 -
Exp(col)+SOC +101 Co 2.72 0.06
OsO6 1.80 0.31
Exp(non_col) 0 Co 2.71 0.25
OsO6 1.69 0.72
AF1 +776 Co 2.71 -
OsO6 1.72 -
AF2 +742 Co 2.71 -
OsO6 1.86 -
AF3 +812 Co 2.76 -
OsO6 1.91 -
AF4 +786 Co 2.74 -
OsO6 2.11 -
AF5 +773 Co 2.74 -
OsO6 1.95 -
AF6 +758 Co 2.74 -
OsO6 1.98 -
AF7 +774 Co 2.73 -
OsO6 1.81 -

126
Figure A.8 Spin- and orbital-resolved density of states of (a) Co and (b) Os in the 16
formula-unit cell of Sr2CoOsO6 with collinear arrangement of independently ordered Co
and Os sublattices obtained using UCo = 4.1 eV and UOs = 2.1 eV.

Figure A.8 shows the orbital- and spin-resolved DOS for the Co and Os atoms for the

ground state configuration. With this particular combination of U, we find the magnetic

moments on the Co and Os ions to be 2.72 and 1.43 µB/ion, respectively. In addition, we

also find the O ions to have a magnetic moment of 0.07 µB/ion, with their sign being

equal to that of the Os ion to which it is connected to, which implies significant charge

sharing between the extended 5d states of the Os ions and the 2p states of the O ions

within the OsO6 octahedra, considering that the six O ions together contribute a moment

of 0.42 µB/OsO6 unit. On adding the magnetic moments of an Os ion and its neighboring

127
O atoms, we get a magnetic moment of 1.82 µB/ion for every OsO6 octahedron. These

values are in good agreement with the experimentally (from neutron powder diffraction)

determined values. Since these calculations were performed without including spin-orbit

coupling effects, the above moments correspond to the spin-only moments.

A.4.2 Non-collinear Magnetic Calculations

Heavy ions like Os, which has an atomic number Z = 76, can have strong spin-orbit

interactions. This is because the magnitude of spin-orbit coupling (SOC) depends on the

nuclear charge and is proportional to Z4 [1]. In order to understand how spin-orbit

interactions affect the properties of Sr2CoOsO6, we have performed DFT calculations on

the 16 f.u. cells by including SOC as implemented within VASP, wherein fully-

relativistic calculations are performed for the core electrons and valence electrons are

treated in a scalar relativistic approximation. Since SOC couples the electron spin to the

crystal structure, it allows carrying out non-collinear magnetic calculations to simulate

the experimentally observed magnetic configuration [2]. We have performed two

different calculations on the 16 f.u. cells by including SOC effects: (a) A collinear

calculation with the Exp(col) configuration as discussed above with all moments aligned

along the c-axis, and (b) a non-collinear calculation with the moments on each ion

oriented along the experimentally observed directions, i.e. the Co and Os moments lying

along a direction in the ac- and ab-plane, respectively. From Table A.3, we can see that

the structure with non-collinear arrangement of moments has the lowest energy, followed

by the Exp(col) configuration with and without SOC, respectively. Figure 4.4 in the

Chapter 4 shows the total and atom-resolved density of states for the non-collinear

128
configuration. The spin- and orbital-moment of the Co atoms and OsO6 units for the two

calculations with SOC are listed in Table A.3.

As expected, the OsO6 units have a larger orbital moment, 0.72 µB, than Co with an

orbital moment of 0.25 µB/ion, due to the higher Z of Os,. The total magnetic moment

(spin+orbital) on the Os ions can be represented by the vector ±(0.97 xˆ − 0.82 yˆ − 0.1zˆ )

µB/ion, where x̂ , ŷ and ẑ are unit-vectors along the three Cartesian axes. Thus, the Os

moments lie predominantly on the ab-plane, similar to the experimental results where the

moments can be described with the vector ±(1.28 xˆ − 1.28 yˆ + 0 zˆ ) µB/ion. On the other

hand, we find two different sets of Co ions with one set having their moments as

±(2.01xˆ − 0.04 yˆ − 1.82 zˆ) µB/ion and the other having ±(2.27 xˆ − 0.45 yˆ − 1.41zˆ) µB/ion.

While the former set of ions have their moments predominantly along the ac plane,

similar to the experimental moments that can be described with the vector

±(2.05 xˆ + 0 yˆ − 2.05 zˆ) µB/ion, the latter set of Co ions have a small component of

magnetization along the b axis. The total moment on the two different sets of Co ions

however is the same. Although the absolute value of the moments is known to change

strongly with the U values, it is important to note that the arrangement of the moments

for the non-collinear calculation match well with the experiments, with the exception that

our calculations suggest two different sets of Co ions, one with its moments lying on the

ac plane, while the other set of ions have a small component along the c axis, whereas

experimentally, only one set of Co ions was observed. Thus overall, the DFT calculations

predict non-collinear magnetic moments with independent antiferromagnetic ordering of

129
Co and Os ions with their ordering pattern similar to that of Exp(col) configuration, to be

the most stable configuration and are in good agreement with the experiments.

The total DOS of the three lowest energy configurations, i.e., Exp(col), Exp(col) + SOC

and the non-collinear configuration Exp(non_col) are shown together for comparison in

Figure A.9. There are some changes in the peak positions near the Fermi energy,

especially between the non-collinear and the collinear calculations. Their overall features

are however very similar. Also the ordering between of the moments within the Co and

Os sublattices is similar for the two configurations. These combined with the fact that

even while excluding SOC effects, Exp(col) is found to be most stable amongst other

collinear configurations, which provides a justification to use collinear calculations over

the computationally expensive non-collinear calculations for estimating the exchange

interactions between the Co and Os ions, as discussed in the following section.

130
Figure A.9 The total DOS for the 16 f.u. cell with (a) Exp(non_col), (b) Exp(col)+SOC
and (c) Exp(col) magnetic configurations.

A.4.3 Exchange Interactions

In order to shed light on the reason behind this highly unusual magnetic coupling in

Sr2CoOsO6, we have examined the strength of different exchange interactions J1 – J7

between the Co and Os ions, with the interaction pathways shown in Figure A.10 [3]. In

addition to FM, FiM (AF1), A-type AFM (AF2), C-type, and Exp(col), we have used

four additional magnetic configurations labeled as AF4-AF7 , all of which were done

using 16 f.u. supercells and are shown schematically in Figure A.11 and listed in Table

131
A.5. We have used a Heisenberg model to express the total energy of each magnetic

configuration, within a 2 f.u. cell, as a sum of a non-magnetic part with energy E0 and

magnetic parts given by the equation

H = E0 − ∑ J ij Si ⋅ S j , (A.2)
i< j

where Jij are the exchange constants for interaction between spin Si and Sj at sites i and j,

respectively. If Jij is positive (negative) it favors parallel (antiparallel) coupling of the two

spins. For simplicity, we have expressed Jij in terms of J1−J7 for different pairs of Co−Co,

Co−Os, and Os−Os interactions, as shown in Figure A.10. As discussed earlier, the

magnetic moments of Co and OsO6 units for the Exp(col) calculation are 2.72 and 1.82

µB/ion, respectively, and remain within ±0.2 µB for all the other configurations, roughly

corresponding to S = 3/2 for Co and S = 1 for Os. These moments are indicative of a

valence state of Co2+ in its high-spin and Os6+ in its low-spin configuration and also agree

well with their experimental values.

132
Figure A.10 Schematic showing the different exchange interactions between Co and Os
ions.

The Hamiltonian above has eight unknown parameters: E0 and J1 – J7. Hence, using the

eight magnetic configurations discussed above, we have solved the set of Eqs. A.3−A.10

using a least-squares fitting method to determine the values of E0 and J1 – J7,

2 2
EFM = (−4 J 4 − 6 J 2 − 8 J 5 ) SCo + (−8 J 6 − 6 J 3 − 4 J 7 ) SOs − 12 J1SCo ⋅ SOs (A.3)

2 2
EAF1 = ( −4 J 4 − 6 J 2 − 8 J 5 ) SCo + ( −8 J 6 − 6 J 3 − 4 J 7 ) SOs + 12 J1SCo ⋅ SOs (A.4)

2
EAF2 = ( −4 J 4 − 6 J 2 + 8 J 5 ) SCo 2
+ (−6 J 3 + 4 J 7 ) SOs − 4 J1SCo ⋅ SOs (A.5)

2
EC-type = (−4 J 4 − 6 J 2 + 8J 5 )SCo 2
+ (−6 J 3 + 4 J 7 ) SOs + 4 J1SCo ⋅ SOs (A.6)

2
EAF4 = ( −2 J 4 − 4 J 2 − 4 J 5 ) SCo 2
+ (−8 J 6 − 6 J 3 − 4 J 7 ) SOs − 9 J1SCo ⋅ SOs (A.7)

133
2
EAF5 = (−4 J 4 − 6 J 2 − 8 J 5 ) SCo 2
− 6 J 3 SOs − 6 J1SCo ⋅ SOs (A.8)

2
EAF6 = ( −2 J 4 − 2 J 2 ) SCo 2
+ (−2 J 6 − 2 J 3 ) SOs − 4 J 1S Co ⋅ S Os (A.9)

2
EAF7 = ( −2 J 4 − 4 J 2 + 4 J 5 ) SCo 2
+ (−4 J 3 + 2 J 7 ) SOs + 2 J1SCo ⋅ SOs (A.10)

In order to compare to the relative strengths of the spin interactions between the Co2+

(S=3/2) and Os6+ (S=1), we use their effective spin exchange constants J ijeff = Si S j J ij ,

which are listed in Table A.4 in addition to E0.

Table A.4 Spin exchange parameters J ijeff = Si S j J ij for Sr2CoOsO6 (in meV).

E0 −131.3458 eV/(2 f.u.)


!""
𝐽! -1.3
!""
𝐽! -47.2
!""
𝐽! 20.2
!""
𝐽! 29.0
!""
𝐽! 2.86
!""
𝐽! 0.8
!""
𝐽! -13.4

134
Table A.5 Comparison of energy per f.u. of the different magnetic configurations

calculated DFT and with the Heisenberg model.

Energy (eV/f.u.) DFT Model


FM −65.621 −65.630
AF1 −65.654 −65.646
AF2 −65.688 −65.663
C-type −65.642 −65.668
AF4 −65.644 −65.644
AF5 −65.657 −65.657
AF6 −65.673 −65.673
AF7 −65.657 −65.657
Exp(col) −65.718 −65.845

135
Figure A.11 Schematic showing the spin arrangements in two neighboring ab planes in
Sr2CoOsO6.

136
References

1. Atkins, P. W. Molecular quantum mechanics (Oxford University Press, New York,


1999).

2. Hobbs, D., Kresse, G. & Hafner, J. Fully unconstrained noncollinear magnetism


within the projector augmented-wave method. Physical Review B 62, 11556–11570
(2000).

3. Tian,C.; Wibowo, A. C.; Zur Loye, H.-C.; Whangbo, M-H. On the Magnetic
Insulating States, Spin Frustration, and Dominant Spin Exchange of the Ordered
Double-Perovskites Sr2CuOsO6 and Sr2NiOsO6: Density Functional Analysis
Inorg. Chem. 2011, 50, 4142-4148

137
Appendix B. Supplementary Information from Chapter 5

B.1 Experimental Details of Sr2CrReO6 Film Deposition

Sr2CrReO6 (001) epitaxial films are deposited on three kinds of substrates or buffer

layers: (LaAlO3)0.3(Sr2AlTaO6)0.7 (LSAT), SrTiO3 (STO), and a relaxed Sr2CrNbO6

(SCNO) buffer layer on LSAT. The Sr2CrReO6 (SCRO) films are grown using an off-

axis ultrahigh vacuum (UHV) sputtering technique we recently developed for epitaxial

growth of a variety of complex materials, including double perovskites (e.g. Sr2FeMoO6

and Sr2CrReO6), Yttrium ion garnet (Y3Fe5O12 or YIG), Heusler compounds (e.g.

Co2FeAl0.5Si0.5). The SCNO buffer layer on LSAT was deposited to a thickness of ~200

nm to allow the SCNO buffer layer to relax to near its bulk lattice parameters away from

the SCNO/LSAT interface. SCRO films were deposited at a substrate temperature of Ts =

700 °C using a DC power source at a constant current of 55 mA. For optimal deposition,

we used a total O2/Ar sputtering gas pressure of 12.5 mTorr with an oxygen partial

pressure of 26.25 µTorr. The Sr2CrNbO6 buffer layer was deposited at Ts = 600 °C using

a radio-frequency (RF) power source at a power of 50 W in a total O2/Ar pressure of 12.5

mTorr; the oxygen partial pressure was 17.5 µTorr. The deposition rates of Sr2CrReO6

and Sr2CrNbO6 layers were 1.10 and 1.25 nm/min, respectively.

138
B.2 Diamagnetic Background Subtraction in Squid Magnetometry
Measurements

First, we show in Supplemental Material Figure B.1 the raw magnetometry data taken at

T = 20 K for the Sr2CrReO6 film on LSAT. It is readily seen that the film gives almost no

distinguishable signal above that of the diamagnetic LSAT substrate for the hard axis

(field applied perpendicular to the film) hysteresis loop. Additionally, the easy axis loop

does not show a linear high field region. This implies that the film on LSAT is not nearly

fully saturated at 7 T. We manually fit the LSAT substrate background to acquire a

ferrimagnetic response for the SCRO film (after diamagnetic substrate background

subtraction) exhibiting a high field magnetization close to the expected value (of which

we have prior knowledge). As a result of the manual fitting of the LSAT substrate

background and low out-of-plane magnetic signal from the SCRO/LSAT film, we are

unable to extrapolate the anisotropy field Hu for SCRO/LSAT and simply estimate it to

be at least 10's of tesla (T).

139
Figure B.1 Raw magnetic data for the 90 nm SCRO/LSAT film taken with a Quantum
Design SQUID for both in-plane and out-of-plane geometries at T = 20 K. The hard axis
loop (field applied perpendicular to the film) shows almost no signal differentiating the
SCRO film from the background of the LSAT substrate, so a reliable anisotropy field
analysis could not be performed. We estimate Hu to be 10's of tesla.

As a demonstration of the analysis of magnetometry data, we show in Supplemental

Material Figure B.2 the raw magnetic data for the 90-nm Sr2CrReO6 film on SrTiO3.

Magnetometry measurements are taken in both in-plane (red) and out-of-plane (blue)

geometries. Raw data includes the magnetic signal from the film as well as the

diamagnetic background from the SrTiO3 substrate. The diamagnetic background of the

substrate is not expected to have orientation dependence. We fit the substrate background

using the high field regions of the easy-axis measurement (in-plane) because at high

fields the SCRO film is almost saturated and thus the high field region data gives the best

estimate of the diamagnetic background slope. Supplemental Material Figure B.2 shows

the substrate background (green) slope matches the SCRO/STO high field region slope.

140
After the substrate background is subtracted from the raw data, we normalize the raw

signal (in µemu) to the sample volume and present the data in units of µB/f.u., where f.u.

is the formula unit cell of Sr2CrReO6.

Figure B.2 Raw magnetic data for 90-nm SCRO/STO film taken with a Quantum Design
SQUID magnetometer for both in-plane and out-of-plane geometries at T = 20 K. The
(linear) diamagnetic background of the substrate is shown.

Application of this analysis to the SCRO/LSAT film [panel (a) of Fig. 5.3 in the Chapter

5] results in negative magnetizations at positive fields, which suggests the film (in the

easy-axis in-plane measurement) is not nearly saturated at fields of 7 T and our fitting

procedure is not applicable. We manually fit the LSAT substrate background to acquire a

ferrimagnetic response exhibiting a high field magnetization close to the expected value.

As a result of the manual fitting of the LSAT substrate background and low out-of-plane

141
magnetic signal from the SCRO/LSAT film, we are unable to extrapolate the anisotropy

field HU for SCRO/LSAT and simply estimate it to be at least 10’s of tesla (T).

For the SCRO/SCNO/LSAT, we use the same analysis procedure to fit the substrate

background, which includes contributions from both the SCNO double perovskite buffer

layer and LSAT substrate. The easy axis in this case is along the axis normal to the film

surface and so we fit the substrate background using the high field regions of the out-of-

plane measurement data.

B.3 DFT-calculated Structural Parameters for Sr2CrReO6 Films on


LSAT, STO, & SCNO/LSAT

See Supplemental Material Table B.1 for DFT-calculated rotation angles of the octahedra

as well as bond lengths.

Table B.1 DFT-calculated rotation angles of the Cr and Re oxygen octahedra as well as
equatorial (deq) and axial (dax) lengths of the Cr–O and Re–O bonds for the Sr2CrReO6
films grown on LSAT, SrTiO3 and Sr2CrNbO6/LSAT.

Substrate or Ɵ Ɵ deq(Cr– dax(Cr– deq(Re–O) dax(Re–


Buffer Layer (CCW) (CW) O) (Å) O) (Å) (Å) O) (Å)

LSAT 5.863° 5.802° 1.933 1.976 1.953 1.987

SrTiO3 4.623° 4.543° 1.950 1.960 1.966 1.970

Sr2CrNbO6/LSA 3.554° 3.532° 1.967 1.945 1.978 1.957


T

142
B.4 Atomically Resolved Spin and Orbital Magnetic Moments for
SCRO Films on LSAT, STO, and SCNO/LSAT

See Supplemental Material Table B.2 for calculated spin and orbital moments for the Cr,

Re, and O atoms in SCRO for each strain state (and both in-plane and out-of-plane).

Table B.2 First principles calculated in-plane (average of [100] and [010] and out-of-
plane ([001]) spin and orbital magnetic moments for Cr, Re, and O atoms in SCRO films
of different strain states.

Spin magnetic moment Orbital magnetic moment


(𝛍B/f.u.) (𝛍B/f.u.)
c/a = 0.99 1.01 1.025 0.99 1.01 1.025
Cr In-plane 1.488 1.488 1.494 -0.021 -0.021 -0.021
Out-of-plane 1.51 1.494 1.465 -0.022 -0.021 -0.021

Re In-plane -0.678 -0.681 -0.69 0.13 0.135 0.146


Out-of-plane -0.704 -0.688 -0.654 0.149 0.138 0.113
O In-plane -0.118 -0.12 -0.123 0.012 0.011 0.011
Out-of-plane -0.124 -0.122 -0.114 0.014 0.014 0.01

B.5 Calculated DOS of Rhenium Suborbitals Varying with Strain

See Supplemental Material Figure A.3 for the calculated DOS of Re t2g (a) xy and (b) yz

and xz suborbitals near the Fermi energy in SCRO for tetragonal disortions c/a = 1.025

and c/a = 0.99.

143
Figure B.3 Calculated DOS of Re t2g (a) xy and (b) yz and xz suborbitals near the Fermi
energy in SCRO for tetragonal distortions c/a = 1.025 and c/a = 0.99.

144
Appendix C. Supplementary Information for Chapter 7

C.1 Electronic Structure Calculations

A comparison of the calculated geometry optimized lattice parameters compared with the

experimental lattice parameters is given in Table C.1.

Table C.1 An HSE hybrid functional with spin orbit coupling was used for the
computations.

Compound Space Group Experimental Calculated


Symmetry
CH3NH3PbCl3 Pm-3m 5.70 5.64359
Cs2AgBiCl6 Fm-3m 10.7774(2) 10.7768
CH3NH3PbBr3 Pm-3m 5.94 5.98378
Cs2AgBiBr6 Fm-3m 11.2711(1) 11.2712

C.2 Kubelka Munk Plots

The diffuse reflectance plots have been transformed into pseudo absorbance using the

Kubelka Munk transformation as shown in Figure C.1.

145
Figure C.1 Kubleka-Munk plots for the silver-bismuth (blue) and lead compounds (red),
with chloride compounds on the left and bromide compounds on the right.

C.3 Light and Moisture Stability

Plots of the X-ray powder diffraction patterns of the Cs2AgBiCl6 and the Cs2AgBiBr6

samples after exposure to air in the dark for 2 weeks, exposure to both air and light for 2

weeks and for 4 weeks are shown in Figure C.2.

146
Figure C.2 XRPD overlays of Cs2AgBiCl6 (upper) and Cs2AgBiBr6 (lower) showing
the light stability of the compounds after a month of exposure to sunlight. The change in
the powder diffractogram for the bromine compound indicates that structural changes are
accompanying the visual darkening of the pigment.

147
References

References from Chapter 2

1. Piskunov, S., Heifets, E., Eglitis, R. . & Borstel, G. Comput. Mater. Sci. 29, 165–
178 (2004).

2. Van Benthem, K., Elsa sser, C. & French, R. H. J. Appl. Phys. 90, 6156 (2001).

3. Johnsson, M. & Lemmens, P. J. Phys. Condens. Matter 20, 264001 (2008).

4. Woodward, P. M. Acta Crystallogr. Sect. B Struct. Sci. 53, 32–43 (1997).

5. Lucy, J. M. et al. 000400, 1–6 (2014).

6. Lufaso, M. W. & Woodward, P. M. Acta Crystallogr. Sect. B Struct. Sci. 60, 10–
20 (2004).

7. Popov, G., Greenblatt, M. & Croft, M. Phys. Rev. B 67, 024406 (2003).

8. Kim, B. G., Hor, Y. S. & Cheong, S. W. Appl. Phys. Lett. 79, (2001).

9. Glazer, A. Acta Crystallogr. Sect. B Struct. Sci. B 28, (1972).

10. Mishra, R., Soliz, J. R., Woodward, P. M. & Windl, W. Chem. Mater. 24, 2757–
2763 (2012).

11. Lufaso, M. W., Barnes, P. W. & Woodward, P. M. Acta Crystallogr. Sect. B


Struct. Sci. 62, 397–410 (2006).

12. Housecroft, C. & Sharpe, A. G. Inorganic Chemistry. (Prentice Hall, 2008).

13. Woodward, P. M., Karen, P., Evans, J. S. . & Vogt, T. Functional Materials.
(Unpublished).

148
14. Mishra, R., Restrepo, O. D., Woodward, P. M. & Windl, W. Chem. Mater. 22,
6092–6102 (2010).

15. Popov, G., Greenblatt, M. & Croft, M. Phys. Rev. B 67, 024406 (2003).

16. Vasala, S. & Karppinen, M. Prog. Solid State Chem. 43 1-36 (2015).

17. Zener, C. Phys. Rev. 82, (1951).

18. Kanamori, J. J. Phys. Chem. Solids 10, 87–98 (1959).

References from Chapter 3

1. Hohenberg, P. & Kohn, W. Physical Review 136, B864−B871 (1964).

2. Kohn, W. & Sham, L. J. Physical Review 140, A1133−A1138 (1965).

3. Sholl, D.S. & Steckel, J.A. Density Functional Theory: A Practical Introduction
(John Wiley and Sons, Hoboken, NJ, 2009).
4. Himmetoglu, B., Floris, A., de Gironcoli, Stefano & Cococcioni, M. Int. J. Quant.
Chem. 114, 14-49 (2014).

References from Chapter 4

1. Goodenough, J. B. Phys. Rev. 1955, 100, 564-573.

2. Kanamori, J. J. Phys. Chem. Solids 1959, 10, 87-98.

3. Lufaso, M.W.; BarnesP. W.;Woodward, P.M. Acta Cryst. B 2006, 62, 397-410.

4. Krockenberger, Y; Mogare, K.; Reehuis, M.; Tovar, M.; Jansen, M.;


Vaitheeswaran, G.; Kanchana, V.; Bultmark, F.; Delin, A.; Wilhelm, F.; Rogalev,
A.; Winkler, A.; Alff, L. Phys. Rev. B. 2007, 75, 020404(R).

5. Rogado, N. S.; Li, J.; Sleight, A. W.; Subramanian, M.A. Adv. Mater. 2005, 17
(18), 2225-2227.

6. Sleight, A. W.; Longo, J.; Ward, R. Inorg. Chem. 1962, 1 (2), 245–250.

149
7. Macquart, R; Kim, S. J.; Gemmill, W. R.; Stalick, J. K.; Lee, Y.; Vogt, T.; Zur
Loye, H. C . Inorg. Chem. 2005, 44, 9676-9683.

8. Lufaso, M. W.; Gemmill, W. R.; Mugavero, S. J.; Kim, S. J.; Lee, Y.; Vogt, T.;
zur Loye, H. C. J. Solid State Chem. 2008, 181 (3), 623-627.

9. Huq A.; Hodges J. P.; Gourdon, O.; Heroux, L. Zeitschrift für Kristallographie
Proceedings 2011, 1, 127-135.

10. Rodriguez-Carvajal, J. Physica B. 1993, 192, 55-69.

11. Wills, A. Physica B. 1991, 276, 680-681.

12. Larson, A. C.; Von Dreele, R. B. Los Alamos National Laboratory Report LAUR
86-748 2000.

13. Toby, B. H. EXPGUI, a graphical user interface for GSAS. J. Appl. Cryst. 1991,
34, 210-213.

14. Kobayashi, K.; Nagao, T.; Ito, M. Acta Crys. A 1991, 67, 473-480.

15. Kresse, G; Hafner, J. Phys. Rev. B 1991, 47, 558−561.

16. Kresse, G.; Hafner, J. Phys. Rev. B 1991, 49, 14251−14269.

17. Blöchl, P. E. Phys. Rev. B 1991, 50, 17953−17979.

18. Wang, Y.; Perdew, J. P. Phys. Rev. B 1991, 44, 13298−13307.

19. Monkhorst, H. J.; Pack, J. D. Phys. Rev. B 1991, 13, 5188−5192.

20. Hobbs, D.; Kresse, G.; Hafner, J. Phys. Rev. B 1991, 62, 11556–11570.

21. Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.; Sutton, A. J.
Phys. Rev. B 1991, 57, 1505–1509.

22. Anisimov, V. I.; Zaanen, J.; Andersen, O. K. Phys. Rev. B 1991, 44, 943-954.

23. Viola, M. C; Martinez-Lope, M. J.; Alonso, J. A.; De Paoli, J. M.; Pagola, S.;
Pedregosa, J. C.; Fernandez-Diaz, M. T.; Carbonio, R. E. Chem. Mater. 2003, 15
(8), 1655-1663.

24. Ivanov, S. A.; Eriksson, S. G.; Tellgren, R.; Rundlöf, H.; Tseggai, M. Mater. Res.
Bull. 2005, 40, 840-849.

150
25. Lopez, C. A.; Saleta, M. E.; Curiale, J.; Sanchez, R. D. Mater. Res. Bull. 2012,
47, 1158-1163.

26. Retuerto, M.; Martínez-Lope, M. J.; García-Hernández M.; Fernández-Díaz, M.


T.; Alonso, J. A. Eur. J. Inorg. Chem. 2008, 4, 588-595.

27. Choy, J. H.; Kim, D. K.; Kim., J. Y. Solid State Ionics. 1998, 108, 159–163.

28. Chmaissem, O.; Kruk, R.; Dabrowski, B.; Brown, D. E.; Xiong, X.; Kolesnik, S.;
Jorgensen, J. D.; Kimball C. W. Phys. Rev. B 2000, 62, 14197-14206.

29. Shannon, R. D. Acta Cryst. A 1976, 32, 751-767.

30. Brese, N. E.; O’Keefe, M. Acta Cryst. B 1991, 47, 192-197.

31. Currie, R. C.; Vente, J. F.; Frikkee, E.; IJdo, D. J. W. J. Solid State Chem. 1995,
116 (1), 199-204.

32. Tian,C.; Wibowo, A. C.; Zur Loye, H.-C.; Whangbo, M-H. Inorg. Chem. 2011,
50, 4142-4148.

33. Wang, J.; Meng, J.; Wu, Z. Chem. Phys. Lett. 2011, 501, 324-329.

References from Chapter 5

1. M. Seifert, V. Neu, and L. Schultz, Appl. Phys. Lett. 94, 022501 (2009).

2. T. Shima, K. Takanashi, Y. K. Takahashi, and K. Hono, Appl. Phys. Lett. 85,


2571 (2004).
3. J. H. Haeni, P. Irvin, W. Chang, R. Uecker, P. Reiche, Y. L. Li, S. Choudhury, W.
Tian, M. E. Hawley, B. Craigo, A. K. Tagantsev, X. Q. Pan, S. K. Streiffer, L. Q.
Chen, S. W. Kirchoefer, J. Levy, and D. G. Schlom, Nature 430, 758-761 (2004).

4. J. Wang, J. B. Neaton, H. Zheng, V. Nagarajan, S. B. Ogale, B. Liu, D. Viehland,


V. Vaithyanathan, D. G. Schlom, U. V. Waghmare, N. A. Spaldin, K. M. Rabe,
M. Wuttig, and R. Ramesh, Science 299, 1719 (2003).

5. P. V. Lukashev, N. Horrell, and R. F. Sabirianov, J. Appl. Phys. 111, 07A318

(2012).

6. D. R. Behrendt, S. Legvold, and F. H. Spedding, Phys. Rev. 109, 1544 (1958).

151
7. F. Y. Yang, C. L. Chien, E. F. Ferrari, X. W. Li, G. Xiao, and A. Gupta, Appl.
Phys. Lett. 77, 286 (2000).

8. G. H. O. Daalderop, P. J. Kelly, and M. F. H. Schuurmans, Phys. Rev. B 44,


12054 (1991).

9. B. Peters, A. Alfonsov, C. G. F. Blum, S. J. Hageman, P. M. Woodward, S.


Wurmehl, B. Büchner, and F. Y. Yang, Appl. Phys. Lett. 103, 162404 (2013).

10. H. Kato, T. Okuda, Y. Okimoto, Y. Tomioka, K. Oikawa, T. Kamiyama, and Y.


Tokura, Phys. Rev. B 69, 184412 (2004).

11. D. Serrate, J. M. De Teresa, and M. R. Ibarra, J. Phys.: Condens. Matter 19,


023201 (2007).

12. K. I. Kobayashi, T. Kimura, H. Sawada, K. Terakura, and Y. Tokura, Nature 395,


677-680 (1998).

13. Y. Krockenberger, K. Mogare, M. Reehuis, M. Tovar, M. Jansen, G.


Vaitheeswaran, V. Kanchana, F. Bultmark, A. Delin, F. Wilhelm, A. Rogalev, A.
Winkler, and L. Alff, Phys. Rev. B 75, 020404(R) (2007).

14. A. J. Hauser, J. R. Soliz, M. Dixit, R. E. A. Williams, M. A. Susner, B. Peters, L.


M. Mier, T. L. Gustafson, M. D. Sumption, H. L. Fraser, P. M. Woodward, and F.
Y. Yang, Phys. Rev. B 85, 161201(R) (2012).

15. J. M. Lucy, A. J. Hauser, H. L. Wang, J. R. Soliz, M. Dixit, R. E. A. Williams, A.


Holcombe, P. Morris, H. L. Fraser, D. W. McComb, P. M. Woodward, and F. Y.
Yang, Appl. Phys. Lett. 103, 042414 (2013).

16. A. J. Hauser, J. M. Lucy, H. L. Wang, J. R. Soliz, A. Holcomb, P. Morris, P. M.


Woodward, and F. Y. Yang, Appl. Phys. Lett. 102, 032403 (2013).

17. F. D. Czeschka, S. Geprägs, M. Opel, S. T. B. Goennenwein, and R. Gross, Appl.


Phys. Lett. 95, 062508 (2009).

18. M. Komelj, Phys. Rev. B 82, 012410 (2010).

19. X. Chen, D. Parker, K. P. Ong, M. H. Du, and D. J. Singh, Appl. Phys. Lett. 102,
102403 (2013).

20. A. J. Hauser, R. E. A. Williams, R. A. Ricciardo, A. Genc, M. Dixit, J. M. Lucy,


P. M. Woodward, H. L. Fraser, and F. Y. Yang, Phys. Rev. B 83, 014407 (2011).

152
21. C. H. Du, R. Adur, H. L. Wang, A. J. Hauser, F. Y. Yang, and P. C. Hammel,
Phys. Rev. Lett. 110, 147204 (2013).

22. H. L. Wang, C. H. Du, Y. Pu, R. Adur, P. C. Hammel, and F. Y. Yang, Phys. Rev.
B 88, 100406(R) (2013).

23. C. H. Du, H. L. Wang, Y. Pu, T. L. Meyer, P. M. Woodward, F. Y. Yang, and P.


C. Hammel, Phys. Rev. Lett. 111, 247202 (2013).

24. See Supplemental Material at http://link.aps.org/supplemental/


10.1103/PhysRevB.90.180401 or Appendix B of this document for: (1) film
growth parameters, (2) methods used for SQUID magnetometry data analysis, (3)
tabularized structure, spin moment, and orbital moment parameters obtained from
DFT calculations, and (4) density of states obtained from DFT calculations.

25. P. M. Woodward, Acta. Cryst. B 53, 44 (1997).

26. J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J.


Singh, and C. Fiolhais, Phys Rev B 46, 6671–87 (1992).

27. G. Kresse and J. Hafner, Phys. Rev. B 49, 14251-14269 (1994).

28. G. Kresse and J. Hafner, Phys. Rev. B 47, 558 (1993).

29. P. E. Blöchl, Phys. Rev. B 50, 17953–17979 (1994).

30. V. I. Anisimov, J. Zaanen, and O. K. Andersen, Phys Rev B 44, 943–954 (1991).

31. H.-T. Jeng and G. Y. Guo, Phys. Rev. B 67, 094438 (2003).

32. J. M. De Teresa, J. M. Michalik, J. Blasco, P. A. Algarabel, M. R. Ibarra, C.


Kapusta, and U. Zeitler, Appl. Phys. Lett. 90, 252514 (2007).

33. P. Bruno, Phys. Rev. B 39, 865 (1989).

34. O. Hjortstam, K. Baberschke, J. M. Wills, B. Johansson, and O. Eriksson, Phys.


Rev. B 55, 15026 (1997).

35. I. Yang, S. Y. Savrasov, and G. Kotliar, Phys. Rev. Lett. 87, 216405 (2001).

36. F. Wilhelm, P. Poulopoulos, P. Srivastava, H. Wende, M. Farle, K. Baberschke,


M. Angelakeris, N. K. Flevaris, W. Grange, J.-P. Kappler, G. Ghiringhelli, and N.
B. Brookes, Phys. Rev. B 61, 8647 (2000).

153
37. S. Gold, E. Goering, C. König, U. Rüdiger, G. Güntherodt, and G. Schütz, Phys.
Rev. B 71, 220404 (2005).

38. H.-T. Jeng and G. Y. Guo, Phys. Rev. B 65, 094429 (2002).

39. A. J. Hauser, J. M. Lucy, M. W. Gaultois, M. R. Ball, J. R. Soliz, Y. Choi, O. D.


Restrepo, W. Windl, J. W. Freeland, D. Haskel, P. M. Woodward, and F. Yang,
Phys. Rev. B 89, 180402 (2014)

References from Chapter 6

1. D. Serrate, J.M. De Teresa, and M.R. Ibarra, J. Phys. Condens. Matter 19, 023201
(2007).

2. H. Kato, T. Okuda, Y. Okimoto, Y. Tomioka, Y. Takenoya, A. Ohkubo, M.


Kawasaki, and Y. Tokura, Appl. Phys. Lett. 81, 328 (2002).

3. A.J. Hauser, J.R. Soliz, M. Dixit, R.E.A. Williams, M.A. Susner, B. Peters, L.M.
Mier, T.L. Gustafson, M.D. Sumption, H.L. Fraser, P.M. Woodward, and F.Y.
Yang, Phys. Rev. B 85, 161201 (2012).

4. H. Asano, N. Kozuka, A. Tsuzuki, and M. Matsui, Appl. Phys. Lett. 85, 263
(2004).

5. G. Vaitheeswaran, V. Kanchana, and A. Delin, Appl. Phys. Lett. 86, 032513


(2005).

6. O.N. Meetei, O. Erten, A. Mukherjee, M. Randeria, N. Trivedi, and P.


Woodward, Phys. Rev. B 87, 165104 (2013).

7. S.A. Wolf, D. Awschalom, R. Buhrman, J. Daughton, S. von Molnár, M. Roukes,


A. Chtchelkanova, and D. Treger, Science (80-. ). 294, 1488 (2001).

8. K. Ando, Science (80-. ). 312, 1883 (2006).

9. H. Das, P. Sanyal, T. Saha-Dasgupta, and D.D. Sarma, Phys. Rev. B 83, 104418
(2011).

10. G. Vaitheeswaran, V. Kanchana, M. Alouani, and A. Delin, EPL (Europhysics


Lett. 84, 47005 (2008).

11. P. Majewski, S. Gepra gs, O. Sanganas, M. Opel, R. Gross, F. Wilhelm, A.


Rogalev, and L. Alff, Appl. Phys. Lett. 87, 202503 (2005).
154
12. J.M. Lucy, A.J. Hauser, H.L. Wang, J.R. Soliz, M. Dixit, R.E.A. Williams, A.
Holcombe, P. Morris, H.L. Fraser, D.W. McComb, P.M. Woodward, and F.Y.
Yang, Appl. Phys. Lett. 103, 042414 (2013).

13. A.J. Hauser, J.M. Lucy, H.L. Wang, J.R. Soliz, A. Holcomb, P. Morris, P.M.
Woodward, and F.Y. Yang, Appl. Phys. Lett. 102, 032403 (2013).

14. B.T. Thole, P. Carra, F. Sette, and G. van der Laan, Phys. Rev. Lett. 68, 1943
(1992).

15. C. Chen, Y. Idzerda, H. Lin, and N. Smith, Phys. Rev. Lett. 75, 152 (1995).

16. E. Goering, Philos. Mag. 85, 2895 (2005).

17. P.E. Blochl, Phys. Rev. B 50, (1994).

18. G. Kresse and J. Hafner, Phys. Rev. B 47, (1993).

19. G. Kresse and J. Hafner, Phys. Rev. B 49, (1994).

20. V.I. Anisimov, J. Zaanen, and O.K. Andersen, Phys. Rev. B 44, (1991).

21. H.-T. Jeng and G. Guo, Phys. Rev. B 67, 094438 (2003).

22. J. M. Lucy, M. R. Ball, O. D. Restrepo, A. J. Hauser, J. R. Soliz, J. W. Freeland,


P. M. Woodward, W. Windl, and F. Y. Yang, Phys. Rev. B 90, (2014).

23. V. Kanchana, G. Vaitheeswaran, and M. Alouani, J. Phys. Condens. Matter 18,


5155 (2006).

24. L. Mattheiss, Phys. Rev. 151, 450 (1966).

25. P.M. Woodward, Acta Crystallogr. Sect. B Struct. Sci. 53, 32 (1997).

26. P.M. Woodward, Acta Crystallogr. Sect. B Struct. Sci. 53, 44 (1997).

27. D. Huang, H.-T. Jeng, C. Chang, G. Guo, J. Chen, W. Wu, S. Chung, S. Shyu, C.
Wu, H.-J. Lin, and C. Chen, Phys. Rev. B 66, 174440 (2002).

28. E. Goering, A. Bayer, S. Gold, G. Schutz, M. Rabe, U. Rudiger, and G.


Guntherodt, 58, 906 (2002).

29. R. Mishra, O.D. Restrepo, P.M. Woodward, and W. Windl, Chem. Mater. 22,
6092 (2010).

155
30. R. Leapman, L. Grunes, and P. Fejes, Phys. Rev. B 26, 614 (1982).

31. J.G. Chen, Surf. Sci. Rep. 30, 1 (1997).

32. T.L. Daulton and B.J. Little, Ultramicroscopy 106, 561 (2006).

33. F. de Groot, Coord. Chem. Rev. 249, 31 (2005).

34. J.P. Clancy, N. Chen, C.Y. Kim, W.F. Chen, K.W. Plumb, B.C. Jeon, T.W. Noh,
and Y.-J. Kim, Phys. Rev. B 86, 195131 (2012).

35. J. van Bokhoven A., H. Sambe, D. Koningsberger C., and D. Ramaker E., J. Phys.
IV Fr. 7, 835 (1997).

36. R. Laskowski and P. Blaha, Phys. Rev. B 82, 205104 (2010).

References from Chapter 7

1. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal halide


perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc.
2009, 131 , 6050–6051.
2. National Renewable Energy Labs (NREL) efficiency chart (2015);
http://www.nrel.gov/ncpv/images/efficiency_chart.jpg (accessed 30 October
2015).
3. Stranks, S. D.; Snaith, H. J. Metal-halide perovskites for photovoltaic and light-
emitting devices. Nat. Nanotechnol. 2015, 10, 391–402.
4. Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon,
S.-J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J. E.; Grätzel, M.; Park, N.-G.
Lead iodide perovskite sensitized all solid state submicron thin film mesoscopic
solar cell with efficiency exceeding 9%. Sci. Rep. 2012, 2, 591.
5. Noh, J. H.; Im, S. H.; Heo, J. H.; Mandal, T. N.; Seok, S. Il. Chemical
management for colorful, efficient and stable inorganic-organic nanostructured
solar cells. Nano Lett. 2013, 13, 1764–1769.
6. Leijtens, T.; Eperon, G. E.; Pathak, S.; Abate, A.; Lee, M. M.; Snaith, H. J.
Overcoming ultraviolet light instability of sensitized TiO2 with meso-
superstructured organometal trihalide perovskite solar cells. Nat. Commun. 2013,
4, 2885.

156
7. Fthenakis, V. M.; Kim, H. C.; Alsema, E. Emissions from photovoltaic life
cycles. Environ. Sci. Technol. 2008, 42, 2168–2174.
8. Held, M.; Ilg, R. Update of environmental idicators and energy payback time of
CdTe PV systems in Europe. Prog. Photovoltaics Res. Appl. 2011, 19, 614–626.
9. Stoumpos, C. C.; Malliakas, C. D.; Kanatzidis, M. G. Semiconducting tin and
lead perovskites with organic cations: Phase transitions, high mobilities, and near-
infrared photoluminescent properties. Inorg. Chem. 2013, 52, 9019–9038.
10. Noel, N. K.; Stranks, S. D.; Abate, A.; Wehrenfennig, C.; Guarnera, S.;
Haghighirad, A.-A.; Sadhanala, A.; Eperon, G. E.; Pathak, S. K.; Johnston, M. B.;
Petrozza, A.; Herz, L. M.; Snaith, H. J. Lead-free organic–inorganic tin halide
perovskites for photovoltaic applications. Energy Environ. Sci. 2014, 7, 3061.
11. Hao, F.; Stoumpos, C. C.; Cao, D. H.; Chang, R. P. H.; Kanatzidis, M. G. Lead-
free solid-state organic–inorganic halide perovskite solar cells. Nat. Photonics
2014, 8, 489–494.
12. Stoumpos, C. C.; Malliakas, C. D.; Peters, J. A.; Liu, Z.; Sebastian, M.; Im, J.;
Chasapis, T. C.; Wibowo, A. C.; Chung, D. Y.; Freeman, A. J.; Wessels, B. W.;
Kanatzidis, M. G. Crystal growth of the perovskite semiconductor CsPbBr3: A
new material for high-energy radiation detection. Cryst. Growth Des. 2013, 13,
2722–2727.
13. Myagkota, S.; Voloshinovskii, A.; Stefanskii, I.; Mikhailik, M.; Pashuk, I.
Reflection and emission properties of lead-based perovskite-like crystals. Radiat.
Meas. 1998, 29, 273–277.
14. Kulbak, M.; Cahen, D.; Hodes, G. Planar How important is the organic part of
lead halide perovskite solar cells? Efficient CsPbBr3 solar cells. J. Phys. Chem.
Lett. 2015, 6, 2452–2456.
15. Kulbak, M.; Gupta, S.; Kedem, N.; Levine, I.; Bendikov, T.; Hodes, G.; Cahen,
D. Cesium enhances long-term stability of lead bromide perovskite-based solar
cells J. Phys. Chem. Lett. 2016, 7, 167−172.
16. Topas Academic, General Profile and Structural Analysis for Powder Diffraction
Data; Bruker AXS: Karlsruhe, Germany 2004.
17. Kresse, G.; Hafner, J. Ab initio Molecular-Dynamics for Liquid- Metals. Phys.
Rev. B 1993, 47, 558–561.
18. Kresse, G.; Hafner, J. Ab Initio Molecular-Dynamics Simulation of the Liquid-
Metal Amorphous-Semiconductor Transition in Germanium. Phys. Rev. B 1994,
49, 14251–14269.

157
19. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B. 1994. 50, 17953-
17979.
20. Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made
simple. Phys. Rev. Lett. 1996, 77, 3865–3868.
21. Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid Functionals Based on a Screened
Coulomb Potential. J. Chem. Phys. 2003, 118, 8207–8215.
22. Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid Functionals Based on a Screened
Coulomb Potential. J. Chem. Phys. 2006, 124, 219906.
23. Paier, J.; Marsman, M.; Hummer, K.; Kresse, G.; Gerber, I. C.; Angyan, J. G.
Screened Hybrid Density Functionals Applied to Solids. J. Chem. Phys. 2006,
125, 249901.
24. Giorgi, G.; Fujisawa, J. I.; Segawa, H.; Yamashita, K. Small photocarrier
effective masses featuring ambipolar transport in methylammonium lead iodide
perovskite: A density functional analysis. J. Phys. Chem. Lett. 2013, 4, 4213–
4216.
25. Melissen, S. T. A. G.; Labat, F.; Sautet, P.; Le Bahers, T. Electronic properties of
PbX3CH3NH3 (X = Cl, Br, I) compounds for photovoltaic and photocatalytic
applications. Phys. Chem. Chem. Phys. 2015, 17, 2199-2209.
26. Yin, W.-J.; Yang, J.-H.; Kang, J.; Yan, Y.; Wei, S.-H. Halide perovskite materials
for solar cells: a theoretical review J. Mater. Chem. A 2015, 3, 8926-8942.
27. Park, B. W.; Philippe, B.; Zhang, X.; Rensmo, H.; Boshchloo, G.; Johansson, E.
M. J. Bismuth Based Hybrid Perovskites A3Bi2I9 (A: Methylammonium or
Cesium) for Solar Cell Application, Adv. Mater. 2015, 27, 6806–6813.
28. Lehner, A. J.; Fabini, D. H.; Evans, H. A.; Hebert, C. A.; Smock, S. R.; Hu, J.;
Wang, H.; Zwanziger, J. W.; Chabinyc, M. L.; Seshadri, R. Crystal and Electronic
Structures of Complex Bismuth Iodides A3Bi2I9 (A = K, Rb, Cs) Related to
Perovskite: Aiding the Rational Design of Photovoltaics, Chem. Mater. 2015, 27,
7137−7148.
29. Saparov, B.; Hong, F.; Sun, J.-P.; Duan, H.-S.; Meng, W.; Cameron, S.; Hill, I.
G.; Yan, Y., Mitzi, D. B. Thin-Film Preparation and Characterization of Cs3Sb2I9:
A Lead-Free Layered Perovskite Semiconductor, Chem. Mater. 2015, 27,
5622−5632.
30. Morss, L. R.; Siegall, M.; Stenger, L.; Edelstein, N. Preparation of cubic chloro
complex compounds of trivalent metals: Cs2NaMCl6 Inorg. Chem.,1970, 9, 1771-
1775.

158
References from Appendix A

1. Atkins, P. W. Molecular quantum mechanics (Oxford University Press, New York,


1999).

2. Hobbs, D., Kresse, G. & Hafner, J. Fully unconstrained noncollinear magnetism


within the projector augmented-wave method. Physical Review B 62, 11556–11570
(2000).

3. Tian,C.; Wibowo, A. C.; Zur Loye, H.-C.; Whangbo, M-H. On the Magnetic
Insulating States, Spin Frustration, and Dominant Spin Exchange of the Ordered
Double-Perovskites Sr2CuOsO6 and Sr2NiOsO6: Density Functional Analysis
Inorg. Chem. 2011, 50, 4142-4148

159

You might also like