European Commission erasmus mundus

A NISOTROPICINTERACTIONSINTRANSITION METALOXIDES

dissertation zur Erlangung des akademischen Grades Doctor rerum naturalium (Dr. rer. nat.)

vorgelegt von nikolay bogdanov geboren am 09.01.1987 in Moskau, Russland

Fachrichtung Physik Fakultät für Mathematik und Naturwissenschaften Technische Universität Dresden 2017 Nikolay Bogdanov: Anisotropic interactions in transition metal oxides, Quantum chemistry study of strongly correlated materials gutachter: Prof. Dr. Jeroen van den Brink Prof. Dr. Ria Broer eingereicht am: 29 September 2017 disputation am: 6 April 2018 Моим родителям, Ирине и Александру.

To my parents, Irina and Alexander.

ACKNOWLEDGEMENTS

First of all, I would like to express my deep gratitude to my direct advisor Liviu Hozoi and my supervisor Jeroen van den Brink for the continuous support of my study and research. Liviu, you not only introduced me to quantum chemistry, practi- cal use of quantum mechanics and many other things, but also were the person I looked up to. Thank you for your constant motivation, patience and guidance during my time at IFW and especially when writing this thesis. Jeroen, thank you for the generous encouragement during my stay in Dresden and during the later stages of finishing the thesis. It was my pleasure to talk to you, as you always had a great project to pro- pose, an interesting idea to check, or a smart explanation of results we obtained. I would like to thank Ria Broer for refereeing the thesis, her sug- gestions and comments helped me to improve it a lot. I am grateful to Hermann Stoll for his help in overcoming diffi- culties I had with quantum chemical methods, both conceptual and practical. I would like to express my gratitude to Peter Fulde for insightful discussions that led to better understanding of the electron correla- tion problem on my side. I have greatly benefited from learning and working together with Vamshi Katukuri. Vamshi, thank you for pointing out so many im- portant papers and for sharing your ideas during our numberless discussions. I knew that I could always count on you. Among people I had the opportunity to work with in Dresden, I learned the most from Judit Romhányi, Ioannis Rousochatzakis and Viktor Yushankhai. Judit, thank you for sharing your knowledge on everything, especially on group theory, spin-wave theory and dif- ferent aspects of magnetism. Ioannis, I appreciate your help in un- derstanding the spin Hamiltonian machinery and its connection to macroscopic properties. Viktor, I am grateful to you for your detailed explanations of the spin-orbit coupling in solids, especially in the iri- dates problem and for suggesting great books on crystal-field theory. I had the luck to collaborate with Rémi Maurice during my short visit to Groningen. Rémi, thank you for showing me the chemist view on model Hamiltonian problems and demonstrating me the concept of effective Hamiltonian. Advice and comments given by Ganesh has been a great help in learning lattice models and reciprocal-space properties.

v I would like to thank all other fellows at ITF who kindly discussed science with me, Alexander, André, Beom Hyun, Ching-Hao, Dai- Ning, Dmitry, Frank, Jörn, Hsiao-Yu, Katia, Krzysztof, Lei, Liudmila, Pasquale, Sahinur, Satoshi, Stefanos, Stefan-Ludwig and Ravi. I am grateful to Ulrike Nitzsche for the IT support and to Grit Rötzer for her help in solving practical issues. I also acknowledge financial support from the Erasmus Mundus Programme of the European Union. I give my special thanks to Anna Efimenko and Maria Bogdanova, without their assistance I would never find my way to Dresden. Finally, I want to thank my family for their constant support and love. Спасибо всем вам!

vi CONTENTS introduction 1 i theoretical foundations 5 1 quantum chemistry approach 7 1.1 Local correlation treatment 9 1.2 Embedding for cluster calculations 14 ii 3d transition-metal compounds 25 2 d-level electronic structure of one-dimensional cuprates 27 2.1 Computational details 28 2.2 Chains of corner-sharing CuO4 plaquettes 29 2.3 Corner-sharing CuF6 octahedra in KCuF3 31 2.4 Chains of edge-sharing CuO4 plaquettes 32 2.5 A ladder cuprate: CaCu2O3 34 2.6 Conclusions 36 3 d–d excitation energies in Ti and V oxychlorides 37 3.1 Structural and computational details 37 3.2 Quasi-one-dimensional titanium oxychloride 39 3.3 Quasi-two-dimensional vanadium oxychloride 41 3.4 Conclusions 43 4 computationofx-ray d-ion spectra in solids 45 4.1 Resonant inelastic X-ray scattering 45 4.2 Electron – field interaction 46 4.3 X-ray absorption cross section 48 4.4 RIXS double differential cross section 49 4.5 Computational scheme for Cu L-edge excitations 50 4.6 Results and discussion 53 4.7 Conclusions 56 iii 5d transition-metal systems 57 5 magnetic properties of pyrochlore Cd2Os2O7 59 5.1 Os d3 electron configuration, single-ion physics 60 5.2 Isotropic and antisymmetric exchange interactions 64 5.3 Conclusions 68 6 anisotropic magnetism in 214 strontium iridate 71 6.1 Model Hamiltonian and ESR measurements 72 6.2 Quantum chemistry calculations of g factors 75 6.3 Anisotropic exchange couplings in Sr2IrO4 78 6.4 Conclusions 84

vii viii contents

7 post perovskite a 1 2 quasi 1d anti CaIrO3 - , jeff ≈ / - - ferromagnet 87 7.1 d–d splittings and magnetic g factors 89 7.2 Intersite exchange interactions 92 7.3 Conclusions 99

summary 101 publications 103 bibliography 105 LISTOFFIGURES

Figure 1.1 Exact solution and limiting cases within the CI paradigm 10 Figure 1.2 (a) Construction of configurations using the CAS scheme; (b) The notation for excited determi- nants 12 Figure 2.1 Chains of CuO4 plaquettes in Sr2CuO3 and SrCuO2 29 Figure 2.2 Chains of edge-sharing CuO4 plaquettes in four studied systems 33 Figure 2.3 MRCI crystal-field splittings in 1D Cu d9 com- pounds 36 Figure 3.1 Layered crystal structure of TiOCl and VOCl 38 Figure 3.2 Singly occupied Ti 3d orbitals for the ground and excited states in TiOCl 40 Figure 4.1 Sketch of the scattering in Li2CuO2 52 Figure 4.2 Calculated versus experimental L3-edge dou- ble differential cross section 54 Figure 4.3 Theoretical RIXS plots for Li2CuO2 using the geometrical setup 55 Figure 5.1 Sketch of (a) the cluster used for the ZFS calcu- lations, (b) the compressive trigonal distortion in Cd2Os2O7 60 Figure 5.2 (a) Single-ion anisotropy parameter D and (b) total energy of the cluster ∆Etot, as functions of the trigonal distortion angle ∆θ 64 Figure 5.3 The cluster used for computing the intersite magnetic couplings for Cd2Os2O7 66 Figure 5.4 Network of corner-shared Os4 tetrahedra in Cd2Os2O7 67 Figure 6.1 TM t2g splittings for tetragonal distortions of the oxygen 72 Figure 6.2 ESR data for Sr2IrO4 73 Figure 6.3 Planar IrO2 network in Sr2IrO4 78 Figure 7.1 Structural details of CaIrO3 88 Figure 7.2 Local coordination and the main AF superex- change path for corner-sharing IrO6 octahedra in CaIrO3 92 Figure 7.3 Sketch of the NN magnetic interactions in CaIrO3 99

ix LISTOFTABLES

Table 2.1 d–d excitation energies in Sr2CuO3 and SrCuO2 30 Table 2.2 d–d excitation energies in KCuF3 32 Table 2.3 d–d excitation energies for edge-sharing chain compounds 34 Table 2.4 d–d excitation energies for the ladder system CaCu2O3 35 Table 3.1 d-d excitation energies in TiOCl 39 Table 3.2 d-d excitation energies in VOCl 42 Table 4.1 Relative energies for Cu2+ 3d9 and 2p53d10 states of Li2CuO2 52 5+ 3 Table 5.1 Os 5d multiplet structure of Cd2Os2O7 61 Table 5.2 Single-site effective interaction matrix obtained by MRCI+SOC for Cd2Os2O7 62 Table 5.3 Interaction matrix for the single-ion anisotropy model Hamiltonian and S=3/2 62 Table 5.4 Os5+ 5d3 multiplet structure in a hypothetical Cd2Os2O7 crystal with no trigonal distortions 63 Table 5.5 Energies of the singlet, triplet, quintet, and septet 3 states for two adjacent Os d sites in Cd2Os2O7 65 Table 5.6 Model interaction matrix in the coupled basis tot tot 69 S , MS for two-sites and S1,2 =3/2 Table 5.7 Effective interaction matrix in the coupled ba- tot tot sis S , MS , obtained from two-site MRCI+SOC calculations for Cd2Os2O7 70 Table 6.1 g factors for Sr2IrO4 and Ba2IrO4 77 Table 6.2 Ir t2g splittings and g factors for Sr2IrO4 and Ba2IrO4 as obtained with the orca program 77 Table 6.3 Matrix elements of the ab initio spin-orbit Hamil- tonian for Sr2IrO4 79 Table 6.4 Matrix elements of the model spin Hamilto- nian diagonal in zero field 79 Table 6.5 Notations used to define anisotropic exchange 80 Table 6.6 Matrix form of the model spin Hamiltonian in the ‘uncoupled’ bond basis 81 Table 6.7 Matrix elements of the model spin Hamilto- nian in the coupled basis 82 Table 6.8 Matrix form of the model spin Hamiltonian with J and Γ¯ exchange brought to diagonal form 82 Table 6.9 NN magnetic couplings and g-factors in Sr2IrO4 as obtained from the SO MRCI 84 7 1 5 90 Table . Relative energies of t2g states in CaIrO3

x Table 7.2 Energies of the singlet and triplet states for two adjacent Ir d5 sites along c and a directionsin CaIrO3 93 Table 7.3 Matrix elements of the ab initio SOC Hamil- tonian for NN octahedra along the c bond in CaIrO3 95 Table 7.4 Matrix form of the model Hamiltonian for the c bond in CaIrO3 95 Table 7.5 NN magnetic couplings and g-factors for corner- sharing octahedra in CaIrO3 96 Table 7.6 NN magnetic couplings and g-factors for edge- sharing a-axis octahedra in CaIrO3 96 Table 7.7 Matrix elements of the ab initio spin-orbit Hamil- tonian for a bond in CaIrO3 98 Table 7.8 Matrix form of the model Hamiltonian for a- axis bonds in CaIrO3 98

ACRONYMS

1D one-dimensional

2D two-dimensional

3D three-dimensional

AF antiferromagnetic

CAS complete active space

CASCI complete-active-space configuration interaction

CASSCF complete-active-space self-consistent-field

CASPT2 complete-active-space second-order perturbation theory

CF crystal field

CC coupled cluster

CI configuration interaction

CSF configuration state function

DFT density-functional theory

DKH Douglas-Kroll-Hess

DM Dzyaloshinskii-Moriya

xi xii acronyms

DMET density-matrix embedding theory

DMFT dynamical mean-field theory

DMRG density-matrix renormalization group

ECP effective core potential

ESR electron spin resonance

FCI full configuration interaction

FTOR freeze-and-thaw orbital relaxation

FCIQMC full configuration interaction quantum Monte Carlo

FM ferromagnetic

FWHM full width at half maximum

GGA generalized gradient approximation

GS ground state

HF Hartree–Fock

KD Kramers doublet

LDA local density approximation

MCSCF multiconfiguration self-consistent-field

ME matrix element

MIT metal-to-insulator-transition

MO molecular orbital

MRCI multireference configuration interaction

NEVPT2 N-electron valence-state second-order perturbation theory

NIXS non-resonant inelastic X-ray scattering

NN nearest neighbour

PAO projected atomic orbital

RASSCF restricted active space self-consistent-field

RHF restricted Hartree–Fock

RIXS resonant inelastic X-ray scattering

RXD resonant X-ray diffraction

SCF self-consistent field acronyms xiii

SD singles and doubles

SIA single-ion anisotropy

SO spin-orbit

SOC spin-orbit coupling

TM transition metal

QC quantum chemistry

WF wavefunction

WO Wannier orbital

XA X-ray absorption

XMCD X-ray magnetic circular dichroism

ZFS zero-field splitting

INTRODUCTION

lenty of remarkable phenomena occur in transition metal (TM) compounds with a partially filled d shell. The main source P of the intriguing physics in these systems is the interplay be- tween electron delocalization and strong Coulomb interactions among d electrons. Electron delocalization caused by intersite orbital over- lap is a generic feature of crystalline solids and can be described within electronic band-structure theory [1]. Its successful application to simple metals and insulators went hand in hand with for instance developments in the field of semiconductors and semiconducting de- vices. According to conventional band theory a material is insulating if the valence (highest occupied) band is completely full. It implies that all materials with an odd number of electrons per unit cell have par- tially filled valence bands, and therefore are metals. However in the case of TM compounds this simple rule is not obeyed experimentally. For example, it was found experimentally that simple TM oxides like CuO and CoO (with 9 and 7 electrons in the d shell) are insulators. The missing ingredient in band theory is the instantaneous electron- electron repulsion that often stabilizes insulating ground states with localized unpaired electrons. This state, called Mott insulator, is qual- itatively different from the closed-shell band insulators. The main characteristic of the Mott insulator is the presence of localized spins that gives rise to intersite magnetic interactions. To describe these in- ^ teractionsP it is often sufficient to use the Heisenberg model HHeis = ~ ~ i,j JSi · Sj, where J denotes the isotropic exchange interaction and Si and Sj are spins at lattice sites indexed by i and j. Mott insulators can host different kinds of magnetic interactions — ferromagnetic, an- tiferromagnetic or combination of these when there are inequivalent magnetic bonds or longer-range magnetic interactions — depending on details of the crystal structure and the number of electrons in the d or f shell [2]. Consequently, many different types of magnetic order can be realized. Examples of interesting magnetic states can be found in so-called low-dimensional systems. They have an anisotropic crystal structure, such that the magnetic interaction is strong only in one or two di- rections, while remaining couplings are negligibly small. This way spin-chain and spin-plane models can be realized in real materials. Examples are layered compounds with Cu2+ ions, which are famous for being the parent compounds of the high-temperature supercon- ductors [3]. Chain systems with the same Cu2+ sites show very pecu-

1 2 introduction

liar novel excited states where spin and charge [4] or spin and orbital [5] of an electron in solid can split apart. Recently a lot of attention has been paid to TM compounds involv- ing 5d and 4d series of transition metals, namely Ir, Os, Ru, Rh. A special feature of these systems is that a Mott-like insulating state forms in the presence of strong spin-orbit coupling (SOC) within the d shell. Because of SOC, magnetic interactions occur not between con- ventional spins, but rather between pseudospins, that have both spin (~S) and orbital angular momentum (~L) components, so that the total angular momentum is ~J = ~S + ~L. Orbital angular momentum follows the symmetry of the crystal, which is always lower as compared to the full rotational symmetry of a pure spin. Therefore, the admixture of orbital part makes pseudospin magnetic interactions anisotropic, i.e. the three different spatial components of two pseudospins ~Ji and ~Jj may have different coupling strengths. In this case the Heisen- berg modelP should be generalized to the anisotropic exchange model ^ ~ ¯ ~ ¯ Haniso = i,j Ji · T · Jj, where T stands for the anisotropic exchange ten- sor. The more complicated magnetic coupling terms allow many more exotic magnetic structures to be realized in the case of pseudospins. For example, it is in principle possible to have bond-dependent in- teractions in honeycomb lattices plane reproducing the recently pro- posed Kitaev spin model [6]. In the Kitaev model each pair of the neighbouring sites in honeycomb lattice is coupled via only one of the three components of the (pseudo-)spin depending on direction of the bond. It turns out that in real honeycomb materials with strong SOC, e.g. Na2IrO3 and Li2IrO3, such Kitaev-type coupling exceeds isotropic Heisenberg interactions [7]. This model is of interest because it represents a non-trivial exactly solvable spin model in two dimen- sions, in which spin excitations fractionalize into two different sets of quasiparticles (Majorana modes and flux-modes). Such property has been argued to be of relevance to topological quantum computing [6]. In this thesis we employ methods of quantum chemistry (QC) for ab initio studies of the TM oxides. In the field of theoretical chemistry the problem of treating strong correlations (many-body electron-electron interactions) has a long history. It was first discussed in connection to the study of the H2 molecule, when the Heitler-London correlated ansatz for the wavefunction (WF)[8] was opposed to the independent- electron molecular orbital (MO) approach by Hund, Slater and others [9, 10]. During the last 80 years QC computational methods were con- stantly developed and tested for tackling different kind of problems. An established hierarchy of approximations for the N-electron WF al- lows to systematically improve the accuracy of the calculation. Start- ing with the single-Slater-determinant Hartree–Fock (HF) method as a first approximation and possibly end with the full configuration in- teraction (FCI) that is exact within a given basis set. One can compare the accuracy of different models for an extensive list of typical appli- introduction 3 cations. When results of the less accurate approximation become very close to the higher level of theory, the former is considered to be ap- propriate for the task. Widely used computational schemes based on the multireference configuration interaction (MRCI) approach for the WF and many-body second-order perturbation theory showed abil- ity to describe well electron correlations in molecular and solid-state systems. In order to treat solid systems, one can use a relatively large cluster embedded in an effective potential that accounts for the environment. The essentially localized character of valence electrons in Mott insula- tors together with the local nature of electron correlations [11] makes such embedded-cluster approach widely applicable. It was success- fully used for calculation of different properties, such as d–d split- tings and strengths of magnetic interactions in many TM oxides with accuracy close to experimental [12–17]. Structurally the thesis is divided into three parts. In Chapter 1 we discuss the WF-based methods suitable for studies of strongly cor- related systems with closer attention to the description of embed- ding techniques relevant for solids. The second part contains the re- sults of studies of the first row TM compounds. Calculations of d–d transitions for a variety of highly anisotropic one-dimensional (1D) cuprates is presented in Chapter 2. In Chapter 3 we report a compli- mentary study of d–d splittings for early 3d, titanium d1 and vana- dium d2 compounds. In Chapter 4 we formulate a computational scheme for absolute intensities of resonant inelastic X-ray scattering (RIXS) spectra based on ab initio QC calculations and apply it to the Cu L-edge in chain compound Li2CuO2. The third part of the the- sis includes studies of 5d TM oxides with characteristic strong SOC. In Chapter 5 we present a study of the Cd2Os2O7 pyrochlore com- pound, an Os5+ 5d3 system, where the effect of SOC leads to large on-site spin anisotropy. We find that type of anisotropy can be var- ied via a fine interplay of the local ligand-cage distortions and the electrostatic field due to the further-neighbour environment. In Chap- ter 6 and Chapter 7 we explore the details of the ground and low- lying excited states, including interactions with magnetic field and anisotropic magnetic exchange, in the prototypical j=1/2 compound Sr2IrO4 and the quasi-1D antiferromagnet CaIrO3 respectively. For quantitative studies of anisotropic exchange we introduce in Chap- ter 6 a procedure for one-to-one mapping of the QC results onto spin model Hamiltonian.

Part I

THEORETICALFOUNDATIONS

QUANTUMCHEMISTRYAPPROACH 1

eliable first principles studies of TM compounds are essential for understanding their very diverse properties. The delicate R interplay between lattice, orbital and spin degrees of free- dom in these materials can lead to major differences even between systems that are chemically or structurally very similar. Therefore semi-empirical approaches that require some prior knowledge about the system can only be used with great care. For the investigation of the electronic structure of solids at the ab initio level two strategies are used: periodic and embedded cluster. Translational symmetry is one of the characteristic features of crys- talline solids. For this reason it is tempting to formulate the electronic- structure calculations on crystalline systems along the same line – model the ‘repeating’ unit cell to obtain the electronic structure of the infinite periodic system. However, difficulties in the practical im- plementation, namely the exponential growth of the Hilbert space with the system size, strongly restrict the number of available ab ini- tio methods for periodic calculations [18–20]. In ab initio condensed matter studies the most used method is the density-functional theory within the local density approximation (LDA) or generalized gradient approximation (GGA)[21]. The Kohn-Sham density-functional theory (DFT) is an efficient ap- proach for the calculation of the ground-state properties of weakly correlated systems, such as lattice constants, bulk modulus or cohe- sive energies [21]. DFT is in principle exact for all properties deter- mined by the ground state electron density, but in practice the exact exchange-correlation potential is not known, which implies certain assumptions and approximations. Moreover, it is common practice to interpret the Kohn-Sham bands, initially built in calculations only to obtain the electron density, as approximated quasiparticle bands. Band gaps computed this way are usually strongly underestimated as compared to the measured values [22]. It is even more problem- atic for strongly correlated systems, such as TM and rare earth com- pounds, where DFT within the LDA (and GGA) fails to predict the correct insulating ground state (GS). To improve the simulation of such materials, two extensions are often used, DFT+U [23, 24] and DFT+dynamical mean-field theory (DMFT)[25]. Both methods, with proper parametrizations, are capable of obtaining qualitatively cor- rect GS and band gaps in TM oxides, but strictly speaking one cannot refer to them as ab initio.

7 8 quantum chemistry approach

Another first principles technique utilizing the translational sym- metry is the periodic HF method [26]. The values of band gaps cal- culated with periodic HF are overestimated because the bands are optimized for the ground state, and polarization and relaxation of the surrounding in the excited state is not accounted for [11]. The fact that the band gap is either underestimated by LDA or overestimated by HF stimulated the construction of the so-called hybrid function- als that mix LDA and HF contributions to obtain band gaps close to experiment [27]. Both the DFT and HF methods employ the independent particle ap- proximation, assuming each electron is moving in an effective mean field generated by all other electrons, and therefore fail to correctly describe systems in which many-body interactions are strong [11]. In the last decade several post-HF algorithms for periodic calculations have been developed. They utilize either many-body perturbation the- ory [28–31] or more advanced coupled cluster (CC) methods [32, 33]. These algorithms were applied to simple covalent and ionic solids [34–36], however their extension to open-shell systems is not trivial. Recently a new method based on quantum Monte Carlo sampling of the Hilbert space of Slater determinants, full configuration interaction quantum Monte Carlo (FCIQMC), was adapted to periodic solids [37]. First studies show satisfying results for the spin gap in NiO, as well as for the formation energies of simple solids [37]. Nevertheless, further testing is required to justify the general applicability of FCIQMC. As it was mentioned in the beginning of the chapter, a different class of computational methods for solids is formulated within the embedded cluster approximation. Its main idea is to handle only a small part of the solid at an advanced level, while the rest of the system can be treated in a simpler manner. For the cluster region it is possible to use the most accurate many-body QC methods ini- tially proposed and well tested by now for molecular systems. The embedding can be formulated in many different ways. It should re- produce effects such as the symmetry of the surroundings and the Madelung field in the cluster region. The simplest way to achieve this is to adopt the picture of a fully ionic model and simulate an infinite array of ionic point charges around the cluster. Already with such point-charge description of the crystalline surroundings, the ab initio cluster calculations can describe local d-d and magnetic excitation en- ergies in strongly correlated TM oxides with rather good accuracy. More advanced embedding schemes can be formulated using prior periodic DFT or HF calculations to further improve the accuracy. In this chapter we concentrate on the main concepts of the em- bedded-cluster QC approach, which is used in the rest of thesis to study the electronic and magnetic properties of TM compounds. First we discuss how the correlated N-electron WF of the cluster can be constructed, further we describe how to build point-charge and HF 1.1 local correlation treatment 9 embedding for cluster calculations and mention several other embed- ding techniques.

1.1 Local correlation treatment

One of the central goals in quantum mechanics is to find a solution of the time-independent Schrödinger equation, in order to describe the electronic structure of the studied material. The starting point is often the Born-Oppenheimer approximation, which separates the nuclear and electronic parts of the Schrödinger equation. Because of the huge difference in masses of nuclei and electrons, the latter can respond al- most instantaneously to any displacements of the former. Therefore, it is often sufficient to focus only on the electronic Schrödinger equa- tion for fixed locations of the nuclei

H^ BO Ψ(~r) = EΨ(~r), (1.1) where the electronic wavefunction (WF) Ψ explicitly depends on the coordinates of all electrons~r and E is the electronic energy. The Hamil- tonian (in atomic units) reads

XN XN XNn XN H^ 1 2 Zα 1 1 1 2 BO = − ∇i − + ,( . ) 2 ~ 2 ~ri −~rj i i α ~ri − Rα i6=j with the first two terms being the one-particle kinetic energy and the electron-nucleus interaction and the last term describing electron- electron interactions. Analytic solutions of Equation 1.1 can be found + only in very simple cases like the H atom or H2 ion, while for more complex systems only approximate and/or numerical solutions can be achieved. In WF-based approaches, one chooses a model for the N- electron WF and solves Equation 1.1 using only the atomic numbers of the nuclei and fundamental constants. In other words, the accuracy of the QC calculation and the computational effort needed for it are determined by the approximation chosen for Ψ. The approximation for the WF can be improved in a systematic way towards the exact solution The exact WF can be written as a linear combination of all ‡ here the m index possible N-electron Slater determinants Φj, or of configuration state stands for both spatial and spin functions Ξk, arising from a complete set of one-electron spin-orbitals ‡ coordinates. ϕm using a hierarchy of standard QC methods developed and exten- sively used over several decades [38]. To understand what CSF is, one can look X X X X X at a simple problem k k 1 3 of two electrons on Ψ = CkΞk = Ck Cj Φj = Ck Cj kϕaϕb ··· ϕNk .( . ) different orbitals. It k k j k j is possible to cons- truct two ‘low-spin’ Each CSF is a spin and symmetry adapted linear combination of determinants Ck Slater determinants. The coefficients j are fully defined by the spin k↑ψa ↓ψbk and k↓ψa ↑ψbk. But only their linear combinations will be eigenfunctions of the total S2 operator. 10 quantum chemistry approach

# of CSF's 1010 Exact Full CI 105 Complete- basis-set limit HF 1 min 50 100... # of basis functions Figure 1.1: Exact solution and limiting cases within the CI paradigm.

and symmetry restrictions, for brevity they will be implicit in the fol- lowing. An ab initio method in which the WF is expressed in the form of Equation 1.3 is called configuration interaction (CI). The full CI expansion is computationally not feasible for most systems because of the two limiting factors summarized in Figure 1.1. First, it is not possible to construct an extremely large number of N-electron CSF’s within the space of all spin-orbitals ϕm. The second difficulty is that the spatial parts of the spin-orbitals ψm themselves are constructed from a set of atomic-like basis orbitals χl, with

XNb m 1 4 ϕm = σmψm = σm cl χl , ( . ) l

and the total number Nb of these basis functions χl has to approach infinity (complete basis set) to represent the WF exactly. Several ex- trapolation schemes were proposed to estimate the complete-basis-set limit, based on performing several sets of calculations gradually ex- panding the basis sets [38]. Fortunately, comparison to experimental results shows that reasonable results can be obtained with finite basis sets. It is common to represent each atomic-like core orbital by only one basis function and use two or more functions for every valence orbital; such basis set is usually quoted as a split-valence basis. De- pending on the number of basis functions per valence orbital the set is called valence-double, valence-triple-zeta basis set and so on. Zeta is a parameter that was in early studies commonly used to represent the radial part of the orbital e−ζr. For small systems it is computationally possible to build WF’s us- ing all CSF’s in a given finite basis set (N-combinations from a set of 2Nb elements). Such calculations, referred to as FCI, are equiva- lent to exact diagonalization and are usually taken as reference for all other more approximate QC methods. The number of possible de- terminants grows exponentially with the number of particles and in practice it becomes too large to be feasible in terms of both computer 1.1 local correlation treatment 11 time and memory. Hence, schemes that restrict the number of consid- ered determinants have to be employed. The simplest limit is to reduce Equation 1.3 to a single CSF (or a single Slater determinant if one wants to avoid working with eigen- functions of S2). The best WF of such type (corresponding to the lowest energy) can be obtained by using the Hartree–Fock (HF) self- consistent field (SCF) method, which relies on variational theory to minimize the expectation value of the Hamiltonian (1.2)

| H^ | EHF = min hΨHF BO ΨHFi . (1.5) l cm

l The variation parameters cm are the orbital expansion coefficients, as defined in Equation 1.4. The energy expression is non-linear in l terms of cm, therefore they have to be optimized within an iterative procedure. The single-determinant WF can describe relatively well the GS of closed-shell systems and some useful information can be extracted from the HF results, e.g. approximate ionization energies with the help of Koopman’s theorem [38]. However, the HF method cannot handle situations in which sev- eral CSF’s are very close in energy and need to be treated on equal footing, i.e. when the WF is multiconfigurational in nature. This hap- pens in TM compounds with partially filled t2g or eg shells or in bond-breaking process in molecules. A question that arises is how to select the configurations that should be included into the WF. The nearly degenerate CSF’s usually differ in the occupations of a rather small set of orbitals, for example five of d character. We can then re- strict ourselves to these relevant (valence) ϕm to build a set of CSF’s for the CI expansion. This idea is realized within the complete ac- tive space (CAS) configuration selection scheme, in which the orbital space is partitioned into three subspaces: active, inactive and virtual. The active orbitals can have any occupation, while the inactive ϕm are doubly occupied and the virtual are kept unoccupied, see Figure 1.2a.

It is possible to use orbitals optimized within the HF calculation together with the CAS scheme and minimize the energy varying only the CI expansion coefficients. Such complete-active-space configura- tion interaction (CASCI) calculation would be biased towards the CSF converged in the HF calculation. To treat all configurations on equal footing, it is necessary to take into account orbital relaxation together with the multiconfigurational expansion, as it is done in the multi- configuration self-consistent-field (MCSCF) approach. Here the vari- l ational parameters are both the orbital expansion parameters cm of Equation 1.4 and the CSF expansion coefficients Ck of Equation 1.3:

| H^ | hΨMCSCF BO ΨMCSCFi EMCSCF = min . (1.6) l | cm,Ck hΨMCSCF ΨMCSCFi 12 quantum chemistry approach

φ Nb virtual ...... φq φp ...... active − φc φb φa inactive ...... φ2 φ1 p pq Ξ1 Ξ2 ... Ψ0 Ψa Ψab (a) CAS (b) Truncated CI

Figure 1.2: (a) Sketch showing the construction of configurations using the complete active space scheme; (b) The notation for excited deter- minants within the truncated CI approach.

There are two commonly used iterative procedures to handle the MCSCF WF optimization. In the Newton-Rapshon method both types of parameters are optimized together based on a Taylor expansion of the energy expression. In the super-CI scheme, the parameters are op- timized each iteration separately in two sub-steps. First, the best CI l coefficients Ck for the current orbital coefficients (cm) are found, then l 38 the cm parameters are optimized for the approximate WF [ ]. Both minimization procedures are computationally expensive and can be used only with a limited number of orbitals , as one of the steps is a full CI within the active space with its limit of 16 electrons in 16 orbitals. The MCSCF method is most often used with the CAS scheme for se- lecting the configurational space and commonly referred to as comp- lete-active-space self-consistent-field (CASSCF). There are also more flexible approaches that partition orbitals into more than three ‘spaces’. In the restricted active space self-consistent-field (RASSCF) method the active space is further subdivided into three subclasses denoted as RAS1, RAS2 and RAS3 [38]. Additional restrictions are imposed on the maximum numbers of holes in RAS1 and on maximum num- bers of electrons in the RAS3 subsets. The advantage of RASSCF is the possibility to treat larger active spaces. Larger number of orbitals can be correlated on top of a MCSCF WF with the help of the multireference CI method. It is possible to 1 3 reformulate the expansion given in Equation . classifyingP all CSF’s 38 as excitations on top of the reference MCSCF state Ψ0 = k CkΞk [ ]. Such an expansion reads X X X p p pq pq pqr pqr 1 7 ΨCI = C0 Ψ0 + Ca Ψa + Cab Ψab + Cabc Ψabc + ... ,( . ) a a

where each term corresponds to a class of CSF’s with a particular number of excitations on top of Ψ0 as explained in Figure 1.2b. As 1.1 local correlation treatment 13 we start from a multiconfigurational WF, the Equation 1.7 defines the multireference configuration interaction (MRCI) method, where the energy expectation values are given by

hΨ | H^ |Ψ i E = min MRCI BO MRCI , (1.8) MRCI @ | C@ hΨMRCI ΨMRCIi @ here C@ is the full set of coefficients that show up in Equation 1.7. Because the Hamiltonian (1.2) contains only up to two-body interac- tions, it is reasonable in first approximation to truncate the expan- sion (1.7) with only singly and doubly excited configurations. Indeed, the MRCI-singles and doubles (SD) results are close to the FCI data in many test calculations and thus seem to capture a major part of dynamical correlations, related to short-distance two-electron interac- tions [38]. However, it is also known that truncated CI suffers from size inconsistency. A method is called size-consistent when the energy Truncated CI is of a system consisting of two infinitely separated fragments isequal neither size-exten- to the fragment energies computed separately with the same method. sive meaning that the energy of a It turns out that for MRCI-SD methods this condition is not fulfilled system is not [38]. For relatively small clusters and molecules close to equilibrium proportional to the the error is small, moreover it can be improved by using approximate number of non- correction schemes [39–41]. interacting electrons. Alternatively, electron correlations on top of MCSCF can be cap- tured by non-variational methods. Second order many-body pertur- bation theory is the simplest of them, with a much smaller com- putational effort compared to variational CI. According to standard Rayleigh-Schrödinger perturbation theory the full Hamiltonian is ex- pressed as a sum of the zeroth-order Hamiltonian H^ (0) and a per- V^ H^ turbation . The full Hamiltonian is BO given by Equation 1.2. The perturbation is formally defined as V^ H^ H^ (0) = BO − , (1.9) and the second-order correction to the energy is D ED E X (1) V^ (0) (0) V^ (1) Ψk Ψ Ψ Ψk E(2) = , (1.10) (0) (0) k E − Ek (1) (0) where Ψ is a multiply excited CSF with an expectation value E = D k E k (1) H^ (0) (1) Ψk Ψk . Different perturbative methods can be built de- pending on the choice of the unperturbed Hamiltonian. Møller and Plesset proposed to use a zeroth-order Hamiltonian that has the Har- tree-Fock WF as an eigenfunction [42]. The correspondent second- order perturbation method is designated MP2 [38]. However, for a multireference WF the Møller-Plesset partitioning of the Hamiltonian is not possible and another definition for H^ (0) has to be formulated. The most widely used multireference perturbation scheme is the comp- lete-active-space second-order perturbation theory (CASPT2)[43, 44], 14 quantum chemistry approach

where the zero-order Hamiltonian is chosen to be Møller-Plesset-like. It is a sum of Fock-type one-electron operators that reduces to the Møller-Plesset operator in case of a single-configurational reference WF. The first-order WF Ψ(1) is constructed by an internally contracted expansion of all CSF related to single and double excitations on the reference CASSCF WF. However, in some cases the one-electron na- ture of the zero-order Hamiltonian can lead to the appearance of so- called intruder states. It happens when one of the configurations in (1) H^ (0) the first-order interacting space Ψk has the expectation value of (0) (0) very close to the reference function, i.e. E ≈Ek . In this case the de- nominator in Equation 1.10 is close to zero and even if the interaction matrix elements in the numerator in Equation 1.10 are very small, an artificially large contribution from such intruder state appears, caus- ing the perturbation expansion to diverge. To overcome this prob- (0) lem a level shift can be applied to Ek , such that the near degener- acy is avoided [38]. N-electron valence-state second-order perturba- tion theory (NEVPT2) is another multireference perturbation scheme with the main advantage of having no intruder states [45, 46]. To define the zero-order energies, it uses the Dyall Hamiltonian [47] con- taining two-electron interactions between electrons on active orbitals. CASPT2 and NEVPT2 give qualitatively similar results and both turn out to be very efficient methods for the description of the electronic structure of insulating TM-based materials [48, 49]. More recently, several new computational methods for building of the N-electron WF have been proposed in the literature. Among the most interesting are the density-matrix renormalization group (DMRG) approach in the QC framework [50, 51] and the FCI opti- mization by stochastic quantum Monte Carlo (FCIQMC)[37, 52]. Both of them can be used as replacements to the FCI. In the MCSCF frame- work it means that larger active spaces become accessible [53–56]. For- mulation of many-body perturbation theory with the help of DMRG and FCIQMC is also possible [57–60]. Other promising directions are the so-called explicitly correlated methods, in which terms that explic- itly depend on the inter-electronic distances r12 are included into the WF, such that a complete-basis-set can be effectively reached [61, 62]. Development of the conventional CAS [63, 64] and CI [65, 66] meth- ods towards better flexibility and higher computational efficiency is also in progress.

1.2 Embedding for cluster calculations

The computational effort of the methods discussed in the previous section grows very fast with increasing system size. The ’manageable’ number of electrons is limited to tens or hundreds, depending on the method. However, the size of the system to be handled can be significantly reduced with the help of additional simplifications for 1.2 embedding for cluster calculations 15 the farther environment. It is the aim of the embedding approach to extend the applicability of accurate QC methods from isolated molecules to complex systems including solids. Over the years var- ious embedding schemes based on both semi-empirical and quan- tum mechanical electronic-structure methods have been developed, see for example reviews [67] and [68]. We will discuss in detail only two of them, that are applicable to strongly-correlated oxides and are used in the present thesis, namely the point-charge and the WF-based cluster-in-solid embedding schemes. Electrostatics is energetically among the largest interactions in solids. Hence if we have chosen a finite region of interest and want to describe the impact of the environment, the Madelung field due to ions that surround the fragment has to be accounted for. In the simplest model every ion is represented by its total charge concen- trated at a lattice point. Then the potential VPC generated by all point- charges can be calculated as an infinite sum over the lattice sites:

X unitX cell qα 1 11 VPC(~r) = , ( . ) ~r − (~Rα + ~k) ~k α where ~k is a vector of the Bravais lattice and α refers to a charge qα located at the position ~Rα of the unit cell. Simple truncation of the real-space sum by taking large but finite ~k in Equation 1.11 converges rather slow with respect to the size of the region, moreover it is diver- gent at some truncation levels. A way to ensure fast convergence of Equation 1.11 is to eliminate the multipolar moments in the unit cell ‡ by applying the Evjen method [69] or its extensions [70, 71]. In the ‡ It guarantees the original Evjen scheme the sum (1.11) stops after a finite number of overall charge crystal unit cell and the ‘terminal-surface’ charges are renormalized neutrality and cancellation of the by factors 1/2, 1/4, or 1/8 if the charge occurs at a , an edge, or dipolar moment. a corner of the charge array. Alternatively, the Madelung field VPC can be calculated exactly using the Ewald summation, in which the sum is split into real-space and reciprocal-space parts [72]. For a real space setting, the resulting potential is commonly represented by a finite array of effective point charges, from which the embedding po- tential can be constructed using Equation 1.11 [73]. It is also possible to directly use the potential obtained from the reciprocal-space sum within the so-called periodic electrostatic embedding scheme [74]. The point-charge embedding potential is fully determined by the crystal structure and the values of the distributed charges. For a given compound with defined lattice parameters, formal ionic charges can be easily assigned by assuming purely ionic bonding. In TM com- pounds with small oxidation numbers, usage of the formal charges is sufficient to provide reasonably good description of the electro- static environment. However, for ions in high oxidation states (+4 and more) the bonds can display substantial covalent character. In some 16 quantum chemistry approach

Generally speaking cases covalency effects can bring sizable corrections to, e.g. the calcu- significant amount lated magnetic exchange [75]. The effect of covalent bonding can be of covalency can taken into account by introducing ‘partial’ charges resembling better occur even for ions in low oxidation the actual electron distribution. These partial charges can be derived states. from a periodic calculation or obtained within the embedded-cluster scheme with an iterative update of charges in the embedding until they correspond to the counterparts in the cluster region. To estimate partial charges one has to analyze the electronic WF or electron density. The determination of these charges is not always straightforward and in general is not unique [76]. The most com- monly used approach is the Mulliken population analysis [77], in which the partial charges are determined via summation of overlaps between the basis functions centered at a given ion site and the N- particle WF. But it is known to be strongly dependent on the choice of the basis set and has problems with overestimating covalency. More sophisticated methods, like the natural population analysis [78] and the atoms-in-molecule model [79] have been proposed to overcome these difficulties. Recently a new technique called intrinsic atomic or- bitals has been used to calculate values of the partial charges in a rather simple and consistent fashion [80]. Point-charge embeddings account reasonably well for the effect of the long-range electrostatic interaction of the cluster with the remain- ing infinite solid. Problems, however, occur at the boundary between the QC cluster and the classical point-charge region. Bare positive charges located next to the cluster cause an artificial polarization of cluster electrons due to the lack of short-range repulsion effects of quantum-mechanical origin. It can be seen as absence of orthogonal- ity between the cluster WF and the embedding [81, 82]. To improve the simplest point-charge description one can introduce a buffer re- gion between the bare charges and the actual cluster. This buffer should to be treated quantum-mechanically, but eventually at the low- est level of approximation. One simple way is to simply increase the size of the QC cluster, perform preliminary HF calculation for the extended cluster, and freeze orbitals centered at the buffer sites. The buffer sites can be also represented by large-core effective core po- tentials (ECP’s) with no associated valence electrons (referred to as total-ion potentials). Such ECP’s can be obtained using either atomic calculations [83] or by optimization for each compound in a series of individual HF SCF calculations [84]. A more accurate description of the infinite surrounding is obtained when both short- and long-range interactions are treated within a single framework. This can be achieved by utilizing prior periodic calculations, such as DFT or HF. Although it is possible to perform WF-based calculations by using a DFT embedding [85], its usage for strongly-correlated materials is problematic because with DFT one of- ten obtains an erroneous metallic ground state. Since in the present 1.2 embedding for cluster calculations 17 thesis the HF-based embedding method is used, we will discuss it in more detail. A comprehensive discussion of DFT-based frozen den- sity embedding one can find, for instance, in references [68] and [86]. Based on the theory of electron separability [81, 82], the WF of the full system Υ(~x1,~x2, ...,~xN) can be written as a generalized product of the subsystems’ WF’s:

 A Υ(~x ,~x , ...,~x ) = MA^ Ψ (~x ,~x , ...,~x ) 1 2 N 1 2 NA 1 12 B  ( . ) ×Ψ (~xNA+1,~xNA+2, ...,~xNA+NB ) × ... , where M and A^ denote a normalization factor and the antisymmetriza- tion operator, respectively. Such construction can be always done for a single-determinant WF with NA = NB = ... = NN = 1. However, for generally defined groups with NR >1 an additional constraint has to be imposed onto the group functions : Z R † S 1 13 Ψr (~x1, ...,~xi,~xj, ...) Ψs (~x1, ...,~xk,~xl, ...)d~x1 = 0 (R6=S),( . )

R where Ψr describes group R in state r. This strong-orthogonality con- dition states that the set of all possible determinants is partitioned into disjoint subsets and every group function can be built only from ‘its own’ subset of spin-orbitals, which are orthogonal to all spin- orbitals belonging to the other subsets. Accordingly, the total Hamil- tonian of the full system can be written as a sum of all internal group Hamiltonians extended with interaction terms between differ- ent groups: X 1 X X h i H^ full = H^ R + ^JRS − K^RS . (1.14) 2 R R S6=R

^ R Each H describes NR electrons alone ignoring all other groups and has non-zero matrix elements only in the spin-orbital subspace of the R group:

XNR XNR XNR R R R R R 1 R hΥ| H^ |Υi = Ψ H^ Ψ = Ψ h^ (i) + g^ (i, j) Ψ . 2 i∈R i∈R j6=i∈R (1.15)

Here and in the following the lowercase h^ (i) stands for the one- electron Hamiltonian and g^ (i, j) = 1/^r(i, j) is the two-electron repul- sion term. The second part of Equation 1.14 is an inter-group interaction de- fined for each pair of group functions as

RS RS R S RS RS R S hΥ|^J − K^ |Υi = Ψ Ψ ^J − K^ Ψ Ψ XNR XNS XNR XNS ^ 1 16 R S 1 P(i, j) R S ( . ) = Ψ Ψ − Ψ Ψ , rij rij i∈R j∈S i∈R j∈S 18 quantum chemistry approach

^ where rij is the interelectronic distance and the operator P(i, j) inter- changes electron i in ΨR with electron j in ΨS. The particular form of Equation 1.16 makes possible to introduce the one-electron Coulomb ^ ^ and exchange group operators JS(i) and KS(i) describing the poten- tial due to electrons in group S:

XNS ^ S 1 S JS(i) = Ψ Ψ , rij j∈S (1.17) XNS ^ ^ S P(i, j) S KS(i) = Ψ Ψ . rij j∈S 1 14 Moreover, the full HamiltonianP ( . ) can be now expressed as a sum H^ full H^ R over groups, = R eff, with

N N N XR 1 XR XR H^ R = h^ (i) + g^ (i, j) eff eff 2 i∈R i∈R j6=i∈R   (1.18) N   N N XR 1 X h i 1 XR XR = h^ (i) + ^J (i) − K^ (i) + g^ (i, j) .  2 S S  2 i∈R S6=R i∈R j6=i∈R  H^ R R Each eff is spanned within the R-group spin-orbital subspace ϕ or in the space of basis functions associated with the group χR 1 4 H^ R R R following Equation . . Therefore each of the effΨ =EΨ equations can be solved individually. ^ ^ Matrix elements of JS(i) and KS(i) are most conveniently defined in terms of the one-electron reduced density matrix ρS of the group S [81]: X R ^ R S R S R S ϕi JS ϕj = ρkl ϕi ϕl g^ ϕj ϕk , k,l X (1.19) R ^ R S R S R S ϕi KS ϕj = ρkl ϕi ϕl g^ ϕkϕj , k,l

where on the right hand side we have linear combinations of two- electron integrals connecting groups R and S, multiplied by matrix elements (ME’s) of ρS in the spin-orbital basis of the group S. The reduced density matrix in the MO basis can be defined in terms of ^ ^ Xkl takes a very excitation operators Xkl that probe if two determinants differ in the † simple form akal in occupation of orbitals k and l: second quantization X language. | ^ | ∗ | ^ | 1 20 ρkl = hΨ Xkl Ψi = CI CJ hΦI Xkl ΦJi . ( . ) I,J

Expanding Equation 1.19 in the finite atomic-like basis using Equa- tion 1.4 is cumbersome and requires four-indices summations. But if we restrict ourselves to the case of the closed-shell HF solution for ΨS, the corresponding reduced density matrix takes the diagonal 1.2 embedding for cluster calculations 19

S form ρkl = δk,l, such that spin and orbital sums can be eliminated. The required operators expressed in terms of basis functions X R ^HF R S R S R S χm JS χn = Ptu χmχu g^ χnχt , t,u X (1.21) R R 1 S R S R S χ K^HF χ = P χ χ g^ χ χ , m S n 2 tu m u t n t,u can be easily evaluated and added to the one-electron Hamiltonian ^R 1 18 S heff in Equation . . Here Ptu are elements of the reduced density matrix in the atomic-like basis. For the closed-shell HF solution they are fully defined by the expansion coefficients of the occupied MO’s

X NXS/2 S l k∗ HF↑↓ k k∗ 1 22 Ptu = ρklctcu = 2 ct cu . ( . ) k,l k σ Equation 1.18 defines an embedding problem, where the effect of one or more groups S onto the group of interest R is completely given by Coulomb and exchange potentials as in Equation 1.19 or equiva- lently by the one-body reduced density matrices Equation 1.22. It is quite general and can be applied to any system that consists of sepa- rable groups of electrons [81]. These ideas were applied to insulating solids within the cluster-in-solid embedding scheme [87–90]. The electronic state of a periodic system is commonly represented in terms of Bloch functions, which are simultaneously eigenfunctions of the periodic Hamiltonian and the lattice translation operator:

f (r)=u (r) ei~k·~r 1 23 ~k,α ~ ~k,α ~ , ( . ) α e (~k) u (r) where labels the bands α , and ~k,α ~ are functions that have the periodicity of the Hamiltonian. Equivalently, the electronic structure can be represented in real space with the help of spatially-localized Wannier orbitals (WO’s)[1]. Bloch waves can be transformed into WO’s by an unitary transformation: Z 1 ~ ~ w (r)= √ e−ik·RL f (r) d3~k 1 24 L,α ~ ~k,α ~ , ( . ) Ω BZ with Ω standing for the volume of the Brillouin zone and ~RL being the lattice vector associated with a unit cell L. The complete orthonormal set of WO’s is defined within a single reference unit cell specified by the lattice vector ~R0. However, each WO can spread beyond the unit cell where it is centered and in principle have infinitely long ‡ radial distribution . The latter point becomes more clear when we ‡ In insulators it E quickly decays after expand a WO in a finite Gaussian basis set µ(~RL) very much few close-neighbour 1 4 alike Equation . . In Dirac notation this expansion reads sites, while in metals WO’s are more solidX XNµ E α E delocalized. wα(~R0) = c µ(~RL) . (1.25) µ(R~ L) R~ L µ(R~ L) 20 quantum chemistry approach

In practical implementations the sum over lattice vectors ~RL is re- stricted to a large set of ~RL using certain cutoff criteria [91]. In the embedded-cluster approach the full system is partitioned into the cluster and the embedding parts in real space. In this case the real-space WO’s representation is a natural choice, as localized WO’s can be easily partitioned into the corresponding groups. WO’s as ob- tained from the periodic HF calculation are used within the cluster-in- solid scheme [87–90] not only for construction of the embedding, but also as starting orbitals for the cluster calculations. The required peri- odic HF solution can be found directly in the basis of WO’s [92, 93], but more commonly it is done in terms of Bloch functions with sub- sequent transformation to WO’s. For some of the investigations pre- sented in this thesis, we used the crystal package for the prior pe- riodic calculations [94] together with the built-in Wannier-Boys local- ization module [91]. From the set of occupied HF WO’s {wκ} one can select a subset associated with ions of a finite cluster C. In general C may not coincide with the crystallographic unit cell or a larger supercell, therefore the C ~ C cluster subset should be defined as wκ ≡ wκ(RL) ∈ . The rest of the occupied WO’s implicitly form an infinite subset associated with ‡ E { } \ C ‡ Embedding set is the HF environment wκ = wκ wκ . This way each WO belongs the full set minus either to the cluster  or the environment, in other words the two sets the cluster one. E C are disjoint wκ ∩ wκ = ∅. Therefore, determinants constructed within each set obey the strong orthogonality condition (1.13) and the single-determinant HF wavefunction of the infinite system can  C E be written as a generalized product Φfull =MA^ Φ Φ according to Equation 1.12 [81]. Furthermore, following Equations 1.14 and 1.18 the full Hamiltonian reads h i H^ full H^ C H^ E ^CE ^CE H^ C H^ E 1 26 = + + J − K = eff + eff. ( . )

An important feature of Equation 1.26 is that effective operators on the right-hand side are defined in nonintersecting WO suspaces, and H^ full is effectively block-diagonal, with ME’s

| H^ C | H^ full C 1 27 hwα eff wβ =hwα wβ , wα, wβ ∈ .( . )

On the other hand, by regrouping terms in Equation 1.18 one can formally write

H^ C H^ C ^emb 1 28 eff = + V , ( . )

where V^emb is the embedding potential operator that incorporates all C terms on top of the bare H^ defined in Equation 1.15. Within the HF method ME’s of the embedding potential V^HF can be obtained using the self-consistent Fock operator from the periodic calculation F^full ^ C and the Fock operator of the cluster subsystem H . The latter is fully C defined by the set of occupied WO’s wκ and the corresponding 1.2 embedding for cluster calculations 21

C reduced density matrix ρ . ME’s of the embedding potential are then expressed as

^HF ^full ^ C  hwα| V wβ =hwα| F wβ − hwα| F ρ wβ wα, wβ ∈C . (1.29)

The one-electron embedding operator V^HF represents the interac- tion between the cluster and the infinite solid at the HF level. But since the cluster and the embedding problems are separated within this approach, it is possible to perform post-HF calculations for the cluster region keeping the embedding frozen. For this purpose ME’s ^HF of V have to be calculated within the basis of all WO’s orbitals as- C C C sociated with , both occupied wκ and virtual wν . Localized vir- tual WO’s are needed to set up the variational space for the local cor- relation treatment. However, conduction bands are usually strongly entangled (have avoided crossings at some points in the Brillouin zone), and their proper localization requires a band-disentanglement procedure [95–97]. A simple way to get the set of localized virtual orbitals is to use the projected atomic orbital (PAO) procedure intro- duced by Pulay [98–101]. Within this formalism the occupied WO’s C wκ are projected out from the atomic basis set associated with the cluster using the Schmidt orthogonalization. The remaining orbitals form, after additional orthonormalization, the required set of virtual C WO’s wν . This HF embedding scheme is exact if it is possible to keep the occupied Wannier orbitals the same in the periodic and cluster cal- culations. In practice, however, clusters are finite and do not include all basis functions associated with the embedding, therefore longer- range tails of WO’s have to be projected out. Such Projection of a WO C wα onto the atomic basis set assigned to the cluster region χ reads

" # X E X X D E E | α C −1 C C 1 30 wαi = ct χt = Stu χu wα χt ,( . ) t t u

C C where wα is the projected WO, χt and χu are basis functions from C −1 χ , and Stu is the corresponding ME of the inverse overlap ma- trix. The ‘approximate’ HF cluster and embedding in terms of wα is p close to the ‘exact’ one (1.29) if the norms of projections hwα|wαi are close to 1. To guarantee this condition for the region of interest Birkenheuer et al. [87–89] proposed to enlarge the cluster region and treat it in a special way. The larger cluster C splits into two parts: C the active region A where the actual correlation treatment will be C performed and the buffer region B. The main role of the buffer is to provide support for the longer-range tails of WO’s centered at C sites in A and to ensure large norms of their projections. Crystal C WO’s assigned to B are also projected onto the cluster basis but their 22 quantum chemistry approach

projection norms are somewhat smaller. However, the effect of this truncation is compensated by keeping the projections of the occupied C WO’s centered in B frozen together with their projected-out coun- terparts incorporated into the embedding potential. Virtual orbitals associated with the buffer should be excluded from the variational space in the subsequent post-HF calculations. Projected functions |wαi as obtained in Equation 1.30 are not or- thogonal and have to be orthonormalized prior to any further treat- ment. Following the idea of the active and buffer regions, it is sensible C to do the orthonormalization such that WO’s in A are modified as little as possible. In the work of Birkenheuer et al. [87–89] the Schmidt C method is first used to orthogonalize WO’s in B with respect to those C in A following by Löwdin orthonormalization within each region.

We will denote the projected orthonormalized WO’s as we α . Projection and normalization of the cluster WO’s change the corre- sponding reduced density matrix and the embedding potential (1.29). It means that by introducing the projected, i.e. truncated, orthonor- mal WO’s we slightly increase the electron density in the cluster region. Again, this effect is small when projection norms are close to 1. In practice, norms of the projected valence-band WO’s centered C within the A are always larger than 0.99 [88–90, 102–105]. Therefore C the effect is indeed small and only related to the buffer B. The ex- pression for the embedding potential in the orthonormalized basis follows Equation 1.29. The only difference is the reduced density ma- trix:

^HF F^full F^ C C we α V we β = we α we β − we α ρe we β we α, we β ∈ . (1.31)

F^full Matrix elements we αP we β can be easily evaluated using the α C expansions we α = t ect χt . For the closed-shell HF case Equa- tion 1.31 can be further simplified, because the reduced density ma- C HF↑↓ trix takes a diagonal form ρekl = 2δk,l. To construct the HF em- bedding potential in this case one only needs the crystal Fock ma- F^full C trix and the set of orthonormalized orbitals we in the atomic C basis associated with the cluster χ . That is exactly what is pro- vided by the crystal-molpro interface program [106]. An extension to open-shell HF problems is straightforward, as the only additional C quantity required is the reduced density matrix ρe in terms of pro- jected orthonormalized WO’s. Cluster-in-solid HF embeddings (in conjunction with high-level cluster calculations) were used for obtain- ing correlated quasiparticle bands in ionic and covalent solids [88– 90, 103, 105] and for the calculation of the local multiplet structure in correlated oxides [102, 104]. Further application of the HF embed- ding to low-dimensional TM oxides is presented in Chapter 2 and Chapter 3 of this thesis. 1.2 embedding for cluster calculations 23

The presented scheme is not the only way to build an embedding that takes into account interactions beyond basic electrostatics [67, 68]. Even if we restrict ourselves to the HF level of theory, we can find other methods alternative to the cluster-in-solid scheme. There are several DFT embedding approaches [68, 86, 107–109] that can in prin- ciple use the HF functional. Several of them are formulated entirely in terms of electron density and have to deal with non-additive kinetic energy term functional either approximately [107, 109], or using addi- tional potential reconstruction [108]. Other employ the orthogonality of the cluster and embedding Kohn-Sham orbitals [86], similar to the cluster-in-solid scheme as described above. Recently an embedding theory that uses the density matrix rather than the electron density to represent the embedding has been proposed [110–112]. The idea behind the density-matrix embedding theory (DMET) is to use the Schmidt decomposition to represent the total WF [113] instead of the generalized product (1.12) used in cluster-in-solid scheme. It allows to represent the interaction of the finite fragment with the environ- ment (even infinite) using just as many additional orbitals as con- tained within the fragment. This ‘doubled’ set of orbitals is half-filled with electrons, that have to be redistributed between the fragment and the additional (bath) orbitals. It means that fractionally-charged fragments, that occur if some bonds are broken during the construc- tion of the embedding problem, can be treated properly. On the other hand, because the transformation to the DMET picture of fragment plus bath orbitals is done via the Schmidt decomposition, all the re- maining orbitals are projected out and the ability to correlate any other virtual or occupied orbital is lost. The applicability of DMET and other embedding schemes mentioned in this paragraph to ab ini- tio calculations for strongly correlated solids is not clear and should be examined in separate studies.

Part II

3 d TRANSITION-METALCOMPOUNDS

d -LEVELELECTRONICSTRUCTUREOF 2 ONE-DIMENSIONALCUPRATES

orrelated electronic structure and material trends in the se- ries of one-dimensional (1D) Cu 3d9 oxides is addressed in C this chapter. Particular attention is paid to fine details con- cerning the d-level relative energies and splittings. The motivation to study the 1D copper compounds is twofold. On one hand, we would like to check the performance of the cluster-in-solid computational scheme described in Chapter 1 for the case of highly anisotropic chain and ladder Cu d9 systems. Quantum chemical calculations on rela- tively small clusters have been earlier performed on two-dimensional (2D) cuprates, e.g. the layered materials La2CuO4 and Sr2CuO2Cl2, and the chain-like 1D system Sr2CuO3 [114]. However, differences as large as 0.5 eV were found between the on-site d–d excitation energies reported for La2CuO4 and Sr2CuO2Cl2 in Ref. [114] and more recent results discussed in Ref. [104]. Those differences seem to be related to the less precise description in Ref. [114] of the adjacent 3d-metal and farther O ions around the CuO4 plaquette at which the d–d excited states are explicitly computed. Similar differences are found as com- pared to earlier QC calculations for the d-level splittings of the chain system Sr2CuO3, which shows that a careful analysis of this issue is indeed well motivated. Secondly, our ab initio data should be helpful for the correct inter- pretation of RIXS and optical spectra in these compounds. For highly anisotropic structures, the degeneracy of both the t2g and eg levels is lifted and the excitation spectra display a very rich structure. Even in the 2D compounds, the sequence of the different excited states cannot be always predicted beforehand. For example, due to the different ra- tios between the in-plane and apical Cu–O bond lengths, the z2 hole state corresponds to the lowest crystal-field excited state in La2CuO4, with a relative energy of 1.4 eV [104, 115], and to the highest crystal- field excitation in HgBa2CuO4, with a relative energy of 2.1 eV [104]. The situation should be even more complex in 1D systems. Reliable ab initio results are therefore desirable for the chain and ladder cuprates. Along this line we undertake a systematic study of three classes of low-dimensional Cu compounds: (i) chains of corner-sharing plaque- ttes/octahedra in Sr2CuO3, SrCuO2, and KCuF3; (ii) systems display- ing edge-sharing plaquettes such as LiVCuO4, CuGeO3, LiCu2O2 and Li2CuO2; (iii) the two-leg ladder compound CaCu2O3.

27 28 d-level electronic structure of one-dimensional cuprates

2.1 Computational details

For the present study we use the cluster-in-solid embedding method- ology introduced in Section 1.2 [106]. In this approach the embed- ding for the correlated cluster calculation is prepared on the basis of prior periodic HF calculations for the infinite crystal. The peri- odic restricted Hartree–Fock (RHF) calculations were performed with the crystal program package [94], while the finite-cluster post-HF treatment is carried out with the molpro quantum chemical software [116]. We applied Gaussian-type atomic basis sets from the crystal library, i.e. basis functions of triple-zeta quality [38] for Cu, O, and F, and basis sets of either double-zeta or triple-zeta quality for metal ions next to the CuO2 chains (e.g. Li, Ca or V). In all computations experimental lattice parameters were used, as reported in Refs. [117– 124]. Post-HF correlation calculations are carried out on finite embed- ded clusters. Each cluster C consists of two distinct regions: an active region CA where the actual correlation treatment is performed and a buffer region CB whose role is to provide support for the longer-range tails of orbitals centred at sites in CA. The active region CA includes one reference Cu site, the nearest neighbour (NN) ligands, and the NN Cu ions. For the studied systems, a given Cu ion may have four, five or six NN ligands, forming L4 plaquettes, L5 pyramids or dis- torted L6 octahedra around a particular Cu site. As concerns the CB region, we include in there each ligand coordination cage around the NN Cu sites and all NN closed-shell metal ions. In CuGeO3, for ex- 4+ 1+ 10 ample, there are 8 Ge NN’s. In LiCu2O2, there are 5 Cu 3d and 8 Li1+ NN’s. The occupied orbitals in the buffer region CB are kept frozen as The orbital basis for obtained from the periodic RHF calculation. In contrast, all valence the correlation orbitals centred at ligand and Cu sites in CA are further reoptimized calculations is a set in multiconfiguration CASSCF calculations. In the latter, the ground- of HF WO’s projected onto the state wavefunction and the crystal-field excited states at the central atomic basis Cu site are computed by state-averaged multiroot optimizations. The functions at sites in Cu d-level splittings are finally obtained from additional MRCI calcu- the cluster region, lations with single and double excitations [11, 38]. The central Cu 3s, see Section 1.2. 3p, 3d, NN ligand 2p, and NN half-filled Cu 3d orbitals are correlated in MRCI. Although the orbitals at the atomic sites of C are derived for each of the compounds by periodic RHF calculations for the Cu 3d9 electron configuration, the embedding potentials are obtained by replacing the Cu2+ 3d9 ions by closed-shell Zn2+ 3d10 species. This is a reasonably good approximation for the farther 3d-metal sites, as the comparison between our results and RIXS data shows [104, 115]. 2.2 chains of corner-sharing CuO4 plaquettes 29

2.2 Chains of corner-sharing CuO4 plaquettes

The 1D compounds Sr2CuO3 and SrCuO2, built of chains of corner- sharing CuO4 plaquettes, display a number of quite unusual proper- ties. First, these materials have the largest NN antiferromagnetic (AF) exchange integrals in the family of Cu d9 oxides. Yet, the interchain couplings are very weak [125–130]. They thus constitute ideal systems for studying the magnetic response of 1D spin-1/2 antiferromagnets. Secondly, high-resolution angle-resolved photoemission and RIXS ex- periments on these compounds have for the first time revealed the realization of spin-charge and spin-orbital separation in 1D electron systems [4, 5, 131, 132]. Results for the d-level splittings of Sr2CuO3 and SrCuO2 are listed in Table 2.1. As concerns the crystallographic details, a major differ- ence between these two compounds is the presence of a network of single CuO2 chains in Sr2CuO3 [117], see Figure 2.1a, whereas SrCuO2 displays a double-chain structure, with CuO4 plaquettes on adjacent chains sharing common edges [118]. The clusters we use in our CASSCF and MRCI calculations include therefore two NN plaquettes in the case of Sr2CuO3 and four adjacent plaquettes for SrCuO2. The coordinate system is chosen such that the x-axis is along the CuO2 chains and z is perpendicular onto the CuO4 plaquettes. In this framework, the ground-state hole orbital has x2 −y2 symmetry (shown in Figure 2.1). The data in Table 2.1 show that the d-level splittings in Sr2CuO3 and SrCuO2 are very similar. In both cases, the relative energy of each particular excited state is not very different from the corresponding value in the Li2CuO2 compound, see Sec- tion 2.4. As in Sr2CuO3 and SrCuO2, there are no apical ligands in Li2CuO2 either. Results for the d-level electronic structure of the lat- ter compound are listed in Table 2.3. Inclusion of single and double excitations on top of the CASSCF wavefunctions brings a nearly uniform upward shift of 0.1 to 0.2 eV for the d–d transitions. By adding Davidson corrections [38], the rel-

z Sr z Sr y Cu y Cu x O x O (a) Sr2CuO3 (b) SrCuO2

Figure 2.1: Chains of CuO4 plaquettes in Sr2CuO3 and SrCuO2. The ground-state hole orbital is dx2−y2 in both cases. 30 d-level electronic structure of one-dimensional cuprates

hole orbital Sr2 CuO3 SrCuO2 casscf/mrci casscf/mrci

x2 − y2 0/0 0/0 xy 1.20/1.50 1.26/1.55 yz 1.83/2.14 1.77/2.03 xz 1.90/2.24 1.88/2.19 z2 2.16/2.55 2.08/2.44

Table 2.1: CASSCF/MRCI d–d excitation energies for corner-sharing chains of CuO4 plaquettes in Sr2CuO3 and SrCuO2 (eV). The MRCI val- ues include Davidson corrections.

ative energies of the excited states further increase by 0.05 to 0.1 eV. The MRCI values presented in Table 2.1 and Table 2.3 include such Davidson correction terms. This is done to account for the lack of quadruple and higher excitations in the MRCI expansion, which leads to size-inconsistency of the method [38]. The relative stabilization of the electronic ground state with respect to the higher lying states in MRCI is mainly related to ligand 2p to metal 3d charge-transfer correlation effects which are more important for the in-plane dx2−y2 hole orbital having σ-type overlap with the NN O 2p functions. As concerns the comparison with other theoretical investigations, there are differences of 0.3 eV or more between the d–d excitation en- ergies listed for Sr2CuO3 in Table 2.1 and the values computed earlier in Ref. [114]. We presume that these large differences, e.g. about 0.5 eV for the yz hole states, are mainly related to the approximations ‡ The virtual orbital used in Ref. [114] for representing the Cu and O ions on NN plaque- space in the MRCI ttes. While here we represent those species at the all-electron level, in calculation cannot the earlier study [114] the NN Cu and O ions were modeled by Mg2+ be presently restricted just to the ions and formal 2− point charges, respectively. Point charges were 114 CA region. It thus also used for the embedding employed in Ref. [ ]. includes virtual We note at this point that in an AF lattice the relative energy of C orbitals in both A a given state within the dn manifold is a sum of a crystal-field con- and CB, which leads to very large MRCI tribution, i.e. an on-site crystal-field splitting, and a magnetic term expansions. To make (see also the discussion in Refs. [104, 133]). As mentioned above, the the computations AF NN spin coupling constant J is remarkably large in Sr2CuO3 and feasible, we restrict SrCuO2, 0.20 to 0.25 eV [125–129]. From the exact Bethe-ansatz solu- our study to FM 1 134 137 alignment of the d tion for the D Heisenberg Hamiltonian [ – ], the AF ground- spins. Even for FM state stabilization energy is J ln 2. On the other hand, from overlap clusters, the MRCI considerations, we conclude that for the crystal-field excited states expansion may the (super)exchange with the NN Cu dx2−y2 spins is either zero or include in some much weaker. For technical reasons‡, the quantum chemical calcu- cases up to ∼1010 Slater determinants. lations were here performed for a ferromagnetic (FM) cluster. For a 2.3 corner-sharing CuF6 octahedra in KCuF3 31 meaningful comparison between the MRCI results and experimen- tal RIXS data, one should therefore subtract a term J ln 2 from the relative RIXS energies, representing the energy stabilization of the AF ground-state with respect to the crystal-field excited states. These considerations are relevant in light of future RIXS measurements on AF 1D cuprates.

2.3 Corner-sharing CuF6 octahedra in KCuF3

KCuF3 is a prototype material for systems with strong coupling among the charge, orbital, and spin degrees of freedom [138–140]. The crys- talline structure of this compound is perovskite-like [124], i.e. three- dimensional (3D). The degeneracy of the Cu eg levels in an ideal perovskite lattice is lifted however through cooperative Jahn-Teller distortions. The latter imply a configuration with alternating, longer and shorter, Cu–F bonds for Cu ions along the y and z axes and orbital ordering in the yz plane. The hole in the Cu 3d shell thus alternately occupies 3dx2−y2 and 3dx2−z2 orbitals. In the yz plane, the magnetic couplings are weak and FM. On the other hand, along the x-axis the NN J is large and AF [141]. Actu- ally this makes KCuF3 a close to ideal 1D Heisenberg antiferromag- net. The predictions for the low-lying spin excitations of the 1D AF Heisenberg chain [136, 137] have been indeed confirmed by inelastic neutron scattering experiments [142]. The magnetic behavior changes from 1D to 3D at the Néel tempera- ture TN =39 K. The emergence of sharp crystal-field absorption peaks in the optical spectra approximately 10 K above TN has been inter- preted as evidence for a symmetry change related with a crossover from dynamic to static displacements of the F ions [143]. In this pic- ture, the orbital and 3D AF ordering are intimately related, in the sense that the former paves the road for the latter. The d–d excitation energies seen in the optical spectra were com- pared in Ref. [143] with the outcome of DFT band-structure calcula- tions. However, strictly speaking, DFT is a ground-state theory. Here, we provide results of excited-state CASSCF and MRCI calculations for the Cu d-level splittings in KCuF3, see Table 2.2. The MRCI treatment brings corrections of 0.1–0.15 eV to the CASSCF splittings, somewhat smaller than for the oxides discussed in Section 2.2. This is related to the more ionic character of the Cu–F bond in KCuF3 and smaller de- gree of Cu 3d and F 2p orbital mixing. As a general trend, the MRCI results tend to slightly underestimate the values corresponding to the maxima of the experimental absorption peaks, by 0.1–0.2 eV. One ob- vious effect of the less anisotropic environment is a much smaller splitting between the two eg components dx2−y2 and dz2 . This partic- ular electronic-structure parameter is in fact the smallest among the Cu d9 compounds investigated here. 32 d-level electronic structure of one-dimensional cuprates

holeorbital casscf mrci experiment*

x2 − y2 0 0 0 z2 0.76 0.85 0.71–1.02 xy 0.89 1.01 1.05–1.15 yz 1.04 1.17 1.21–1.37 xz 1.11 1.25 1.34–1.46

* Optical absorption, Ref. [143]. The two numbers correspond to the onsets and the maxima of the absorption bands, respectively.

Table 2.2: Cu d–d excitation energies in KCuF3 (eV). The ground-state t6 d2 d1 Cu 2g z2 x2−y2 configuration is taken as reference. The Jahn- Teller distortions occur within the yz plane. The MRCI values in- clude Davidson corrections.

2.4 Chains of edge-sharing CuO4 plaquettes

Cuprates in which the CuO4 plaquettes are edge-sharing, are char- acterized by weak superexchange interactions between NN Cu spins, which is due to Cu–O–Cu bond angles that are close to 90◦. Depend- ing on the detailed crystal structure, the NN interaction is usually FM, like in LiVCuO4 and Li2CuO2)[144–146]. It may also be however AF, as it happens in CuGeO3 [147, 148]. The in-chain next-NN exchange is always AF and causes frustration independently of the sign and size of the NN coupling. Further, the longer-range intrachain and in- terchain couplings are sizeable as well in some of these compounds and lead to an intricate competition of magnetic interactions. This has resulted in an appreciable scientific interest in the edge-sharing chain systems. In LiVCuO4 the ratio of the FM NN coupling and the AF next- NN exchange – the frustration parameter – is controversial [144, 145]. For LiCu2O2 even the sign of the NN coupling constant gave rise to debate [149–151]. It has been also argued that a large fourth-nearest- neighbour AF coupling constant is responsible for the incommensu- rate helimagnetic ground state of LiCu2O2 [149]. In CuGeO3 the NN exchange is AF and this system is famous to undergo the spin-Peierls transition [147, 148] to an AF state with a gapped excitation spec- trum. When deriving the values of the shorter- and longer-range ex- change interactions in terms of downfolding techniques within the DFT framework, also the 3d-level excitation energies come into play. The calculated CASSCF and MRCI results for the Cu d-level split- tings in LiVCuO4, CuGeO3, LiCu2O2, and Li2CuO2 are listed in Ta- ble 2.3. The finite embedded clusters on which the correlation treat- ment is carried out include one central and two NN CuO4 plaquettes for the single-chain compounds LiVCuO4, CuGeO3, and Li2CuO2, see Figures 2.2a, 2.2b and 2.2d. For the zigzag double-chain system 2.4 chains of edge-sharing CuO4 plaquettes 33

Li z V z Ge Cu Cu y x y x O O

(a) LiVCuO4 (b) CuGeO3

z Li z Li Cu Cu y x O y x O (c) LiCu2O2 (d) Li2CuO2

Figure 2.2: Chains of edge-sharing CuO4 plaquettes in four studied systems. The ground-state hole orbital dx2−y2 is shown for each com- pound.

LiCu2O2 [122], the cluster C includes four adjacent plaquettes, see Figure 2.2c. The coordinate framework is chosen such that z is per- pendicular to the CuO4 plaquettes, and the x and y axes are along the in-plane Cu–O bonds. As for the corner-sharing plaquettes in Sr2CuO3 and SrCuO2 (see Section 2.2), inclusion of single and double excitations on top of the CASSCF wavefunctions brings a nearly uniform upward shift of 0.1 to 0.2 eV for the d–d transitions. By adding Davidson corrections, the relative energies of the excited states further increase by 0.05 to 0.1 eV. The MRCI values listed in Table 2.3 include the Davidson correction terms. A second effect which deserves attention is the large variations found for the relative energy of the dz2 excited hole state. It is known that the length of the apical Cu–O bond varies widely in Cu d9 oxides. The strong influence of the apical bond length on the dz2 hole excita- tion was previously stressed for 2D cuprates in, e.g. Refs. [104, 115]. Using simple electrostatic arguments, when the negative apical ions are closer to the Cu site, less energy is required for exciting the in- plane Cu 3d hole to the dz2 orbital pointing toward those apical lig- ands. As concerns the chain-like compounds from Table 2.3, there are two apical O ions in CuGeO3, one apical in LiCu2O2, and no apical ligand in Li2CuO2. The relative energy of the dz2 hole state conse- quently increases from 1.71 eV in CuGeO3 to 1.98 in LiCu2O2 and 34 d-level electronic structure of one-dimensional cuprates

hole orbital LiVCuO4 CuGeO3 LiCu2 O2 Li2 CuO2 casscf/mrci

x2 − y2 0/0 0/0 0/0 0/0 xy 1.08/1.28 1.19/1.41 1.13/1.43 1.20/1.52 xz 1.25/1.47 1.42/1.61 1.58/1.88 1.72/2.04 yz 1.29/1.52 1.45/1.64 1.64/1.94 1.72/2.04 z2 0.98/1.18 1.50/1.71 1.67/1.98 1.93/2.30

Table 2.3: Calculated d–d excitation energies for edge-sharing chains of CuO4 plaquettes (eV). The MRCI values include Davidson cor- rections.

to 2.30 eV in Li2CuO2. In LiVCuO4 there are two apical O ions as in CuGeO3 but the apical bond lengths are much shorter as compared 119 121 to CuGeO3, 2.49 versus 2.76 Å (see Refs. [ , ]). The dx2−y2 to dz2 excitation energy in LiVCuO4 is thus substantially lower, 1.18 eV, even lower than in La2CuO4, where the apical bond length is 2.40 Å 104 115 and the transition to the dz2 orbital occurs at about 1.4 eV [ , ]. However, one additional effect which comes into play in LiVCuO4 is the strong covalency between the V ions and both in-plane and apical O neighbours of the Cu sites. Such covalency effects on the VO4 tetra- hedra, see Figure 2.2a, give rise to large deviations from the formal V5+ and O2− valence states employed in a fully ionic picture. When large deviations from the formal value of 2− occur for the effective charges of all adjacent O ions, the strongest affected is actually the ground-state energy because there are four O neighbours to which the lobes of the Cu dx2−y2 orbital are directed. This ‘destabilization’ of the ground-state energy of LiVCuO4 explains the nearly uniform downward shift of the dxy, dyz, and dxz excited hole states, by about 0.15 eV as compared to CuGeO3 (see Table 2.3).

2.5 A ladder cuprate: CaCu2O3

In the so-called ladder compounds, the transition-metal ions are ar- ranged in a planar network of ladders which can display two or more legs. These systems became subject of intense study when Dagotto et al. [152, 153] found theoretical evidence that the isolated S = 1/2 AF two-leg ladder has a finite spin gap, i.e. a finite energy is needed to create a spin excitation. Physical realizations are found in the vanadium oxide compound CaV2O5 [154, 155] and in cuprates such as CaCu2O3, SrCu2O3 [156], and (Sr,Ca)14Cu24O41 [157, 158]. The latter system also displays su- perconductivity [159]. 2.6 conclusions 35

hole orbital casscf mrci

x2 − y2 0 0 xy 1.09 1.32 yz 1.46 1.69 xz 1.61 1.87 z2 1.67 1.96

Table 2.4: CASSCF and MRCI d–d excitation energies for the ladder system CaCu2O3 (eV). The MRCI values include Davidson corrections.

We analyze in this section the d-level electronic structure of CaCu2O3. For most of the two-leg ladder cuprates, including CaCu2O3, octahedra on the same ladder share common corners while adjacent octahedra on different ladders share common edges. One peculiar feature of CaCu2O3 is that on each rung of a ladder the Cu–O–Cu angle is 123◦, much smaller than in other ladder com- pounds [123]. This is the reason the intersite magnetic couplings across the rungs are AF and small, about 10 meV, while the superex- change along the leg of the ladder is AF and one order of magnitude larger, about 135 meV [160]. In other words, from the magnetic point of view the system is rather 1D, with weakly interacting AF CuO2 chains. Choosing the x-axis parallel to the rung and y along the leg of the ladder, the ground-state hole orbital is dx2−y2 . The quantum chemical calculations were carried out on a cluster whose active region CA is defined by one reference CuO4 plaquette and the five adjacent Cu ions (one Cu NN on the same rung, two leg NN’s, and two Cu NN’s on an adjacent ladder). The buffer region CB includes the two apical O ions of the reference Cu site, the nearest four ligands for each adjacent Cu ion, and six Ca neighbours. The apical O ligands are at rather large distance from the central Cu site, 3.0 Å[123]. CASSCF and MRCI results for the Cu d-level splittings in CaCu2O3 are listed in Table 2.4. The MRCI treatment and the Davidson correc- tions bring shifts of 0.2 to 0.3 eV to the CASSCF d–d excitation ener- gies. The excitation energies to the dxz, dyz, and dz2 orbitals are some- what smaller than in Sr2CuO3, SrCuO2, and Li2CuO2 because in the latter compounds there are no apical ligands. At the same time, the splitting between dx2−y2 and dz2 is larger than in the other cuprates we are investigating in this chapter, see Figure 2.3, because the apical Cu–O bond length is the largest in CaCu2O3. 36 d-level electronic structure of one-dimensional cuprates

3.0

2.5 Hole 2.0 orbital d 1.5 xy dxz 1.0 dyz 0.5 dz2 0.0 Sr2CuO3 SrCuO2 KCuF3 LiVCuO4 CuGeO3 LiCu2O2 Li2CuO2 CaCu2O3

Figure 2.3: MRCI crystal-field splittings in 1D Cu d9 compounds. Ground state configurations with a hole in the dx2−y2 orbital are taken as reference.

2.6 Conclusions

It has been shown in this chapter that wavefunction-based electronic- structure calculations are well-suited for computing local charge ex- citations as observed in transition-metal oxides by optical and RIXS measurements. While this was demonstrated previously for a num- ber of layered cuprates [104], we have extended those applications to 1D and ladder compounds. Results and trends for the d–d exci- tations in low-dimensional cuprates are summarized in Figure 2.3. The ab initio data compare well with the results of optical absorption measurements on KCuF3 [143]. Future RIXS experiments on other 1D cuprates included in Figure 2.3 will constitute a direct test for the symmetries and energies of the different d–d excitations that we predict here. From a more general perspective, the results reported here clearly indicate the significant potential of the ab initio wavefunction-based methods in the context of correlated electronic materials and the need to further develop such techniques in parallel with approaches based on density functional theory. The latter has revolutionized the field of electronic-structure calculations but its limits are also clear. Computa- tions of excited states of strongly correlated electrons are in a way the worst case for DFT. One is then forced to avail oneself to alternatives, i.e. wavefunction-based methods. An attractive feature is that strong correlations can be handled by CASSCF calculations even when the full system is infinite. In all approximations which are being made, the accuracy of the wavefunction-based methods can be progressively increased. Here we supplemented the CASSCF calculations by MRCI, yet, alternative supplementary methods are possible. d – d EXCITATIONENERGIESIN 3 LOW-DIMENSIONAL Ti AND V OXYCHLORIDES

n this chapter we study the local d-d excitation energies for two strongly correlated Mott insulators, the low-dimensional oxy- chlorides TiOCl and VOCl, using a quantum chemical cluster- I 1 in-solid computational scheme. The Ti d single-electron configura- tion is to a great extent complimentary to the single-hole Cu d9 case studied in the Chapter 2, having in mind electron-hole symmetry. TiOCl harbours quasi-one-dimensional spin chains made of S = 1/2 Ti3+ ions, while the electronic structure of VOCl displays a more two-dimensional character. The d-level electronic structure of TiOCl has been previously addressed by optical absorption experiments corroborated with model-Hamiltonian configuration-interaction sim- ulations of the optical spectra [161, 162], DFT-based investigations [163–167], ab initio quantum chemical cluster calculations [168], and RIXS measurements [169]. The V oxychloride compound, on the other hand, is much less investigated. On the experimental side, basic in- formation concerning its d-level electronic structure is available from optical absorption data [170]. We here perform a detailed quantum chemical investigation to better understand the nature and symmetry of the different excited states identified by optical experiments [170].

3.1 Structural and computational details

TiOCl and VOCl display the same highly anisotropic crystalline struc- ture, with a bilayer network of metal and oxygen sites well separated along the crystallographic c axis by double Cl layers (see Figure 3.1). Each metal ion lies in the center of a distorted O4Cl2 octahedron. Ad- jacent octahedra share edges along the b axis and corners along a. The unit cell is orthorhombic, with space group Pmmn (number 59) [171, 172]. Our computational investigation implies a sequence of several steps. Periodic restricted HF calculations are first performed with the crystal package [94]. We used the experimental lattice parameters and all-electron Gaussian-type basis sets [38]. Further post-HF corre- lation calculations are carried out on finite embedded clusters. The same type of cluster is employed for both TiOCl and VOCl. It con- sists of nine MO4Cl2 octahedra, i.e. one reference octahedron plus eight neighbouring octahedra, where M is either Ti or V. The ref-

37 38 d–d excitation energies in Ti and V oxychlorides

z Ti/V O y x Cl

Figure 3.1: Layered crystal structure of TiOCl and VOCl. A distorted O4Cl2 octahedron is highlighted on the right hand side. The coordinate framework is chosen such that the y and z axes are not along the metal–ligand bonds, see text.

erence octahedron in the center of this cluster C defines the active region CA where local d-d excitations are explicitly computed. The adjacent eight octahedra constitute a buffer region CB. The role of the sites in CB is to provide support for the tails of WO’s in the CA region (see Chapter 1 for more details). With the help of the cluster-in-solid scheme, the HF data is used to generate an effective embedding potential for the nine-octahedra fragment C. This potential is obtained from the Fock operator in the restricted HF calculation as described in Section 1.2 and models the surroundings of the finite cluster, i.e. the remaining of the crystalline lattice. The orbital basis for the correlation calculations is a set of pro- jected HF Wannier functions. Localized HF WO’s are obtained with the Wannier-Boys localization module [91] of the crystal package and subsequently projected onto the set of Gaussian basis functions associated with the atomic sites of the cluster C. For our choice of CB, the norms of the projected WO’s centered within the active region CA are not lower than 99.8% of the original HF WO’s. All valence orbitals centered at ligand and metal sites in CA (in- cluding their tails in CB) are further reoptimized in multiconfigura- tion CASSCF calculations [38]. In the latter, the ground-state wave- function and the crystal-field excited states at the central Ti/V site are computed in state-averaged multiroot calculations. The d-level splittings are finally obtained by additional single and double MRCI calculations [38]. The CASSCF and MRCI investigations are carried out with the molpro package [116]. The effective embedding poten- tial is added to the one-electron Hamiltonian in the CASSCF/MRCI computations with the help of the crystal-molpro interface [106]. We used O and Cl basis sets of triple-zeta quality as designed by Towler et al. [173] and Prencipe et al. [174]. For the Ti3+ and V3+ species, we employed the triple-zeta s and p basis sets of Mackrodt et al. [175] and Dovesi et al. [176], respectively, plus triple-zeta d basis functions as designed by Ahlrichs et al. [177]. 3.2 quasi-one-dimensional titanium oxychloride 39

state casscf* /caspt2* casscf/mrci rixs† optics‡

1 2 2 t2g, y −z 0/0 0/0 0 0 1 t2g, xy 0.29/0.29 0.36/0.34 0.32 — 1 t2g, xz 0.66/0.68 0.65/0.63 0.55 0.65 1 eg, yz 1.59/1.68 1.51/1.55 1.41 1.50 1 2 eg, x 2.30/2.37 2.35/2.36 2.01 — * † Ref. [168]; Ref. [169] data with Jb ln 2 = 0.04 subtracted from each of the RIXS values, see text; ‡ Ref. [162].

Table 3.1: Ti3+ d-d excitation energies in TiOCl, quantum chemical vs. exper- imental results. A non-standard local reference system is chosen for the octahedrally coordinated Ti ion, see Figure 3.1, with a ro- 45◦ x d1 tation of around . The ground-state y2−z2 configuration is taken as reference. All energies are given in eV.

3.2 Quasi-one-dimensional titanium oxychloride

The magnetic properties of TiOCl have been the topic of extensive investigations over the last decade. The largest intersite interactions are AF, occur between NN Ti3+ S=1/2 sites along the b axis, and give rise to AF spin chains along that direction. The interchain couplings are also sizable and responsible for geometrical frustration. The doubling of the unit cell in the low-temperature nonmagnetic phase [178, 179] suggests a spin-Peierls dimerized ground state at low temperatures. The observed deviations from conventional spin- Peierls behavior, e.g. the existence of two distinct phase transitions at 91 and 67 K, were attributed to the above mentioned frustration of the interchain interactions [161]. Despite the actual As concerns the electronic structure of TiOCl, DFT calculations orthorhombic C2v within the LDA indicate sizable occupation numbers for each of the symmetry at the 163 166 Ti/V position, we Ti t2g levels [ – ]. Early interpretations of the unusual physical use the more properties of this material thus invoked the presence of strong orbital common notations of fluctuations [163, 164, 180]. Inclusion of local Mott-Hubbard-type cor- the cubic Oh group. relations within DMFT and LDA+DMFT calculations effectively pulls one of the t2g levels below the other two components and reduces the occupation of the latter to nearly zero [165, 166]. On the theoretical side, this is confirmed by earlier quantum chemical cluster calcula- tions [168] as well by our present results. In Table 3.1, CASSCF and MRCI results for TiOCl are listed. The active orbital set contains all five 3d functions at the central Ti3+ site and the singly occupied 3dy2−z2 orbitals at the eight Ti NN sites. We considered high-spin intersite couplings, i.e. a FM configuration of Ti d spins. Although the WO’s at the atomic sites of C are derived from periodic restricted HF calculations for the Ti3+ 3d1 electron config- uration, the embedding potential is obtained by replacing the Ti3+ 40 d–d excitation energies in Ti and V oxychlorides

y2-z2 (t ) xy (t ) xz (t ) Ti 2g 2g 2g O Cl

2 yz (eg) x (eg) z

y x

Figure 3.2: Plots of the singly occupied Ti 3d orbitals for the ground (left) and each of the crystal-field excited states in TiOCl, as obtained in the five-root CASSCF calculation. All sites of the cluster C are shown on the left side of the figure.

3d1 ions by closed-shell V3+ 3d2 species. This is a reasonably good approximation for the farther 3d-metal sites, as the comparison be- tween our results and RIXS data shows. In addition to the active Ti 3d functions in CA and CB, all O 2p and Cl 3p orbitals at the central octahedron are included in the MRCI calculations. Ti 3d orbitals as obtained for each Ti 3d1 electron configuration in the state-averaged CASSCF calculation are plotted in Figure 3.2. The ‘tails’ at the NN ligand sites are also visible in the plots. The asymmetric shape of the 3d functions reflects the distorsion of the ligand octahedron and the presence of two distinct ligand species. As expected for a d1 system, the MRCI treatment brings minor corrections to the d-d CASSCF excitation energies. Our results are compared in Table 3.1 with earlier CASSCF and CASPT2 data from Ref. [168]. A major difference between the present and earlier quan- tum chemical calculations is the size of the embedded clusters, i.e nine octahedra here and one octahedron in Ref. [168]. In contrast 9 to other systems, e.g. layered Cu d oxides such as La2CuO4 and Sr2CuO2Cl2 where differences of about 0.5 eV were found as func- tion of the cluster size (see the discussion in Ref. [104]), the agreement between the two sets of results is good for TiOCl. Good agreement is also found with RIXS and optical absorption experimental data, with deviations not larger than 0.1 eV. The only exception is the dy2−z2 to dx2 excitation, which is predicted to oc- cur at 2.36 eV in the quantum chemical calculations and observed at 2.01 eV in RIXS, see Table 3.1. In the optical absorption measurements 161 162 [ , ], this particular transition as well as the dy2−z2 to dxy exci- tation were not identified, which is related to the different selection rules for RIXS and optical absorption. As concerns the comparison with other theoretical approaches, model-Hamiltonian cluster CI calculations [162] put the lowest d-d excitation at 0.25 eV, somewhat lower than our data and the RIXS 3.3 quasi-two-dimensional vanadium oxychloride 41 results listed in Table 3.1. A value of about 0.25 eV has also been pre- dicted for the splitting between the dy2−z2 and dxy levels by DFT cal- culations [167]. These DFT results understimates somewhat the size One has to keep in 167 169 mind, however, that of the dy2−z2 -dxz splitting, 0.46 [ ] versus 0.55 eV in RIXS [ ], while the numbers obtained with model-Hamiltonian CI underesti- strictly speaking, in 162 its initial formula- mate the dy2−z2 -dyz splitting, 1.24 in Ref. [ ] versus 1.41 eV in tion, DFT is only a RIXS [169]. Each of those theoretical approaches provides very good ground-state theory. estimates for the relative energy of the dx2 level. The DFT calcula- tions in [167], nevertheless, predict a metallic ground state for TiOCl. An onsite Coulomb repulsion U of about 2 eV can turn the metal into an insulator [167]. However, the dependence of the d-d splittings on U is not discussed in Ref. [167]. The model-Hamiltonian CI calcu- lations [162] are obviously even more sensitive to the choice of the input parameters because not only the Coulomb repulsion U but also the Slater integrals, p-d hoppings, and p-d charge-transfer energies enter as adjustable quantities in that methodology. In contrast, all in- teractions are computed from scratch, fully ab initio, in the quantum chemical scheme. It is known, that the NN spin interactions along Ti chains (parallel to the b axis) in TiOCl are quite large and AF, Jb ≈ 57 meV [163]. But due to technical reasons it is only possible to perform the quan- tum chemical calculations with FM-aligned NN Ti spins. For a mean- ingful comparison between the MRCI and RIXS data, we subtracted in Table 3.1 from the relative RIXS energies reported in Ref. [169] a term Jb ln 2 ≈ 0.04 eV representing the energy stabilization of the AF ground state with respect to the crystal-field excited states (see the discussion in the last paragraph of Section 2.2).

3.3 Quasi-two-dimensional vanadium oxychloride

In the family of correlated transition-metal oxide materials, the vana- dium compounds display an impressive variety of vanadium valence states, crystalline structures, and physical properties. The valence con- 2+ 3 figuration, for example, may vary from V d in VO1−x [181, 182] 4+ 1 to V d in CaV2O5 [183] and NaV2O5 [184]. While the latter vana- dates display a layered lattice with ladders of O-bridged V sites, VO1−x has a rocksalt crystalline structure. At the level of ab initio wave-function calculations, the V oxides are rather unexplored. To our knowledge, only the ladder vanadates have been investigated by advanced quantum chemical calculations [183–185]. Besides having two electrons instead of one within the t2g set of or- bitals, VOCl also displays in comparison to TiOCl a more pronounced two-dimensional character as concerns the electronic structure and magnetic properties [186]. In Table 3.2,V d-d excitation energies for the VOCl compound are given. Two different sets of CASSCF calcu- lations were carried out, for two different electron configurations at 42 d–d excitation energies in Ti and V oxychlorides

dominant config. casscf* casscf† mrci† optics‡

2 1 1 t2g, S=1 dy2−z2 dxy 0 0 0 0 1 1 dy2−z2 dxz 0.36 0.36 0.34 0.3 1 1 dxydxz 0.43 0.44 0.46 0.4

1 1 1 1 t2geg, S=1 dxydyz 1.61 1.60 1.73 1.5–1.9 1 1 dy2−z2 dx2 1.74 1.71 1.82 1.5–1.9 1 1 dxzdyz 1.75 1.73 1.88 1.5–1.9 1 1 dy2−z2 dyz 2.85 2.82 2.55 — 1 1 dxzdx2 3.53 3.52 3.17 — 1 1 dxydx2 3.53 3.53 3.19 —

2 2 t2g, S=0 dy2−z2 1.27 1.47 1.1–1.4 1 1 dy2−z2 dxy 1.28 1.47 1.1–1.4 1 1 dy2−z2 dxz 1.56 1.75 1.5–1.9 2 dxz 1.79 1.97 1.5–1.9 1 1 dxydxz 1.81 2.00 1.5–1.9 * 3+ 1 1 † 3+ 2 High-spin V dy2−z2 dxy NN, see text; Closed-shell V dy2−z2 NN; ‡ Ref. [170].

Table 3.2:V3+ d-d excitation energies in VOCl. A non-standard local refer- ence system is chosen, see Figure 3.1, with a 45◦ rotation around x. d1 d1 The dominant GS configuration is y2−z2 xy with a weight of 81 %. For each of the excited states, the weight of the dominant configuration is not smaller than 64 %. All energies are given in eV.

each of the V NN’s. Preliminary CASSCF calculations were first per- formed with eighteen orbitals in the active space, i.e. two t2g orbitals at each of the V sites in the nine-octahedra cluster. As for TiOCl, we considered high-spin intersite couplings, i.e. a FM cluster. It turns out that the dominant ground-state configuration at a given V site is 1 1 3+ dy2−z2 dxy. The crystal-field excited states for the central V ion are then obtained in state-averaged multiroot computations [38] where a group of three additional orbitals was added to the active space. To be sure that the three additional orbitals are localized at the central V site and to allow only for on-site d-d excitations, the occupation of the NN dy2−z2 and dxy functions was restricted to 1. Crystal-field splittings obtained from such SCF calculations are listed in the third column of Table 3.2. To avoid complications related to states that display low- spin intersite couplings, only the high-spin on-site excitations were computed in this case. 3.4 conclusions 43

We note that the ground-state electron configuration in the CASSCF 1 1 calculation is dy2−z2 dxy but in the construction of the HF embedding 3+ 2 potential we imposed a closed-shell V dy2−z2 configuration for the remaining part of the lattice. In a second set of CASSCF calculations, 2 we imposed a closed-shell dy2−z2 electron configuration for the V NN’s in the CB region of the cluster as well. Only tiny variations were found as compared to the CASSCF d-d excitation energies with high- 1 1 spin dy2−z2 dxy NN’s, i.e. less than 0.01 eV for the lowest three roots and less than 0.03 eV for the higher roots. The CASSCF V crystal-field splittings for closed-shell V NN’s are given in the fourth column of Table 3.2. The latter CASSCF wavefunctions were used as reference wavefunctions in further MRCI calculations. Those MRCI results are listed in the fifth column of Table 3.2. Only the V 3d,O 2p, and Cl 3p orbitals at the central octahedron were correlated in the MRCI treatment. The agreement between our quantum chemical data and optical absorption results reported in Ref. [170] is very good, especially for the lowest two excitations (see Table 3.2). The small difference be- tween those excitation energies indicates that the splitting between the dy2−z2 and dxy levels, singly occupied in the ground-state con- figuration, is also small, about 0.1 eV. The DFT calculations predict a value of 0.02–0.03 eV for that splitting [167]. On the other hand, the energy separation between the dxy and dxz levels is substantially larger, about 0.3 eV (see Table 3.2), and indicates that orbital fluctua- tions are unlikely to be important in VOCl. A splitting of about 0.3 eV between the dxy and dxz levels has also been predicted by ear- 167 2 1 1 lier DFT calculations [ ]. For the t2g → t2geg transitions, we find that each excited state acquires sizeable multiconfigurational charac- ter. The CASSCF/MRCI results are therefore not directly comparable to a one-electron description as that provided by DFT [167]. As con- cerns the S = 0 singlet excitations, they acquire very low intensity in the optical spectra. On the experimental side, a better characterization of those states call for high-resolution RIXS measurements.

3.4 Conclusions

With a quantum chemical cluster-in-solid computational scheme, we have determined the d-d excitation energies of TiOCl and VOCl. The 1 2 ground-state electron configuration is t2g for Ti and high-spin t2g for V in these compounds. The lowest-energy d-d excitations are for both materials within the t2g subshell, starting at 0.34 eV. We conclude that therefore orbital degeneracies are lifted to a similar and significant extent in the two systems, which excludes the presence of strong or- bital fluctuations in the ground state. In vanadium oxychloride, spin triplet to singlet excitations start at an excitation energy of 1.47 eV. The computed d-level electronic structure and the symmetries of the 44 d–d excitation energies in Ti and V oxychlorides

wavefunctions are in very good agreement with RIXS results and op- tical absorption data for TiOCl. For VOCl, future RIXS experiments will constitute a direct and stringent test of the symmetry and energy of about a dozen of different d-d excitations that we predict here. Our computational scheme and present results indicate a promising route for the ab initio modeling of RIXS spectra in 3d-metal compounds. COMPUTATIONOFX-RAY d -ION 4 SPECTRAINSOLIDS

ome of the most interesting properties of a molecule or a solid concern the response to an external electromagnetic field. This S response also provides clues on the actual electronic structure of the system and is therefore a main aspect that is analyzed in ex- perimental labs. For high-energy incoming photons as used in X-ray absorption or photoemission measurements, a challenging task on the theoretical and computational side is a reliable description of the changes that occur in the electronic environment after creating a lo- calized electron vacancy, be it in a core or valence-shell level. Such changes are also referred to as charge relaxation, orbital breathing, or screening effects and can be classified into intra-atomic and extra- atomic contributions [11, 187]. In this chapter we outline an ab initio wave-function-based methodology that allows to explicitly describe the readjustment of the charge distribution in the ‘vicinity’ of an ex- cited electron by performing separate SCF optimizations for the dif- ferent electron configurations. While this idea of using individually optimized wave functions has been earlier employed for the interpre- tation of X-ray absorption (XA)[188, 189] and resonant inelastic X-ray scattering (RIXS)[190, 191] spectra of small organic molecules [188] and of Fe-based complexes in solution [189–191], we here apply it to the calculation of XA and RIXS excitations and cross sections of a transition-metal ion in a solid-state matrix and compare the com- putational results to fresh experimental data. Li2CuO2, a strongly correlated d-electron system already studied in Chapter 2, is cho- sen as a test case. We find large orbital relaxation effects of & 10 eV for the Cu core-hole excited states. In the RIXS spectra computed us- ing many-body wave functions, all trends found experimentally for the incoming-photon incident-angle and polarization dependence are faithfully reproduced. This also allows us to determine RIXS intensi- ties on an absolute scale, from which we compute a resonant scatter- ing enhancement of the order of 105.

4.1 Resonant inelastic X-ray scattering

RIXS is a powerful and fast-developing technique to measure a di- versity of different elementary excitations in correlated electron sys- tems [133, 187], for example, dispersive magnetic modes [192], or- bital excitations [5, 115, 193] and phonons [194, 195]. As the interac-

45 46 computation of x-ray d-ion spectra in solids

tion between X-ray photons and matter is relatively strong — much stronger than for instance the interaction of neutrons with matter — these modes can be measured on very small sample volumes, for example, cuprate nanostructures that are only a few unit cells thick [196, 197]. Also, the scattered X-ray photon leaves the material under study behind with a low-energy elementary excitation that may be directly compared to the elementary response of other inelastic scat- tering techniques. One complication in RIXS is however posed by the resonant character of the technique, which implies the presence of a resonant intermediate state. The intermediate state is transient but strongly perturbed with respect to the ground state due to the absorp- tion of a high-energy X-ray photon which results in the creation of a hole in the electronic core. The difficulties in the calculation and in- terpretation of RIXS spectra arise from the fact that the intermediary electron configuration determines all the transition probabilities from the initial to the final state and as such affects the scattering cross sections of the different excitations. Attempts to compute full RIXS spectra for correlated, d-electron open-shell systems have been made with both model-Hamiltonian ap- proaches [133, 187, 198–205] and ab initio wave-function-based quan- tum chemistry methods [190, 191, 206–208], to address either the d- shell multiplet structure [190, 191, 203–207] or Mott-Hubbard physics [198–202]. Yet, in the solid-state context, one important aspect that is either missing or described just phenomenologically in earlier com- putational work is the valence-shell charge relaxation in response to the creation of a core hole in the intermediate RIXS state. To treat such effects in a reliable way, we choose to carry out independent SCF optimizations for the many-body wave functions describing the reference dn and core-hole 2p5dn+1 configurations. This is achieved through MCSCF calculations [38] on relatively large but nevertheless finite clusters with appropriate solid-state embedding. The individ- ual MCSCF optimizations obviously lead to sets of nonorthogonal or- bitals. ME’s between wave functions expressed in terms of nonorthog- onal orbitals can be however computed with dedicated quantum chemistry algorithms [188, 209–211]. Not only is our approach un- biased this way with regard to strong charge readjustment effects, it also allows to calculate absolute RIXS cross sections — core informa- tion for any scattering technique. Our results for the latter can directly be tested experimentally.

4.2 Electron – field interaction

For an electromagnetic field interacting with electrons, assuming a system with fixed nuclei, the total Hamiltonian can be written as:

H H H H 0 H H 0 4 1 = field + el + = 0 + . ( . ) 4.2 electron – field interaction 47

H H Here field describes the quantized electromagnetic field, el is the electronic part, and H 0 stands for the interaction between them. We are interested in the latter term, which can be written quite generally [212] as:

2 X X e 2 e H 0 = A~ (~r ) − A~ (~r ) · ~p 2m c2 j m c j j e j e j X 2 X eh h i eh e h~ i − ~σ · ∇~ × A~ (~r ) − ~σ · A˙ (~r ) × A~ (~r ) , m c j j 2(m c)2 c2 j j j e j e j (4.2) where A~ is the vector potential, ~rj, ~pj, and ~σj are the position, mo- mentum, and spin operators acting on an electron, while e, me, c, h are fundamental constants. The vector potential can be expanded in plane waves:

X s 2 2πhc  ~ ~  A~ (~r ) = ~ a eik·~rj + ~∗ a† e−ik·~rj ,(4.3) j V ~kε ~kε ~kε ~kε ω~k ~kε V a a† here is the system phase volume, ~kε and ~kε are annihilation ~ and creation operators of a photon with wave vector k, energy hω~k, and polarization ε = {σ, π} with the corresponding unit polarization vector ~~kε. The dominant contribution to resonant absorption and scattering 0 processes is given by the optical transition part of H ( A~ (~rj) ·~pj terms) [212] and we can therefore continue by keeping only the second term in (4.2). Using expansion (4.3) we obtain X s e 2πh  ~ ~  H 0 = ~ a eik·~rj + ~∗ a† e−ik·~rj .(4.4) op V ~kε ~kε ~kε ~kε me ω~k j,~kε

The ME’s for the absorption of one photon characterized by ~k and

ε, which would decrease the number of such photons n~kε by one and excite the electronic system from state |ai to state |bi, have the following form: H 0 b; n~kε − 1 op a; n~kε = s XN 2πh e  i~k·~rj = b; n~ − 1 ~~ · ~pj a~ e a; n~ m ω V kε kε kε kε j=1 e ~k s XN 2πh e  i~k·~rj = hb| ~~ · ~pj e |ai · n~ − 1 a~ n~ m ω V kε kε kε kε (4.5) j=1 e ~k XN r2πhω = i ~k hb| ~ ·~r  |ai · h0| a |1i e V ~kε j ~kε j=1 r 2πhω   = i ~k hb| ~ · ~R |ai · 1 e V ~kε . 48 computation of x-ray d-ion spectra in solids

i~k·~r Here we used the electric dipoleP approximation e j ≈1, introduced ~ the total position operator R = j ~rj, and assumed that the radiation is incoherent, i.e. every photon can be dealt with individually, with

the number of photons in a mode n~kε =1.

4.3 X-ray absorption cross section

It is instructive to first consider the XA, the first step of the scattering process. Since our main concern is RIXS, we will consider here the In X-ray spectro- case of resonance, when an electron is excited from a core level but scopy the absorption does not leave the solid. resonance is also Not every photon passing through the solid is absorbed, and one referred to as an absorption edge. At has to analyze the transition rate of the process: the resonance transition probability 2π 0 2 photon energy one W = = hB| H |Ai ρ 4 6 B. ( . ) sees a sharp step in unit time h intensity on the XA The expression on the right hand side can be derived using time- plot. dependent perturbation theory [213, 214]. The initial |Ai and final |Bi states are eigenfunctions of the joint electron-plus-field Hamiltonian (4.1), with the total energy conserved (EA−EB =0), and ρB = dN/dE being the density of states to which transitions can occur. Equation 4.6 is general, and we can use it for a single photon ab- sorption. When a photon ~kε is absorbed, the electronic subsystem is excited from the ground state |gsi to a core-hole state |chi with  the density of states defined by the delta function δ Egs + hω~k − Ech . Equation (4.6) is Taking this into account together with the relation for the ME (4.5) commonly known as we obtain Fermi’s second 2π golden rule, but in | H 0 |gs 2  W~ = hch op i × δ Egs + hω~ − Ech fact it was first kε h k (4.7) obtained by Dirac 2 2 4π αchω~k  = h |~ ·~R |gsi × δ Egs + hω − E [215] in the frame of V ch ~kε ~k ch , time-dependent perturbation theory, where α = e2/(hc ) ≈ 1/137 is the fine structure constant. Equation 4.7 see [216]. still contains the normalization volume V introduced to represent the electromagnetic field. This is quite natural, because the transition rate should be proportional to the intensity of the incoming light. The latter can be represented by the incident flux # of photons c Φ0 = = . (4.8) unit time · area V From the ratio between the transition rate (4.7) and the incident The cross section σ flux (4.8) we obtain the expression for the absorption cross section: has the dimension of 2 area and is conventi- 2 | ~ |gs  4 9 σ~ = 4π αhω~ hch ~~ ·R i × δ Egs + hω~ − Ech .( . ) onally measured in kε k kε k barn = 10−24 cm2. One can see that σ~kε depends on energy, polarization, and the prop- agation direction of the incoming X-rays, as well as on the nature of the electronic ground and excited states. 4.4 rixs double differential cross section 49

The excited state |chi does not exist infinitely long, but tends to decay via various channels. Apart from direct X-ray emission, other possible decay channels are the fluorescent X-ray emission when an electron from the outer shell fills the core-hole level creating an out- going photon, and the two-electron Auger process with one electron from a filled level goes down to the core and a second electron emit- ted to the continuum. The presence of all these processes leads to a finite lifetime broadening of the core-hole state, which is equivalent to introducing a damping force in time-dependent perturbation theory [217, 218]. As a result, we need to substitute the delta function in (4.9) by the normalized Cauchy-Lorentz distribution [218]. The full width Γ = hW tot at half maximum (FWHM) of the distribution ch |chi is related to the total transition probability per unit time from the |chi state and can be estimated from experiment [219, 220] . Additionally, the XA cross section (directly comparable with measured data) should involve summation over all possible initial and final states: X 1 X 2 σXA = 4π2α hω hch|~ ·~R |gsi ~k,ε ~k ~kε gs ggs ch (4.10) Γch/2π × 1 , gs 2 2 (E +hω~k−Ech) + 4 Γch where ggs indicates the degeneracy of the ground state.

4.4 RIXS double differential cross section

The dominant part of the electron–radiation interaction Hamiltonian H 0 4 4 ( op in Equation . ) includes only terms that create or annihilate a single photon. Scattering, however, is a photon-in–photon-out pro- cess. Thus to obtain an expression for the scattering transition rate, one has to use at least the second-order order in the expansion: 2 X hF| H 0 |Ni hN| H 0 |Ii 2π | H 0 | op op 4 11 W = hF op Ii + ρF.( . ) h EI − EN N | gs | | | Where Ii = ; 1~kε , Ni = ch; 0i, and Fi = fs; 1~k0ε0 are initial, in- hω 0ρdω0dΩ0 termediate, and final states, respectively; ρF is the density of final 6 @I  states. The first-order term vanishes within the dipole approximation, hω P@}- 4 4 © @@R as can be seen from the form of Equation . , and will be omitted in ? the following. The scattering process involves absorption of a single photon with The scattering known energy, momentum and polarization {hω ,~k, ε} and emission process includes { 0 ~ 0 0} absorption of a of another photon characterized by the parameters hω , k , ε . The single photon with density of states that enters (4.11) in this case can be written as known {hω ,~k, ε} and emission of 0 0 0 ρF =δ Egs + hω − Efs − hω ρhω 0 dΩ dhω another photon with V 02 4 12 some other 0 ω 0 0 ( . ) =δ Egs + hω − E − hω dΩ dhω {hω 0, k~0, ε0} fs 3 3 , in any (2π) hc direction. 50 computation of x-ray d-ion spectra in solids

0 where ρhω 0 is the density of photons with energy hω . For brevity 0 we introduce the notation hω = hω~kε, ~ = ~~kε and hω = hω~k0ε0 , 0 ~ = ~k~0ε0 for the energies and polarization vectors of the incoming and scattered photons, respectively. To write down the expression for the RIXS double differential cross section, we plug in the ME (4.5) and the density of states (4.12) into the expression for the transition rate (4.11) and further divide it by the incident flux (4.8) and dΩ0dhω 0. This way the common form of the Kramers-Heisenberg dispersion formula [213, 221] is obtained:

2 X 2 d σ~ α2h 2 hfs|~0·~R |chi hch|~·~R |gsi k,ε = ωω03 0 0 2 dΩ dhω c Egs + hω − Ech (4.13) ch 0 ×δ Egs + hω − Efs − hω . We should further introduce finite lifetimes for the excited core-hole and final states, characterized by natural breadths Γch and Γfs, respec- tively. The summation should be done over all available final states and the possible degeneracy of the ground state. We then reach the final formula for the RIXS double differential cross section that we are going to use for the calculations:

2 RIXS d σ~ α2h 2 k,ε = ωω03 dΩ0dhω 0 c2 X X X 2 1 hfs|~0·~R |chi hch|~·~R |gsi × 4 14 ( . ) ggs Egs + hω − Ech + iΓch/2 gs fs ch Γfs/2π × 2 . ( gs 0) 2 E + hω − Efs − hω +Γfs/4

4.5 Computational scheme for Cu L-edge excitations

Li2CuO2 is being extensively investigated in the context of low- dimensional quantum magnetism. We here address the Cu L-edge (2p to 3d) excitations of this compound. Quantum chemistry calcu- lations were carried out on a [Cu3O8Li16] cluster consisting of one reference CuO4 plaquette for which Cu 2p–3d and 3d–3d excitations are explicitly computed, two other NN CuO4 plaquettes (see Fig- ure 4.1) and 16 adjacent Li ions. The surrounding solid-state matrix was modeled by an array of point charges optimized to reproduce the ionic Madelung field [73]. To reduce the computational effort and the complexity of the subsequent analysis, the NN Cu2+ S = 1/2 3d9 ions were represented by closed-shell Zn2+ 3d10 species. We used all- electron relativistic basis sets in conjunction with the second-order Douglas-Kroll-Hess scalar relativistic operator [222, 223]. For the cen- tral Cu site, we utilized the s and p functions of the cc-pwCVQZ-DK basis set together with d and f functions from the cc-pVQZ-DK com- pilation [224]. For O ligands at the central CuO4 plaquette, the s and 4.5 computational scheme for Cu L-edge excitations 51 p functions of the cc-pVQZ-DK basis set were used, along with polar- ization d functions of the cc-pVTZ-DK kit [225]. For other ions in the neighbourhood of the central CuO4 plaquette, valence double-zeta basis functions of cc-pVDZ-DK type were employed (spd functions for Cu and sp functions for O and Li) [224–226]. All computations were performed with the molpro quantum chemistry package [116]. Many-body wave functions, eigenvalues and dipole transition ME’s needed for evaluation of Equation 4.14 were computed at both the MCSCF and MRCI levels of theory [38, 227]. The MCSCF calcula- tions were carried out by keeping ‘frozen’ all orbitals centered at O and metal sites beyond the reference CuO4 plaquette as in the ini- tial Hartree-Fock computation. These orbitals were included in the cluster for better description of the central plaquette, see discussion in Section 1.2. To separate the transition-metal 3d and O 2p valence orbitals into different groups, i.e. central-plaquette and adjacent-ion orbitals, we used the Pipek-Mezey localization module [228] available in molpro. The active orbital space includes all five Cu 3d orbitals for the Cu 3d9 states and five Cu 3d orbitals plus three Cu 2p functions for the intermediate 2p53d10 states. In order to prevent for the latter MCSCF states 2p53d10 → 2p63d9 ‘de-excitation’, the orbital optimiza- tion was carried out in two steps in the MCSCF calculations, follow- ing a freeze-and-thaw orbital relaxation (FTOR) scheme : (i) optimize first the core-level Cu 2p orbitals, while freezing all other orbitals; (ii) freeze next the Cu 2p functions and optimize the rest. Repeating these two steps until there is no further change in energy (averaged over all 2p53d10 states) ‘keeps’ the hole down in the core and al- lows the valence and semi-valence shells to relax and polarize, both on-site and at neighbouring sites. Appropriate changes were further operated in the configuration-interaction module of molpro to pre- clude 2p53d10 → 2p63d9 de-excitation in the MRCI wave functions. The dipole transition ME’s between the independently optimized 3d9 and 2p53d10 groups of states were derived according to the proce- dure described in [211]. An approach in the same spirit, based on separate multiconfiguration SCF optimizations for the different dn states but a subsequent treatment of further correlation effects by per- turbational methods, has been recently applied for the computation of XA and RIXS spectra of iron hexacyanides in solution [189–191]. In earlier work in the solid-state context, this idea of using individually optimized wave functions has been employed for the interpretation of X-ray photoelectron spectra [229, 230], the computation of quasiparti- cle band structures [89, 105, 231] and the analysis of superexchange virtual excitations [17, 232]. The scattering geometry and details on how the polarization analy- sis is carried out are specified in Figure 4.1. The unit cell of Li2CuO2 is defined in terms of skew axes, where the angle between the (100) and (101) planes is γ≈69◦, not 45◦ as in the cubic case. The samples could 52 computation of x-ray d-ion spectra in solids

ℏω'

π' σ ~z ℏω π ~y α=50˚ ~x σ z(a) y(c) θ γ=69˚ x(b)

Figure 4.1: Sketch of the scattering geometry in Li2CuO2. The small spheres depict a sequence of three edge-sharing CuO4 plaquettes.

only be cleaved perpendicular to the [101] axis [233]. The scattering plane thus incorporates the crystallographic b axis and is perpendic- ular to the (101) plane. The b axis matches the direction of the chain of edge-sharing CuO4 plaquettes, see Figure 4.1. The ‘included’ scat- tering angle (between incoming and outgoing light beams) was fixed to α = 180◦−130◦= 50◦, while the incident angle θ (between the sam- ple plane and the incoming photon direction) was varied from 10◦ to 125◦. The incoming light was linearly polarized, either perpendicular to the scattering plane (σ polarization) or within the scattering plane (π polarization). Since no polarization analysis was performed for the outgoing radiation, we carry out for the latter an ‘incoherent’ summa- tion over the two independent polarization directions (the vector π0 is introduced to this end). For the analysis of the quantum chemistry wave functions we use a local coordinate frame {x, y, z} having the z axis perpendicular to the

hole orbital mcscf+soc mrci+soc experiment

dxy 0 0 0

dx2−y2 1.24 1.55 1.7

dzx 1.73 2.02 2.1

dyz 1.80 2.09 2.1

dz2 2.54 2.85 2.6 † p3/2 933.6; 933.8 932.5; 932.6 931.6 † p1/2 954.4 953.3 — † 944.9, 945.1 and 965.2 eV by using orbitals optimized for the 3d9 configuration.

Table 4.1: Relative energies (eV) for Cu2+ 3d9 and 2p53d10 states in Li2CuO2. The MRCI d-level splittings include Davidson correc- tions [38]. The x axis is taken along the chain of CuO4 plaquettes and z perpendicular to the plaquette plane, see Figure 4.1. Each theoretical value stands for a spin-orbit Kramers doublet. 4.6 results and discussion 53

CuO4 plaquette and x along the chain. The transformation to the set- ting used on the experimental side ({x˜,y ˜ ,z ˜}, see Figure 4.1) is however straightforward. The rotation of the σ, π and π0 vectors as function of the angle θ is also described by simple geometrical relations : ~ ~ Dσ = Dy˜ , ~ ~ ~ 4 15 Dπ = Dx˜ sin θ + Dz˜ cos θ , ( . ) ~ ~ ~ Dπ0 = Dx˜ sin (θ + α) + Dz˜ cos (θ + α) , E D~ ij = Ψi e ·~R Ψj where ch {gs,fs} stands for dipole transition ME’s, with superscripts dropped in Equation 4.15. By summing up over the out- going polarization directions, Equation 4.14 changes to

d2σRIXS 2 2 ~k ε α h I(hω ,hω 0, ε, θ)= , = ωω03 dΩ0dhω 0 e4c2 X X X X 2 1 hfs| D~ 0 |chi hch| D~ |gsi × ε ε 4 16 ( . ) ggs Egs + hω − Ech + iΓch/2 ε0 gs fs ch Γfs/2π × 2 . ( gs 0) 2 E + hω − Efs − hω +Γfs/4

The natural widths Γch and Γfs are here set to 1 and 0.1 eV, respectively, typical values for Cu L-edge RIXS [115, 133].

4.6 Results and discussion

Computed excitation energies are compared to experimental data in Table 4.1. The Cu 2p core-hole states are strongly affected by spin- orbit couplings (SOC’s): the 2p5 j = 3/2 quartet and j = 1/2 doublet are separated by ≈20 eV. The effect of SOC is on the other hand fairly small for the Cu 3d9 valence states and brings only tiny corrections 6 9 5 10 to those relative energies. The p3/2 2p 3d → 2p 3d excitation en- ergy as computed at the MCSCF level overestimates by about 2 eV the experimental result. But without a separate SCF optimization of the 2p53d10 states and using orbitals optimized for the ground-state 2p63d9 configuration, as in previous L-edge electronic-structure cal- culations [206–208, 234, 235], this energy difference is larger by a fac- tor of 5, i.e. about 11 eV (see footnote in Table 4.1). This strong effect that we quantify here has to do with readjustment of the electronic charge within the Cu second and third atomic shells upon the 2p → 3d transition and further to orbital polarization at neighbouring O sites. It is even stronger than the short-range relaxation and polarization effects for (N1) processes (i.e. complete removal of a e charge) within the valence levels of, e.g. MgO or diamond [89, 105]. Obvi- ously, such strong charge readjustment can in principle significantly affect the dipole transition ME’s employed for deriving the RIXS in- tensities. 54 computation of x-ray d-ion spectra in solids

IRIXS (Å2eV-1) IRIXS (arb.units) 0.0006 σ polarization θ σ polarization θ ab initio experiment 125 125 0.0005 110 110 0.0004 90 90 0.0003 65 65 0.0002 30 30 0.0001 10 10 0.0000 -3 -2 -1 -3 -2 -1 energy loss (eV) energy loss (eV) IRIXS (Å2eV-1) IRIXS (arb.units) 0.0006 π polarization θ π polarization θ ab initio experiment 125 125 0.0005 110 110 0.0004 90 90 0.0003 65 65 0.0002 30 30 0.0001 10 10 0.0000 -3 -2 -1 -3 -2 -1 energy loss (eV) energy loss (eV)

Figure 4.2: Calculated (MRCI+SOC) versus experimental L3-edge double differential cross sections as function of the incident angle θ.

As seen in Table 4.1, the MRCI treatment yields additional correc- 6 9 5 10 tions of ≈ 1 eV to the p3/2 2p 3d → 2p 3d excitation energies, which brings the ab initio values within 1 eV of the measured p3/2 XA edge. It is worth pointing out that longer-range polarization ef- fects are not very important for relatively localized excitations that conserve the number of electrons in the system such as the L-edge 2p → 3d transitions. This is one feature that makes the accurate mod- eling (and the interpretation) of L-edge d-metal RIXS spectra eas- ier than of, e.g. photoemission spectra [11, 236–238], once a reliable many-body scheme is set up for describing the ‘local’ multiplet struc- ture. Ab initio results for full Cu L3-edge RIXS spectra are shown in Fig- ure 4.2 along with experimental data. The theoretical spectra are ob- tained on an absolute intensity scale, an aspect on which we shall 4.6 results and discussion 55

Eloss, eV Eloss, eV 0.0 0.0 I RIXS, Å2eV-1 -0.5 -0.5 0.0007 -1.0 -1.0 -1.5 -1.5 0.0006 -2.0 -2.0

-2.5 -2.5 0.0005 -3.0 -3.0 930 935 940 945 950 955 Einc, eV 930 935 940 945 950 955 Einc, eV θ = 65◦, σ polarization θ = 125◦, σ polarization 0.0004

Eloss, eV Eloss, eV 0.0 0.0 0.0003 -0.5 -0.5 -1.0 -1.0 0.0002 -1.5 -1.5 -2.0 -2.0 0.0001 -2.5 -2.5 -3.0 -3.0 930 935 940 945 950 955 Einc, eV 930 935 940 945 950 955 Einc, eV 0.0000 θ = 65◦, π polarization θ = 125◦, π polarization 4 3 Figure . : Theoretical I(Einc, Eloss) RIXS plots for Li2CuO2 using the geo- metrical setup employed in the measurements. In addition to the Cu L3-edge excitations, the L2 ‘signal’ is also visible. elaborate in the following. The evolution of the RIXS intensities when modifying the incident angle θ is different for each of the main peaks 4 3 — see also the I(Einc, Eloss) plots in Figure . — and this specific de- pendence can be used to unmistakably identify the origin of each particular feature. The quantum chemistry data agree well with the experiment, with all major trends nicely reproduced: the intensities of the two peaks in the range of 2–3 eV, in particular, clearly display distinct behavior with increasing incident angle — while the higher- energy excitation loses spectral weight, the one at about 2 eV grad- ually acquires more and more intensity. The calculations also reveal that the experimental feature at ≈ 2 eV has to do with two different possible final states, with the 3d hole in either the Cu dyz or dzx or- bital. Due to x–y anisotropy, the yz and zx components are slightly split apart but not strongly enough to be resolved as two individual peaks in experiment. The anisotropic environment splits as well the computed p3/2 states (Table 4.1), into two sets of doublets. With re- gard to peak positions, the agreement with the experimental data is within 0.15 eV for the dx2−y2 , dyz and dzx hole states. For the dz2 hole state, the MRCI calculations overestimate the relative energy by 0.25 eV, ≈10%. One achievement with the ab initio methodology is that absolute values for the dipole transition ME’s are obtained, not just relative estimates as in semi-empirical schemes. In particular, it is possible to calculate absolute cross sections for the RIXS process using Equa- tion 4.16. The theoretical plots provided in Figure 4.2 and Figure 4.3 56 computation of x-ray d-ion spectra in solids

are based on such absolute values. Having the latter, one can quan- tify specific ‘scattering properties’ of the material, e.g. how many pho- tons of particular energy and polarization will be scattered by a given atomic site in a given direction. No results have been reported so far for absolute experimental RIXS intensities, although such measure- ments should in principle be possible. Absolute intensities have been reported however for non-resonant inelastic X-ray scattering (NIXS) [239, 240]. Since the RIXS intensities are expected to be a few or- ders of magnitude larger than in NIXS [133], it is then instructive to compare our computational results with the available NIXS data. For direct comparison with the dynamical structure factor per vol- ume unit measured by NIXS, s, we must divide the cross sections from Equation 4.16 by the product of the unit cell volume per Cu site, the squared classical electron radius and the ratio ω0/ω≈1 and addi- tionally account for the fact that in RIXS we integrate over all ε0. By doing so, the initial value IRIXS ∼10−4 Å2eV−1 translates into intensi- ties ∼103 Å−3eV−1, about 105 times larger than typical values of s in NIXS (∼10−2 Å−3eV−1)[240].

4.7 Conclusions

In this chapter we introduce a multireference configuration- interaction-type of methodology for the ab initio calculation of transition-metal L-edge excitations in solids, i.e. both XA energies and RIXS cross sections for the 2p → 3d → 2p transitions. The results ob- tained with this computational scheme show good agreement with experimental data, for both excitation energies and polarization- and angle-dependent intensities. Orbital relaxation effects in the presence of the core hole are taken into account by performing two indepen- dent MCSCF calculations and utilising afterwards a nonorthogonal configuration-interaction type of approach for the computation of the dipole transition matrix elements [188, 209–211]. The use of separately optimized orbitals brings an energy stabilization effect of 11 eV for the core-hole states as compared to the case where ground-state or- bitals are employed. While here we test this computational scheme for the more convenient case of a Cu2+ 3d9 ground state — giving rise to a simple, closed-shell 3d10 valence electron configuration subse- quent to the initial absorption process — the procedure is applicable to other types of correlated electron systems, dn or fn. Having the new methodology in hand, such investigations and more involved calculations for arbitrary n will be the topic of future studies. Part III

5 d TRANSITION-METALSYSTEMS

MAGNETICPROPERTIESOF 5 PYROCHLORE Cd2Os2O7

y many-body quantum-chemical calculations, we investigate in this chapter the role of two structural effects – local ligand B distortions and the anisotropic Cd-ion coordination – on the magnetic state of Cd2Os2O7, a spin S = 3/2 pyrochlore. We find that these effects strongly compete, rendering the magnetic interac- tions and ordering crucially depends on geometrical features. With- out trigonal distortions a large easy-plane magnetic anisotropy devel- ops. Their presence, however, reverses the sign of the zero-field split- ting and causes a large easy-axis anisotropy (D≈−6.8 meV), which in conjunction with the antiferromagnetic exchange interaction (J ≈ 6.4 meV) stabilizes an all-in/all-out magnetic order. The uncovered com- petition is a generic feature of pyrochlore magnets. Oxide compounds of the 5d elements are at the heart of inten- sive experimental and theoretical investigations in condensed-matter physics and materials science. The few different structural varieties displaying Ir ions in octahedral coordination and tetravalent 5d5 va- lence states, in particular, have set the playground for new insights into the interplay between strong spin-orbit interactions and electron correlation effects [241–243]. Equally intriguing but less investigated 3 are pentavalent Os 5d oxides such as the pyrochlore Cd2Os2O7 [244, 245] and the perovskite NaOsO3 [246, 247]. Just as for the iri- dates, major open questions with respect to the osmates are the mech- anism of the metal-to-insulator-transition (MIT) and the nature of the low-temperature AF configuration. While initially a Slater type pic- ture has been proposed for the MIT’s in both Cd2Os2O7 [244] and NaOsO3 [246, 247], alternative scenarios such as a Lifshitz transition in the presence of sizeable electron-electron interactions have been suggested as well very recently [248, 249]. So far theoretical investigations of the electronic structure of the Os 5d3 oxides have been carried out only as mean-field GS calculations within the LDA [250] or the LDA+U model [246, 248]. In this chap- ter we present results of many-body quantum-chemical calculations 3 for Cd2Os2O7. We describe the local Os d multiplet structure, the precise mechanism of 2nd-order SOC and zero-field splitting (ZFS), and determine the parameters of the model spin Hamiltonian, i.e. the single-ion anisotropy (SIA), NN Heisenberg exchange as well as the Dzyaloshinskii-Moriya (DM) interactions. The results indicate that ba- sic electronic-structure parameters such as the SIA crucially depend on geometrical details concerning both the O and adjacent Cd ions.

59 60 magnetic properties of pyrochlore Cd2Os2O7

In particular, the anisotropic Cd-ion coordination lifts the degeneracy of the Os t2g levels even in the absence of O-ion trigonal distortions and yields large ZFS’s and easy-plane magnetic anisotropy. A config- uration with trigonally compressed O octahedra is however energeti- cally more favorable, changes the sign of the ZFS, and thus gives rise to easy-axis anisotropy at the Os sites, which in conjunction with the AF NN exchange stabilizes the so-called all-in/all-out spin order. The competition between the two structural effects is a generic feature in 227 pyrochlores and opens new perspectives on the basic magnetism in these materials.

5.1 Os d3 electron configuration, single-ion physics

To investigate in detail the Os d-level electronic structure, CASSCF and MRCI calculations (see Chapter 1 and Refs. [11, 38, 227]) were first performed on seven-octahedra embedded clusters made of one reference OsO6 octahedron, six adjacent Cd sites, and six NN OsO6 octahedra (see Figure 5.1a). The farther solid-state environment was modeled as an one-electron effective potential which in a ionic pic- ture reproduces the Madelung field in the cluster region. Valence ba- sis sets of quadruple-zeta quality [38] from the molpro [116] library were used for the central Os ion [251] and triple-zeta basis functions were applied for the ligands [225] of the central octahedron and the NN 5d sites [251]. For the central Os ion we also employed two po- larization f functions [251]. For farther ligands in our clusters we applied minimal atomic-natural-orbital basis sets [252]. All occupied shells of the NN Cd2+ ions were incorporated in the large-core pseu- dopotentials and each of the Cd 5s orbitals was described by a sin- gle contracted Gaussian function [253]. We accounted for all single and double excitations from the Os 5d and O 2p orbitals at the cen-

(a) (b)

Figure 5.1: Sketch of (a) the cluster used for the ZFS calculations, (b) the com- pressive trigonal distortion in Cd2Os2O7. Os, O, and Cd ions are shown in blue, beige, and green, respectively. 5.1 Os d3 electron configuration, single-ion physics 61

5d3 splittings mrci mrci+soc (×2)

4 3 −3 A2 (t2g) 0.00 0.00; 13.5 · 10 2 3 E (t2g) 1.51; 1.51 1.40; 1.53 2 3 T1 (t2g) 1.61; 1.62; 1.62 1.63; 1.66; 1.76 2 3 T2 (t2g) 2.46; 2.49; 2.49 2.63; 2.76; 2.87 4 2 1 T2 (t2geg) 5.08; 5.20; 5.20 5.14; ... ; 5.45 4 2 1 T1 (t2geg) 5.89; 6.01; 6.01 6.02; ... ; 6.33 4 1 2 T1 (t2geg) 10.29; 10.63; 10.63 10.41; ... ; 11.00

Table 5.1: MRCI and MRCI+SOC relative energies (eV) for the Os5+ 5d3 multiplet structure of Cd2Os2O7. Since cubic symmetry is lifted, the T states are split even without SOC. Each MRCI+SOC value stands for a spin-orbit doublet; for the 4T states only the lowest and highest components are shown.

tral OsO6 octahedron in the MRCI treatment. To separate the O 2p valence orbitals into two different groups, i.e. at sites of the central octahedron and at NN octahedra, we used the Pipek-Mezey orbital localization procedure [228]. All calculations were performed with the molpro quantum-chemical software [116]. From simple considerations based on crystal field (CF) theory [254, 255], the orbital GS for a d3 ion in octahedral coordination is a singlet. Three singly occupy the t2g orbitals with their spins aligned due to the Hund’s coupling. Thus, neglecting SOC, the lowest elec- 4 3 tron configuration should be A2(t2g). This is indeed found in the quantum-chemical calculations if SOC is not included, see Table 5.1. 4 The components of the spin-quartet A2 GS can interact, however, ‡ via SOC with higher lying terms and in noncubic axial environment ‡See, for example, split into two Kramers doublets, mS = 1/2 and mS = 3/2, respec- Table A34 in Ref. [255]. tively [256–258]. Excited T1 and T2 states are themselves split in non- cubic CF’s. In the recent work of Matsuda et al. [245], for instance, the 4 4 interaction of the A2 and lowest T2 states was analyzed. The ab ini- tio quantum-chemical calculations yield, nevertheless, a rather large 4 4 A2– T2 excitation energy, 5.4 eV by CASSCF calculations including in the active orbital space all 5d functions at the central Os site and the t2g orbitals of the six Os NN’s, which is not surprising in fact for extended 5d orbitals that feel very effectively the O 2p charge distri- bution. Additionally, the ab initio investigation actually shows that the 2 2 3 excited E and T t2g spin doublets occur at much lower energies, 1 4 to 2.5 eV (see Table 5.1), and the ZFS of the A2 levels is mainly related to the interaction with those states. To make the MRCI calculations feasible and the analysis of the results limpid, we further replaced the ‡‡ the size of the six Os5+ d3 NN’s by closed-shell Ta5+ d0 species‡‡. We checked, for MRCI expansion is that way reduced by example, that the 5d charge distribution and the t2g–eg splitting at a few orders of magnitude. 62 magnetic properties of pyrochlore Cd2Os2O7

E E E E H^ 3 3 3 1 3 1 3 3 eff 2 , − 2 2 , − 2 2 , 2 2 , 2 D 3 3 2 , − 2 6.7738 3.9110 + 3.9110i −3.9106i 0.0000 D 3 1 2 , − 2 3.9110 − 3.9110i 6.7741 0.0000 −3.9106i D 3 1 2 , 2 3.9106i 0.0000 6.7738 −3.9110 − 3.9110i D 3 3 2 , 2 0.0000 3.9106i −3.9110 + 3.9110i 6.7738

Table 5.2: Single-site effective interaction matrix obtained by MRCI+SOC cal- culations for Cd2Os2O7 on seven-octahedra cluster (meV).

E E E E H^ 3 3 3 1 3 1 3 3 mod 2 , − 2 2 , − 2 2 , 2 2 , 2 D √ √ 3 3 3 9 3 2 , − 2 4 (Dxx+Dyy)+ 4 Dzz − 3(Dxz+iDyz) 2 (Dxx−Dyy+2iDxy) 0 D √ √ 3 1 7 1 3 2 , − 2 − 3(Dxz−iDyz) 4 (Dxx+Dyy)+ 4 Dzz 0 2 (Dxx−Dyy+2iDxy) D √ √ 3 1 3 7 1 2 , 2 2 (Dxx−Dyy−2iDxy) 0 4 (Dxx+Dyy)+ 4 Dzz 3(Dxz+iDyz) D √ √ 3 3 3 3 9 2 , 2 0 2 (Dxx−Dyy−2iDxy) 3(Dxz−iDyz) 4 (Dxx+Dyy)+ 4 Dzz

5 3 H^ ~ ¯ ~ 3 Table . : Interaction matrix for the single-site mod = S · D · S and S = 2 .

the central Os site are not affected by this substitution. In the MRCI computations, we included on top of the CASSCF wavefunctions all single and double excitations from the Os 5d and O 2p orbitals at the central octahedron. The reference multiconfiguration active space is given by five 5d orbitals at the central Os site and three electrons and the results are listed in Table 5.1. To extract the ZFS, we compute the spin-orbit interaction matrix for the lowest four quartet and three doublet spin states, see Table 5.1. 4 The MRCI+SOC data of Table 5.1 show that the ZFS of the A2 GS is rather large, 13.5 meV. This gives a first estimate of the splitting parameter |D| ≈ 6.75 meV, one to two orders of magnitude larger than for 3d ions in standard coordination [256–258]. However, for explicitly deriving the full SIA tensor D¯ , we further employ the ef- fective Hamiltonian methodology described in Ref. [259]. Here, SOC 4 and the mixing of the A2 components with higher lying CF excited states are treated as small perturbations and the spin-orbit wavefunc- 3 tions related to the high-spin t2g configuration are projected onto 4 the space spanned by the A2 |S, MSi states. Using the orthonor- 4 malized projections Ψ˜ k of the low-lying A2 quartet wave functions and the correspondingP eigenvalues Ek we then construct the effective H^ 259 Hamiltonian eff = k Ek Ψ˜ k Ψ˜ k [ ]. The orthonormalization is done using the formalism introduced by des Cloizeaux [260], i.e. S−1/2 | S | Ψ˜ k = ov Ψki, where ov is the overlap matrix and Ψki are the projections of the wavefunctions onto the model space. A one-to-one correspondence can be now drawn between the matrix elements of H^ 5 2 eff (see Table . ) and the model Hamiltonian for the anisotropic 5.1 Os d3 electron configuration, single-ion physics 63

H^ ~ ¯ ~ 5 3 single site problem, mod = S · D¯ · S (shown in Table . ). It gives rise to a system of equations that can be easily solved to obtain the single-ion anisotropy tensor in the original (crystallographic) coordi- nate frame. The latter reads, in meV,     Dxx Dxy Dxz 1.806 −2.258 −2.258     D¯ = D D D = −2.258 1.806 −2.258 .(5.1)  xy yy yz    Dxz Dyz Dzz −2.258 −2.258 1.806

Diagonalizing the D¯ tensor     DXX 0 0 4.064 0 0 ¯     5 2 D¯ mag = 0 D 0 = 0 4.064 0 ,( . )  YY    0 0 DZZ 0 0 −2.710 we finally obtain on the basis of the ab initio quantum-chemical data an axial parameter D=DZZ − (DXX + DYY)/2=−6.77 meV, a rhombic −5 parameter E = (DXX − DYY)/2 = 2 · 10 meV ≈ 0, and an univocally defined easy axis of magnetization along the h111i direction, perpen- dicular to the plane of Cd NN’s. To gain deeper insight into the origin of the strong axial anisotropy, we performed additional calculations for an ideal pyrochlore struc- ture with no trigonal distortions of the O octahedra. It turns out that the 227 pyrochlore lattice is fully defined by just three parameters: the space group, the cubic lattice constant a, and the coordinate x of the O at the 48f site, denoted O(1) in Ref. [244]. For x=x0 =5/16, the oxygen cage around each Os site forms an undistorted, regular octa- hedron [261]. In Cd2Os2O7, however, x is slightly larger than x0 [244], which translates into a compressive trigonal distortion of the OsO6

5d3 splittings mrci mrci+soc (×2)

4 3 −3 A2 (t2g) 0.00 0.00; 5.00 · 10 2 3 E (t2g) 1.53; 1.53 1.43; 1.50 2 3 T1 (t2g) 1.61; 1.61; 1.63 1.64; 1.64; 1.75 2 3 T2 (t2g) 2.42; 2.45; 2.45 2.59; 2.75; 2.81 4 2 1 T2 (t2geg) 5.34; 5.41; 5.41 5.36; ... ; 5.65 4 2 1 T1 (t2geg) 6.13; 6.20; 6.20 6.23; ... ; 6.52 4 1 2 T1 (t2geg) 10.85; 10.98; 10.98 10.93; ... ; 11.38

Table 5.4: MRCI and MRCI+SOC relative energies (eV) for the Os5+ 5d3 multiplet structure in a hypothetical Cd2Os2O7 crystal with no trigonal distortions. Each MRCI+SOC value stands for a spin-orbit doublet; for the 4T states only the lowest and highest components are shown. 64 magnetic properties of pyrochlore Cd2Os2O7

D (meV) 1.2 ∆Etot(eV) 10 (a) (b) 1 5 0.8 Δθ (°) 0.6 -2 -1 1 2 3 4 0.4 -5 0.2 -10 Δθ (°) -2 -1 1 2 3 4 -15 -0.2

Figure 5.2: (a) Single-ion anisotropy parameter D and (b) total energy of the cluster ∆Etot, as functions of the trigonal distortion angle ∆θ, see text. MRCI+SOC results for clusters as that depicted in Fig- ure 5.1b. No trigonal distortions are applied within the embed- ding, i.e. x=x0 =5/16 for the surrounding lattice.

octahedra. Interestingly, for undistorted O octahedra (i.e. x = x0) we still found E=0 and the same orientation of the magnetic z axis. The axial parameter, however, changes its sign to D=2.50 meV. A positive D indicates that the z axis is now a hard axis. The basic mechanism responsible for this change of sign is apparent when comparing the Os 5d3 multiplet structure for the actual lattice (Table 5.1) with that for a idealized crystal with no trigonal distortions (Table 5.4). In par- 2 ticular, the order of the split components of the T1 states, i.e. one singly-degenerate and one doubly-degenerate term, changes in Ta- ble 5.4 as compared to Table 5.1, which modifies the interaction of 4 the A2 terms with the excited states. We find in the MRCI calculations a linear relation between D and the distortion angle ∆θ, see Figure 5.2. The effect of a trigonal field alone on the energy levels and on the ZFS of d3 ions has been previ- ously investigated in the framework of a simplified CF model and per- turbation theory [257, 258, 262]. A linear relation is also established in alum salts [262]. In contrast to the model assumed in Ref. [262], how- ever, we find that the D(∆θ) line does not go through origin. This is due to the noncubic field generated by the Cd NN’s, which form a highly anisotropic planar structure arround each Os site. Similar competing stuctural effects have been evidenced in NiII–YIII molecu- lar complexes [263]. We also plot in Figure 5.2 the total-energy land- scape as function of the distortion angle ∆θ. It is seen that at the MRCI level the minimum of the parabola corresponds to a trigonal compressive distortion ∆θ=1.7◦ for fixed Os-O bond length, close to the experimental value of 1.9◦ [244].

5.2 Isotropic and antisymmetric exchange interactions

To investigate the NN magnetic couplings, we designed ten- octahedra embedded clusters. The analysis of the intersite interac- tions is here carried out for only two magnetically active Os5+ d3 5.2 isotropic and antisymmetric exchange interactions 65

casscf mrci mrci+soc

Stot =0 0.00 0.00 0.00

Stot =1 2.51 5.98 6.04, 6.04, 6.44

Stot =2 7.75 18.38 18.66, 18.66, 19.01, 19.01, 19.12

Stot =3 16.24 37.93 38.51, 38.51, 38.81, 38.81, 38.99, 38.99, 39.05

Table 5.5: Relative energies (meV) of the singlet, triplet, quintet, and septet 3 states for two adjacent Os d sites in Cd2Os2O7. In each case, the singlet is taken as reference. ions (see Figure 5.3), while for simplicity the eight Os5+ NN’s were modeled as closed-shell Ta5+ d0 ions. The basis functions used here were as described in Section 5.1. The only exception was the O ligand bridging the two magnetically active Os sites, for which we employed quintuple-zeta valence basis sets and four polarization d functions [225]. Multiconfiguration wavefunctions were first generated by state- averaged CASSCF optimizations for the lowest singlet, triplet, quin- tet, and septet states in the two-site problem. This manifold gives rise in the spin-orbit calculations to sixteen spin-orbit states. For each spin multiplicity the CASSCF active space is defined by the set of six Os t2g orbitals accommodating a total number of six electrons. We then further accounted for single and double excitations from the Os t2g and bridging-O 2p orbitals on top of the CASSCF reference wavefunctions. Such MRCI calculations yield magnetic coupling con- stants in very good agreement with experimental data in the 5d5 iri- date Sr2IrO4 [264]. Similar strategies of explicitly dealing only with selected groups of localized ligand orbitals were adopted in earlier studies on 3d compounds [12, 16, 17]. Relative energies for the lowest singlet, triplet, quintet, and septet are listed in Table 5.5. From the calculations without SOC, we see that the energy splittings between the different spin states follow the sequence J, 3J, 6J and can be fitted to an AF isotropic exchange model H^ ~ ~ AF = J S1 ·S2, with a root-mean- error of 0.3 meV. Adding a 2 biquadratic term K(~S1 · ~S2) improves the accuracy of the fit to 0.01 meV and yields a first MRCI estimate of J=6.42 meV. Thus J is of the same order of magnitude as the SIA while K=0.07 meV. Anisotropic, non-Heisenberg terms can be further obtained from 3 spin-orbit calculations within the manifold defined by a ttg S = 3/2 4 4 electron configuration at each active Os site. For this A2⊗ A2 mani- fold, we do not include second-order couplings through higher lying crystal-field excited states. As a result, we describe only the diagonal interactions and the first-order contribution to the DM anisotropic term. For 3d oxides, it has been shown, nevertheless, that the di- agonal couplings are the most important, see Refs. [265, 266]. For the same reason, the above treatment does not include the single-ion 66 magnetic properties of pyrochlore Cd2Os2O7

Figure 5.3: The cluster used for computing the intersite magnetic couplings for Cd2Os2O7. Os, O, and Cd sites are shown in blue, beige, and green, respectively. The NN Os5+ ions around the two central octahedra are modeled as Ta5+ d0 species.

anisotropies, which we have obtained from the single-site calculations described above. As expected, including SOC does not bring significant corrections to the effective superexchange: J is 6.43 meV by MRCI+SOC cal- culations. The spin-orbit interactions, however, lift the degeneracy of the initial spin multiplets and brings in antisymmetric contribu- tions. To describe the latter, we adopt a two-site model Hamiltonian H^ 0 ~ ~ ~ ~ 2 ~ ~ ~ 5 6 mod =JS1 · S2 + K(S1 · S2) + d · [S1 × S2] (see also Table . ) contain- ing, in addition to the Heisenberg and biquadratic terms, a DM anti- symmetric exchange contribution defined by a pseudovector ~d with components dx, dy, dz. This model is then compared with the ab initio H^ 0 16×16 effective matrix eff derived from the MRCI+SOC calculations One should notice shown in Table 5.7. We found perfect one-to-one correspondence be- that only the dif- tween the two sets of matrix elements Table 5.6 and Table 5.7, con- ferences between the H^0 firming that our choice of mod is appropriate and could extract the diagonal elements ~ can be used in the numerical parameters J = 6.43, K = 0.07, and d = (1.17, −1.17, 0) meV. analysis because the According to these results, the DM vector is oriented along the h110i trace of an effective directions (see Figure 5.4 and Ref. [267]) and its norm is |~d| = 1.65 Hamiltonian is in meV. The sign of the DM vector for a given Os-Os link, however, can- general arbitrary. not be determined from our ab intio data. Small off-diagonal elements on the order of 0.01 meV are neglected in our analysis. Whereas the absolute value of the MRCI DM parameter is simi- lar to estimates derived for certain values of the on-site Coulomb repulsion U by LDA+U calculations [248], the superexchange J and the SIA D turn out to be two (J) to seven (D) times smaller in our quantum-chemical study. It is known that quantum-chemical calcu- 5.2 isotropic and antisymmetric exchange interactions 67 lations based on either configuration-interaction techniques or 2nd- order perturbation theory are able to reproduce the experimentally derived J coupling constants with an accuracy better than 20% in insulating 3d-metal oxides, see, e.g. Refs. [13–15] and [12, 16, 17]. For more extended 5d electronic states in 5d-metal oxides, the Hubbard U is neither small nor large compared to the bandwidth, as corrob- orated from the MIT’s [244, 247] and from constrained LDA calcu- lations [246, 248, 268]. Electron itineracy may therefore pose addi- tional technical problems and limit the applicability of a finite cluster approach. It has been shown, however, that accurate J’s [264] and d-d excitation energies [269–271] can be indeed computed for 5d ox- ides such as the iridates. Those findings in [264] and [269–271] indi- cate that approaching the essential physics in the 5d oxides from a more localized perspective and within a finite-fragment framework certainly is a good starting point. Additional potentially problematic points concern the size of the clusters, as discussed, e.g. in Ref. [104]. The clusters we employ here are, however, large enough to ensure an accurate description of both the charge distribution of NN octa- hedra around the ‘active’ d sites and the tails at those neighbouring sites of Wannier-like orbitals in the active region (see discussion in Section 1.2. As concerns the SIA at d-metal sites in solids, extensive quantum-chemical investigations are missing. The present investiga- tion therefore constitutes a step toward filling this gap. The pure AF Heisenberg model on the pyrochlore lattice has an extensive number of classical GS’s which prevent any Néel type or- dering down to zero temperature [272–275]. The easy-axis anisotropy lifts this macroscopic degeneracy and selects the so-called all-in/all- out (or it’s time reversed) state, whereby all four spins sharing a given point toward or away from its center. This GS is

Figure 5.4: Network of corner-shared Os4 tetrahedra (blue color) in Cd2Os2O7 and O ligands plus NN Cd ions around two adjacent Os sites. The two possible orientations of the DM vector [267] for two adjacent Os ions is also shown, along with their respective easy axes. 68 magnetic properties of pyrochlore Cd2Os2O7

selected irrespective of the relative strength of the anisotropy and AF exchange energy [276]. Now, since the macroscopic GS degeneracy is already lifted by the SIA, including the smaller DM interactions is not expected to drastically modify the physics of the system. In fact, if the sign of the ~d vector is that of the direct type (as defined by Elha- jal et al. [267]), the DM interaction also favors the all-in/all-out state and thus does not compete with the SIA. The competition arises for the opposite sign of the ~d vector, labelled as of indirect type in [267], where the DM couplings favor a continuous family of coplanar and non-coplanar GS’s [267].

5.3 Conclusions

While trigonal distortions of the ligand cage are believed to open the door to topologically insulating states in j = 1/2 227 iridate py- rochlores [242, 268, 277], we here show that the lattice degrees of freedom can fundamentally modify the nature of the magnetic inter- actions and ground states in 5d 227 compounds with larger magnetic moments. For this we employ ab initio multireference configuration- interaction methods, the results of which are further mapped onto model Hamiltonians including both isotropic and anisotropic terms. To extract the DM couplings, we have expanded the formalism pro- posed earlier for S = 1/2 cuprates [265, 266, 278]. For the experimen- tally determined crystal structure of Cd2Os2O7 [244], we find an AF superexchange coupling J=6.43 meV and easy-axis anisotropy of the same magnitude, D = −6.77 meV, while the DM parameter is 3 to 4 times smaller. The dominant magnetic interactions, J and D, give rise to the all-in/all-out AF order [267, 276]. Most remarkably, the hexagonal Cd-ion coordination induces electrostatic fields that com- pete with the compressive trigonal distortion of the O octahedra that is present in the system. If the latter is removed the SIA turns pos- itive, i.e. D reverses sign and the system develops easy-plane mag- netic anisotropy. This obviously implies qualitatively different mag- netic properties [279, 280]. Our findings thus indicate that there is in 227 pyrochlores a transition from unfrustrated to highly frustrated magnetism controlled by the amount of trigonal distortion. The lat- ter can in principle be varied by changing either the A-site or B-site ions in the A2B2O7 compounds, replacing for instance Cd by Hg or Os by Re [261], and is in addition tunable by applying pressure. For a correct understanding of such magnetic interactions, ordering tenden- cies, and transitions the subtle interplay between oxygen- and A-ion anisotropic electrostatic effects is essential. 5.3 conclusions 69 . 2 3 = 1 , 2 S and ) 2 S ~ × 1 S ~ d ( ~ ) ) + x x K 2 z id id ) 25 i 16 − + 2 5d 2 0 y y √ 0 0 0 0 0 0 0 0 0 0 0 0 S ~ + · 1 2 J 0 , ( d ( d | 5 2 5 2 1 4 15 − i S ~ q q − 1 2 2 1 K ( ) ) ) x x K x + z id 6 id id 21 2 d i 1 − 1 + 1 S ~ 3 5 − y · 0 0 0 0 0 0 0 0 0 0 0 y + y q J 1 , ( d 1 ( d | ( d i2 5 2 4 3 5 S ~ 11 2 5 − J q q − q 1 2 = ) ) ) x K x x id K 6 z z id id 21 mod i i 1 1 6 d − 0 1 3 0 − + 8 1 5d 5 y ^ 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 2 0 0 0 0 4 + y y H √ √ + J 3 , 1 , ( d i | | 1 2 J ( d ( d 3 2 i 4 − 9 4 11 6 5 6 5 q − q q 3 2 ) ) ) ) x x K x x z id K 6 id id 21 id d i i 1 + 1 z − 2 3 5 + 81 16 − d y − 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 y + y y q 3 2 + J 3 , ( d 1 , i ( d | J | ( d 5 2 4 3 5 9 4 11 2 5 ( dy − i2 for the two-site q 3 2 q − q 1 2

) ) ) ) ) x x tot S x x x K id id id id K id z M z i i 9 + − − + d , 16 − 1 2 d 81 16 2 5 y y y y 0 0 0 0 0 0 0 0 0 2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 y 3 2 + tot + i J q ( d ( d 3 , 2 , ( d ( d ( d | | J S − 3 4 0 0 3 5 3 2

i3 3 5 9 4 1 1 − 3 3 q q q √ √ 2 3 3 2 2 2 2 ) ) ) x x x ) ) x id id id K x z K z z id d − + − i i 9 id d d 16 0 1 2 5 − 81 16 y y y 3 5 5 + 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 + y q 9 + √ ( d ( d ( d y J q 3 , 2 , | | 2 J 0 0 0 ( d 3 4 3 3 3 i i2 ( d 9 4 1 1 1 6 5 − i3 − 3 2 q q q q 3 3 2 2 3 2 ) ) ) ) ) ) x x x x x x K id z id id id K z id id 6 d z i i 9 1 6 − + − + d 1 0 d 5 − + 1 8 2 5 y y y y − 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 5 9 + √ y y 4 + 2 J q ( d √ 2 , ( d ( d ( d 3 , | J | ( d ( d i 3 4 3 5 3 5 3 5 i3 9 4 − i 2 5 2 5 10 − 3 q q q √ q q 3 3 2 3 2 2 2 ) x ) ) ) x id K x x z K z id 6 d + i i 9 1 6 id id d 1 z 2 5 + 1 8 y d 3 5 + − − 2 − 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 + y q 3 2 + ( d y y J q 3 , 2 , i J | | ( d i3 4 3 3 i2 ( d ( d 9 4 10 6 5 − 3 2 3 2 q q 2 3 ) − ) ) ) x x x x K id id id K id 6 z i i 9 + + − 1 + 81 16 id y y y − 3 − 2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 y + 3 2 + 5 . 6 : Model interaction matrix in the coupled basis J ( d ( d ( d 3 , 2 , ( d J | | − 3 4 0 3 2 3 2 3 5 9 4 1 − 3 q q q √ 2 3 2 3 2 2 able | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | T 0 0 1 0 0 0 1 2 3 1 3 2 1 1 2 0 1 2 0 0 − 1 − 2 − 2 − 1 − 2 − 2 − 3 − 3 − 1 − 1 − 1 − 1 mod mod 0 0 3 , 1 , 1 , 0 , 2 , 2 , 2 , 3 , 3 , 3 , 3 , 2 , 3 , 1 , 3 , 3 , 2 , 2 , 1 , 0 , 3 , 3 , 3 , 3 , 2 , 2 , 3 , 3 , 2 , 2 , 1 , 1 , ^ ^ h h h h h h h h h h h h h h h h h h h h h h h h h h h h h h h h H H 70 magnetic properties of pyrochlore Cd2Os2O7 (meV). 7 O 2 Os 2 i 0 0 . 936i 0 . 936i + − 0 , 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 | . 936 . 936 i 1 0 . 745i1 . 827i 0 . 000 0 . 000 0 . 936i 0 . 579 + − + 1 , 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 0000 . 000 0 0 . 000 6 . 528 0 | . 745 . 827 . 936 i i 2 . 120i 0 3 1 . 293i1 . 293i 0 . 000 0 . 000 − . 525 + − 1 , 3 , 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 0000 . 000 6 . 528 1 0 . 000 0 . 000 0 . 000 0 . 000 0 0 . 000 0 | | 0 . 016i 38 − 0 . 011 . 293 . 293 1 1 − 2 . 120 i i 1 . 727i 0 . 000 2 1 . 827i 0 . 745i 0 . 936i − − 1 . 505 + − − . 000 . 000 . 000 . 000 . 000 . 000 . 000 3 , 1 , 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 6 . 528 0 . 000 0 0 0 | | 38 . 827 . 936 1 0 − 1 . 727 , obtained from two-site MRCI+SOC calculations for Cd

tot S i i 1 . 339i 0 0 . 550i 0 0 . 550i 0 2 . 120i 0 2 1 1 . 827i − − + + M . 511 . 954 + . 000 , 2 , 3 , 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 | | 0 . 010i 38 18 − 0 . 011 − 0 . 016i − 0 . 019i 0 . 745 tot . 827 1 S − 1 . 339 − 0 . 550 − 0 . 550 − 2 . 120

i i 0 . 947i0 . 947i 0 . 000 0 . 000 0 . 947i1 . 727i 0 . 000 0 . 000 1 0 1 . 293i − + − + . 977 . 507 + . 000 . 000 . 000 . 000 . 000 2 , 3 , 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 0 0 0 0 | | 18 38 − 0 . 023i . 293 1 − 0 . 947 − 0 . 947 − 0 . 947 − 1 . 727 i i 0 . 550i1 . 339i 0 . 000 0 . 000 1 . 339i1 . 339i 0 . 000 0 . 000 0 0 . 745i 0 . 745i − + − + − 1 . 511 . 985 + − . 000 2 , 3 , 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 | 0 . 016i 0 . 019i | 38 18 − 0 . 01i − 0 . 011 − 0 . 019i . 745 . 745 0 0 − 0 . 550 − 1 . 339 − 1 . 339 − 1 . 339 i i 1 . 727i 0 . 000 1 . 727i0 . 947i 0 . 000 0 . 000 1 . 293i + − + − 2 − 1 . 505 . 977 − . 000 . 000 . 000 3 , 2 , 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 . 000 0 0 0 0 . 023i | | 38 18 . 293 1 − 1 . 727 − 1 . 727 − 0 . 947 i i 1 5 4 6i 2 . 120i 0 . 000 2 . 120i0 . 550i 0 . 000 0 . 000 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 9i 1 . 827i + − + − 3 − 2 . 52 . 95 − 0 0 0 3 , 2 , 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 0 . 00 7 0 . 01 | | 38 18 − 0 . 01 − 0 . 01 5 . 7 : Effective interaction matrix in the coupled basis 1 . 82 − 2 . 12 − 2 . 12 − 0 . 55 able | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | f f T 0 1 0 1 2 3 0 1 1 2 3 0 1 2 0 1 0 0 2 0 ef ef 0 0 − 1 − 1 − 2 − 1 − 1 − 2 − 1 − 3 − 3 − 2 − 2 − 1 ^ ^ 3 , 2 , 3 , 3 , 3 , 3 , 1 , 1 , 3 , 3 , 3 , 2 , 2 , 2 , 1 , 1 , 0 , 2 , 2 , 0 , H H 2 , 3 , 3 , 1 , 3 , 3 , 2 , 3 , 3 , 2 , 2 , 1 , h h h h h h h h h h h h h h h h h h h h h h h h h h h h h h h h ANISOTROPIC MAGNETISM IN 214 STRONTIUMIRIDATE 6

heir unique diversity of transport and magnetic properties endows TM oxides with a long-term potential for applica- T tions in microelectronics and electrical engineering. Nowa- days the search for new or superior properties goes beyond known bulk phases and includes oxide interfaces and stacked superlat- tices [281, 282]. A variety of intrinsically stacked crystalline oxides is presently known. The high-temperature cuprate superconductors [283], for example, fall in this category but also iridates of the type 2+ 2+ A2IrO4 (A=Sr , Ba ) which closely resemble undoped cuprates, both structurally and magnetically [284–288]. Sr2IrO4 has quasi 2D crystalline structure displaying stacked IrO2 and double SrO layers. It is the prototype compound in the class of spin-orbit (SO)-driven, Mott-like magnetic insulators [284, 285]. The single s = 1/2 hole present in the nearly degenerate t2g orbitals car- ries an angular moment leff = 1 and is subject to a large SOC, which in first approximation results in an effective Ir4+ moment, or pseu- 6 284 289 dospin, jeff =leff−s=1/2 [ , , ]. In this chapter we compare the experimental properties of these pseudospins to the ones we have cal- culated by ab initio quantum chemistry methods. Namely, we obtain and discuss the spatial anisotropies of the g factors in Sr2IrO4 and the structurally related Ba2IrO4 compound and further the detailed microscopic exchange interactions in strontium iridate. Prediction of isotropic exchange couplings in 3d TM oxides with QC WF-based methods is a well-established procedure not only for the simplest case of s = 1/2 NN’s [14, 16, 17], but also for next-NN couplings [15, 127, 290] and for s>1/2 systems [12, 13, 291]. Calculations of the anisotropic exchange arising due to SOC are on the other hand rather scarce. Studies based on the effective Hamil- tonian mapping procedure were applied to systems with quenched orbital moment and second-order SOC [265, 266, 278], see also Chap- ter 5. Mapping of the ab initio excitation energies to eigenvalues of a model Hamiltonian was successfully used to explain the anisotropic exchange in iridates with strong first-order SOC [7, 292]. An alter- native approach of treating intersite exchange in presence of SOC based on the Lines model [293] was recently applied to polynuclear TM molecular complexes [294, 295]. In the Lines scheme one projects the Heisenberg model onto the SOC pseudospin basis and obtains the rescaled isotropic coupling ignoring the anisotropic part of the exchange.

71 72 anisotropic magnetism in 214 strontium iridate

xy xz,yz

δ δ xz,yz C4 xy C4

δ<0, g⊥<20, g||<2

Figure 6.1: TM t2g splittings for tetragonal distortions of the oxygen octahe- dron and corresponding S˜ = 1/2 hole profiles. (a) z-axis compres- sion of the octahedron causes an orbital doublet with occupied xy orbital to be lowest in energy. (b) Elongation of the octahedron causes an orbital singlet with occupied xz and yz orbitals to be lowest in energy. Purple dashed lines indicate the conventional zero level used to define the sign of δ.

The computational approach, explored in this chapter on a material with strong first-order spin-orbit coupling for the first time, provides direct access to both isotropic and anisotropic magnetic exchange in- teractions. It is an extension of the method employed in [7, 292] to cases in which ab initio energy spectrum provides lesser values than parameters in the model. The good agreement between the electron spin resonance (ESR) data and the outcome of the computational methodology we describe and employ in this chapter establishes the latter as a reliable tool for the investigation of nontrivial electronic structures and magnetic cou- plings.

6.1 Model Hamiltonian and ESR measurements

Mott-Hubbard physics in d-metal compounds has been traditionally associated with first-series (3d) TM oxides. However, recently, one more ingredient entered the TM-oxide ‘Mottness’ paradigm – large SOC’s in 5d systems. SOC in 5d and to some extent 4d anisotropic oxides modifies the very nature of the correlation hole of an electron, by admixing the different t2g components [6, 289], changes the condi- tions for localization [284], the criteria of Mottness and further gives rise to new types of magnetic ground states and excitations [286, 288]. While various measurements indicate that indeed spin-orbit-coupled 284 286 296 jeff ≈1/2 states form in A2IrO4 [ , , ], it has been also pointed out that off-diagonal SOC’s may mix into the GS wavefunction sub- 4 1 297 300 5 4 1 stantial amounts of t2geg character [ – ]. Such t2g–t2geg many- body interactions were shown to produce remarkable effects in X- ray absorption and X-ray magnetic circular dichroism (XMCD): the branching ratio between the L3 and L2 Ir 2p absorption edges reaches values as large as 4, nearly 50 % higher than the 2.75 value for a ‘pure’ 301 cubic jeff = 1/2 system [ ]. Additionally, low-symmetry noncubic 6.1 model hamiltonian and esr measurements 73

35 35 4 g|| 30 30 3 2.31

) ) 2 g⊥=2 4 25 4 25 g⊥ 1 g -factor

10 10 ca 2 ν=270 2 0 20 GHz 20 0 π/6 π/3

ESR intensity (arb. u) 2 4 6 8 α 15 μ0hz (T) 15 (GHz (GHz 2 2 ν ν 10 g||=2 10 5 5 T=4K T=4K 0 0 0 2 4 6 8 10 12 14 0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 μ0h|| (T) μ0h⊥ (T)

(a) Field parallel to the C4 axis (b) Field perpendicular to the C4 axis

Figure 6.2: ESR data for Sr2IrO4. (a) Out-of-plane magnetic-field geome- tries; the inset shows a representative AF resonance spectrum. (b) In-plane magnetic fields; the inset demonstrates the g-factor anisotropy as function of the tetragonal distortion parameter α. Symbols denote experimental data points, solid lines are theo- retical curves obtained using Equations 6.2, 6.3, and 6.4 together with the quantum chemically computed g factors. Dashed lines are obtained by assuming isotropic g factors.

fields produce sizable splittings of the 5d t2g levels, in some cases close to or even larger than 0.5 eV [269, 271] (see also Chapter 7), and 5 289 297 therefore admix the jeff ≈1/2 and jeff ≈3/2 t2g components [ , ]. Obviously, the structure of the spin-orbit GS depends on both the strength and sign of these splittings. Interestingly, the best fits of the X-ray absorption and XMCD data are achieved in Sr2IrO4 with a neg- ative t2g tetragonal splitting [301], although the oxygen octahedra in this material display a distinct positive tetragonal distortion – the IrO6 octahedra are substantially elongated [302] (a negative tetrago- nal splitting should occur when the IrO6 octahedra are compressed [254, 289], see Figure 6.1). The interactions between a pair hiji of NN 1/2 pseudospins in the presence of an external magnetic field ~h is given by the model Hamil- tonian h i X Hi,j ~˜ ~˜ ~ ~˜ ~˜ ~˜ ¯ ~˜ ~ ¯ ~˜ 6 1 mod = JSi · Sj + d · Si × Sj + Si · Γ · Sj + µB h · g¯k · Sk,( . ) k=i,j ~ ~ where S˜i, S˜j are pseudospin (jeff ≈ 1/2) operators, J is the isotropic Heisenberg exchange, ~d=(0, 0, d) defines the antisymmetric DM cou- pling, Γ¯ is a symmetric traceless second-rank tensor describing the symmetric anisotropy and due to the staggered rotation of the IrO6 octahedra the g¯ tensor splits for each of the two sites into uniform u s 298 303 and staggered components g¯k  g¯k (see, e.g. Refs. [ , ]). This model spin Hamiltonian is of direct relevance to the interpretation of the ESR data. For a single crystal of Sr2IrO4 one observes AF resonance modes in the sub-THz frequency domain [304] as displayed in Figure 6.2. 74 anisotropic magnetism in 214 strontium iridate

~ There are two modes if hkz: a gapless Goldstone mode νk1 =0 and a gapped excitation 2 2 2 2 6 2 νk2 = ∆ + 2gkhkeJ/(2(J + Γzz) +eJ), ( . ) q p 2 2 where ∆ = 2eJ(eJ − 2(J + Γzz)), eJ = 4d + (2J − Γzz) and Γzz cou- ˜z ˜z ples the Si and Sj components (along the c axis, perpendicular to the ab-plane IrO2 layers [302]) in the third term of Equation 6.1. Exper- imental results are shown for νk2 in Figure 6.2a. The data comprise at T  TN = 240 K a group of overlapping resonances (Figure 6.2a, in- set), possibly due to some distribution of internal fields in the sample. 2 Though revealing some scatter, the νk2 AF resonance data follow ap- proximately a parabolic dependence on h and, most importantly, they lie substantially above the curve corresponding to the free-electron Landé factor ge = 2 (dashed line in Figure 6.2a). The experimental 2 2 dependence νk2(h ) can be reasonably well modeled with gk values of 2.3–2.45. The solid line in Figure 6.2a is obtained by using gk =2.31 and magnetic exchange couplings as derived from quantum chem- istry calculations that will be discussed in Section 6.2 and Section 6.3. Sizable deviations to values larger than 2 of gk is clear indication of the presence of low-symmetry, noncubic crystal fields. In the sim- 4+ 5 plest approximation, i.e. restricting ourselves to the Ir t2g mani- fold, the anisotropic g factors in tetragonal environment can be ex- g = g = (2+2k) 2α−2 2α pressed up to the sign as√ k c cos sin and 2 g⊥ = ga = gb = 2 sin α+2 2k cos α sin α, where k is a covalency re-  √  duction factor [305], α = (1/2) arctan 2 2λ/(λ−2δ) parameterizes the deviation from octahedral symmetry, λ is the SOC constant and δ the Ir t2g splitting [289]. A plot for the dependence of the diagonal g factors on the distortion parameter α is shown in the inset to Fig- ure 6.2b. For simplicity, k =1 is for the moment assumed but smaller values of k do not bring qualitative changes. In cubic symmetry δ=0, ◦ αcub =35.26 and the g¯ matrix is isotropic with gk =g⊥ =2. According to standard textbooks on ligand-field theory [254, 255, 289, 306], an elongation of the out-of-plane Ir-O bond induces a positive tetragonal splitting of the Ir t2g levels, with δ > 0 and α > αcub, whereas a bond 6 1 compression yields δ < 0 and α < αcub (see Figure . ). As in Sr2IrO4 the IrO6 octahedra are substantially elongated in the z direction [302], the g factors are expected to correspond to the case of positive split- ting α > αcub, see the area to the right of the crossing point shown in the inset to Figure 6.2b. It follows from the plot that gk should be <2, which obviously contradicts the AF resonance data for ~hkz (Fig- ure 6.2a). The value gk = 2.31 used to draw the curve connecting the 6 2 ◦ open circles in Figure . a in fact corresponds to α=32.05 <αcub (see the inset in Figure 6.2b) and indicates that, despite the positive c-axis tetragonal distortion, a counterintuitive negative tetragonal splitting of the 5d t2g levels is present in Sr2IrO4 according to the ESR mea- surements [304]. 6.2 quantum chemistry calculations of g factors 75

It should be noted that there is no Goldstone mode for finite in- plane magnetic fields. The canting angle depends in this case both on the strength of the DM interaction and the applied field. The two modes are 2 2 2 2 0 6 3 ν⊥1 = m g⊥h⊥ + mg⊥h⊥(−4J − Γzz + mJ ) ( . ) and ν2 = m2g2 h2 + mg h (−4J + 5Γ + 3mJ0) ⊥2 ⊥ ⊥ ⊥ ⊥ zz 6 4 0 0 ( . ) + (−4J + 2Γzz + 2mJ )(3Γzz + mJ ) , √ 0 2 where J = −4d 1 − m + 2m(2J − Γzz) and m is the in-plane ferro- magnetic component of the effective moments. To first order in mag- netic field, m can be expressed as s eJ − 2J + Γzz g⊥h⊥ m = − − , (6.5) p 2 2 2eJ 2 6d + (2J − Γzz) which holds for weak fields where m remains small. The first term corresponds to the zero-field canting, arising from the DM interac- tion, while the second term shows how this canting evolves with in- creasing h⊥. Plots based on Equation 6.3 and Equation 6.4 plus the quantum chemically derived interaction parameters (see Section 6.2 and Section 6.3) are displayed in Figure 6.2b together with the ESR data. Details of the derivation of Equations 6.2, 6.3, and 6.4 for the AF resonance modes by use of spin-wave theory is reported in [NB5].

6.2 Quantum chemistry calculations of g factors

The g factors were obtained by computations on clusters which con- tain one central IrO6 octahedron, the four NN IrO6 octahedra and the nearby ten Sr/Ba ions. The solid-state surroundings were mod- eled as an array of point charges fitted to reproduce the crystal Madelung field in the cluster region with help of the ewald tool [72, 73]. To obtain a clear picture on crystal-field effects and spin-orbit interactions at the central Ir site, we cut off the magnetic couplings with the adjacent Ir ions by replacing the tetravalent open-shell d5 4+ 6 NN’s with tetravalent closed-shell Pt t2g species. This is a usual procedure in quantum chemistry studies on TM systems, see e.g. Refs. [104, 184, 307–309]. We used energy-consistent relativistic pseu- dopotentials and valence basis sets of quadruple-zeta quality supple- mented with f polarization functions for the central Ir ion [251] and all-electron triple-zeta basis sets for the six adjacent ligands [225]. For the TM NN’s, we applied energy-consistent relativistic pseudopoten- tials and triple-zeta basis functions [251] along with minimal atomic- natural-orbital basis sets [252] for the O’s coordinating those TM sites 76 anisotropic magnetism in 214 strontium iridate

but not shared with the central octahedron. The Sr and Ba species were modeled by divalent total-ion effective potentials supplemented with a single s function [310]. All O 2p and metal t2g electrons at the central octahedron were correlated in the MRCI calculations [11, 38]. To separate the metal 5d and O 2p valence orbitals into different groups, i.e. central-octahedron and adjacent-octahedra orbitals, we used the Pipek-Mezey localization module [228] available in molpro. The spin-orbit treatment was performed according to the procedure Absence of the described in Ref. [311]. In a first step, the scalar relativistic Hamilto- orbital relaxation or nian is used to calculate correlated wavefunctions for a finite number the wavefunction of low-lying states, either at the CASSCF or MRCI level. In a sec- optimization after taking into account ond step, the spin-orbit part is added to the initial scalar relativistic SOC in principle Hamiltonian, matrix elements of the aforementioned states are evalu- can be a source of ated for this extended Hamiltonian and the resulting matrix is finally error. Variational diagonalized to yield spin-orbit coupled wavefunctions. SOC methods were g proposed [312, 313], Our computational scheme for factors follows the prescription of but their application Bolvin [315] and Vancoillie et al. [316] (for an alternative formulation, is limited to very see also Ref. [317]). It maps the ME’s of the ab initio Zeeman Hamil- small systems. H^ = −~µ^ · ~h = µ (~L^ + g ~S^) · ~h Nevertheless, the tonian Z B e onto the ME’s of the model H^ ~ ~ ~ ~^ ~^ ~ two-step approach pseudospin Hamiltonian S = µBh · g¯ · S˜, where µ^, L, S and S˜ are for SOC is quite magnetic moment, orbital angular-momentum, spin and pseudospin accurate even in the operators, respectively. case of fourth- and  fifth-row elements For the Kramers doublet (KD) GS ψ, ψ¯ one can write the [314]. Abragam-Bleaney tensor G¯ =g¯g¯T [289, 318]. In matrix form it reads X ^ ^ ^ ^ Gkl = 2 hu| Lk + geSk |vi hv| Ll + geSl |ui u,v=ψ,ψ¯ X (6.6) = (Λkm + geΣkm)(Λlm + geΣlm) , m=x,y,z where Λ¯ and Σ¯ tensors are defined by h ^ i h ^ i Λkx =2Re ψ¯ Lk |ψi , Σkx = 2Re ψ¯ Sk |ψi , h ^ i h ^ i 6 7 Λky =2Im ψ¯ Lk |ψi , Σky = 2Im ψ¯ Sk |ψi , ( . ) ^ ^ Λkz =2 hψ| Lk |ψi , Σkz = 2 hψ| Sk |ψi . ^ ^ The ME’s of Lk are provided by molpro while those of Sk can be introduced using the expressions for the generalized Pauli matrices : ^ (Sz)MM0 = MδMM0 , 1p (S^ ) 0 = (S + M)(S − M + 1)δ 0 x MM 2 M−1,M 1p + (S − M)(S + M + 1)δ 0 , 6 8 2 M+1,M ( . ) i p (S^ ) 0 = − (S + M)(S − M + 1)δ 0 y MM 2 M−1,M i p + (S − M)(S + M + 1)δ 0 . 2 M+1,M 6.2 quantum chemistry calculations of g factors 77

casscf mrci states considered g⊥ gk g⊥ gk

Sr2IrO4 : 5 3 t2g ( KD’s) 1.67 2.25 1.60 2.35 5 4 1 3 2 27 t2g, t2geg, t2geg ( KD’s) 1.81 2.27 1.76 2.31

Ba2IrO4 : 5 3 t2g ( KD’s) 2.00 1.61 2.01 1.60 5 4 1 3 2 27 t2g, t2geg, t2geg ( KD’s) 2.09 1.77 2.10 1.76

Table 6.1: g factors for Sr2IrO4 and Ba2IrO4 as obtained on the basis of mol- pro quantum chemistry calculations. The left column displays the electron configurations entering the spin-orbit treatment. Only the 3 2 high-spin sextet state is considered out of the t2geg manifold. Fi- nal results are shown in red.

G¯ is next diagonalized and the g factors are obtained as the positive square roots of the three eigenvalues. The corresponding eigenvectors specify the rotation matrix to the main magnetic axes. In the case of Ir 214 compounds, the magnetic Z axis is along the crystallographic c coordinate, while X and Y are ‘degenerate’ and can be any two per- pendicular directions in the ab plane. It is possible to further obtain the sign of the product gx·gy·gz or even the signs of the individual g factors following the procedure described in Ref. [317]. Results of ab initio quantum chemistry calculations for the g factors in Sr2IrO4 and in the structurally related material Ba2IrO4 are listed in Table 6.1. In a first set of calculations, only the three t2g orbitals at a given Ir site and five electrons were considered in the active space. The self-consistent-field optimization was carried out for the 2 5 corresponding T2g(t2g) state. Inclusion of SOC yields in this case a We use the more set of three KD’s, see Table 6.1. convenient notations O Subsequently we performed calculations with larger active spaces, associated to h 5 2 4 1 2 symmetry, although including also the Ir eg orbitals. One t2g ( T2g) plus four t2geg ( A2g, the calculations were performed for the actual experimental geometry, with Sr2IrO4 Ba2IrO4 point-group sym- casscf nevpt2 casscf nevpt2 metry lower than octahedral. δ, meV −127 −199 30 70

g⊥ 1.66 1.55 1.93 2.01

gk 2.23 2.41 1.74 1.58

Table 6.2: Ir t2g splittings and g factors for Sr2IrO4 and Ba2IrO4 as obtained orca 319 2 5 with the program [ ]. Only the T2g (t2g) states were in- cluded in the CASSCF optimization and in the SO treatment. 78 anisotropic magnetism in 214 strontium iridate

Ir σ σ'

O b y C a 2v x (a) 2D unit cell (b) 2-site bond

Figure 6.3: Planar IrO2 network in Sr2IrO4. (a) Coordination of the Ir site. Dashed lines show the boundaries of the crystallographic unit cell within a given IrO2 layer. (b) The point-group symmetry of the [Ir2O11] block is C2v, associated symmetry elements are indicated in the figure.

2 2 2 4 1 4 T1g, Eg, and T2g) spin doublets, two t2geg spin quartets ( T1g and 4 3 2 6 T2g) and one t2geg spin sextet ( A1g) entered the SO treatment. The orbitals were optimized for an average of all these terms. 4 1 The effect of enlarging the active space to include t2geg terms in the reference wavefunction is in the range of 10 %, in line with ear- lier semi-empirical estimates for 4d5 and 5d5 systems [297–300]. The calculations yield g factors characteristic to negative tetragonal split- tings in Sr2IrO4 and positive t2g splittings in Ba2IrO4 (see Table 6.1). Actual values for the Ir t2g splittings in these two compounds are provided in Table 6.2 and Ref. [320]. To cross-check the g-factor values computed with our subroutine, we further performed g-factor calculations using the module avail- able within the orca quantum chemistry package [319]. We applied all-electron Douglas-Kroll-Hess (DKH)-optimized basis sets of triple- zeta quality for the TM ions [321], triple-zeta basis functions for the ligands of the central octahedron [225] and double-zeta basis func- tions for additional O’s at the NN octahedra [225]. Dynamical corre- lation effects were accounted for by NEVPT2 [46, 322]. CASSCF and NEVPT2 results are listed in Table 6.2, for both Sr2IrO4 and Ba2IrO4. It is seen that the data in Table 6.1 and Table 6.2 compare very well and indicate the same overall trends.

6.3 Anisotropic exchange couplings in Sr2IrO4

To obtain ab initio quantum chemistry values for the intersite effec- tive magnetic couplings in introduced in Equation 6.1, we carried out additional calculations on larger clusters that incorporate two IrO6 octahedra as magnetically active units. Similarly to the calculation of single-site magnetic properties, to accurately describe the charge distribution in the immediate neigh- 6.3 anisotropic exchange couplings in Sr2 IrO4 79

Hi,j |si |t i |t i |t˜ i ab initio ˜ x y z | hs˜ 0 0.2308i µBhy −0.1768i µBhx 0 | htx −0.2308i µBhy 48.3328 2.3083i µBhz −1.6854i µBhy | hty 0.1768i µBhx −2.3083i µBhz 48.9626 1.6266i µBhx | ht˜z 0 1.6854i µBhy −1.6266i µBhx 49.0630

Table 6.3: Matrix elements of the ab initio spin-orbit Hamiltonian for Sr2IrO4, as described by Equation 6.9. Results of spin-orbit MRCI calcula- tions are shown (meV).

2+ bourhood, the adjacent IrO6 octahedra (6) and the closest Sr ions (16) were also incorporated in the actual cluster. We used energy- consistent relativistic pseudopotentials along with quadruple-zeta basis sets for the valence shells of the two magnetically active Ir ions [251], all-electron quintuple-zeta basis sets for the bridging lig- and [225] and triple-zeta basis functions for the other O’s associated with the two reference octahedra [225]. We further employed polar- ization functions at the two central Ir sites and for the bridging anion, namely two Ir f and four O d functions [225, 251]. Additional ions defining the NN octahedra, the nearby Sr2+ species and the farther crystalline surroundings were modeled as in the single-site study, see Section 6.2. 2 5 2 5 For two adjacent magnetic sites, the T2g(t2g)⊗ T2g(t2g) manifold entails nine singlet and nine triplet states. The CASSCF optimiza- tion was carried out for an average of these nine singlet and nine triplet eigenfunctions of the scalar-relativistic Hamiltonian. In the sub- sequent MRCI treatment, only the Ir t2g and the O 2p electrons at the bridging ligand site were correlated. Results in good agreement with the experimental data were recently obtained with this computational approach for related 5d5 iridates [7], see also Chapter 7.

Hi,j |si |t i |t i |t˜ i mod ˜ x y z     Je Je gyyd−gxy /2+Jx+y gxxd+gxy /2+Jx+y hs˜| 0 ihy r −ihx r 0 J  J  eJ e/2+Jx+y eJ e/2+Jx+y J  J  (gyyd−gxy e/ +Jx+y   gyyd+gxy e/ −Jx+y | 2 1 eJ 2 htx −ihy r 2 /2 + Jx+y − Γx−y + Γz −igzzhz −ihy r J  J  eJ e/2+Jx+y eJ e/2−Jx+y J  J  gxxd+gxy e/ +Jx+y   gxxd−gxy e/ −Jx+y | 2 1 eJ 2 hty ihx r igzzhz 2 /2 + Jx+y + Γx−y + Γz ihx r J  J  eJ e/2+Jx+y eJ e/2−Jx+y J  J  gyyd+gxy e/ −Jx+y gxxd−gxy e/ −Jx+y 2 2 eJ ht˜ z| 0 ihy r −ihx r /2 J  J  eJ e/2−Jx+y eJ e/2−Jx+y

Table 6.4: Matrix elements of the model spin Hamiltonian given by Equa- tion 6.1. For additional notations used here, see Table 6.5. 80 anisotropic magnetism in 214 strontium iridate

J, Γxx =−Γyy−Γzz Jx, Jy, Jz J, δJxy, δJz (x bond) (Table 6.4) (Ref. [323]) (Refs. [6, 323, 324])

Γ (J +J ) δJ Jx+y J − zz/2 x y /2 J + xy/2 q q q J 2 d2 + J2 2 d2 + J2 2 d2 + J2 e x+y x+y  x+y | H | (Γ −Γ ) (J −J ) δJxy for htx txi Γx−y xx yy /2 x y /2 0 for hty| H |tyi 3 Γz /2Γzz Jz − Jx+y δJz

Table 6.5: Notations used to define the anisotropic exchange couplings in Sr2IrO4. Other conventions employed in the literature are shown for comparison.

By one-to-one correspondence between the ME’s of the ab initio Hamiltonian

Hi,j = Hi,j + Hi,j + Hi,j 6 9 ab initio SR SO Z ( . )

and the ME’s of the model spin Hamiltonian (Equation 6.1) in the basis of the lowest four spin-orbit states defining the magnetic spec- trum of two NN octahedra, we can derive in addition to the g factors the strengths of the Heisenberg and anisotropic intersite couplings. In 6 9 Hi,j Equation . , SR is the scalar-relativistic Born-Oppenheimer Hamil- Hi,j 311 Hi,j tonian, SO describes spin-orbit interactions [ ] and Z is the two- site Zeeman Hamiltonian. Hi,j Hi,j Diagonalization of the Hamiltonian SR + SO in the basis of the lowest nine singlet and nine triplet states provides a total of 36 spin- i j 1 1 orbit-coupled S˜ , S˜ eigenfunctions, namely, four /2, /2 , eight 1 3 3 1 3 3 /2, /2 , eight /2, /2 and sixteen /2, /2 states. In the simplest 1 1 picture, the lowest four /2, /2 roots imply either singlet or triplet coupling of the spin-orbit jeff ≈ 1/2 on-site objects and are separated ‡ ‡ It is the jeff ≈1/2 from higher-lying states by a gap of ≈ 0.5 eV , much larger than the to jeff ≈3/2 excito- strength of the intersite exchange. We use here the more general pseu- nic gap [284, 286]. dospin S˜ = 1/2 notation, because the jeff = 1/2 picture does not hold in a strict sense due to tetragonal distortions and interaction with ex- 4 1 cited-state t2geg configurations. It is this set of lowest four spin-orbit MRCI roots that we map onto the eigenstates of the model two-site (pseudo)spin Hamiltonian (6.1). The Zeeman interaction shows up on the quantum chemistry side Hi,j ~^ ~ ~^i,j ~^i,j ~ ~^i,j ~^i,j as Z = −µ · h = µB(L + geS ) · h, where L and S are trans- formed to the spin-orbit-coupled basis using the spin-orbit wavefunc- tions as unitary transformation matrix. ME’s of the ab initio Hamilto- nian given in Equation 6.9 are shown in Table 6.3. Diagonal compo- nents show the energies of the zero-field states, while the off-diagonal ME’s describe the coupling to magnetic field. 6.3 anisotropic exchange couplings in Sr2 IrO4 81

Hi,j 1 1 1 1 mod 2 , 2 2 , − 2

1 1 J+Γzz hx(gxx+igxy)−hy(igyy+gxy) 2 , 2 4 + gzzhz 2 1 1 hx(gxx−igxy)−hy(−igyy+gxy) J+Γzz 2 , − 2 2 − 4 1 1 hx(gxx+igxy)−hy(−igyy−gxy) 2J−2iD−Γzz − 2 , 2 2 4 1 1 2Γyy+Γzz hx(gxx+igxy)−hy(−igyy−gxy) − 2 , − 2 − 4 2 Hi,j 1 1 1 1 mod − 2 , 2 − 2 , − 2

1 1 hx(gxx−igxy)−hy(igyy−gxy) 2Γyy+Γzz 2 , 2 2 − 4 1 1 2J+2iD−Γzz hx(gxx−igxy)−hy(igyy−gxy) 2 , − 2 4 2 1 1 J+Γzz hx(gxx+igxy)−hy(gyy+gxy) − 2 , 2 − 4 2 1 1 hx(gxx−igxy)−hy(−igyy+gxy) J+Γzz − 2 , − 2 2 4 − gzzhz

Table 6.6: Matrix form of the model spin Hamiltonian in the ‘uncoupled’ E ˜z ˜z bond basis Si , Sj . The zero-field couplings are shown in red.

For the experimentally determined crystal structure of Sr2IrO4 [302], the two-octahedra [Ir2O11] cluster displays C2v symmetry, see Figure 6.3b. Having the x axis along the h110i crystallographic di- rection [302] and z k c, the effective anisotropic couplings read ~d = [0, 0, d],   Γxx 0 0 (x)   Γ¯ = 0 Γ 0  (6.10)  yy  0 0 Γzz for Ir-Ir links along x and   Γyy 0 0 (y)   Γ¯ = 0 Γ 0  (6.11)  xx  0 0 Γzz for Ir-Ir links along y, with Γxx + Γyy + Γzz = 0. The uniform and staggered components of the g¯ tensor take for individual Ir sites the following form :     gxx 0 0 0 gxy 0 u s     g¯i = g¯  g¯ = 0 g 0  g 0 0  .(6.12) i i  yy   xy  0 0 gzz 0 0 0 gzz =gk while gxx and gyy are directly related to g⊥ but not restricted to be equal due to the lower symmetry of the two-octahedra cluster as compared to the IrO6 unit discussed in Section 6.2. To perform now the actual mapping, we need to transform the model spin Hamiltonian given by Equation 6.1 to the same form 82 anisotropic magnetism in 214 strontium iridate

Hi,j | | | | mod 0, 0i 1, −1i 1, 1i 1, 0i g (h −ih ) g (−ih −h ) h0 0| − 3 J xy √y x xy √ x y i D , 4 2 2 2 g (ih +h ) Γ g h +ig h h1 −1| xy √x y J + Γzz − 4g h − yy − Γzz xx x√ yy y , 2 4 4 zz z 2 4 2 g (ih −h ) Γ g h −ig h h1 1| xy √x y − yy − Γzz J + Γzz + 4g h xx x√ yy y , 2 2 4 4 4 zz z 2 g h −ig h g h +ig h h1 0| − i D xx x√ yy y xx x√ yy y J − Γzz , 2 2 2 4 4

Table 6.7: Matrix elements of the model spin Hamiltonian in the coupled ˜ ˜z basis Stot, Stot . The zero-field couplings are shown in red.

Hi,j | | | | mod si txi tyi tzi | 3 i hs − 4 J −gxyhy −gxyhx 2 D | J Γyy Γzz htx −gxyhy 4 + 2 + 2 −igzzhz −igyyhy | J Γyy hty gxyhx igzzhz 4 − 2 igxxhx | i J Γzz htz − 2 D igyyhy −igxxhx 4 − 2

Table 6.8: Matrix form of the model spin Hamiltonian with the isotropic and symmetric anisotropic exchange brought to diagonal form. The zero-field couplings are shown in red.

as the ab initio Hamiltonian shown in Table 6.3, i.e. diagonal in zero magnetic field. The result of such a transformation is shown in Ta- ble 6.4, while the additional parameters introduced there are defined in Table 6.5. Diagonalization of the spin Hamiltonian provides the expected singlet |s˜i and three (split) triplet components |txi, |tyi and |t˜zi. Due to the DM interaction, |s˜i and |t˜zi are admixtures of ‘pure’ |0, 0i and |1, 0i spin functions. To better illustrate the nature of these {|s˜i , |txi , |tyi , |t˜zi} states forming the basis of the effective Hamiltonian shown in Table 6.4 and the different steps followed to arrive to that particular matrix form, we also provide in this section ME’s in (i) the ‘uncoupled’ E ˜z ˜z 6 6 ˜ ˜z Si , Sj basis (see Table . ), (ii) the ‘coupled’ Stot, Stot basis in which the isotropic exchange is diagonal (Table 6.7) and (iii) the basis {|si , |txi , |tyi , |tzi} that also diagonalizes the symmetric anisotropic exchange (Table 6.8). In the latter case, |si = |0, 0i and |tzi = |1, 0i while |txi and |tyi are antisymmetric and symmetric combinations of |1, −1i and |1, 1i. The unitary matrix describing the transformation to the form given in Table 6.8 reads   1 0 0 0   0 − √1 − √i 0  2 2  6 13 UΓ =  . ( . ) 0 √1 − √i 0  2 2  0 0 0 1 6.3 anisotropic exchange couplings in Sr2 IrO4 83

Finally, |si and |tzi are to some degree admixed due to the antisym- metric exchange. The resulting states |s˜i and |t˜zi imply a ‘rescaling’ of the Zeeman couplings. The Hamiltonian thus takes the final form shown in Table 6.4. The corresponding transformation matrix reads

 iD iD  − q 0 0 q ˜ J˜ ˜ J˜  J( /2−Jx+y) J( /2+Jx+y)    0 1 0 0    6 14 UD =  . ( . )  0 0 1 0   r r   J˜ J˜  /2−Jx+y 0 0 /2+Jx+y J˜ J˜

Direct correspondence between homologous ME’s in the two ma- trices yields a set of eight independent equations that finally allow to derive hard values for all coupling constants that enter Equation 6.1 (with symmetry constrains given by Equations 6.10, 6.11 and 6.12). These values of the effective couplings are shown in Table 6.9. The procedure described in this chapter resembles the effective Hamiltonian extraction scheme [259, 266] employed in Chapter 5, where we also compared ME’s of model and ab initio Hamiltonians. The difference is that in Chapter 5 we transformed the diagonal ab ini- tio Hamiltonian to the basis of the model with a special unitary trans- formation. Such transformation can be straightforwardly constructed in case of second-order SOC, because the basis of the model span just the pure-spin states. But when the first-order SO interaction plays the 5 major role, as in 5d iridates, the basis of the effective model, i.e. S˜i, can not be easily found. The model Hamiltonian can be transformed to the form of the ab initio Hamiltonian by diagonalization in zero magnetic field, as it was done in previous studies [7, 292]. However, in the case of C2v symmetry number of diagonal ME’s is not enough to perform the mapping to the effective model. We have to perform supplemental calculation of the Zeeman Hamiltonian both at ab initio and model level to have additional ME’s to compare. Our mapping procedure yields J≈48 meV, somewhat lower than J values of 55–60 meV derived from experiment [286, 325], and a ratio between the DM and Heisenberg couplings d/J = 0.25, in agreement with estimates based on effective superexchange models [6, 323, 324] and large enough to explain the nearly rigid rotation of magnetic mo- ments that is observed when the IrO6 octahedra revolve [326, 327]. Ab initio results for the NN anisotropic couplings Γ¯, also relevant for a detailed understanding of the magnetic properties of Sr2IrO4, are shown as well in Table 6.9. In our convention the x axis is taken along the Ir-Ir link, i.e. it coincides with the h110i crystallographic direction [302], and z k c. We obtain Γxx ≈ Γzz, which then allows to recast the Heisenberg and symmetric anisotropic terms in (6.1) as ~˜ ~˜ ˜y ˜y (J − Γyy/2) Si · Sj + (3Γyy/2) Si Sj , with Γxx = Γzz = −Γyy/2. Equally interesting, for no rotation of the IrO6 octahedra and straight Ir-O-Ir 84 anisotropic magnetism in 214 strontium iridate

magnetic couplings, meV g factors

J d Γxx Γyy Γzz gxx gyy gzz gxy

47.8 11.9 0.42 −0.84 0.42 1.64 1.70 2.31 0.02

Table 6.9: Nearest-neighbour magnetic couplings and g-factors in Sr2IrO4 as obtained from the spin-orbit MRCI calculations on two-octahedra clusters. Γxx+Γyy+Γzz =0 since Γ¯ is traceless.

bonds in Ba2IrO4, it is Γyy and Γzz which are approximately the same, providing a realization of the compass-Heisenberg model [292, 323] since the DM coupling is by symmetry zero in that case. The two-site magnetic Hamiltonian (6.1) features in-plane symmetric-anisotropy couplings Γxx and Γyy which were not con- sidered in earlier studies [6, 304]. In the presence of two-sublattice order, terms containing these couplings cancel each other in the mean-field energy but they are in general relevant for pseudospin fluctuations and excitations. Using spin-wave theory and effective parameters derived from the quantum chemistry calculations, we nicely reproduce the correct GS and character of the modes, as shown in Figure 6.2. To reproduce the experimental zero-field gap, in particular, we used J, d and g-factor values as listed in Table 6.9 and a somewhat larger Γzz parameter of 0.98 meV. To leading order, 2 2 6 3 the dependence of ν⊥1 and ν⊥2 on h is linear, see Equation . and Equation 6.4 The slope is proportional to m, which strongly depends on the DM coupling. At low fields (61 T), m can be actually replaced with its field independent value derived in Ref. [304]. Using the MRCI coupling constants, the first term in Equation 6.5 then yields a moment m≈0.12µB, in good agreement with the experiment [285, 328].

6.4 Conclusions

In this chapter we have provided an integrated picture on the d-level structure and magnetic anisotropies in Sr2IrO4, a prototype spin-orbit driven magnetic insulator. Both the single-site g¯ tensor and intersite effective exchange interactions are analyzed in detail. To access the lat- ter, we build on an earlier computational scheme for deriving intersite matrix elements in mixed-valence spin-orbit coupled systems [329]. While the ratio d/J of the antisymmetric Dzyaloshinskii-Moriya and isotropic Heisenberg couplings is remarkably large in Sr2IrO4 and concurs with an in-plane rotation pattern of the Ir magnetic moments that follows nearly rigidly the staggered rotation of the IrO6 octahe- dra [326, 327], the most prominent symmetric anisotropic terms are according to the quantum chemistry data in-plane, perpendicular to 6.4 conclusions 85 the Ir-Ir links. The structure of the g¯ tensor, as computed with first- principles electronic-structure methods, consistent with ESR measure- ments [304] and such that gk = gc > g⊥ = ga = gb. It distinctly indi- cates a negative tetragonal splitting of the Ir t2g levels, in spite of sizable positive local distortions in Sr2IrO4.

CaIrO3 POST-PEROVSKITE,A je f f ≈ 1/2 7 QUASI-1D ANTIFERROMAGNET

ridium oxide compounds are at the heart of intensive experimen- tal and theoretical investigations in solid state physics. The few different structural varieties displaying Ir ions in octahedral co- I 5 ordination and tetravalent 5d valence states, in particular, have re- cently become a fertile ground for studies of new physics driven by the interplay between strong spin-orbit interactions and electron cor- relations [241]. In the simplest picture, the SOC’s are sufficiently large 5 to split the Ir t2g manifold into well separated jeff ≈3/2 and jeff ≈1/2 states. The former are fully occupied while the latter are only half ∼ filled. It appears that the widths of the jeff ≈1/2 bands, w N·t, where t is an effective Ir-Ir hopping integral and N is the coordination num- ber, are comparable to the strength of the on-site Coulomb repulsion U. For smaller coordination numbers, such as N=3 in the layered hon- eycomb systems Li2IrO3 and Na2IrO3 and N = 4 for square-lattice Sr2IrO4 and Ba2IrO4 and the post-perovskite CaIrO3, the undoped compounds are insulating at all temperatures and a jeff ≈ 1/2 Mott- Hubbard type picture seems to be appropriate [241, 330, 331]. For larger coordination numbers like N = 6 in the pyrochlore structure, 5 the t2g Ir oxides are metallic at high T’s and display transitions to AF insulating states on cooling [332]. Furthermore, a Slater type picture in which the gap opening is intimately related to the onset of AF or- der might be more appropriate, as proposed for example for Os oxide perovskite systems with N = 6 [247]. Also, more exotic ground states such as a Mott topological insulator [242] or a Weyl semimetal [268] have been proposed for the pyrochlore iridates. Besides the SOC’s and Coulomb interactions, one important addi- tional energy scale is the magnitude of the CF splittings within the Ir t2g shell. Such splittings may arise from distortions of the O octa- hedron around the Ir site and also from the anisotropy of the farther surroundings [271, 320], see also Chapter 6. In this respect, highly anisotropic systems such as the post-perovskite CaIrO3 provide an ideal playground for new insights into the interplay between SOC’s and lattice distortions. In the layered post-perovskite structure, IrO6 octahedra in the ac layers share corners along the crystallographic c axis, see Figure 7.1a, and have common edges along a [333, 334]. The local symmetry at the Ir sites is described by the monoclinic C2h point group, with the C2 axis along a. The Ir-O bonds along the corner-sharing chains are substantially shorter than for the edge- sharing links, 1.94 vs. 2.07 Å[333] (see Figure 7.1a), which suggests

87 88 post perovskite a 1 2 quasi 1d antiferromagnet CaIrO3 - , jeff ≈ / -

y Ca 4.8˚ z x Ir O b 1.94Å 2.07Å 80.8˚ a c

(a) Crystal structure of CaIrO3 (b) IrO6 octahedron

Figure 7.1: Structural details of CaIrO3. (a) Layered crystal structure and the global crystalline axes. (b) The IrO6 octahedron belongs to the C2h point group. Local axes for the IrO6 octahedron and distortions away from C4h symmetry are illustrated.

a tetragonal-like stabilization of the xy-like orbitals with respect to the xz/yz-like, as illustrated in Figure 6.1. Additionally, small split- tings between the xz/yz-like components arise due to the a-b crystal anisotropy. The precise nature of the relativistic 5d5 ground-state wavefunction obviously depends on the size of the Ir t2g splittings. Further, the on- site 5d5 electron configuration and the kind of Ir-O-Ir path along a given direction determine the sign and magnitude of the magnetic ex- change couplings. Ab initio calculations where all these effects and in- teractions are treated on equal footing are therefore highly desirable. The post-perovskite phase of CaIrO3 (space group 63, Cmcm) has received significant attention from geologists in recent years because the same crystalline structure is adopted by MgSiO3 in the lower Earth’s mantle [335, 336]. In contrast to MgSiO3, the post-perovskite phase of CaIrO3 can be synthesized even at ambient pressure. Since measurements are less complicated in ambient conditions, CaIrO3 has been used extensively as an analogue of MgSiO3 post-perovskite for structural, mechanical, and transport studies [337–340]. To unveil the valence electronic structure of CaIrO3 we carried out ab initio wavefunction quantum chemical calculations which are dis- cussed in detail in this chapter. With no SOC’s, splittings as large 5 as 0.8 eV are found for the t2g levels of CaIrO3. The 5d relativistic ground-state wavefunction therefore displays a highly uneven admix- ture of xy, xz, and yz character, which contradicts the interpretation of recent resonant X-ray diffraction (RXD) experiments [341]. For the corner-sharing octahedra, the quantum chemical calculations further predict large AF superexchange couplings [342] of 120 meV, about the same size with the magnetic interactions in the parent compounds of the high-Tc cuprate superconductors. The lowest d-d SO excited states are two doublets at approximately 0.7 and 1.4 eV, originating from the jeff = 3/2 quartet in an ideal cubic environment. This is the 7.1 d–d splittings and magnetic g factors 89

5 largest splitting reported so far for the jeff ≈ 3/2 states in d systems and singles out CaIrO3 as an unique material in which both the SOC and the non-cubic crystal field splittings are of the order of half eV or larger, leading to strong deviations from the simple cubic picture with pure jeff =1/2 doublet and jeff =3/2 quartet states.

7.1 d–d splittings and magnetic g factors

To investigate the local Ir d-level splittings, CASSCF and MRCI calculations [38] were performed on embedded clusters made of one reference IrO6 octahedron, ten adjacent Ca sites, and four NN IrO6 octahedra. The farther solid-state environment was modeled as a one-electron effective potential which in an ionic picture repro- duces the Madelung field in the cluster region [72]. All calculations were performed with the molpro quantum chemical software [116]. We employed energy-consistent relativistic pseudopotentials for Ir [251] and Ca [310] along with Gaussian-type valence basis functions. Quadruple-zeta basis sets from the molpro library were applied for the valence shells of the central Ir ion [251] and triple-zeta basis sets for the ligands [225] of the central octahedron and the nearest-neigbor Ir sites [251]. For the central Ir ion we also used two polarization f functions [251]. For farther ligands in our clusters we applied mini- mal atomic-natural-orbital basis sets [252]. All occupied shells of the Ca2+ ions were incorporated in the large-core pseudopotentials and each of the Ca 4s orbitals was described by a single contracted Gaus- sian function [310]. For the ground-state calculations, the orbitals within each finite cluster were variationally optimized at the CASSCF level [38]. All Ir t2g functions were included in the active orbital space. Excitations within the Ir t2g shell were afterwards computed just for the central IrO6 octahedron while the occupation of the NN Ir valence shells was held frozen as in the CASSCF ground-state configuration. The subsequent MRCI treatment includes all single and double exci- tations from the O 2p orbitals at the central octahedron and the Ir 5d functions [38]. To partition the O 2p valence orbitals into two differ- ent groups, i.e. at sites of the central octahedron and at NN octahedra, we employed the Pipek-Mezey orbital localization module available with molpro [228]. Spin-orbit interactions were accounted following the procedure described in Ref. [311]. In a first step, the scalar rela- tivistic Hamiltonian is used to calculate correlated wavefunctions for a finite number of low-lying states, either at the CASSCF or MRCI level. In a second step, the SO part is added to the initial scalar rel- ativistic Hamiltonian, matrix elements of the aforementioned states are evaluated for this extended Hamiltonian and the resulting matrix is finally diagonalized to yield SO coupled wavefunctions. Crystallo- graphic data as reported by Rodi and Babel [333] was used. 90 post perovskite a 1 2 quasi 1d antiferromagnet CaIrO3 - , jeff ≈ / -

hole (spin-)orbital casscf mrci

xzf 0.00 0.00 yzf 0.13 0.15 xy 0.76 0.83

|1/2i (KD1) 0.00–0.12 0.00–0.16 |3/2, αi (KD2) 0.62–0.70 0.66–0.76 |3/2, βi (KD3) 1.25–1.27 1.36–1.39

7 1 5 Table . : Relative energies of t2g states in CaIrO3 (eV). CASSCF and MRCI results with and without SOC’s for 5-plaquette clusters, see text. KD stands for Kramers doublet.

It is convenient to define local coordination framework for the IrO6 octahedron consistently with the more studied C4h symmetry, as shown in Figure 7.1b. The actual C2h octahedron has the follow- ing distortions away from tetragonal case: (i) O-Ir-O angles within ◦ the ‘basal’ IrO4 plaquette are not 90 , with the smaller one equals to ◦ 80.8 , (ii) the plane defined by the IrO4 plaquette is not perpendicular to axis defined by the Ir and two ‘apical’ O ions. Therefore the 4.8◦ an- gle between y axis and one of the Ir-O bond shown in Figure 7.1b has components both within xy plane and perpendicular to it. A draw- back of using local axes consistent with the tetragonal symmetry is that the xz and yz orbitals defined in such framework do not trans- form like irreducible representations in the C2h point group.√ To fix it and keep the√ notation short, we introduce orbitals xzf =1/ 2(xz+yz) and yzf =1/ 2(xz−yz) that transform like irreducible representations in the C2h point group. 5 With no SOC’s, we find that for the lowest-energy t2g state, the hole occupies the Ir xzf orbital. Excited states with holes in the yzf and xy orbitals have relative energies of 0.13 and 0.76 eV by CASSCF calculations and slightly larger values of 0.15 and 0.83 eV at the MRCI level (see the first lines in Table 7.1). In electron picture, the lowest t2g level is xy, while the splittings are 0.63, 0.76 eV by CASSCF and 0.68, 0.83 eV by MRCI, close to the situation shown in Figure 6.1a. The spin- orbit ground-state wavefunction therefore has dominant xzf character, as much as 75% by MRCI calculations with SOC (MRCI+SOC). The four components of the jeff ≈ 3/2 quartet are split into groups of two 7 1 by 0.6–0.7 eV (lower lines of Table . ). The lower two jeff ≈3/2 terms have 72% yzf character, the highest jeff ≈3/2 components have 90% xy character. It would be therefore more correct to label these SO states as KD1, KD2 and KD3, as it was done in Ref. [343]. When SO interactions are not accounted for, the coupling to spin moments at NN Ir sites gives rise for each particular orbital occu- pation to one sextet, four quartet, and five doublet states. All those 7.1 d–d splittings and magnetic g factors 91 configuration state functions entered our spin-orbit calculations. 32 different SO states are therefore generated for each KD ‘configura- tion’ at the central Ir site and this causes a finite energy spread in Table 7.1. A simpler and more transparent picture can be obtained by 4+ 5 4+ 6 replacing the four Ir t2g NN’s by closed-shell Pt t2g ions. In that case, the MRCI excitation energies for the jeff ≈ 3/2 states (KD2 and KD3) at the central Ir site are 0.67 and 1.38 eV. An instructive com- putational experiment which we can further make is to replace the distorted post-perovskite structure by a hypothetical cubic-perovskite lattice in which the Ir-O bond length is the average of the three dif- ferent bond lengths in post-perovskite CaIrO3, i.e. 2.03 Å[333]. This allows to directly extract the effective SOC strength λ for the Ir4+ 5 H 5d ion ( SOC = λLeff·S) because the splitting between the spin-orbit jeff = 1/2 doublet and the four-fold degenerate jeff = 3/2 state is 3λ/2 in cubic symmetry [289]. The doublet-quartet splitting comes out as 0.70 eV by MRCI+SOC calculations for the idealized cubic perovskite structure. This yields an effective SO coupling constant λ=0.47 eV, in agreement with values of 0.39–0.49 eV earlier extracted for Ir4+ ions from electron spin resonance measurements [299]. Using the methodology explained in Section 6.2 we are also able to calculate the magnetic g¯ tensor for CaIrO3. For computing the ME’s of the magnetic moment operator we employed MRCI calcu- lations including the SO treatment with all 5d5 states with excitation 2 5 4 4 energy less than 5.5 eV. Those states are the T1g (t2g), T2g and T1g 4 1 6 3 2 (t2geg) plus A1g (t2geg) multiplets (split in the actual C2h symme- try), which give rise to 36 SO-coupled states. The calculated g¯ ten- sor in the local {x, y, z} and crystallographic {a, b, c} axes frameworks reads     1.14 −0.86 −0.20 0.28 0 0 {x y z}   {a b c}   g¯ , , =−0.86 1.14 0.20  ; g¯ , , = 0 1.98 0.20 .     −0.20 0.20 3.31 0 0.20 3.34 (7.1)

{ } It is seen that the tensor g¯ x,y,z has diagonal part typical for tetragonally-compressed octahedra, but there are also additional off- diagonal ME’s due to the monoclinic C2h symmetry. The calculated, strongly anisotropic, g¯ tensor should be useful for the interpretation of future ESR measurements on CaIrO3. The results of our ab initio calculations, i.e. the unusually large t2g splittings and the highly uneven admixture of xzf, yzf, and xy charac- 5 ter for the SO t2g states, are in sharp contrast to the interpretation of recent RXD experiments on CaIrO3 [341]. In particular, the au- thors of Ref. [341] conclude from the RXD data that the deviations from an isotropic xz-yz-xy picture are marginal, as also found for in- stance in the layered square-lattice material Sr2IrO4 [241]. While for 92 post perovskite a 1 2 quasi 1d antiferromagnet CaIrO3 - , jeff ≈ / -

Sr2IrO4 one finds indeed by QC calculations rather small t2g split- tings of 0.15 eV, comparable weights of the xz, yz, xy components in 264 320 the jeff ≈ 1/2 ground-state wavefunction [ , ] (see also Chap- ter 6), and NN superexchange interactions in good agreement with RIXS data, for the post-perovskite CaIrO3 the quantum chemical re- sults point in a different direction. These conflicting findings moti- vated more recently Ir L-edge RIXS measurements on CaIrO3 [343], which turned out to be in perfect agreement with our QC results.

7.2 Intersite exchange interactions

The presence of both corner-sharing and edge-sharing octahedra and the large splittings within the t2g shell have important implications on the nature and the magnitude of the magnetic interactions. To de- termine the latter, we designed 8-octahedra clusters including two ac- 4+ 5 tive Ir t2g sites. To make the analysis of the intersite spin couplings 4+ 5 straightforward, the six Ir t2g NN’s were modeled as closed-shell 4+ 6 Pt t2g ions. The basis functions used here were as described in Section 7.1. The only exception was the O ligand(s) bridging the two magnetically active Ir sites, for which we employed quintuple-zeta valence basis sets and four polarization d functions [225]. All possi- ble occupations were allowed within the set of t2g orbitals at the two active Ir sites in the CASSCF calculations, which gives rise to nine S = 0 and nine S = 1 states with no SOC. The CASSCF wavefunctions were expressed in terms of orbitals optimized for an average of those singlet and triplet states. All eighteen states entered the spin-orbit calculations, both at the CASSCF and MRCI levels. In MRCI, single and double excitations from the Ir t2g orbitals and the 2p orbitals of the bridging ligand(s) were included. Such a computational scheme

Figure 7.2: Local coordination and the main, π-type 5d -2p -5d , AF su- xzf xe xzf perexchange path for corner-sharing IrO6 octahedra along the c axis in CaIrO . The O 2p tails of the active Ir d orbitals are 3 xzf overlapping at the bridging O site. 7.2 intersite exchange interactions 93

state casscf mrci mrci+soc

Corner-sharing octahedra along c: Singlet 0.0 0.0 0.0 Triplet 91.2 170.5 124.8, 124.9, 127.0 Edge-sharing octahedra along a: Triplet 0.0 0.0 0.0, 0.40, 0.63, 7.2 Singlet 2.3 −5.1

Table 7.2: Relative energies (meV) of the singlet and triplet states for two 5 adjacent Ir t2g sites along the c and a directions in CaIrO3. provides J values in very good agreement with RIXS measurements in the layered iridate Sr2IrO4 [7, 264]. Relative energies for the low- est Stot = 0 and jtot = 0 singlet and Stot = 1 and jtot = 1 triplet states for both corner-sharing and edge-sharing octahedra are listed in Ta- ble 7.2. In the spin-orbit calculations, the triplet components display small splittings that indicate the presence of non-Heisenberg terms. Remarkably, with no SOC’s, the NN AF exchange constant J(c) for octahedra sharing corners along the c axis is as large as 91.2 meV by CASSCF calculations and 170.5 meV by MRCI, see Table 7.2. The t2g hole orbitals are here the xzf components, as discussed in the previ- ous section. Although the IrO6 octahedra are tilted somewhat due to rotations about the a axis, see Figure 7.2, this does not affect much the π 5d 2p 5d 342 -type xzf - xe- xzf overlap and superexchange interactions [ ]. An AF J(c) of 170.5 meV is even larger than the values for σ-type superexchange paths in two-dimensional S = 1/2 3d Cu oxide super- conductors [344] and points to the role of the larger spatial extent of 5d functions in iridates. With SOC’s accounted for, the effective AF coupling J(c) can be roughly estimated to be ≈ 126 meV (using the last column in Table 7.2). The spin-orbit ground-state wavefunction now acquires sizable yzf character and since the tilting of the IrO6 5d 2p 5d octahedra does affect the out-of-plane yzf - ye - yzf superexchange path, the effective magnetic coupling between the jeff ≈ 1/2 sites is somewhat reduced. The case of the edge-sharing octahedra along the a axis is quite dif- ferent. The Heisenberg exchange coupling changes from weak and FM by CASSCF to weak and AF by MRCI, and finally evolve to anisotropic exchange by MRCI+SOC, see Table 7.2. Given the struc- ture of the two-site magnetic spectrum, with three lower-energy states spread over a narrow window of ≈ 0.6 meV and a higher-lying state with an excitation energy of several meV, one may be tempted to in- terpret the data in terms of a dominant FM Heisenberg coupling and much weaker additional anisotropic magnetic interactions. In the fol- 94 post perovskite a 1 2 quasi 1d antiferromagnet CaIrO3 - , jeff ≈ / -

lowing, we carry out a detailed analysis and show that the situation is actually somewhat more subtle. More accurate values for J(c) and J(a) together with all relevant anisotropic couplings can be obtained with help of the mapping procedure described in Section 6.3. One important detail is that anisotropic exchange couplings allow the eigenstates of the isotropic model to interact, e.g. the DM coupling allows mixing of the singlet and triplet states, while the symmetric anisotropic exchange Γ¯ causes mixing of the different triplet components. Such mixing effects of the zero-field eigenstates modify their interaction with external magnetic field. We can therefore trace back the values of all zero-field-model couplings via one-to-one correspondence between the ME’s of the model spin Hamiltonian and the ab initio Hamiltonian. The latter is assosiated with the quantum-chemically calculated energies of the four low-lying ‘magnetic’ states and with the ME’s of the Zeeman H ~ ~ ~ Hamiltonian Z = µB(L + geS) · h. We employ a model Hamiltonian quadratic in (pseudo)spin operators to describe the interactions be- tween a pair hiji of NN 1/2 pseudospins in the presence of magnetic field. Keeping in mind the presence of two different bond directions we can write it as h i X H(γ) (γ)~˜ ~˜ ~ (γ) ~˜ ~˜ ~˜ ¯ (γ) ~˜ ~ ¯(γ) ~˜ mod = J Si · Sj + d · Si × Sj + Si · Γ · Sj + µB h · g¯k · Sk, k=i,j (7.2)

where the γ index stands for the bond direction, γ=c for j=i +~c and γ=a for j=i + a~. The corner-sharing Ir-Ir ‘bond unit’ along the c direction has lo- cal symmetry of the C2v point group, same as for the two-octahedra unit in Sr2IrO4 (see Section 6.3). The symmetry constraints on the anisotropic couplings should be therefore identical in this two cases. However, because of the difference in notations—bond along c and C2 axis along b for CaIrO3 vs. bond along x and C2 axis along y for Sr2IrO4—the couplings have similar, but not exactly the same form (compare the following with Equation 6.10 and Equation 6.12):   Γ (c) 0 0 h i aa ~d(c) = d(c) 0 0 Γ¯ (c) = (c)  7 3 a , , ,  0 Γbb 0  ,( . )  (c)  0 0 Γcc

(c) (c) (c) ¯ (c) with Γaa + Γbb + Γcc =0 to ensure a traceless form of the Γ tensor. The g¯(c) tensor has uniform and staggered components due to the staggered rotation of the IrO6 octahedra along the c direction :

 (c)    gaa 0 0 0 0 0 g¯(c) =  (c)  (c)  7 4 ¯k  0 gbb 0   0 0 gbc  .( . )  (c)   (c)  0 0 gcc 0 gbc 0 7.2 intersite exchange interactions 95

H(c) |si |t˜ i |t i |t i ab initio ˜ x y z

hs˜| 0 0 0.1916i hz −0.2105i hy

ht˜x| 0 124.839 −3.2431i hz −1.7312i hy

hty| −0.1916i hz 3.2431i hz 124.934 0.1398i hx

htz| 0.2105i hy 1.7312i hy −0.1398i hx 127.006

Table 7.3: Matrix elements of the ab initio SOC Hamiltonian for NN octahe- dra along the c bond in CaIrO3. Results of MRCI+SOC calcula- tions are shown (meV).

H(c) H(c) Ready-to-compare matrix forms of ab initio and mod in the basis of four lowest SOC states {|s˜i , |t˜xi , |tyi , |tzi} are given in Table 7.3 and Table 7.4 respectively. This representation of the numerical ab initio Hamiltonian is explicitly obtained from the MRCI+SOC calculation. The assignment of the states is made by analysis of the ME’s of the Zeeman Hamiltonian and also of the ME’s of the electric dipole and quadrupole operators. The model Hamiltonian given in Equation 7.2 can be easily expanded in the bond basis and then further trans- formed to the requested basis by the corresponding unitary trans- (c) (c) (c) formation. We set Γbb = −Γaa − Γcc and introduce the additional notations

(c) (c) (c) (c) (c) (c) (J + Γ ) + (J + Γcc ) Γaa J = bb = J(c) − , b+c 2 2 (7.5) q (c) (c) 2 (c) 2 eJ = (da ) + (Jb+c) to make expressions in Table 7.4 a bit more concise. One-to-one correspondence between ME’s in the two matrices al- lows us to extract the actual values of the effective couplings and g factors. One should notice that the g factors obtained within the cal- culation for two magnetic sites are somewhat different as compare to

H(c) |si |t˜ i |t i |t i mod ˜ x y z (c) (c) (c) (c) (c)  (c) (c) (c) (c) (c)  gcc da −gbc eJ +Jb+c gbbda +gbc eJ +Jb+c hs˜| 0 0 ihz r −ihy r (c) (c) (c)  (c) (c) (c)  2eJ eJ +Jb+c 2eJ eJ +Jb+c (c) (c)  (c) (c) (c)  (c)  (c) (c) g J(c)+J +g d g J(c)+J −g d (c) cc e b+c bc a bb e b+c bc a ht˜ x| 0 eJ ihz r −ihy r (c) (c) (c)  (c) (c) (c)  2eJ eJ +Jb+c 2eJ eJ +Jb+c (c) (c) (c) (c) (c)  (c) (c) (c)  (c) (c) gcc da −g eJ +J gcc eJ +J +g da   | bc b+c b+c bc 1 (c) (c) (c) (c) (c) hty −ihz r −ihz r 2 eJ +Jb+c+2Γaa +Γcc −igaahx (c) (c) (c)  (c) (c) (c)  2eJ eJ +Jb+c 2eJ eJ +Jb+c (c) (c) (c) (c) (c)  (c)  (c) (c)  (c) (c) g da +g eJ +J g eJ +J −g da   | bb bc b+c bb b+c bc (c) 1 (c) (c) (c) (c) htz ihy r ihy r igaahx 2 eJ +Jb+c+Γaa −Γcc (c) (c) (c)  (c) (c) (c)  2eJ eJ +Jb+c 2eJ eJ +Jb+c

Table 7.4: Matrix form of the model Hamiltonian given by Equation 7.2 for the c bond in CaIrO3. Notations of Equations 7.3, 7.4 and 7.5 are used. 96 post perovskite a 1 2 quasi 1d antiferromagnet CaIrO3 - , jeff ≈ / -

magnetic couplings, meV g factors (c) (c) (c) (c) (c) (c) (c) (c) (c) J d Γaa Γbb Γcc gaa gbb gcc gbc 125.59 0.95 1.51 1.32 −2.83 0.14 1.73 3.24 0.20

Table 7.5: Nearest-neighbour magnetic couplings and g-factors for corner- sharing octahedra in CaIrO3 as obtained from spin-orbit MRCI calculations. Because Γaa+Γbb+Γcc = 0, only two components are independent.

‡ ‡Because the g¯ the computation for a single magnetic centre . Reasons behind that is tensor is essentially the different symmetry displayed by the two-octahedra cluster (also a single-site in Section 6.3) and the stronger delocalization within the bond direc- property one should refer to data in tion. From the results of the mapping procedure for the c bond shown Equation 7.1. in Table 7.5, it is clear that our initial guess of mapping to an isotropic Heisenberg model was correct. Accidentally even the prediction of J(c) = 125.6 meV is reproduced after the accurate extraction. Despite the very strong deviation from 180◦ (the Ir-O-Ir angle is 140.2◦ here ◦ versus 157 in Sr2IrO4), the DM interaction turns out to be signifi- cantly weaker as compared to Sr2IrO4, less than 1% of the isotropic exchange. It might be related to the precise details of crystallographic distortions in the two compounds: the staggered rotation takes place in CaIrO3 around the local [110¯ ] axis of each octahedron, not around [001] as in Sr2IrO4. In the latter case, the NN octahedra have the same local z axis, while in the former situation there is no common local axis. Such effect was previously discussed in the context of Cu2+ bin- uclear complex [265]. An alternative possible cause of the relatively weak DM coupling may be the strong lattice distortions, that dramati- cally change details of the jeff ≈ 1/2 GS in CaIrO3, as discussed in Sec- tion 7.1. The symmetric anisotropy couplings in the range of 1 − 2% describe small deviations from the isotropic exchange and make the selected c component ‘weaker’ AF. It is clearly seen if we recast the 7 5 (c) (c) (c) numbers in Table . into Ja = 127.1 ≈ Jb = 126.9 and Jc = 122.8 meV. The situation is different for the Ir-Ir link along the a direction in which NN octahedra share edges. The two-octahedra [Ir2O10] cluster ‡ ‡ As well as the unit displays C2h symmetry. This two-octahedra block has a centre 56-site [Ir2O10 Ir6O24Ca14] cluster used for the magnetic couplings, meV g factors calculations. (a) (a) (a) (a) (a) (a) (a) (a) (a) J Γaa Γbb Γcc Γbc gaa gbb gcc gbc 2.21 5.20 3.61 −8.82 2.06 0.37 1.84 3.21 0.14

Table 7.6: Nearest-neighbour magnetic couplings and g-factors for edge- sharing a-axis octahedra in CaIrO3 as obtained from spin-orbit (a) (a) (a) MRCI calculations. The convention Γaa +Γbb +Γcc =0 is used. 7.2 intersite exchange interactions 97 of inversion, therefore the DM coupling is absent here. All symmetry constrains to the model Hamiltonian shown in Equation 7.2 can be obtained using the irreducible representation projector analysis [345]. Within the C2h point-group symmetry the effective couplings read

 (a)  Γaa 0 0 ~d(a) = 0 Γ¯ (a) = (a) (a)  7 6 ,  0 Γbb Γbc  , ( . )  (a) (a)  0 Γbc Γcc

(a) (a) (a) with Γaa + Γbb + Γcc =0, and

 (a)  gaa 0 0 g¯(a) =  (a) (a)  7 7 ¯k  0 gbb gbc  . ( . )  (a) (a)  0 gbc gcc

One can notice that now there is an additional off-diagonal term (a) Γbc in the symmetric anisotropy tensor. To understand how it mod- ifies the spectra of the lowest four states, we in a first stage assume it is zero. The only non-isotropic contribution is then the diagonal ¯ | | Γ. The eigenstates of the model are the√ ‘detached’ singlet si = 0√, 0i and triplet states |txi = (|1, 1i−|1, −1i)/ 2, |tyi = (|1, 1i+|1, −1i)/ 2, | | (a) and tzi= 1, 0i). The Γbc coupling brings further interaction between |tyi and |tzi with resulting |t˜yi and |t˜zi admixed states. Interestingly, the ab initio data of Table 7.7 show that the singlet state is not the highest in energy, as we were tempted initially to assume from the relatively large energy gap between the lowest three states and the higher-lying state, but one of the lowest, i.e the isotropic J(a) turns out to be antiferromagnetic. The full set of anisotropic parameters can be again obtained with help of the mapping procedure between the ab initio and model Hamiltonians shown in Table 7.7 and Table 7.8 respectively. For writing the latter more compactly we introduce sup- plementary notations:

(a) (a) (a) (a) (a) Γc−b = Γcc − Γbb = 2Γcc + Γaa , q (a) (a) eΓ (a) = 4(Γ )2 + (Γ )2, bc c−b (7.8) s (a) s (a) eΓ (a) + Γ eΓ (a) − Γ P = c−b , M = c−b . 2eΓ (a) 2eΓ (a) Although at first sight there are only 6 independent ME’s in the Hamiltonian, which looks not enough to extract all 8 model parame- ters (4 independent exchange couplings and 4 g factors), the mapping can still be made. An additional detail is that from some of the matrix elements, i.e. those in the first lines in Table 7.7 and Table 7.8, we ob- tain two equations instead of one, because there are two components of magnetic field. 98 post perovskite a 1 2 quasi 1d antiferromagnet CaIrO3 - , jeff ≈ / -

H(a) |t i |si |t˜ i |t˜ i ab initio x y z

0.4290i hy −1.800i hy htx| 0 0 −3.191i hz −0.3746i hz hs| 0 0.3945 0 0

−0.4290i hy ht˜y| 0 0.6261 0.3691i hx +3.191i hz

1.800i hy ht˜z| 0 −0.3691i hx 7.176 +0.3746i hz

Table 7.7: Matrix elements of the ab initio spin-orbit Hamiltonian for a bond in CaIrO3. Results of MRCI+SOC calculations are shown (meV).

H(a) | | |˜ |˜ mod txi si tyi tzi  (a) (a)   (a) (a)  i hy gbc M − gbb P i hy gbb M + gbc P htx| 0 0  (a) (a)   (a) (a)  +i hz gcc M − gbc P +i hz gbc M + gcc P hs| 0 −J(a) + Γaa 0 0   2 (a) (a)   | −i hy gbc M − gbb P 1 (a) (a) (a) (a) ht˜y 0 J + Γaa − eΓ igaa hx  (a) (a)  4 −i hz gcc M − gbc P   (a) (a)   | −i hy gbb M + gbc P (a) 1 (a) (a) (a) ht˜z 0 −igaa hx J + Γaa + eΓ  (a) (a)  4 −i hz gbc M + gcc P

Table 7.8: Matrix form of the model Hamiltonian given by Equation 7.2 for a-axis bonds in CaIrO3. The notations introduced in Equations 7.6 and 7.8 are used.

The final results are presented in Table 7.6. Couplings along the a direction are very anisotropic and much weaker as compared to the Heisenberg exchange along the c-axis bond. Symmetric anisotropic exchange couplings are AF the for a and b components of the pseu- dospin and FM for the c component. Together with the isotropic J(a) they can be recast into an anisotropic Heisenberg model with (a) (a) (a) (a) Ja = 7.41, Jb = 5.82, and Jc = −6.61 meV. The off-diagonal Γbc coupling is the smallest, but almost reaches the value of the isotropic part. The picture for the a bond in CaIrO3 is to large extent similar to the edge-sharing honeycomb iridates [7, 243], where one of the FM anisotropic couplings (referred to in that structure as Kitaev coupling) overcomes the strength of the AF isotropic exchange. The present ab initio results, i.e. strong AF interactions along c and (a) weak anisotropic couplings along a with a FM Γcc component (see Figure 7.3), are consistent with the observed very large Curie-Weiss temperature and striped AF magnetic order with spins lying nearly 7.3 conclusions 99

Figure 7.3: Sketch of the NN magnetic interactions in CaIrO3. collinear to c [341]. The MRCI+SOC calculations together with the employed mapping procedure provide a firm quantitative basis in understanding the magnetic properties of this material.

7.3 Conclusions

We have employed, in summary, a set of many-body quantum chem- ical calculations to unravel the 5d electronic structure of the post- perovskite iridate CaIrO3. While electronic structure calculations have been earlier performed within the local density approximation to density functional theory and the role of electron correlation ef- fects beyond the LDA/GGA framework has been pointed out [346], we here make clear quantitative predictions for the Ir t2g splittings, character of the entangled spin-orbit wave function, and magnitude of the NN magnetic exchange interactions. Our results single out 5 CaIrO3 as a 5d system in which lattice distortions and spin-orbit couplings compete on the same energy scale to give rise to highly anisotropic electronic structure and magnetic correlations. In partic- ular, the present ab initio investigation yields a different picture than previously derived on the basis of resonant X-ray diffraction exper- iments for the relativistic ground-state wavefunction. The large t2g splittings that we have found from the calculations give rise to a ‘non- cubic jeff ≈ 1/2’ GS with dominant xzf hole character and remarkably strong AF interactions along the c axis. Further, the large AF cou- plings along c and weak anisotropic exchange along a characterize CaIrO3 as a jeff =1/2 quasi-one-dimensional antiferromagnet.

SUMMARY

his thesis covers different problems that arise due to crystal and pseudospin anisotropy present in 3d and 5d transition T metal oxides. We demonstrate that the methods of compu- tational quantum chemistry can be fruitfully used for quantitative studies of such problems. In Chapter 2, Chapter 3, and Chapter 7 we show that it is possible to reliably calculate local multiplet splittings fully ab initio, and there- fore help to assign peaks in experimental spectra to corresponding electronic states. In a situation of large number of peaks due to low local symmetry such assignment using semi-empirical methods can be very tedious and non-unique. Moreover, in Chapter 4 we present a computational scheme for cal- culating intensities as observed in the resonant inelastic X-ray scatter- ing and X-ray absorption experiments. In our scheme highly-excited core-hole states are calculated explicitly taking into account corre- sponding orbital relaxation and electron polarization. Computed Cu L-edge spectra for the Li2CuO2 compound reproduce all features present in experiment. Unbiased ab initio calculations allow us to unravel a delicate inter- play between the distortion of the local ligand cage around the tran- sition metal ions and the anisotropic electrostatic interactions due to second and farther coordination shells. As shown in Chapter 5 and Chapter 6 this interplay can lead to the counter intuitive multiplet structure, single-ion anisotropy, and magnetic g factors. The effect is quite general and may occur in compounds with large difference between charges of metal ions that form anisotropic environment around the transition metal, like Ir4+ in plane versus Sr2+ out of plane in the case of Sr2IrO4. An important aspect of the presented study is the mapping of the quantum chemistry results onto simpler physical models, namely ex- tended Heisenberg model, providing an ab initio parametrization. In Chapter 5 we employ the effective Hamiltonian technique for extract- ing parameters of the anisotropic Heisenberg model with single-ion anisotropy in the case of quenched orbital moment and second-order spin-orbit coupling. Calculated strong easy-axis anisotropy of the same order of magnitude as the symmetric exchange is consistent with experimentally-observer all-in/all-out magnetic order. In Chapter 6 we introduce new flavour of the mapping procedure applicable to systems with first-order spin-orbit coupling, such as 5d5 iridates based on analysis of the wavefunction and interaction with magnetic field. In Chapter 6 and Chapter 7 we use this new proce-

101 102 summary

dure to obtain parameters of the pseudospin anisotropic Heisenberg model. We find large antisymmetric exchange leading to the canted antiferromagnetic state in Sr2IrO4 and nearly ideal one-dimensional Heisenberg behaviour of the CaIrO3, both agree very well with exper- imental findings. PUBLICATIONS

[NB1] H.-Y. Huang, N. A. Bogdanov, L. Siurakshina, P. Fulde, J. van den Brink, and L. Hozoi, ‘Ab initio calculation of d–d excitations in quasi- one-dimensional Cu d9 correlated materials,’ Phys. Rev. B 84, 235125 (2011).

[NB2] N. A. Bogdanov, J. van den Brink, and L. Hozoi, ‘Ab initio computa- tion of d–d excitation energies in low-dimensional Ti and V oxychlo- rides,’ Phys. Rev. B 84, 235146 (2011).

[NB3] N. A. Bogdanov, V. M. Katukuri, H. Stoll, J. van den Brink, and L. Hozoi, ‘Post-perovskite CaIrO3: a j = 1/2 quasi-one-dimensional antiferromagnet,’ Phys. Rev. B 85, 235147 (2012).

[NB4] N. A. Bogdanov, R. Maurice, I. Rousochatzakis, J. van den Brink, and L. Hozoi, ‘Magnetic state of pyrochlore Cd2Os2O7 emerging from strong competition of ligand distortions and longer-range crystalline anisotropy,’ Phys. Rev. Lett. 110, 127206 (2013).

[NB5] N. A. Bogdanov, V. M. Katukuri, J. Romhányi, V. Yushankhai, V. Kataev, B. Büchner, J. van den Brink, and L. Hozoi, ‘Orbital recon- struction in nonpolar tetravalent transition-metal oxide layers,’ Nat. Commun. 6, 7306 (2015).

[NB6] S. Calder, J. G. Vale, N. A. Bogdanov, X. Liu, C. Donnerer, M. H. Up- ton, D. Casa, A. H. Said, M. D. Lumsden, Z. Zhao, J. Q. Yan, D. Man- drus, S. Nishimoto, J. van den Brink, J. P. Hill, D. F. McMorrow, and A. D. Christianson, ‘Spin-orbit-driven magnetic structure and excita- tion in the 5d pyrochlore Cd2Os2O7,’ Nat. Commun. 7, 11651 (2016).

[NB7] R. Yadav, N. A. Bogdanov, V. M. Katukuri, S. Nishimoto, J. van den Brink, and L. Hozoi, ‘Kitaev exchange and field-induced quantum spin-liquid states in honeycomb α-RuCl3,’ Sci. Rep. 6, 37925 (2016).

[NB8] L. Xu, N. A. Bogdanov, A. Princep, P. Fulde, J. van den Brink, and L. Hozoi, ‘Covalency and vibronic couplings make a nonmagnetic j=3/2 ion magnetic,’ Npj Quant. Mater. 1, 16029 (2016).

[NB9] N. A. Bogdanov, V. Bisogni, R. Kraus, C. Monney, K. Zhou, T. Schmitt, J. Geck, A. O. Mitrushchenkov, H. Stoll, J. van den Brink, and L. Ho- zoi, ‘Orbital breathing effects in the computation of x-ray d-ion spec- tra in solids by ab initio wave-function-based methods,’ J. Phys.: Con- dens. Matter 29, 035502 (2017).

[NB10] S. Calder, J. G. Vale, N. Bogdanov, C. Donnerer, D. Pincini, M. Moretti Sala, X. Liu, M. H. Upton, D. Casa, Y. G. Shi, Y. Tsuji- moto, K. Yamaura, J. P. Hill, J. van den Brink, D. F. McMorrow, and A. D. Christianson, ‘Strongly gapped spin-wave excitation in the in- sulating phase of NaOsO3,’ Phys. Rev. B 95, 020413(R) (2017)

103

BIBLIOGRAPHY

[1] N. W. Ashcroft and N. D. Mermin, Solid state physics (Holt, Rinehart and Win- ston, New York, 1976).

[2] D. I. Khomskii, Transition Metal Compounds (Cambridge University Press, 2014).

[3] J. G. Bednorz and K. A. Müller, ‘Possible high Tc superconductivity in the Ba-La-Cu-O system,’ Z. Phys. B 64, 189 (1986).

[4] C. Kim, Z.-X. Shen, N. Motoyama, H. Eisaki, S. Uchida, T. Tohyama, and S. Maekawa, ‘Separation of spin and charge excitations in one-dimensional SrCuO2,’ Phys. Rev. B 56, 15589 (1997).

[5] J. Schlappa, K. Wohlfeld, K. J. Zhou, M. Mourigal, M. W. Haverkort, V. N. Stro- cov, L. Hozoi, C. Monney, S. Nishimoto, S. Singh, A. Revcolevschi, J.-S. Caux, L. Patthey, H. M. Ronnow, J. van den Brink, and T. Schmitt, ‘Spin-orbital sep- aration in the quasi-one-dimensional Mott insulator Sr2CuO3,’ Nature 485, 82 (2012).

[6] G. Jackeli and G. Khaliullin, ‘Mott insulators in the strong spin-orbit coupling limit: from Heisenberg to a quantum compass and Kitaev models,’ Phys. Rev. Lett. 102, 017205 (2009).

[7] V. M. Katukuri, S. Nishimoto, V. Yushankhai, A. Stoyanova, H. Kandpal, S. K. Choi, R. Coldea, I. Rousochatzakis, L. Hozoi, and J. van den Brink, ‘Kitaev interactions between j = 1/2 moments in honeycomb Na2IrO3 are large and ferromagnetic: insights from ab initio quantum chemistry calculations,’ New J. Phys. 16, 013056 (2014).

[8] W. Heitler and F. London, ‘Wechselwirkung neutraler Atome und homöopolare Bindung nach der Quantenmechanik,’ Z. Phys. 44, 455 (1927).

[9] F. Hund, ‘Zur Deutung der Molekelspektren. IV,’ Z. Phys. 51, 759 (1928).

[10] J. C. Slater, ‘Note on Hartree’s method,’ Phys. Rev. 35, 210 (1930).

[11] P. Fulde, Correlated electrons in quantum matter (World Scientific, 2012).

[12] K. Fink, R. Fink, and V. Staemmler, ‘Ab Initio calculation of the magnetic ex- change coupling in linear oxo-bridged binuclear complexes of titanium(III), vanadium(III), and chromium(III),’ Inorg. Chem. 33, 6219 (1994).

[13] C. de Graaf, R. Broer, and W. C. Nieuwpoort, ‘Comparison of the superex- change interaction in NiO and in a NiO (100) surface,’ Chem. Phys. Lett. 271, 372 (1997).

[14] D. Muñoz, I. de P. R. Moreira, and F. Illas, ‘Effective t–J model Hamiltonian parameters of monolayered cuprate superconductors from ab initio electronic structure calculations,’ Phys. Rev. B 65, 224521 (2002).

[15] L. Hozoi, C. Presura, C. de Graaf, and R. Broer, ‘Electronic structure α0– NaV2O5: Wave-function-based embedded-cluster calculations,’ Phys. Rev. B 67, 035117 (2003).

[16] C. J. Calzado, S. Evangelisti, and D. Maynau, ‘Local orbitals for the truncation of inactive space: application to magnetic systems,’ J. Phys. Chem. A 107, 7581 (2003).

105 106 bibliography

[17] R. Broer, L. Hozoi, and W. C. Nieuwpoort, ‘Non-orthogonal approaches to the study of magnetic interactions,’ Mol. Phys. 101, 233 (2003).

[18] C. Pisani, ed., Quantum-mechanical ab initio calculation of the properties of crystalline materials (Springer, Berlin, 1996).

[19] R. Dovesi, B. Civalleri, C. Roetti, V. R. Saunders, and R. Orlando, ‘Ab initio quantum simulation in solid state chemistry,’ in Reviews in computational chem- istry vol. 21, edited by K. B. Lipkowitz, R. Larter, and T. R. Cundari (Wiley, Hoboken, New Jersey, 2005) pp. 1–125.

[20] R. A. Evarestov, Quantum shemistry of solids (Springer, Berlin, 2012).

[21] R. O. Jones and O. Gunnarsson, ‘The density functional formalism, its applica- tions and prospects,’ Rev. Mod. Phys. 61, 689 (1989).

[22] J. P. Perdew, ‘Density functional theory and the band gap problem,’ Int. J. Quan- tum Chem. 28, 497 (1985).

[23] V. I. Anisimov, J. Zaanen, and O. K. Andersen, ‘Band theory and Mott insula- tors: Hubbard U instead of Stoner I,’ Phys. Rev. B 44, 943 (1991).

[24] L. Vaugier, H. Jiang, and S. Biermann, ‘Hubbard U and Hund exchange J in transition metal oxides: screening versus localization trends from constrained random phase approximation,’ Phys. Rev. B 86, 165105 (2012).

[25] A. Georges, G. Kotliar, W. Krauth, and M. J. Rozenberg, ‘Dynamical mean- field theory of strongly correlated fermion systems and the limit of infinite dimensions,’ Rev. Mod. Phys. 68, 13 (1996).

[26] C. Pisani, R. Dovesi, and C. Roetti, Hartree-Fock ab initio treatment of crystalline systems (Springer, Berlin, 1988).

[27] A. D. Becke, ‘A new mixing of Hartree-Fock and local density-functional theo- ries,’ J. Chem. Phys. 98, 1372 (1993).

[28] P. Y. Ayala, K. N. Kudin, and G. E. Scuseria, ‘Atomic orbital Laplace- transformed second-order Møller-Plesset theory for periodic systems,’ J. Chem. Phys. 115, 9698 (2001).

[29] L. Maschio, D. Usvyat, F. R. Manby, S. Casassa, C. Pisani, and M. Schütz, ‘Fast local-MP2 method with density-fitting for crystals. I. Theory and algorithms,’ Phys. Rev. B 76, 075101 (2007).

[30] C. Pisani, L. Maschio, S. Casassa, M. Halo, M. Schütz, and D. Usvyat, ‘Periodic local MP2 method for the study of electronic correlation in crystals: Theory and preliminary applications,’ J. Comput. Chem. 29, 2113 (2008).

[31] M. Marsman, A. Grüneis, J. Paier, and G. Kresse, ‘Second-order Møller- Plesset perturbation theory applied to extended systems. I. Within the projector- augmented-wave formalism using a plane wave basis set,’ J. Chem. Phys. 130, 184103 (2009).

[32] Y. ya Ohnishi and S. Hirata, ‘Hybrid coupled-cluster and perturbation method for extended systems of one-dimensional periodicity,’ J. Chem. Phys. 135, 094108 (2011).

[33] A. Grüneis, G. H. Booth, M. Marsman, J. Spencer, A. Alavi, and G. Kresse, ‘Natural orbitals for wave function based correlated calculations using a plane wave basis set,’ J. Chem. Theory Comput. 7, 2780 (2011). bibliography 107

[34] C. Pisani, M. Busso, G. Capecchi, S. Casassa, R. Dovesi, L. Maschio, C. Zicovich- Wilson, and M. Schuütz, ‘Local-MP2 electron correlation method for noncon- ducting crystals,’ J. Chem. Phys. 122, 094113 (2005).

[35] A. Grüneis, M. Marsman, and G. Kresse, ‘Second-order Møller-Plesset pertur- bation theory applied to extended systems. II. Structural and energetic proper- ties,’ J. Chem. Phys. 133, 074107 (2010).

[36] A. Erba, M. Itou, Y. Sakurai, R. Yamaki, M. Ito, S. Casassa, L. Maschio, A. Teren- tjevs, and C. Pisani, ‘Beyond a single-determinantal description of the density matrix of periodic systems: experimental versus theoretical Compton profiles of crystalline silicon,’ Phys. Rev. B 83, 125208 (2011).

[37] G. H. Booth, A. Grüneis, G. Kresse, and A. Alavi, ‘Towards an exact description of electronic wavefunctions in real solids,’ Nature 493, 365 (2012).

[38] T. Helgaker, P. Jørgensen, and J. Olsen, Molecular electronic-structure theory (Wi- ley, Chichester, 2000).

[39] P. G. Szalay, T. Müller, and H. Lischka, ‘Excitation energies and transition mo- ments by the multireference averaged quadratic coupled cluster (MR-AQCC) method,’ Phys. Chem. Chem. Phys. 2, 2067 (2000).

[40] P. G. Szalay, ‘Multireference averaged quadratic coupled-cluster (MR-AQCC) method based on the functional of the total energy,’ Chem. Phys. 349, 121 (2008).

[41] A. Venkatnathan, A. B. Szilva, D. Walter, R. J. Gdanitz, and E. A. Carter, ‘Size ex- tensive modification of local multireference configuration interaction,’ J. Chem. Phys. 120, 1693 (2004).

[42] C. Møller and M. S. Plesset, ‘Note on an approximation treatment for many- electron systems,’ Phys. Rev. 46, 618 (1934).

[43] K. Andersson, P. Å. Malmqvist, B. O. Roos, A. J. Sadlej, and K. Wolinski, ‘Second-order perturbation theory with a CASSCF reference function,’ J. Phys. Chem. 94, 5483 (1990).

[44] K. Andersson, P. Å. Malmqvist, and B. O. Roos, ‘Second-order perturbation theory with a complete active space self-consistent field reference function,’ J. Chem. Phys. 96, 1218 (1992).

[45] C. Angeli, R. Cimiraglia, S. Evangelisti, T. Leininger, and J.-P. Malrieu, ‘Intro- duction of N-electron valence states for multireference perturbation theory,’ J. Chem. Phys. 114, 10252 (2001).

[46] C. Angeli, R. Cimiraglia, and J.-P. Malrieu, ‘N-electron valence state perturba- tion theory: a fast implementation of the strongly contracted variant,’ Chem. Phys. Lett. 350, 297 (2001).

[47] K. G. Dyall, ‘The choice of a zeroth-order Hamiltonian for second-order pertur- bation theory with a complete active space self-consistent-field reference func- tion,’ J. Chem. Phys. 102, 4909 (1995).

[48] C. de Graaf, C. Sousa, I. d. P. R. Moreira, and F. Illas, ‘Multiconfigurational perturbation theory: an efficient tool to predict magnetic coupling parameters in biradicals, molecular complexes, and ionic insulators,’ J. Phys. Chem. A 105, 11371 (2001).

[49] C. Angeli and C. J. Calzado, ‘The role of the magnetic orbitals in the calculation of the magnetic coupling constants from multireference perturbation theory methods,’ J. Chem. Phys. 137, 034104 (2012). 108 bibliography

[50] K. H. Marti and M. Reiher, ‘New electron correlation theories for transition metal chemistry,’ Phys. Chem. Chem. Phys. 13, 6750 (2011).

[51] G. K.-L. Chan and S. Sharma, ‘The density matrix renormalization group in quantum chemistry,’ Ann. Rev. Phys. Chem. 62, 465 (2011).

[52] G. H. Booth, A. J. W. Thom, and A. Alavi, ‘Fermion Monte Carlo without fixed nodes: a game of life, death, and annihilation in Slater determinant space,’ J. Chem. Phys. 131, 054106 (2009).

[53] D. Zgid and M. Nooijen, ‘The density matrix renormalization group self- consistent field method: Orbital optimization with the density matrix renormal- ization group method in the active space,’ J. Chem. Phys. 128, 144116 (2008).

[54] D. Ghosh, J. Hachmann, T. Yanai, and G. K.-L. Chan, ‘Orbital optimization in the density matrix renormalization group, with applications to polyenes and beta-carotene,’ J. Chem. Phys. 128, 144117 (2008).

[55] R. E. Thomas, Q. Sun, A. Alavi, and G. H. Booth, ‘Stochastic multiconfigura- tional self-consistent field theory,’ J. Chem. Theory Comput. 11, 5316 (2015).

[56] G. Li Manni, S. D. Smart, and A. Alavi, ‘Combining the complete active space self-consistent field method and the full configuration interaction quan- tum monte carlo within a super-CI framework, with application to challenging metal-porphyrins,’ J. Chem. Theory Comput. 12, 1245 (2016).

[57] S. Guo, M. A. Watson, W. Hu, Q. Sun, and G. K.-L. Chan, ‘N-electron valence state perturbation theory based on a density matrix renormalization group ref- erence function, with applications to the chromium dimer and a trimer model of poly(p-phenylenevinylene),’ J. Chem. Theory Comput. 12, 1583 (2016).

[58] S. Wouters, V. Van Speybroeck, and D. Van Neck, ‘DMRG-CASPT2 study of the longitudinal static second hyperpolarizability of all-trans polyenes,’ J. Chem. Phys. 145, 054120 (2016).

[59] S. Sharma, G. Jeanmairet, and A. Alavi, ‘Quasi-degenerate perturbation theory using matrix product states,’ J. Chem. Phys. 144, 034103 (2016).

[60] G. Jeanmairet, S. Sharma, and A. Alavi, ‘Stochastic multi-reference perturba- tion theory with application to the linearized coupled cluster method,’ J. Chem. Phys. 146, 044107 (2017).

[61] W. Klopper, F. R. Manby, S. Ten-No, and E. F. Valeev, ‘R12 methods in explicitly correlated molecular electronic structure theory,’ Int. Rev. Phys. Chem. 25, 427 (2006).

[62] T. Shiozaki and H.-J. Werner, ‘Multireference explicitly correlated F12 theories,’ Mol. Phys. 111, 607 (2013).

[63] D. Ma, G. Li Manni, and L. Gagliardi, ‘The generalized active space concept in multiconfigurational self-consistent field methods,’ J. Chem. Phys. 135, 044128 (2011).

[64] G. Li Manni, D. Ma, F. Aquilante, J. Olsen, and L. Gagliardi, ‘SplitGAS method for strong correlation and the challenging case of Cr2,’ J. Chem. Theory Com- put. 9, 3375 (2013).

[65] H. Lischka, R. Shepard, R. M. Pitzer, I. Shavitt, M. Dallos, T. Müller, P. G. Szalay, M. Seth, G. S. Kedziora, S. Yabushita, and Z. Zhang, ‘High-level multireference methods in the quantum-chemistry program system columbus: analytic MR- CISD and MR-AQCC gradients and MR-AQCC-LRT for excited states, GUGA spin–orbit CI and parallel CI density,’ Phys. Chem. Chem. Phys. 3, 664 (2001). bibliography 109

[66] F. Neese, ‘A spectroscopy oriented configuration interaction procedure,’ J. Chem. Phys. 119, 9428 (2003).

[67] A. S. P. Gomes and C. R. Jacob, ‘Quantum-chemical embedding methods for treating local electronic excitations in complex chemical systems,’ Annu. Rep. Prog. Chem., Sect. C: Phys. Chem. 108, 222 (2012).

[68] T. A. Wesolowski, S. Shedge, and X. Zhou, ‘Frozen-density embedding strategy for multilevel simulations of electronic structure,’ Chem. Rev. 115, 5891 (2015).

[69] H. M. Evjen, ‘On the stability of certain heteropolar crystals,’ Phys. Rev. 39, 675 (1932).

[70] A. Gellé and M.-B. Lepetit, ‘Fast calculation of the electrostatic potential in ionic crystals by direct summation method,’ J. Chem. Phys. 128, 244716 (2008).

[71] P. V. Sushko and I. V. Abarenkov, ‘General purpose electrostatic embedding potential,’ J. Chem. Theory Comput. 6, 1323 (2010).

[72] P. P. Ewald, ‘Die Berechnung optischer und elektrostatischer Gitterpotentiale,’ Ann. Phys. 369, 253 (1921).

[73] M. Klintenberg, S. Derenzo, and M. Weber, ‘Accurate crystal fields for embed- ded cluster calculations,’ Comput. Phys. Commun. 131, 120 (2000).

[74] A. M. Burow, M. Sierka, J. Döbler, and J. Sauer, ‘Point defects in CaF2 and CeO2 investigated by the periodic electrostatic embedded cluster method,’ J. Chem. Phys. 130, 174710 (2009).

[75] M.-B. Lepetit, N. Suaud, A. Gelle, and V. Robert, ‘Environment effects on effec- tive magnetic exchange integrals and local spectroscopy of extended strongly correlated systems,’ J. Chem. Phys. 118, 3966 (2003).

[76] J. Meister and W. H. E. Schwarz, ‘Principal components of ionicity,’ J. Chem. Phys. 98, 8245 (1994).

[77] R. S. Mulliken, ‘Electronic population analysis on LCAO–MO molecular wave functions. I,’ J. Chem. Phys. 23, 1833 (1955).

[78] A. E. Reed, R. B. Weinstock, and F. Weinhold, ‘Natural population analysis,’ J. Chem. Phys. 83, 735 (1985).

[79] R. F. W. Bader, ‘A quantum theory of molecular structure and its applications,’ Chem. Rev. 91, 893 (1991).

[80] G. Knizia, ‘Intrinsic atomic orbitals: an unbiased bridge between quantum the- ory and chemical concepts,’ J. Chem. Theory Comput. 9, 4834 (2013).

[81] R. McWeeny, Methods of molecular quantum mechanics, 2nd ed. (Academic Press, London, 1992).

[82] S. Huzinaga and A. A. Cantu, ‘Theory of separability of many-electron systems,’ J. Chem. Phys. 55, 5543 (1971).

[83] M. Dolg, U. Wedig, H. Stoll, and H. Preuss, ‘Energy-adjusted ab initio pseu- dopotentials for the first row transition elements,’ J. Chem. Phys. 86, 866 (1987).

[84] Z. Barandiarán and L. Seijo, ‘The ab initio model potential representation of the crystalline environment. theoretical study of the local distortion on NaCl:Cu+,’ J. Chem. Phys. 89, 5739 (1988). 110 bibliography

[85] T. Klüner, N. Govind, Y. A. Wang, and E. A. Carter, ‘Periodic density functional embedding theory for complete active space self-consistent field and configura- tion interaction calculations: Ground and excited states,’ J. Chem. Phys. 116, 42 (2002).

[86] F. R. Manby, M. Stella, J. D. Goodpaster, and T. F. Miller, ‘A simple, exact density-functional-theory embedding scheme,’ J. Chem. Theory Comput. 8, 2564 (2012).

[87] U. Birkenheuer, C. Willnauer, M. von Arnim, W. Alsheimer, and D. Izo- tov, Wavefunction-based correlation calculations for hole and electron attach- ment states in solids and polymers, Annual report (Max Planck Institute for the Physics of Complex Systems, Dresden, Germany, 2000–2002) http://www.mpipks-dresden.mpg.de/pdf/institutsbericht00_01_short.pdf.

[88] U. Birkenheuer, P. Fulde, and H. Stoll, ‘A simplified method for the compu- tation of correlation effects on the band structure of semiconductors,’ Theor. Chem. Acc. 116, 398 (2006).

[89] L. Hozoi, U. Birkenheuer, P. Fulde, A. Mitrushchenkov, and H. Stoll, ‘Ab initio wave function-based methods for excited states in solids: Correlation correc- tions to the band structure of ionic oxides,’ Phys. Rev. B 76, 085109 (2007).

[90] A. Stoyanova, L. Hozoi, P. Fulde, and H. Stoll, ‘Correlation-induced correc- tions to the band structure of boron nitride: A wave-function-based approach,’ J. Chem. Phys. 131, 044119 (2009).

[91] C. M. Zicovich-Wilson, R. Dovesi, and V. R. Saunders, ‘A general method to obtain well localized Wannier functions for composite energy bands in linear combination of atomic orbital periodic calculations,’ J. Chem. Phys. 115, 9708 (2001).

[92] A. Shukla, M. Dolg, H. Stoll, and P. Fulde, ‘An ab initio embedded-cluster approach to electronic structure calculations on perfect solids: a Hartree–Fock study of lithium hydride,’ Chem. Phys. Lett. 262, 213 (1996).

[93] A. Shukla, M. Dolg, P. Fulde, and H. Stoll, ‘Obtaining Wannier functions of a crystalline insulator within a Hartree–Fock approach: Applications to LiF and LiCl,’ Phys. Rev. B 57, 1471 (1998).

[94] R. Dovesi, R. Orlando, B. Civalleri, C. Roetti, V. R. Saunders, and C. M. Zicovich- Wilson, ‘crystal: a computational tool for the ab initio study of the electronic properties of crystals,’ Z. Kristallogr. 220, 571 (2005).

[95] I. Souza, N. Marzari, and D. Vanderbilt, ‘Maximally localized Wannier func- tions for entangled energy bands,’ Phys. Rev. B 65, 035109 (2001).

[96] K. S. Thygesen, L. B. Hansen, and J. K. W., ‘Partly occupied Wannier functions,’ Phys. Rev. Lett. 94, 026405 (2005).

[97] U. Birkenheuer and D. Izotov, ‘Localization of Wannier functions for entangled energy bands,’ Phys. Rev. B 71, 125116 (2005).

[98] P. Pulay, ‘Localizability of dynamic electron correlation,’ Chem. Phys. Lett. 100, 151 (1983).

[99] S. Sæbø and P. Pulay, ‘Local configuration interaction: An efficient approach for larger molecules,’ Chem. Phys. Lett. 113, 13 (1985).

[100] S. Sæbø and P. Pulay, ‘Local treatment of electron correlation,’ Ann. Rev. Phys. Chem. 44, 213 (1993). bibliography 111

[101] C. Hampel and H.-J. Werner, ‘Local treatment of electron correlation in cou- pled cluster theory,’ J. Chem. Phys. 104, 6286 (1996).

[102] L. Hozoi and P. Fulde, ‘Correlated electrons in fe-as compounds: A quantum chemical perspective,’ Phys. Rev. Lett. 102, 136405 (2009).

[103] A. Stoyanova, L. Hozoi, P. Fulde, and H. Stoll, ‘Wave-function-based approach to quasiparticle bands: insight into the electronic structure of c-ZnS,’ Phys. Rev. B 83, 205119 (2011).

[104] L. Hozoi, L. Siurakshina, P. Fulde, and J. van den Brink, ‘Ab Initio determina- tion of Cu 3d orbital energies in layered copper oxides,’ Sci. Rep. 1, 65 (2011).

[105] A. Stoyanova, A. O. Mitrushchenkov, L. Hozoi, H. Stoll, and P. Fulde, ‘Electron correlation effects in diamond: A wave-function quantum-chemistry study of the quasiparticle band structure,’ Phys. Rev. B 89, 235121 (2014).

[106] C. Roetti, R. Dovesi, M. von Armin, W. Alsheimer, and U. Birkenheuer, ‘The crystal-molpro interface,’ Max Planck Institute for the Physics of Complex Systems, Dresden, Germany (2001).

[107] T. A. Wesolowski and A. Warshel, ‘Frozen density functional approach for ab initio calculations of solvated molecules,’ J. Phys. Chem. 97, 8050 (1993).

[108] S. Fux, C. R. Jacob, J. Neugebauer, L. Visscher, and M. Reiher, ‘Accurate frozen- density embedding potentials as a first step towards a subsystem description of covalent bonds,’ J. Chem. Phys. 132, 164101 (2010).

[109] C. Huang, M. Pavone, and E. A. Carter, ‘Quantum mechanical embedding theory based on a unique embedding potential,’ J. Chem. Phys. 134, 154110 (2011).

[110] G. Knizia and G. K.-L. Chan, ‘Density matrix embedding: A simple alternative to dynamical mean-field theory,’ Phys. Rev. Lett. 109, 186404 (2012).

[111] I. W. Bulik, W. Chen, and G. E. Scuseria, ‘Electron correlation in solids via density embedding theory,’ J. Chem. Phys. 141, 054113 (2014).

[112] Q. Sun and G. K.-L. Chan, ‘Quantum embedding theories,’ Acc. Chem. Res. 49, 2705 (2016).

[113] I. Peschel, ‘Special review: Entanglement in solvable many-particle models,’ Braz. J. Phys. 42, 267 (2012).

[114] C. de Graaf and R. Broer, ‘Midinfrared spectrum of undoped cuprates: d–d transitions studied by ab initio methods,’ Phys. Rev. B 62, 702 (2000).

[115] M. Moretti Sala, V. Bisogni, C. Aruta, G. Balestrino, H. Berger, N. B. Brookes, G. M. de Luca, D. Di Castro, M. Grioni, M. Guarise, P. G. Medaglia, F. Miletto Granozio, M. Minola, P. Perna, M. Radovic, M. Salluzzo, T. Schmitt, K. J. Zhou, L. Braicovich, and G. Ghiringhelli, ‘Energy and symmetry of dd excitations in undoped layered cuprates measured by Cu L3 resonant inelastic x-ray scattering,’ New J. Phys. 13, 043026 (2011).

[116] H.-J. Werner, P. J. Knowles, G. Knizia, F. R. Manby, and M. Schütz, ‘molpro: a general-purpose quantum chemistry program package,’ WIREs Comput. Mol. Sci. 2, 242 (2012).

[117] D. Lines, M. Weller, D. Currie, and D. Ogbome, ‘Alkaline earth copper oxides,’ Mat. Res. Bull. 26, 323 (1991). 112 bibliography

[118] M. Heinau, R. Baumann, B. Nick, M. Hartweg, and L. Walz, ‘Single crys- tal refinements of seven Sr1−xCaxCuO2 structures (x = 0–0.573) and of Ca2−ySryCuO3 (y = 0 and 0.134),’ Z. Kristallogr. 209, 418 (1994).

[119] M. Braden, G. Wilkendorf, J. Lorenzana, M. Aïn, G. J. McIntyre, M. Behruzi, G. Heger, G. Dhalenne, and A. Revcolevschi, ‘Structural analysis of CuGeO3 : relation between nuclear structure and magnetic interaction,’ Phys. Rev. B 54, 1105 (1996).

[120] F. Sapiña, J. Rodríguez-Carvajal, M. Sanchis, R. Ibáñez, A. Beltrán, and D. Bel- trán, ‘Crystal and magnetic structure of Li2CuO2,’ Solid State Commun. 74, 779 (1990).

[121] M. A. Lafontaine, M. Leblanc, and G. Ferey, ‘New refinement of the room- temperature structure of LiCuVO4,’ Acta Crystallogr. Sect. C 45, 1205 (1989).

[122] W. Paszkowicz, M. Marczak, A. M. Vorotynov, K. A. Sablina, and G. A. Pe- trakovskii, ‘Powder diffraction study of LiCu2O2 crystals,’ Powder Diffr. 16, 30 (2001).

[123] V. Kiryukhin, Y. J. Kim, K. J. Thomas, F. C. Chou, R. W. Erwin, Q. Huang, M. A. Kastner, and R. J. Birgeneau, ‘Magnetic properties of the S = 1/2 quasi- one-dimensional antiferromagnet CaCu2O3,’ Phys. Rev. B 63, 144418 (2001).

[124] R. H. Buttner, E. N. Maslen, and N. Spadaccini, ‘Structure, electron density and thermal motion of KCuF3,’ Acta Crystallogr. Sect. B 46, 131 (1990).

[125] H. Suzuura, H. Yasuhara, A. Furusaki, N. Nagaosa, and Y. Tokura, ‘Singulari- ties in optical spectra of quantum spin chains,’ Phys. Rev. Lett. 76, 2579 (1996).

[126] J. Lorenzana and R. Eder, ‘Dynamics of the one-dimensional Heisenberg model and optical absorption of spinons in cuprate antiferromagnetic chains,’ Phys. Rev. B 55, 3358(R) (1997).

[127] C. de Graaf and F. Illas, ‘Electronic structure and magnetic interactions of the spin-chain compounds Ca2CuO3 and Sr2CuO3,’ Phys. Rev. B 63, 014404 (2000).

[128] N. Motoyama, H. Eisaki, and S. Uchida, ‘Magnetic susceptibility of ideal spin 1/2 Heisenberg antiferromagnetic chain systems, Sr2CuO3 and SrCuO2,’ Phys. Rev. Lett. 76, 3212 (1996).

[129] I. A. Zaliznyak, H. Woo, T. G. Perring, C. L. Broholm, C. D. Frost, and H. Tak- agi, ‘Spinons in the strongly correlated copper oxide chains in SrCuO2,’ Phys. Rev. Lett. 93, 087202 (2004).

[130] H. Rosner, H. Eschrig, R. Hayn, S.-L. Drechsler, and J. Málek, ‘Electronic structure and magnetic properties of the linear chain cuprates Sr2CuO3 and Ca2CuO3,’ Phys. Rev. B 56, 3402 (1997).

[131] H. Fujisawa, T. Yokoya, T. Takahashi, S. Miyasaka, M. Kibune, and H. Takagi, ‘Angle-resolved photoemission study of Sr2CuO3,’ Phys. Rev. B 59, 7358 (1999).

[132] B. J. Kim, H. Koh, E. Rotenberg, S.-J. Oh, H. Eisaki, N. Motoyama, S. Uchida, T. Tohyama, S. Maekawa, Z.-X. Shen, and C. Kim, ‘Distinct spinon and holon dispersions in photoemission spectral functions from one-dimensional SrCuO2,’ Nat. Phys. 2, 397 (2006).

[133] L. J. P. Ament, M. van Veenendaal, T. P. Devereaux, J. P. Hill, and J. van den Brink, ‘Resonant inelastic X-ray scattering studies of elementary excitations,’ Rev. Mod. Phys. 83, 705 (2011). bibliography 113

[134] H. Bethe, ‘Zur theorie der metalle,’ Z. Physik 71, 205 (1931).

[135] L. Hulthén, ‘Über das austauchsproblem eines kristalles,’ Arkiv Mat., Astron. Fysik 26A, 1 (1938), doctoral theisis.

[136] J. des Cloizeaux and J. J. Pearson, ‘Spin-wave spectrum of the antiferromag- netic linear chain,’ Phys. Rev. 128, 2131 (1962).

[137] L. Faddeev and L. Takhtajan, ‘What is the spin of a spin wave?’ Phys. Lett. A 85, 375 (1981).

[138] S. Kadota, I. Yamada, S. Yoneyama, and K. Hirakawa, ‘Formation of one- dimensional antiferromagnet in KCuF3 with the perovskite structure,’ J. Phys. Soc. Jpn. 23, 751 (1967).

[139] K. I. Kugel' and D. I. Khomskii, ‘The Jahn-Teller effect and magnetism: transi- tion metal compounds,’ Usp. Fiz. Nauk 136, 621 (1982).

[140] K. I. Kugel' and D. I. Khomskii, ‘The Jahn-Teller effect and magnetism: transi- tion metal compounds,’ Sov. Phys. Usp. 25, 231 (1982).

[141] D. A. Tennant, S. E. Nagler, D. Welz, G. Shirane, and K. Yamada, ‘Effects of coupling between chains on the magnetic excitation spectrum of KCuF3,’ Phys. Rev. B 52, 13381 (1995).

[142] D. A. Tennant, T. G. Perring, R. A. Cowley, and S. E. Nagler, ‘Unbound spinons in the S=1/2 antiferromagnetic chain KCuF3,’ Phys. Rev. Lett. 70, 4003 (1993).

[143] J. Deisenhofer, I. Leonov, M. V. Eremin, C. Kant, P. Ghigna, F. Mayr, V. V. Iglamov, V. I. Anisimov, and D. van der Marel, ‘Optical evidence for symmetry changes above the néel temperature of KCuF3,’ Phys. Rev. Lett. 101, 157406 (2008).

[144] M. Enderle, B. Fåk, H.-J. Mikeska, R. K. Kremer, A. Prokofiev, and W. Assmus, ‘Two-spinon and four-spinon continuum in a frustrated ferromagnetic spin-1/2 chain,’ Phys. Rev. Lett. 104, 237207 (2010).

[145] S.-L. Drechsler, S. Nishimoto, R. O. Kuzian, J. Málek, W. E. A. Lorenz, J. Richter, J. van den Brink, M. Schmitt, and H. Rosner, ‘Comment on “two-spinon and four-spinon continuum in a frustrated ferromagnetic spin-1/2 chain”,’ Phys. Rev. Lett. 106, 219701 (2011).

[146] S. Nishimoto, S.-L. Drechsler, R. O. Kuzian, J. van den Brink, J. Richter, W. E. A. Lorenz, Y. Skourski, R. Klingeler, and B. Büchner, ‘Saturation field of frustrated chain cuprates: broad regions of predominant interchain coupling,’ Phys. Rev. Lett. 107, 097201 (2011).

[147] M. Hase, I. Terasaki, and K. Uchinokura, ‘Observation of the spin-Peierls 2+ transition in linear Cu (spin-1/2) chains in an inorganic compound CuGeO3,’ Phys. Rev. Lett. 70, 3651 (1993).

[148] K. Hirota, D. E. Cox, J. E. Lorenzo, G. Shirane, J. M. Tranquada, M. Hase, K. Uchinokura, H. Kojima, Y. Shibuya, and I. Tanaka, ‘Dimerization of CuGeO3 in the spin-Peierls state,’ Phys. Rev. Lett. 73, 736 (1994).

[149] T. Masuda, A. Zheludev, B. Roessli, A. Bush, M. Markina, and A. Vasiliev, ‘Spin waves and magnetic interactions in LiCu2O2,’ Phys. Rev. B 72, 014405 (2005). 114 bibliography

[150] A. A. Gippius, E. N. Morozova, A. S. Moskvin, A. V. Zalessky, A. A. Bush, M. Baenitz, H. Rosner, and S.-L. Drechsler, ‘NMR and local-density- approximation evidence for spiral magnetic order in the chain cuprate LiCu2O2,’ Phys. Rev. B 70, 020406 (2004).

[151] S.-L. Drechsler, J. Málek, J. Richter, A. S. Moskvin, A. A. Gippius, and H. Ros- ner, ‘Comment on “competition between helimagnetism and commensurate quantum spin correlations in LiCu2O2 ”,’ Phys. Rev. Lett. 94, 039705 (2005).

[152] E. Dagotto, J. Riera, and D. Scalapino, ‘Superconductivity in ladders and coupled planes,’ Phys. Rev. B 45, 5744 (1992).

[153] T. Barnes, E. Dagotto, J. Riera, and E. S. Swanson, ‘Excitation spectrum of Heisenberg spin ladders,’ Phys. Rev. B 47, 3196 (1993).

[154] M. Onoda and N. Nishiguchi, ‘Letter to the editor: crystal structure and spin gap state of CaV2O5,’ J. Solid State Chem. 127, 359 (1996).

[155] H. Iwase, M. Isobe, Y. Ueda, and H. Yasuoka, ‘Observation of spin gap in CaV2O5 by NMR,’ J. Phys. Soc. Jpn. 65, 2397 (1996).

[156] M. Azuma, Z. Hiroi, M. Takano, K. Ishida, and Y. Kitaoka, ‘Observation of a spin gap in SrCu2O3 comprising spin-½ quasi-1D two-leg ladders,’ Phys. Rev. Lett. 73, 3463 (1994).

[157] E. McCarron, M. Subramanian, J. Calabrese, and R. Harlow, ‘The incommensu- rate structure of (Sr14−xCax)Cu24O41 (0 < x < 8) a superconductor byproduct,’ Mat. Res. Bull. 23, 1355 (1988).

[158] T. Siegrist, L. Schneemeyer, S. Sunshine, J. Waszczak, and R. Roth, ‘A new layered cuprate structure-type, (A1−xA0x)14Cu24O41,’ Mat. Res. Bull. 23, 1429 (1988).

[159] M. Uehara, T. Nagata, J. Akimitsu, H. Takahashi, N. Môri, and K. Kinoshita, ‘Superconductivity in the ladder material Sr0.4Ca13.6Cu24O41.84,’ J. Phys. Soc. Jpn. 65, 2764 (1996).

[160] E. Bordas, C. de Graaf, R. Caballol, and C. J. Calzado, ‘Electronic structure of CaCu2O3 : spin ladder versus one-dimensional spin chain,’ Phys. Rev. B 71, 045108 (2005).

[161] R. Rückamp, J. Baier, M. Kriener, M. W. Haverkort, T. Lorenz, G. S. Uhrig, L. Jongen, A. Möller, G. Meyer, and M. Grüninger, ‘Zero-field incommensurate spin-Peierls phase with interchain frustration in TiOCl,’ Phys. Rev. Lett. 95, 097203 (2005).

[162] R. Rückamp, E. Benckiser, M. W. Haverkort, H. Roth, T. Lorenz, A. Freimuth, L. Jongen, A. Müller, G. Meyer, P. Reutler, B. Büchner, A. Revcolevschi, S.-W. Cheong, C. Sekar, G. Krabbes, and M. Grüninger, ‘Optical study of orbital excitations in transition-metal oxides,’ New J. Phys. 7, 144 (2005).

[163] A. Seidel, C. A. Marianetti, F. C. Chou, G. Ceder, and P. A. Lee, ‘S=1/2 chains and spin-Peierls transition in TiOCl,’ Phys. Rev. B 67, 020405 (2003).

[164] T. Saha-Dasgupta, R. Valentí, H. Rosner, and C. Gros, ‘TiOCl, an orbital- ordered system?’ Europhys. Lett. 67, 63 (2004).

[165] T. Saha-Dasgupta, A. Lichtenstein, and R. Valentí, ‘Correlation effects on the electronic structure of TiOCl: A NMTO+DMFT study,’ Phys. Rev. B 71, 153108 (2005). bibliography 115

[166] L. Craco, M. S. Laad, and E. Müller-Hartmann, ‘Metallizing the Mott insulator TiOCl by electron doping,’ J. Phys.: Condens. Matter 18, 10943 (2006).

[167] S. Glawion, M. R. Scholz, Y.-Z. Zhang, R. Valentí, T. Saha-Dasgupta, M. Klemm, J. Hemberger, S. Horn, M. Sing, and R. Claessen, ‘Electronic structure of the two-dimensional Heisenberg antiferromagnet VOCl: A multiorbital Mott insu- lator,’ Phys. Rev. B 80, 155119 (2009).

[168] D. Fausti, T. T. A. Lummen, C. Angelescu, R. Macovez, J. Luzon, R. Broer, P. Rudolf, P. H. M. van Loosdrecht, N. Tristan, B. Büchner, S. van Smaalen, A. Möller, G. Meyer, and T. Taetz, ‘Symmetry disquisition on the TiOX phase diagram (X=Br,Cl),’ Phys. Rev. B 75, 245114 (2007).

[169] S. Glawion, J. Heidler, M. W. Haverkort, L. C. Duda, T. Schmitt, V. N. Strocov, C. Monney, K. J. Zhou, A. Ruff, M. Sing, and R. Claessen, ‘Two-spinon and orbital excitations of the spin-Peierls system TiOCl,’ Phys. Rev. Lett. 107, 107402 (2011).

[170] E. Benckiser, R. Rückamp, T. Möller, T. Taetz, A. Möller, A. A. Nugroho, T. T. M. Palstra, G. S. Uhrigm, and M. Grüninger, ‘Collective orbital excitations in or- bitally ordered YVO3 and HoVO3,’ New J. Phys. 10, 053027 (2008).

[171] E. M. Snigireva, S. I. Troyanov, and V. B. Rybakov, ‘Specification of crystalline- structure of TiOCl,’ Zh. Neorg. Khim. 35, 1945 (1990), ICSD code 39314.

[172] A. Haase and G. Brauer, ‘Vanadium oxychloride,’ Acta Crystallogr. B 31, 2521 (1975).

[173] M. D. Towler, N. M. Harrison, and M. I. McCarthy, ‘Ab initio study of the surface and interfacial properties of a layered MgO/NiO film,’ Phys. Rev. B 52, 5375 (1995).

[174] M. Prencipe, A. Zupan, R. Dovesi, E. Aprà, and V. R. Saunders, ‘Ab initio study of the structural properties of LiF, NaF, KF, LiCl, NaCl, and KCl,’ Phys. Rev. B 51, 3391 (1995).

[175] W. C. Mackrodt, E.-A. Simson, and N. M. Harrison, ‘An ab initio Hartree–Fock study of the electron-excess gap states in oxygen-deficient rutile TiO2,’ Surf. Sci. 384, 192 (1997).

[176] M. Catti, G. Sandrone, and R. Dovesi, ‘Periodic unrestricted Hartree–Fock study of corundumlike Ti2O3 and V2O3,’ Phys. Rev. B 55, 16122 (1997).

[177] A. Schäfer, H. Horn, and R. Ahlrichs, ‘Fully optimized contracted Gaussian basis sets for atoms Li to Kr,’ J. Chem. Phys. 97, 2571 (1992).

[178] M. Shaz, S. van Smaalen, L. Palatinus, M. Hoinkis, M. Klemm, S. Horn, and R. Claessen, ‘Spin-Peierls transition in TiOCl,’ Phys. Rev. B 71, 100405 (2005).

[179] T. Sasaki, T. Nagai, K. Kato, M. Mizumaki, T. Asaka, M. Takata, Y. Matsui, H. Sawa, and J. Akimitsu, ‘Crystal symmetry and superlattice reflections in spin-Peierls system TiOBr,’ Sci. Technol. Adv. Mater. 7, 17 (2006).

[180] P. Lemmens, K. Y. Choi, G. Caimi, L. Degiorgi, N. N. Kovaleva, A. Seidel, and F. C. Chou, ‘Giant phonon softening in the pseudogap phase of the quantum spin system TiOCl,’ Phys. Rev. B 70, 134429 (2004).

[181] M. D. Banus, T. B. Reed, and A. J. Strauss, ‘Electrical and magnetic properties of TiO and VO,’ Phys. Rev. B 5, 2775 (1972).

[182] A. D. Rata, V. Kataev, D. Khomskii, and T. Hibma, ‘Giant positive magnetore- sistance in metallic VOx thin films,’ Phys. Rev. B 68, 220403 (2003). 116 bibliography

[183] C. de Graaf, L. Hozoi, and R. Broer, ‘Magnetic interactions in calcium and sodium ladder vanadates,’ J. Chem. Phys. 120, 961 (2004).

[184] L. Hozoi, A. H. de Vries, A. B. van Oosten, R. Broer, J. Cabrero, and C. de Graaf, ‘Theoretical characterization of the ground and optically excited 0 states of α -NaV2O5,’ Phys. Rev. Lett. 89, 076407 (2002).

[185] N. Suaud and M.-B. Lepetit, ‘Ab Initio evaluation of the charge ordering in 0 α -NaV2O5,’ Phys. Rev. Lett. 88, 056405 (2002).

[186] A. Wiedenmann, J. P. Venien, P. Palvadeau, and J. Rossat-Mignod, ‘Magnetic ordering of the quasi-two-dimensional system VOCl,’ J. Phys. C: Solid State Phys. 16, 5339 (1983).

[187] M. van Veenendaal, Theory of Inelastic Scattering and Absorption of X-rays (Cam- bridge University Press, 2015).

[188] L. Yang, H. Ågren, V. Carravetta, and L. G. M. Pettersson, ‘Static exchange and quantum defect analysis of x-ray absorption spectra of carbonyl compounds,’ Phys. Scr. 54, 614 (1996).

[189] R. V. Pinjari, M. G. Delcey, M. Guo, M. Odelius, and M. Lundberg, ‘Restricted active space calculations of L-edge X-ray absorption spectra: From molecular orbitals to multiplet states,’ J. Chem. Phys. 141, 124116 (2014).

[190] K. Kunnus, W. Zhang, M. G. Delcey, R. V. Pinjari, P. S. Miedema, S. Schreck, W. Quevedo, H. Schröder, A. Föhlisch, K. J. Gaffney, M. Lundberg, M. Odelius, and P. Wernet, ‘Viewing the valence electronic structure of ferric and ferrous hexacyanide in solution from the Fe and cyanide perspectives,’ J. of Phys. Chem. B 120, 7182 (2016).

[191] M. Guo, E. Källman, L. K. Sørensen, M. G. Delcey, R. V. Pinjari, and M. Lund- berg, ‘Molecular orbital simulations of metal 1s2p resonant inelastic X-ray scat- tering,’ J. Phys. Chem. A 120, 5848 (2016).

[192] L. Braicovich, L. J. P. Ament, V. Bisogni, F. Forte, C. Aruta, G. Balestrino, N. B. Brookes, G. M. De Luca, P. G. Medaglia, F. M. Granozio, M. Radovic, M. Sal- luzzo, J. van den Brink, and G. Ghiringhelli, ‘Dispersion of magnetic excitations in the cuprate La2CuO4 and CaCuO2 compounds measured using resonant X- ray scattering,’ Phys. Rev. Lett. 102, 167401 (2009).

[193] C. Ulrich, L. J. P. Ament, G. Ghiringhelli, L. Braicovich, M. Moretti Sala, N. Pez- zotta, T. Schmitt, G. Khaliullin, J. van den Brink, H. Roth, T. Lorenz, and B. Keimer, ‘Momentum dependence of orbital excitations in Mott-insulating titanates,’ Phys. Rev. Lett. 103, 107205 (2009).

[194] W. S. Lee, S. Johnston, B. Moritz, J. Lee, M. Yi, K. J. Zhou, T. Schmitt, L. Patthey, V. Strocov, K. Kudo, Y. Koike, J. van den Brink, T. P. Devereaux, and Z. X. Shen, ‘Role of lattice coupling in establishing electronic and magnetic properties in quasi-one-dimensional cuprates,’ Phys. Rev. Lett. 110, 265502 (2013).

[195] S. Johnston, C. Monney, V. Bisogni, K.-J. Zhou, R. Kraus, G. Behr, V. N. Strocov, J. Málek, S.-L. Drechsler, J. Geck, T. Schmitt, and J. van den Brink, ‘Electron- lattice interactions strongly renormalize the charge-transfer energy in the spin- chain cuprate Li2CuO2,’ Nat. Commun. 7, 10563 (2016).

[196] M. P. M. Dean, R. S. Springell, C. Monney, K. J. Zhou, J. Pereiro, I. Božovi´c, B. Dalla Piazza, H. M. Rønnow, E. Morenzoni, J. van den Brink, T. Schmitt, and J. P. Hill, ‘Spin excitations in a single La2CuO4 layer,’ Nature Mat. 11, 850 (2012). bibliography 117

[197] M. Minola, D. Di Castro, L. Braicovich, N. B. Brookes, D. Innocenti, M. Moretti Sala, A. Tebano, G. Balestrino, and G. Ghiringhelli, ‘Magnetic and ligand field properties of copper at the interfaces of (CaCuO2)n/(SrTiO3)n su- perlattices,’ Phys. Rev. B 85, 235138 (2012).

[198] L. J. P. Ament, F. Forte, and J. van den Brink, ‘Ultrashort lifetime expansion for indirect resonant inelastic x-ray scattering,’ Phys. Rev. B 75, 115118 (2007).

[199] N. Pakhira, J. K. Freericks, and A. M. Shvaika, ‘Resonant inelastic x-ray scat- tering in a Mott insulator,’ Phys. Rev. B 86, 125103 (2012).

[200] J. Igarashi and T. Nagao, ‘Elementary excitations probed L-edge resonant in- elastic x-ray scattering in systems with weak and intermediate electron correla- tions,’ Phys. Rev. B 88, 014407 (2013).

[201] T. Tohyama, K. Tsutsui, M. Mori, S. Sota, and S. Yunoki, ‘Enhanced charge excitations in electron-doped cuprates by resonant inelastic x-ray scattering,’ Phys. Rev. B 92, 014515 (2015).

[202] D. Benjamin, I. Klich, and E. Demler, ‘Probing competing and intertwined orders with resonant inelastic x-ray scattering in the hole-doped cuprates,’ Phys. Rev. B 92, 035151 (2015).

[203] L. A. Wray, W. Yang, H. Eisaki, Z. Hussain, and Y.-D. Chuang, ‘Multiplet resonance lifetimes in resonant inelastic x-ray scattering involving shallow core levels,’ Phys. Rev. B 86, 195130 (2012).

[204] H. Ikeno, T. Mizoguchi, and I. Tanaka, ‘Ab initio charge transfer multiplet calculations on L{2, 3}, XANES and ELNES of 3d transition metal oxides,’ Phys. Rev. B 83, 155107 (2011).

[205] M. W. Haverkort, M. Zwierzycki, and O. K. Andersen, ‘Multiplet ligand-field theory using Wannier orbitals,’ Phys. Rev. B 85, 165113 (2012).

[206] I. Josefsson, K. Kunnus, S. Schreck, A. Föhlisch, F. de Groot, P. Wernet, and M. Odelius, ‘Ab Initio calculations of X-ray spectra: Atomic multiplet and molec- ular orbital effects in a multiconfigurational SCF approach to the L-edge spectra of transition metal complexes,’ J. Phys. Chem. Lett. 3, 3565 (2012).

[207] K. Kunnus, I. Josefsson, S. Schreck, W. Quevedo, P. S. Miedema, S. Techert, F. M. F. de Groot, M. Odelius, P. Wernet, and A. Föhlisch, ‘From ligand fields to molecular orbitals: Probing the local valence electronic structure of Ni2+ in aqueous solution with resonant inelastic X-ray scattering,’ J. Phys. Chem. B 117, 16512 (2013).

[208] D. Manganas, P. Kristiansen, L.-C. Duda, A. Knop-Gericke, S. DeBeer, R. Schlögl, and F. Neese, ‘Combined experimental and ab Initio multirefer- ence configuration interaction study of the resonant inelastic X-ray scattering spectrum of CO2,’ J. Phys. Chem. C 118, 20163 (2014).

[209] R. Broer and W. C. Nieuwpoort, ‘Broken orbital-symmetry and the description − of hole states in the tetrahedral [cro4] anion. I. Introductory considerations and calculations on oxygen 1s hole states,’ Chem. Phys. 54, 291 (1981).

[210] P. Å. Malmqvist, ‘Calculation of transition density matrices by nonunitary or- bital transformations,’ Int. J. Quantum Chem. 30, 479 (1986).

[211] A. O. Mitrushchenkov and H.-J. Werner, ‘Calculation of transition moments between internally contracted MRCI wave functions with non-orthogonal or- bitals,’ Mol. Phys. 105, 1239 (2007). 118 bibliography

[212] M. Blume, ‘Magnetic scattering of X-rays (invited),’ J. Appl. Phys. 57, 3615 (1985).

[213] P. A. M. Dirac, The principles of quantum mechanics, fourth (revised) ed. (Oxford University Press, 1958).

[214] L. I. Schiff, Quantum mechanics, third (internatinal student) ed. (McGraw-Hill Companies, 1968).

[215] P. A. M. Dirac, ‘On the theory of quantum mechanics,’ Proc. R. Soc. A 112, 661 (1926).

[216] T. D. Visser, ‘Whose golden rule is it anyway?’ Am. J. Phys. 77, 487 (2009).

[217] V. Weisskopf and E. Wigner, ‘Berechnung der natürlichen Linienbreite auf Grund der Diracschen Lichttheorie,’ Z. Phys. 63, 54 (1930).

[218] W. Heitler, Quantum theory of radiation, 3rd ed. (Oxford University Press, 1954).

[219] K. Hamalainen, S. Manninen, P. Suortti, S. P. Collins, M. J. Cooper, and D. Laundy, ‘Resonant raman scattering and inner-shell hole widths in Cu, Zn and Ho,’ J. Phys.: Condens. Matter 1, 5955 (1989).

[220] W. Schülke, Electron dynamics by inelastic X-Ray scattering (Oxford University Press, 2007).

[221] H. Kramers and W. Heisenberg, ‘Über die Streuung von Strahlung durch Atome,’ Z. Phys. 31, 681 (1925).

[222] M. Douglas and N. M. Kroll, ‘Quantum electrodynamical corrections to the fine structure of helium,’ Ann. Phys. 82, 89 (1974).

[223] B. A. Hess, ‘Relativistic electronic-structure calculations employing a two- component no-pair formalism with external-field projection operators,’ Phys. Rev. A 33, 3742 (1986).

[224] N. B. Balabanov and K. A. Peterson, ‘Systematically convergent basis sets for transition metals. I. All-electron correlation consistent basis sets for the 3d ele- ments Sc–Zn,’ J. Chem. Phys. 123, 064107 (2005).

[225] T. H. Dunning, ‘Gaussian basis sets for use in correlated molecular calcula- tions. I. the atoms boron through neon and hydrogen,’ J. Chem. Phys. 90, 1007 (1989).

[226] B. P. Prascher, D. E. Woon, K. A. Peterson, T. H. Dunning Jr., and A. K. Wilson, ‘Gaussian basis sets for use in correlated molecular calculations. VII. Valence, core-valence, and scalar relativistic basis sets for Li, Be, Na, and Mg,’ Theor. Chem. Acc. 128, 69 (2011).

[227] B. O. Roos, ed., Lecture notes in quantum chemistry I (European summer school in quantum chemistry) (Springer Verlag, Berlin, 1992).

[228] J. Pipek and P. G. Mezey, ‘A fast intrinsic localization procedure applicable for ab initio and semiempirical linear combination of atomic orbital wave functions,’ J. Chem. Phys. 90, 4916 (1989).

[229] P. S. Bagus, R. Broer, C. de Graaf, and W. Nieuwpoort, ‘The electronic struc- ture of NiO for ni 3s-hole states including full orbital relaxation and localiza- tion,’ J. Electron. Spectrosc. 98-99, 303 (1999).

[230] L. Hozoi, A. de Vries, R. Broer, C. de Graaf, and P. Bagus, ‘Ni 3s-hole states in NiO by non-orthogonal configuration interaction,’ Chem. Phys. 331, 178 (2006). bibliography 119

[231] L. Hozoi and M. S. Laad, ‘Quasiparticle bands in cuprates by quantum- chemical methods: Towards an Ab initio description of strong electron corre- lations,’ Phys. Rev. Lett. 99, 256404 (2007).

[232] A. van Oosten, R. Broer, and W. Nieuwpoort, ‘Heisenberg exchange enhance- ment by orbital relaxation in cuprate compounds,’ Chem. Phys. Lett. 257, 207 (1996).

[233] N. Wizent, N. Leps, G. Behr, R. Klingeler, B. Büchner, and W. Löser, ‘The effect of process parameters on floating zone crystal growth of selected cuprates,’ J. Cryst. Growth 401, 596 (2014).

[234] S. Carniato, R. Guillemin, W. C. Stolte, L. Journel, R. Taïeb, D. W. Lindle, and M. Simon, ‘Experimental and theoretical investigation of molecular field effects by polarization-resolved resonant inelastic X-ray scattering,’ Phys. Rev. A 80, 032513 (2009).

[235] M. Kavˇciˇc,M. Žitnik, K. Buˇcar, A. Miheliˇc,S. Carniato, L. Journel, R. Guillemin, and M. Simon, ‘Electronic state interferences in resonant X-ray emission after K-shell excitation in HCl,’ Phys. Rev. Lett. 105, 113004 (2010).

[236] J. E. Hirsch, ‘Dynamic Hubbard model for solids with hydrogen-like atoms,’ Phys. Rev. B 90, 104501 (2014).

[237] J. S. Zhou, J. J. Kas, L. Sponza, I. Reshetnyak, M. Guzzo, C. Giorgetti, M. Gatti, F. Sottile, J. J. Rehr, and L. Reining, ‘Dynamical effects in electron spectroscopy,’ J. Chem. Phys. 143, 184109 (2015).

[238] J. J. Kas, F. D. Vila, J. J. Rehr, and S. A. Chambers, ‘Real-time cumulant ap- proach for charge-transfer satellites in x-ray photoemission spectra,’ Phys. Rev. B 91, 121112 (2015).

[239] J. Z. Tischler, B. C. Larson, P. Zschack, A. Fleszar, and A. G. Eguiluz, ‘Interplay between inelastic X-ray scattering and ab initio density-response calculations: Insight into the electronic correlations in aluminum,’ Phys. Status Solidi B 237, 280 (2003).

[240] I. G. Gurtubay, J. M. Pitarke, W. Ku, A. G. Eguiluz, B. C. Larson, J. Tischler, P. Zschack, and K. D. Finkelstein, ‘Electron-hole and plasmon excitations in 3d transition metals: Ab initio calculations and inelastic x-ray scattering measure- ments,’ Phys. Rev. B 72, 125117 (2005).

[241] B. J. Kim, H. Ohsumi, T. Komesu, S. Sakai, T. Morita, H. Takagi, and T. Arima, ‘Phase-sensitive observation of a spin-orbital Mott state in Sr2IrO4,’ Science 323, 1329 (2009).

[242] D. Pesin and L. Balents, ‘Mott physics and band topology in materials with strong spin-orbit interaction,’ Nat. Phys. 6, 376 (2010).

[243] J. Chaloupka, G. Jackeli, and G. Khaliullin, ‘Kitaev-Heisenberg model on a honeycomb lattice: Possible exotic phases in iridium oxides A2IrO3,’ Phys. Rev. Lett. 105, 027204 (2010).

[244] D. Mandrus, J. R. Thompson, R. Gaal, L. Forro, J. C. Bryan, B. C. Chak- oumakos, L. M. Woods, B. C. Sales, R. S. Fishman, and V. Keppens, ‘Continuous metal-insulator transition in the pyrochlore Cd2Os2O7,’ Phys. Rev. B 63, 195104 (2001).

[245] Y. H. Matsuda, J. L. Her, S. Michimura, T. Inami, M. Suzuki, N. Kawamura, M. Mizumaki, K. Kindo, J. Yamauara, and Z. Hiroi, ‘Orbital magnetism in Cd2Os2O7 studied by X-ray magnetic circular dichroism,’ Phys. Rev. B 84, 174431 (2011). 120 bibliography

[246] Y. G. Shi, Y. F. Guo, S. Yu, M. Arai, A. A. Belik, A. Sato, K. Yamaura, E. Takayama-Muromachi, H. F. Tian, H. X. Yang, J. Q. Li, T. Varga, J. F. Mitchell, and S. Okamoto, ‘Continuous metal-insulator transition of the antiferromag- netic perovskite NaOsO3,’ Phys. Rev. B 80, 161104 (2009).

[247] S. Calder, V. O. Garlea, D. F. McMorrow, M. D. Lumsden, M. B. Stone, J. C. Lang, J.-W. Kim, J. A. Schlueter, Y. G. Shi, K. Yamaura, Y. S. Sun, Y. Tsuji- moto, and A. D. Christianson, ‘Magnetically driven metal-insulator transition in NaOsO3,’ Phys. Rev. Lett. 108, 257209 (2012).

[248] H. Shinaoka, T. Miyake, and S. Ishibashi, ‘Noncollinear magnetism and spin- orbit coupling in 5d pyrochlore oxide Cd2Os2O7,’ Phys. Rev. Lett. 108, 247204 (2012).

[249] J. Yamaura, K. Ohgushi, H. Ohsumi, T. Hasegawa, I. Yamauchi, K. Sugimoto, S. Takeshita, A. Tokuda, M. Takata, M. Udagawa, M. Takigawa, H. Harima, T. Arima, and Z. Hiroi, ‘Tetrahedral magnetic order and the metal-insulator transition in the pyrochlore lattice of Cd2Os2O7,’ Phys. Rev. Lett. 108, 247205 (2012).

[250] D. J. Singh, P. Blaha, K. Schwarz, and J. O. Sofo, ‘Electronic structure of the pyrochlore metals Cd2Os2O7 and Cd2Re2O7,’ Phys. Rev. B 65, 155109 (2002).

[251] D. Figgen, K. A. Peterson, M. Dolg, and H. Stoll, ‘Energy-consistent pseu- dopotentials and correlation consistent basis sets for the 5d elements Hf–Pt,’ J. Chem. Phys. 130, 164108 (2009).

[252] K. Pierloot, B. Dumez, P.-O. Widmark, and B. O. Roos, ‘Density matrix av- eraged atomic natural orbital (ANO) basis sets for correlated molecular wave functions,’ Theor. Chim. Acta 90, 87 (1995).

[253] F. Schautz, H.-J. Flad, and M. Dolg, ‘Quantum Monte Carlo study of Be2 and group 12 dimers M2 (M = Zn, Cd, Hg),’ Theor. Chem. Acc. 99, 23 (1998).

[254] C. J. Ballhausen, Introduction to ligand field theory (McGraw-Hill, New York, 1962).

[255] J. S. Griffith, The theory of transition-metal ions, 5th ed. (Cambridge University Press, 1961).

[256] R. Boˇca,‘Zero-field splitting in metal complexes,’ Coord. Chem. Rev. 248, 757 (2004).

257 3 47 [ ] R. M. Macfarlane, ‘Zero-field splittings of t2 cubic terms,’ J. Chem. Phys. , 2066 (1967).

[258] R. M. Macfarlane, ‘Analysis of the spectrum of d3 ions in trigonal crystal fields,’ J. Chem. Phys. 39, 3118 (1963).

[259] R. Maurice, R. Bastardis, C. de Graaf, N. Suaud, T. Mallah, and N. Gui- héry, ‘Universal theoretical approach to extract anisotropic spin Hamiltonians,’ J. Chem. Theory Comput. 5, 2977 (2009).

[260] J. des Cloizeaux, ‘Extension d'une formule de Lagrange à des problèmes de valeurs propres,’ Nucl. Phys. 20, 321 (1960).

[261] M. A. Subramanian, G. Aravamudan, and G. V. Subba Rao, ‘Oxide py- rochlores — a review,’ Prog. Solid State Chem. 15, 55 (1983). bibliography 121

[262] Y.-F. Li, X.-Y. Kuang, M.-L. Gao, Y.-R. Zhao, and H.-Q. Wang, ‘Effect of local structure on electron paramagnetic resonance spectra for trigonal 3+ [Cr(H2O)6] coordination complex in the sulfate alums series: a ligand field theory study,’ Chin. Phys. B 18, 2967 (2009).

[263] R. Maurice, L. Vendier, and J.-P. Costes, ‘Magnetic anisotropy in NiII –YIII binuclear complexes: on the importance of both the first coordination sphere of the NiII ion and the YIII ion belonging to the second coordination sphere,’ Inorg. Chem. 50, 11075 (2011).

[264] V. M. Katukuri, H. Stoll, J. van den Brink, and L. Hozoi, ‘Ab initio determi- nation of excitation energies and magnetic couplings in correlated quasi-two- dimensional iridates,’ Phys. Rev. B 85, 220402 (2012).

[265] R. Maurice, A. M. Pradipto, N. Guihéry, R. Broer, and C. de Graaf, ‘Anti- symmetric magnetic interactions in oxo-bridged copper(II) bimetallic systems,’ J. Chem. Theory Comput. 6, 3092 (2010).

[266] A.-M. Pradipto, R. Maurice, N. Guihéry, C. de Graaf, and R. Broer, ‘First- principles study of magnetic interactions in cupric oxide,’ Phys. Rev. B 85, 014409 (2012).

[267] M. Elhajal, B. Canals, R. Sunyer, and C. Lacroix, ‘Ordering in the pyrochlore antiferromagnet due to Dzyaloshinskii-Moriya interactions,’ Phys. Rev. B 71, 094420 (2005).

[268] X. Wan, A. M. Turner, A. Vishwanath, and S. Y. Savrasov, ‘Topological semimetal and Fermi-arc surface states in the electronic structure of pyrochlore iridates,’ Phys. Rev. B 83, 205101 (2011).

[269] X. Liu, V. M. Katukuri, L. Hozoi, W.-G. Yin, M. P. M. Dean, M. H. Upton, J. Kim, D. Casa, A. Said, T. Gog, T. F. Qi, G. Cao, A. M. Tsvelik, J. van den Brink, and J. P. Hill, ‘Testing the validity of the strong spin-orbit-coupling limit for octahedrally coordinated iridate compounds in a model system Sr3CuIrO6,’ Phys. Rev. Lett. 109, 157401 (2012).

[270] H. Gretarsson, J. P. Clancy, X. Liu, J. P. Hill, E. Bozin, Y. Singh, S. Manni, P. Gegenwart, J. Kim, A. H. Said, D. Casa, T. Gog, M. H. Upton, H.-S. Kim, J. Yu, V. M. Katukuri, L. Hozoi, J. van den Brink, and Y.-J. une Kim, ‘Crystal- field splitting and correlation effect on the electronic structure of A2IrO3,’ Phys. Rev. Lett. 110, 076402 (2013).

[271] L. Hozoi, H. Gretarsson, J. P. Clancy, B.-G. Jeon, B. Lee, K. H. Kim, V. Yushankhai, P. Fulde, D. Casa, T. Gog, J. Kim, A. H. Said, M. H. Upton, Y. J. Kim, and J. van den Brink, ‘Longer-range lattice anisotropy strongly compet- ing with spin-orbit interactions in pyrochlore iridates,’ Phys. Rev. B 89, 115111 (2014).

[272] J. N. Reimers, ‘Absence of long-range order in a three-dimensional geometri- cally frustrated antiferromagnet,’ Phys. Rev. B 45, 7287 (1992).

[273] J. N. Reimers, A. J. Berlinsky, and A.-C. Shi, ‘Mean-field approach to magnetic ordering in highly frustrated pyrochlores,’ Phys. Rev. B 43, 865 (1991).

[274] R. Moessner and J. T. Chalker, ‘Properties of a classical spin liquid: the Heisen- berg pyrochlore antiferromagnet,’ Phys. Rev. Lett. 80, 2929 (1998).

[275] R. Moessner and J. T. Chalker, ‘Low-temperature properties of classical geo- metrically frustrated antiferromagnets,’ Phys. Rev. B 58, 12049 (1998). 122 bibliography

[276] R. Moessner, ‘Relief and generation of frustration in pyrochlore magnets by single-ion anisotropy,’ Phys. Rev. B 57, 5587 (1998).

[277] B.-J. Yang and Y. B. Kim, ‘Topological insulators and metal-insulator transition in the pyrochlore iridates,’ Phys. Rev. B 82, 085111 (2010).

[278] R. Maurice, A.-M. Pradipto, C. de Graaf, and R. Broer, ‘Magnetic interactions in LiCu2O2: single-chain versus double-chain models,’ Phys. Rev. B 86, 024411 (2012).

[279] S. T. Bramwell, M. J. P. Gingras, and J. N. Reimers, ‘Order by disorder in an anisotropic pyrochlore lattice antiferromagnet,’ J. Appl. Phys. 75, 5523 (1994).

[280] J. D. M. Champion and P. C. W. Holdsworth, ‘Soft modes in the easy plane pyrochlore antiferromagnet,’ J. Phys. Condens. Matter 16,S665 (2004).

[281] J. Heber, ‘Materials science: enter the oxides,’ Nature 459, 28 (2009).

[282] J. Mannhart and D. G. Schlom, ‘Oxide interfaces – an opportunity for electron- ics,’ Science 327, 1607 (2010).

[283] A. J. Leggett, ‘What DO we know about high Tc?’ Nature Phys. 2, 134 (2006).

[284] B. J. Kim, H. Jin, S. J. Moon, J.-Y. Kim, B.-G. Park, C. S. Leem, J. Yu, T. W. Noh, C. Kim, S.-J. Oh, J.-H. Park, V. Durairaj, G. Cao, and E. Rotenberg, ‘Novel 1 2 jeff = / Mott state induced by relativistic spin-orbit coupling in Sr2IrO4,’ Phys. Rev. Lett. 101, 076402 (2008).

[285] B. J. Kim, H. Ohsumi, T. Komesu, S. Sakai, T. Morita, H. Takagi, and T. Arima, ‘Phase-sensitive observation of a spin-orbital Mott state in Sr2IrO4,’ Science 323, 1329 (2009).

[286] J. Kim, D. Casa, M. H. Upton, T. Gog, Y.-J. Kim, J. F. Mitchell, M. van Veenen- daal, M. Daghofer, J. van den Brink, G. Khaliullin, and B. J. Kim, ‘Magnetic excitation spectra of Sr2IrO4 probed by resonant inelastic x-ray scattering: es- tablishing links to cuprate superconductors,’ Phys. Rev. Lett. 108, 177003 (2012).

[287] Y. K. Kim, O. Krupin, J. D. Denlinger, A. Bostwick, E. Rotenberg, Q. Zhao, J. F. Mitchell, J. W. Allen, and B. J. Kim, ‘Fermi arcs in a doped pseudospin-1/2 Heisenberg antiferromagnet,’ Science 345, 187 (2014).

[288] J. Kim, M. Daghofer, A. H. Said, T. Gog, J. van den Brink, G. Khaliullin, and B. J. Kim, ‘Excitonic quasiparticles in a spin-orbit Mott insulator,’ Nat. Commun. 5, 4453 (2014).

[289] A. Abragam and B. Bleaney, Electron paramagnetic resonance of transition ions (Clarendon Press, Oxford, 1970).

[290] C. J. Calzado and J.-P. Malrieu, ‘Origin and evaluation of the four-spin opera- tors in magnetic lattices,’ Phys. Rev. B 69, 094435 (2004).

[291] R. Bastardis, N. Guihéry, and C. de Graaf, ‘Microscopic origin of isotropic non-Heisenberg behavior in S = 1 magnetic systems,’ Phys. Rev. B 76, 132412 (2007).

[292] V. M. Katukuri, V. Yushankhai, L. Siurakshina, J. van den Brink, L. Hozoi, and I. Rousochatzakis, ‘Mechanism of basal-plane antiferromagnetism in the spin-orbit driven iridate Ba2IrO4,’ Phys. Rev. X 4, 021051 (2014).

[293] M. E. Lines, ‘Orbital angular momentum in the theory of paramagnetic clus- ters,’ J. Chem. Phys. 55, 2977 (1971). bibliography 123

[294] S. Petit, G. Pilet, D. Luneau, L. F. Chibotaru, and L. Ungur, ‘A dinuclear cobalt(ii) complex of calix[8]arenes exibiting strong magnetic anisotropy,’ Dal- ton Trans. , 4582 (2007).

[295] L. F. Chibotaru, L. Ungur, C. Aronica, H. Elmoll, G. Pilet, and D. Luneau, II III ‘Structure, magnetism, and theoretical study of a mixed-valence Co3 Co4 hep- tanuclear wheel: Lack of SMM behavior despite negative magnetic anisotropy,’ J. Am. Chem. Soc. 130, 12445 (2008).

[296] S. Boseggia, R. Springell, H. C. Walker, H. M. Rønnow, C. Rüegg, H. Okabe, M. Isobe, R. S. Perry, S. P. Collins, and D. F. McMorrow, ‘Robustness of basal- 1 2 plane antiferromagnetic order and the Jeff = / state in single-layer iridate spin-orbit Mott insulators,’ Phys. Rev. Lett. 110, 117207 (2013).

2− [297] J. H. M. Thornley, ‘The magnetic properties of (IrX6) complexes,’ J. Phys. C (Proc. Phys. Soc.) 1, 1024 (1968).

[298] N. J. Hill, ‘Electron paramagnetic resonance of osmium-doped trichlorotris (diethylphenylphosphine) rhodium (III),’ J. Chem. Soc., Faraday Trans. 2 68, 427 (1972).

[299] B. Andlauer, J. Schneider, and W. Tolksdorf, ‘Optical absorption, fluorescence, 4+ and electron spin resonance of Ir on octahedral sites in Y3Ga5O12,’ Phys. Stat. Sol. B 73, 533 (1976).

[300] A. Stebler, J. M. Ammeter, U. Fürholz, and A. Ludi, ‘Electronic properties 3+ of hexaaquaruthenium (III): EPR and optical spectra of Ru(H2O)6 in alum crystals,’ Inorg. Chem. 23, 2764 (1984).

[301] D. Haskel, G. Fabbris, M. Zhernenkov, P. P. Kong, C. Jin, G. Cao, and M. van Veenendaal, ‘Pressure tuning of the spin-orbit coupled ground state in Sr2IrO4,’ Phys. Rev. Lett. 109, 027204 (2012).

[302] M. K. Crawford, M. A. Subramanian, R. L. Harlow, J. A. Fernandez-Baca, Z. R. Wang, and D. C. Johnston, ‘Structural and magnetic studies of Sr2IrO4,’ Phys. Rev. B 49, 9198 (1994).

[303] M. Oshikawa and I. Affleck, ‘Field-induced gap in S = 1/2 antiferromagnetic chains,’ Phys. Rev. Lett. 79, 2883 (1997).

[304] S. Bahr, A. Alfonsov, G. Jackeli, G. Khaliullin, A. Matsumoto, T. Takayama, H. Takagi, B. Büchner, and V. Kataev, ‘Low-energy magnetic excitations in the spin-orbital Mott insulator Sr2IrO4,’ Phys. Rev. B 89, 180401 (2014).

[305] K. W. H. Stevens, ‘On the magnetic properties of covalent XY6 complexes,’ Proc. R. Soc. A 219, 542 (1953).

[306] S. Sugano, Y. Tanabe, and H. Kamimura, Multiplets of transition-metal ions in crystals (Academic Press, New York, 1970).

[307] C. de Graaf, C. Sousa, and R. Broer, ‘Ionization and excitation energies in CuCl and NiO within different embedding schemes,’ J. Mol. Struct. (Theochem) 458, 53 (1999).

[308] A. Domingo, A. Rodríguez-Fortea, M. Swart, C. de Graaf, and R. Broer, ‘Ab initio absorption spectrum of NiO combining molecular dynamics with the em- bedded cluster approach in a discrete reaction field,’ Phys. Rev. B 85, 155143 (2012). 124 bibliography

[309] R. Maurice, P. Verma, J. M. Zadrozny, S. Luo, J. Borycz, J. R. Long, D. G. Truh- lar, and L. Gagliardi, ‘Single-ion magnetic anisotropy and isotropic magnetic couplings in the metal-organic framework Fe2(dobdc),’ Inorg. Chem. 52, 9379 (2013).

[310] P. Fuentealba, L. von Szentpaly, H. Preuss, and H. Stoll, ‘Pseudopotential calculations for alkaline-earth atoms,’ J. Phys. B 18, 1287 (1985).

[311] A. Berning, M. Schweizer, H.-J. Werner, P. J. Knowles, and P. Palmieri, ‘Spin- orbit matrix elements for internally contracted multireference configuration in- teraction wavefunctions,’ Mol. Phys. 98, 1823 (2000).

[312] S. Yabushita, Z. Zhang, and R. M. Pitzer, ‘Spin-orbit configuration interaction using the graphical unitary group approach and relativistic core potential and spin-orbit operators,’ J. Phys. Chem. A 103, 5791 (1999).

[313] D. Ganyushin and F. Neese, ‘A fully variational spin-orbit coupled complete active space self-consistent field approach: Application to electron paramag- netic resonance g-tensors,’ J. Chem. Phys. 138, 104113 (2013).

[314] S. Mai, T. Müller, F. Plasser, P. Marquetand, H. Lischka, and L. González, ‘Perturbational treatment of spin-orbit coupling for generally applicable high- level multi-reference methods,’ J. Chem. Phys. 141, 074105 (2014).

[315] H. Bolvin, ‘An alternative approach to the g-matrix: theory and applications,’ ChemPhysChem 7, 1575 (2006).

[316] S. Vancoillie, P. Å. Malmqvist, and K. Pierloot, ‘Calculation of EPR g tensors for transition-metal complexes based on multiconfigurational perturbation the- ory (CASPT2),’ ChemPhysChem 8, 1803 (2007).

[317] L. F. Chibotaru and L. Ungur, ‘Ab initio calculation of anisotropic magnetic properties of complexes. I. unique definition of pseudospin hamiltonians and their derivation,’ J. Chem. Phys. 137, 064112 (2012).

[318] M. Gerloch and R. F. McMeeking, ‘Paramagnetic properties of unsymmetrical transition-metal complexes,’ J. Chem. Soc., Dalton Trans. , 2443 (1975).

[319] F. Neese, ‘The ORCA program system,’ Wiley Interdiscip. Rev. Comput. Mol. Sci. 2, 73 (2012).

[320] V. M. Katukuri, K. Roszeitis, V. Yushankhai, A. Mitrushchenkov, H. Stoll, M. van Veenendaal, P. Fulde, J. van den Brink, and L. Hozoi, ‘Electronic struc- ture of low-dimensional 4d5 oxides: interplay of ligand distortions, overall lat- tice anisotropy, and spin-orbit interactions,’ Inorg. Chem. 53, 4833 (2014).

[321] D. A. Pantazis, X. Y. Chen, C. R. Landis, and F. Neese, ‘All-electron scalar relativistic basis sets for third-row transition metal atoms,’ J. Chem. Theory Comput. 4, 908 (2008).

[322] I. Schapiro, K. Sivalingam, and F. Neese, ‘Assessment of N-electron valence state perturbation theory for vertical excitation energies,’ J. Chem. Theory Com- put. 9, 3567 (2013).

[323] N. B. Perkins, Y. Sizyuk, and P. Wölfle, ‘Interplay of many-body and single- particle interactions in iridates and rhodates,’ Phys. Rev. B 89, 035143 (2014).

[324] B. H. Kim, G. Khaliullin, and B. I. Min, ‘Magnetic couplings, optical spectra, and spin-orbit exciton in 5d electron Mott insulator Sr2IrO4,’ Phys. Rev. Lett. 109, 167205 (2012). bibliography 125

[325] M. A. Subramanian, M. K. Crawford, R. L. Harlow, T. Ami, J. A. Fernandez- Baca, Z. R. Wang, and D. C. Johnston, ‘Sr2RhO4 and Sr2IrO4: Structural and magnetic studies of 4d and 5d transition metal analogs of La2CuO4,’ Physica C 235-240, 743 (1994).

[326] S. Boseggia, H. C. Walker, J. Vale, R. Springell, Z. Feng, R. S. Perry, M. M. Sala, H. M. Rønnow, S. P. Collins, and D. F. McMorrow, ‘Locking of iridium magnetic moments to the correlated rotation of oxygen octahedra in Sr2IrO4 revealed by x-ray resonant scattering,’ J. Phys.: Condens. Matter 25, 422202 (2013).

[327] F. Ye, S. Chi, B. C. Chakoumakos, J. A. Fernandez-Baca, T. Qi, and G. Cao, ‘Magnetic and crystal structures of Sr2IrO4: A neutron diffraction study,’ Phys. Rev. B. 87, 140406(R) (2013).

[328] G. Cao, J. Bolivar, S. McCall, J. E. Crow, and R. P. Guertin, ‘Weak ferro- magnetism, metal-to-nonmetal transition, and negative differential resistivity in single-crystal Sr2IrO4,’ Phys. Rev. B 57, 11039(R) (1998).

[329] H. Bolvin, ‘Theoretical determination of the excited states and of g-factors of 5+ the Creutz-Taube ion, [(NH3)5-Ru-pyrazine-Ru-(NH3)5] ,’ Inorg. Chem. 46, 417 (2007).

[330] Y. Singh, S. Manni, J. Reuther, T. Berlijn, R. Thomale, W. Ku, S. Trebst, and P. Gegenwart, ‘Relevance of the Heisenberg-Kitaev model for the honeycomb lattice iridates A2IO3,’ Phys. Rev. Lett. 108, 127203 (2012).

[331] K. Ohgushi, H. Gotou, T. Yagi, Y. Kiuchi, F. Sakai, and Y. Ueda, ‘Metal- insulator transition in Ca1−xNaxIrO3 with post-perovskite structure,’ Phys. Rev. B 74, 241104 (2006).

[332] K. Matsuhira, M. Wakeshima, Y. Hinatsu, and S. Takagi, ‘Metal–insulator tran- sitions in pyrochlore oxides Ln2Ir2O7,’ J. Phys. Soc. Jpn. 80, 094701 (2011).

[333] F. Rodi and D. Babel, ‘Ternäre oxide der Übergangsmetalle. iv. erdalkaliiridium(iv)-oxide: Kristallstruktur von CalrO3,’ Z. Anorg. Allg. Chem. 336, 17 (1965).

[334] S. Hirai, M. D. Welch, F. Aguado, and S. A. T. Redfern, ‘The crystal structure of CaIrO3 post-perovskite revisited,’ Z. Kristallogr. 224, 345 (2009).

[335] M. Murakami, K. Hirose, K. Kawamura, N. Sata, and Y. Ohishi, ‘Post- perovskite phase transition in MgSiO3,’ Science 304, 855 (2004).

[336] A. R. Oganov and S. Ono, ‘Theoretical and experimental evidence for a post- 00 perovskite phase of MgSiO3 in Earth’s D layer,’ Nature (London) 430, 445 (2004).

[337] N. Miyajima, K. Ohgushi, M. Ichihara, and T. Yagi, ‘Crystal morphology and dislocation microstructures of CaIrO3: A TEM study of an analogue of the MgSiO3 post-perovskite phase,’ Geophys. Res. Lett. 33,L12302 (2006).

[338] D. C. Martin, K. W. Chapman, P. J. Chupas, V. Parakapenka, P. L. Lee, S. D. Shastri, and J. B. Parise, ‘Compression, thermal expansion, structure, and in- stability of cairo3, the structure model of mgsio3 post-perovskite,’ Am. Mineral. 92, 1048 (2007).

[339] T. B. Ballaran, R. G. Tronnes, and D. J. Frost, ‘Equations of state of CaIrO3 perovskite and post-perovskite phases,’ Am. Mineral. 92, 1760 (2007). 126 bibliography

[340] S. A. Hunt, D. R. Davies, A. M. Walker, R. J. McCormack, A. S. Wills, D. P. Dob- son, and L. Li, ‘On the increase in thermal diffusivity caused by the perovskite to post-perovskite phase transition and its implications for mantle dynamics,’ Earth. Planet. Sci. Lett. 319-320, 96 (2012).

[341] K. Ohgushi, J.-i. Yamaura, H. Ohsumi, K. Sugimoto, S. Takeshita, A. Tokuda, H. Takagi, M. Takata, and T.-h. Arima, ‘Resonant x-ray diffraction study of the strongly spin-orbit-coupled Mott insulator CaIrO3,’ Phys. Rev. Lett. 110, 217212 (2013).

[342] P. W. Anderson, ‘New approach to the theory of superexchange interactions,’ Phys. Rev. 115, 2 (1959).

[343] M. Moretti Sala, K. Ohgushi, A. Al-Zein, Y. Hirata, G. Monaco, and M. Krisch, 1 2 ‘CaIrO3: A spin-orbit Mott insulator beyond the jeff= / ground state,’ Phys. Rev. Lett. 112, 176402 (2014).

[344] D. Muñoz, F. Illas, and I. de P. R. Moreira, ‘Accurate prediction of large an- tiferromagnetic interactions in high- Tc HgBa2Can−1CunO2n+2+δ (n = 2, 3) superconductor parent compounds,’ Phys. Rev. Lett. 84, 1579 (2000).

[345] M. S. Dresselhaus, G. Dresselhaus, and A. Jorio, Group theory. Application to the physics of condensed matter (Springer Verlag, Berlin, 2008).

[346] A. Subedi, ‘First-principles study of the electronic structure and magnetism of CaIrO3,’ Phys. Rev. B 85, 020408 (2012) EIGENSTÄNDIGKEITSERKLÄRUNG

Hiermit versichere ich, dass ich die vorliegende Arbeit ohne unzuläs- sige Hilfe Dritter und ohne Benutzung anderer als der angegebenen Hilfsmittel angefertigt habe; die aus fremden Quellen direkt oder in- direkt übernommenen Gedanken sind als solche kenntlich gemacht. Die Arbeit wurde bisher weder im Inland noch im Ausland in gle- icher oder ähnlicher Form einer anderen Prüfungsbehörde vorgelegt. Die vorliegende Arbeit wurde unter der wissenschaftlichen Betreu- ung von Prof. Dr. Jeroen van den Brink am IFW Dresden angefertigt. Es haben keine früheren erfolglosen Promotionsverfahren stattge- funden. Ich erkenne die Promotionsordnung der Fakultät Mathematik und Naturwissenschaften der Technischen Universität Dresden an.

Stuttgart, 29 September 2017

Nikolay Bogdanov

Colophon

This document was typeset with pdfLATEX and BIBTEX using the ty- pographical look-and-feel classicthesis developed by André Miede and Ivo Pletikosi´cinspired by Robert Bringhurst’s ‘The Elements of Typographic Style’ book. The text and mathematics are set with Palatino and AMS Euler type- faces respectively, both designed by Hermann Zapf.