This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Charge disproportionation and Hund’s insulating behavior in a five-orbital Hubbard model applicable to d4d^{4} perovskites

Maximilian E. Merkel maximilian.merkel@mat.ethz.ch Materials Theory, ETH Zürich, Wolfgang-Pauli-Strasse 27, 8093 Zürich, Switzerland    Claude Ederer claude.ederer@mat.ethz.ch Materials Theory, ETH Zürich, Wolfgang-Pauli-Strasse 27, 8093 Zürich, Switzerland
(August 2, 2025)
Abstract

We explore the transition to a charge-disproportionated insulating phase in a five-orbital cubic tight-binding model applicable to transition-metal perovskites with a formal d4d^{4} occupation of the transition-metal cation, such as ferrates or manganites. We use dynamical mean-field theory to obtain the phase diagram as a function of the average local Coulomb repulsion UU and the Hund’s coupling JJ. The main structure of the phase diagram follows from the zero band-width (atomic) limit and represents the competition between high-spin and low-spin homogeneous and an inhomogeneous charge-disproportionated state. This results in two distinct insulating phases: the standard homogeneous Mott insulator and the inhomogeneous charge-disproportionated insulator, recently also termed Hund’s insulator. We characterize the unconventional nature of this Hund’s insulating state. Our results are consistent with previous studies of two- and three-orbital models applicable to isolated t2gt_{2g} and ege_{g} subshells, respectively, with the added complexity of the low-spin/high-spin transition. We also test the applicability of an effective two-orbital (ege_{g}-only) model with disordered S=3/2S=3/2 t2gt_{2g} core spins. Our results show that the overall features of the phase diagram in the high-spin region are well described by this simplified two-orbital model but also that the spectral features exhibit pronounced differences compared to the full five-orbital description.

I Introduction

Many of the fascinating properties of transition-metal oxides can be understood as a result of the strong Coulomb interaction, experienced by electrons in the rather localized dd orbitals, in combination with their mixed ionic-covalent character, which is due to hybridization between the transition-metal dd orbitals and the pp orbitals of the surrounding oxygen ligands. The resulting electronic structure is then often in the border regime between itinerant and localized, where the electrons experience strong correlations that can give rise to a number of intriguing properties such as, e.g., metal-insulator transitions [1].

Such correlation effects and the emergence of a Mott-insulating state are usually associated with a large value of the Hubbard parameter UU, which describes the repulsion between electrons located on the same site in the seminal Hubbard model [2; 3]. However, in multi-band Hubbard models, strong correlations, i.e., pronounced deviations from the behavior of (effectively) non-interacting particles, can also be caused by the Hund’s coupling JJ, even if the corresponding system is not particularly close to a Mott-insulating phase [4]. Such materials have recently been termed Hund’s metals [5; 6; 7].

Furthermore, a strong Hund’s coupling can even lead to the formation of an inhomogeneous, charge-ordered or charge-disproportionated, insulating phase, distinct from the usual (homogeneous) Mott-insulator. The existence of such a charge-disproportionated insulator (CDI) has been demonstrated both for a two-orbital and a three-orbital Hubbard model [8; 9; 10], which are applicable to materials with partially filled ege_{g} and t2gt_{2g} sub-shells. Similar physics have also been discussed in the context of valence-skipping metals and insulators resulting from a negative effective Coulomb interaction [11].

The CDI phase has also been termed Hund’s insulator [9; 10; 12], to emphasize the crucial role of the Hund’s interaction for stabilizing the insulating state and to distinguish it from the usual (homogeneous) Mott-insulator. We use the terms CDI and Hund’s insulating phase interchangeably throughout this article, but we also note that earlier papers have used the term Hund’s insulator in slightly different contexts, e.g., to describe situations with half-filled shells where the Hund’s coupling JJ cooperates with the Hubbard UU and the value of UU alone is not sufficient to reach an insulating state [13] or in specific cases where Hund’s coupling leads to the stabilization of a topological insulating phase [14].

The parameter regime where a CDI or Hund’s insulator is observed corresponds to an electronic structure where the Coulomb interaction is strongly screened, which leads to a moderate value of the Hubbard parameter UU, whereas the Hund’s parameter JJ is less affected by the screening. This regime is expected to occur in transition-metal oxides with a pronounced charge transfer character, i.e., strong hybridization between the transition-metal dd and oxygen pp bands. This can generally be expected to occur towards the end of the 3dd transition-metal series or for systems with high oxidation numbers of the transition-metal cation.

The formation of a charge-disproportionated insulating phase has been observed, e.g., in the series of rare earth nickelates, RRNiO3 [15; 16; 17; 18; 8; 19; 20]. Here, the Ni cations are in a formal d7d^{7} electron configuration, which then disproportionates according to 2d7d6+d82d^{7}\rightarrow d^{6}+d^{8} (or, equivalently, 2d8L¯d8L¯2+d82d^{8}\underline{L}\rightarrow d^{8}\underline{L}^{2}+d^{8}, where L¯\underline{L} denotes a ligand hole in the surrounding oxygen network [21; 22]). In this case the t2gt_{2g} subshell is always completely filled with six electrons, and the remaining Ni-dd electron occupies the ege_{g} subshell. It has been shown that the underlying physics can be well described within a minimal two-orbital ege_{g} Hubbard model with an average quarter-filling [8]. Furthermore, the charge disproportionation lowers the symmetry of the crystal structure and couples strongly with a lattice distortion corresponding to a breathing mode of the octahedral network with R1+R_{1}^{+} symmetry, resulting in a three-dimensional checkerboard-like pattern of alternating small and large oxygen octahedra surrounding the nominal d6d^{6} and d8d^{8} Ni sites, respectively.

Similar behavior has also been observed for the alkaline earth ferrate CaFeO3 [23; 24; 25; 26; 27] and in PbCrO3 [28; 29]. The case of PbCrO3 is potentially more complex due to the presence of the stereochemically active Pb2+ cation. However, the nominal d2d^{2} configuration of the Cr4+ cation and the suggested charge disproportionation according to 3Cr4+2Cr3++Cr6+3\rm{Cr}^{4+}\rightarrow 2\rm{Cr}^{3+}+\rm{Cr}^{6+} is consistent with recent theoretical predictions for the three-orbital Hubbard model [9]. On the other hand, the Fe4+ cation in CaFeO3 is assumed to be in a nominal d4d^{4} high spin configuration [30], with partial filling of both t2gt_{2g} and ege_{g} subshells, and thus requires a description using the full five-orbital dd manifold. Additionally, the competition between high-spin and low-spin states might introduce further complexity in the underlying physics.

Motivated by this, we study a full five-orbital Hubbard model with d4d^{4} electron filling. We note that, while our main motivation is to study the possible emergence of a charge disproportionated state in a rather generic five-orbital TB model, we fit the parameters of this model to the actual bandstructure of CaFeO3 in order to work with realistic values for bandwidth and crystal-field splittings. Thus, we consider the physically relevant case with cubic symmetry by incorporating a crystal-field splitting between the three-fold degenerate lower-lying t2gt_{2g} states and the two-fold degenerate higher-lying ege_{g} states. We use dynamical mean-field theory (DMFT) to calculate expectation values of this model around room temperature. This allows us to identify all relevant phases as function of the local interaction parameters UU and JJ, describing the average Coulomb repulsion and Hund’s coupling within the dd-shell.

In particular, we establish the presence of a spontaneously charge-disproportionated insulating phase in this five-orbital model, in addition to the usual homogeneous Mott-phase. The overall structure of the phase diagram resembles analogous results for the two- and three-orbital models, but with the additional feature of low-/high-spin transitions. We analyze the character of the insulating state, and we explore the effect of a structurally induced energy difference between the ege_{g} states on the two different sub-lattices as well as the effect of a reduced ege_{g} band width. Furthermore, we compare the phase diagram obtained for the full five-orbital model, with an effective two-orbital model, applicable to the high-spin limit [31; 32]. This simplified model is able to quantitatively reproduce many features of the full five-orbital model, including, e.g, the phase boundaries as function of the interaction parameters, but also exhibits pronounced differences in the spectral properties.

II Model and methods

In this section, we first introduce the models that we are studying: the complete five-orbital model for the full dd shell and the effective two-orbital ege_{g} model with localized t2gt_{2g} spins. We then describe how the parameters of the models are fitted to the band structure of CaFeO3 obtained from density functional theory (DFT). Finally, we describe the details of our DMFT calculations and introduce the observables used to characterize the different phases.

II.1 The Hamiltonian

We study the following model, which consists of a single-particle tight-binding (TB) part and a local interaction on each site RR:

H=HTB+RHint(R).H=H_{\text{TB}}+\sum_{R}H_{\text{int}}^{(R)}\quad. (1)

The TB part consists of inter-site hopping terms as well as an on-site crystal-field splitting, Δegt2g\Delta_{{e_{g}}-{t_{2g}}}, between eg{e_{g}} and t2g{t_{2g}} orbitals. We include hopping between nearest neighbors (NN) and next-nearest neighbors (NNN) on a cubic lattice, representing the sites of the transition-metal cations in the perovskite structure. The hopping amplitudes are assumed to have the cubic symmetry corresponding to t2g{t_{2g}} and eg{e_{g}} orbitals. Thus, between nearest neighbors there is only hopping between t2g{t_{2g}} orbitals of the same type and no hopping connecting the eg{e_{g}} and t2g{t_{2g}} orbitals. For simplicity, we neglect next-nearest neighbor hopping between different t2gt_{2g} and between t2gt_{2g} and ege_{g} orbitals, and we assume the same fixed ratio tNNN/tNNt_{\mathrm{NNN}}/t_{\mathrm{NN}} between the NNN and NN hopping amplitudes for both eg{e_{g}} and t2g{t_{2g}} orbitals. For the ege_{g} orbitals, tNNt_{\text{NN}} and tNNNt_{\text{NNN}} correspond to the hopping between d3z2r2d_{3z^{2}-r^{2}} orbitals along zz and along x+zx+z or y+zy+z, respectively, see Ref. 33.

We also perform calculations where we introduce an intrinsic “site splitting” of the ege_{g} levels, by raising, respectively lowering, the local on-site energy of the eg{e_{g}} orbitals on adjacent sites by ±Δeg/2\pm\Delta_{e_{g}}/2. This mimics the effect of the structural R1+R_{1}^{+} breathing mode distortion that typically accompanies the charge disproportionation in perovskites (see, e.g., Ref. 34). As described in Sec. I, this distortion consists in a spatially alternating expansion and compression of the oxygen octahedra surrounding the transition-metal sites, resulting in a three-dimensional checkerboard-like pattern of large and small octahedra.

To describe the local electron-electron interaction within the five-orbital dd shell, we use the so-called Slater parametrization in the density-density approximation (see, e.g., Ref. 35):

Hint=\displaystyle H_{\mathrm{int}}= 12mm,σUmmnmσnmσ¯\displaystyle\frac{1}{2}\sum_{mm^{\prime},\sigma}U_{mm^{\prime}}n_{m\sigma}n_{m^{\prime}\bar{\sigma}}
+\displaystyle+ 12mm,σ(UmmJmm)nmσnmσ.\displaystyle\frac{1}{2}\sum_{m\neq m^{\prime},\sigma}(U_{mm^{\prime}}-J_{mm^{\prime}})n_{m\sigma}n_{m^{\prime}\sigma}\quad. (2)

This term is purely local and identical for each transition-metal site (we therefore suppress the site index RR in Eq. (II.1)). Thereby, nmσn_{m\sigma} is the occupation number operator for orbital mm with spin σ\sigma (on site RR) and σ¯\bar{\sigma} is short for σ-\sigma. The matrix elements UmmU_{mm^{\prime}} and JmmJ_{mm^{\prime}} are parametrized in terms of the Slater integrals F0F_{0}, F2F_{2}, and F4F_{4} in the usual form [35]. The Slater integrals themselves are expressed in terms of the average Coulomb repulsion U=F0U=F_{0} and Hund’s rule interaction J=(F2+F4)/14J=(F_{2}+F_{4})/14, using a fixed ratio F4/F2=0.63F_{4}/F_{2}=0.63 [35].

II.2 The effective two-orbital approximation

We also compare the full five-orbital dd-shell model to a simplified effective two-orbital eg{e_{g}} model, applicable to the high-spin limit. In this model, the occupation of the t2g{t_{2g}} states on each site is fixed to three electrons, giving rise to an S=3/2S=3/2 core spin, and hopping in and out of these t2g{t_{2g}} states is neglected [31; 32]. The remaining electron is thus constrained to the ege_{g} orbitals, but can still hop from site to site. Then, for every site, the t2g{t_{2g}} core spin defines a local spin quantization axis, and the spins of the eg{e_{g}} electrons are defined relative to this local axis, denoted by σ=\sigma=\Uparrow/\Downarrow.

The ege_{g} electrons interact with the t2g{t_{2g}} core spins through an effective Hund’s rule interaction, which leads to a local spin splitting of the eg{e_{g}} orbitals:

H=hm(nmnm),\displaystyle H=-h\sum_{m}(n_{m\Uparrow}-n_{m\Downarrow}), (3)

where hh is the effective magnetic field created by the t2g{t_{2g}} core spin. The strength of this effective magnetic field can be related to the parameters of the full interaction Hamiltonian, Eq. (II.1), in particular to the Hund’s interaction parameter JJ, by considering the energy gain (penalty) of an eg{e_{g}} spin (anti-)parallel to the t2g{t_{2g}} core spin, i.e., by considering the energy difference between (t2g)3(eg)1(t_{2g})^{3}_{\uparrow}(e_{g})^{1}_{\uparrow} and (t2g)3(eg)1(t_{2g})^{3}_{\uparrow}(e_{g})^{1}_{\downarrow}:

2h=649F2+25147F41.97J.\displaystyle 2h=\frac{6}{49}F_{2}+\frac{25}{147}F_{4}\approx 1.97J\quad. (4)

Furthermore, assuming a completely random, i.e., paramagnetic orientation of the t2g{t_{2g}} core spins on the different sites results in a renormalization of the hopping amplitudes, enabling also hopping between \Uparrow and \Downarrow states corresponding to different sites. After averaging over all possible relative spin orientations, this can be described by a matrix uσ,σR,Ru^{R,R^{\prime}}_{\sigma,\sigma^{\prime}}, which does not affect the on-site Hamiltonian (uσ,σR,R=δσ,σu^{R,R}_{\sigma,\sigma^{\prime}}=\delta_{\sigma,\sigma^{\prime}}) but renormalizes and mixes the inter-site hopping terms (uσ,σR,RR=2/3u^{R,R^{\prime}\neq R}_{\sigma,\sigma^{\prime}}=2/3) according to [31; 32]:

tm,mR,Rtm,mR,Ruσ,σR,R.\displaystyle t^{R,R^{\prime}}_{m,m^{\prime}}\rightarrow t^{R,R^{\prime}}_{m,m^{\prime}}u^{R,R^{\prime}}_{\sigma,\sigma^{\prime}}\quad. (5)

To enable a consistent comparison with the full five-orbital model, the interaction in the effective two-orbital model is described using the eg{e_{g}} subspace of the full interaction Hamiltonian, Eq. (II.1). This is equivalent to the density-density approximation of the so-called Kanamori Hamiltonian [3], but with the corresponding interaction parameters defined in terms of the Slater integrals (or in terms of the average Coulomb repulsion UU and the corresponding Hund’s rule interaction JJ).

II.3 Fitting the parameters of the TB model

The purpose of this work is to study the physics of charge disproportionation in a relatively generic cubic five-orbital model. Nevertheless, we focus on realistic parameter regimes, applicable to perovskite transition-metal oxides and closely related materials. We therefore obtain parameters for the single-particle part of the Hamiltonian, HTBH_{\text{TB}}, by comparison with the full band structure of CaFeO3, calculated within density functional theory (DFT).

We perform DFT calculations using the “Vienna Ab-initio Simulation Package” (VASP) [36; 37] employing the PBEsol exchange-correlation functional [38], both for the high-temperature PbnmPbnm structure, where all Fe sites are equivalent, and the charge-disproportionated P21/nP2_{1}/n crystal structure observed below 290 K290\text{\,}\mathrm{K} [39; 27]. We use a 7×7×57\times 7\times 5 kk-point mesh for the 20-atom unit cell and a plane-wave energy cutoff of 600 eV600\text{\,}\mathrm{e}\mathrm{V} to ensure well converged results.

We first relax unit-cell parameters as well as atomic positions within PbnmPbnm symmetry. In order to stabilize the high-spin state of the Fe cation, a Hubbard-corrected DFT+UU treatment and the presence of antiferromagnetic (AFM) order is required. We find that moderate values of U=4 eVU=$4\text{\,}\mathrm{e}\mathrm{V}$ and J=1 eVJ=$1\text{\,}\mathrm{e}\mathrm{V}$ and A-type antiferromagnetic order (as proxy for the more complicated helical spin structure in CaFeO3 [40]) result in structural parameters that are in good agreement with experimental data (e.g., the calculated unit cell volume is 210.34 Å3210.34\text{\,}{\mathrm{\SIUnitSymbolAngstrom}}^{3}[39; 26]. We then re-relax the atomic positions within non-spin-polarized DFT (to exclude any structural effects or symmetry lowering stemming from the magnetic order) but with lattice parameters fixed to the ones obtained from the magnetic +U+U calculations. We note that a non-magnetic DFT calculation leads to a low-spin state of the Fe cation, which would result in an unrealistically low unit-cell volume.

To obtain a realistic estimate of the site-splitting Δeg\Delta_{e_{g}}, we manually add a breathing mode of R1+=0.2 Å{R_{1}^{+}}=$0.2\text{\,}\mathrm{\SIUnitSymbolAngstrom}$, very similar to the experimentally observed amplitude R1+=0.18 Å{R_{1}^{+}}=$0.18\text{\,}\mathrm{\SIUnitSymbolAngstrom}$ [26], while keeping all other structural parameters unchanged. A full DFT+UU study of CaFeO3 has been presented in Ref. 41.

Refer to caption
Figure 1: Plots of the band structure of CaFeO3 obtained using DFT (black lines) and the fitted TB bands (blue lines), in (a) PbnmPbnm symmetry (R1+=0.0 Å{R_{1}^{+}}=$0.0\text{\,}\mathrm{\SIUnitSymbolAngstrom}$) and (b) P21/nP2_{1}/n symmetry (R1+=0.2 Å{R_{1}^{+}}=$0.2\text{\,}\mathrm{\SIUnitSymbolAngstrom}$). Zero energy indicates the Fermi level. Additionally, the DFT total density of states (DOS) and Fe-projected DOS are shown in the corresponding right panels.

The calculated band structure and corresponding densities of states (DOS) in the relevant energy window around the Fermi level are shown in Fig. 1. It can be seen that a group of bands with dominant Fe character, located between 0.5 eV and 4 eV above the Fermi level, is separated from all other bands at higher and lower energies. Further inspection confirms that these are the Fe-eg{e_{g}} bands. The bands between approximately -1 eV and 0.5 eV also have mostly Fe character and correspond to the Fe-t2g{t_{2g}} states. Towards lower energies, there is some entanglement of these Fe-t2g{t_{2g}} bands with other bands with dominant O-pp character. Comparing the band structures obtained with and without the structural R1+R_{1}^{+} breathing mode, one can see that its main effect is to introduce a splitting throughout the middle of the eg{e_{g}} bands, analogously to what has been observed for nickelates [8; 42]. In contrast, the Fe-t2g{t_{2g}} bands are not much affected by the structural distortion.

We fit the parameters of our TB model such that the corresponding band dispersion matches the bands obtained from our DFT calculations. Thereby, the ratio between NNN and NN hopping, tNNN/tNNt_{\text{NNN}}/t_{\text{NN}}, is obtained by fitting the well-separated eg{e_{g}} bands, whereas the t2g{t_{2g}} NN hopping is then mainly adjusted to obtain the correct t2g{t_{2g}} band width. Note that the two cases with and without breathing mode are fitted with the same hopping amplitudes, whereas the average eg{e_{g}}-t2g{t_{2g}} splitting Δegt2g\Delta_{{e_{g}}-{t_{2g}}} is slightly altered by the R1+{R_{1}^{+}} breathing mode. The resulting TB band structure is also shown in Fig. 1, and the corresponding parameters are listed in Table 1. It can be seen that the TB model results in an excellent fit of the Fe-eg{e_{g}} bands, both in the PbnmPbnm and P21/nP2_{1}/n cases. For the t2g{t_{2g}} bands, the overall band width is well described, while the dispersion of the individual bands shows notable deviations. This could be improved by introducing a separate tNNN/tNNt_{\text{NNN}}/t_{\text{NN}} for the t2g{t_{2g}} hopping or other additional intra- and inter-site TB parameters. However, for simplicity, and since the specific form of the t2g{t_{2g}} bands is not relevant in the context of this work, we refrain from introducing further parameters into our TB model.

Table 1: The parameters of our TB model, obtained by fitting to the DFT band structures of CaFeO3 with and without the R1+{R_{1}^{+}} breathing mode, together with the corresponding ege_{g} and t2gt_{2g} total band widths, WegW_{e_{g}} and Wt2gW_{t_{2g}}, respectively.
R1+=0.0{R_{1}^{+}}{}=0.0 Å R1+=0.2{R_{1}^{+}}=0.2 Å
Δegt2g\Delta_{{e_{g}}-{t_{2g}}} (eV\mathrm{e}\mathrm{V}) 2.494 2.519
tegt_{e_{g}} (eV\mathrm{e}\mathrm{V}) -0.511 -0.511
tt2gt_{t_{2g}} (eV\mathrm{e}\mathrm{V}) -0.220 -0.220
tNNN/tNNt_{\text{NNN}}/t_{\text{NN}} 0.065 0.065
Δeg\Delta_{e_{g}} (eV\mathrm{e}\mathrm{V}) 0 0.436
WegW_{e_{g}} (eV\mathrm{e}\mathrm{V}) 3.069
Wt2gW_{t_{2g}} (eV\mathrm{e}\mathrm{V}) 1.761

II.4 DMFT calculations

We obtain the phase diagram of the model, Eq. (1), as function of the interaction parameters UU and JJ using dynamical mean-field theory (DMFT) [43]. Within DMFT, the self-energy of the full lattice problem is approximated by a frequency-dependent but local self-energy, which is obtained by mapping each symmetry-inequivalent site on an effective impurity problem, defined via a self-consistency condition requiring that the impurity Green’s function is identical to the corresponding local Green’s function of the full lattice problem.

The calculations are implemented using the TRIQS and DFTTools libraries [44; 45; 46]. To solve the effective impurity problem, we use a continuous-time Quantum-Monte-Carlo solver (CT-HYB), also implemented in the TRIQS library [47], and an inverse temperature β=1/(kBT)=40 eV1\beta=1/(k_{\text{B}}T)=$40\text{\,}\mathrm{e}\mathrm{V}^{-1}$, corresponding to approximately room temperature. Green’s functions are represented by 40 Legendre coefficients [48] for the full five-orbital model (50 for the effective two-orbital model), which results in good convergence and a smooth behavior of the self-energy in imaginary frequency Σ(iωn)\Sigma(\mathrm{i}\omega_{n}).

We average over both spin channels and also over the eg{e_{g}} and t2g{t_{2g}} manifolds separately to ensure a paramagnetic solution without orbital polarization in the two different subshells. Furthermore, to allow for a charge disproportionated solution, we divide the sites of the simple cubic lattice into two interpenetrating fcc lattices, (I) and (II), corresponding to the expected 3D checkerboard-like arrangement, and solve the two resulting effective impurity problems separately.

From the local Green’s function Gmm(τ)=Tcm(τ)cm(0)G_{mm^{\prime}}(\tau)=-\langle Tc_{m}(\tau)c^{\dagger}_{m^{\prime}}(0)\rangle, where TT is the imaginary-time ordering operator and cm(τ)c^{\dagger}_{m}(\tau) is the creation operator for an electron in orbital mm at the local site at imaginary time τ\tau, we obtain the total eg{e_{g}} occupation per site, neg=nz2+nx2y2n_{\mathrm{e_{g}}}=n_{\mathrm{z^{2}}}+n_{\mathrm{x^{2}-y^{2}}}, the (positively-defined) occupation difference between the two sub-lattices, δn=|mnm(I)mnm(II)|\delta n=|\sum_{m}n_{m}^{\mathrm{(I)}}-\sum_{m}n_{m}^{\mathrm{(II)}}|, and the averaged spectral weight around the Fermi level, A(ω=0)=(β/π)TrG(τ=β/2)A(\omega=0)=-(\beta/\pi)\mathrm{Tr}\,G(\tau=\beta/2). Furthermore, we obtain an estimate for the (orbital-dependent) quasiparticle weight Zm=[1ImΣm(iω)(iω)|iω=0]1Z_{m}=\left[1-\left.\frac{\partial\imaginary\Sigma_{m}(\mathrm{i}\omega)}{\partial(\mathrm{i}\omega)}\right|_{\mathrm{i}\omega=0}\right]^{-1} by fitting a fourth-order polynomial to the imaginary part of the local self-energy, ImΣm(iωn)\imaginary\Sigma_{m}(\mathrm{i}\omega_{n}), for the eight lowest positive Matsubara frequencies iωn\mathrm{i}\omega_{n} and extrapolate the slope at zero frequency from this fit, c.f. Appendix C in Ref. 49. We use the maximum-entropy method [50], implemented in the TRIQS library [51], to obtain both kk-averaged and kk-resolved spectral functions on the real frequency axis.

III Results

III.1 Atomic-limit predictions

Before presenting our results from the full DMFT calculations, we outline the general features of the expected phase diagram by considering the “atomic limit” of Eq. (1), i.e., the limit where all hopping amplitudes and thus the associated band widths go to zero, and the ground state configuration of the system is solely determined by the local part of the Hamiltonian, i.e., the interaction term HintH_{\text{int}} plus the local crystal-field and eg{e_{g}} site splitting.

Refer to caption
Figure 2: Schematic phase diagram for our TB model corresponding to equivalent sub-lattices, Δeg=0\Delta_{e_{g}}=0. Thick black lines separate regions with different electronic ground states in the atomic limit: homogeneous HS d4d^{4}, homogeneous LS d4d^{4}, and inhomogeneous HS d3d^{3} + d5d^{5}, as indicated by the energy level diagrams. The colored regions indicate the expected insulating regions, with the homogeneous Mott phase in blue and the charge disproportionated (CD) insulating region in red. The white area represents the expected homogeneous metallic state. The red-hatched area denotes a potential CD metallic region based on simple band-width estimates (see main text). The shifts of the phase boundaries resulting from a site splitting Δeg=0.436 eV\Delta_{e_{g}}=$0.436\text{\,}\mathrm{e}\mathrm{V}$ and an 80 % reduction of the ege_{g} band with WegW_{e_{g}} are indicated by small black arrows, with the length of the arrow corresponding to the actual shift.

We first compare the energies of the three most relevant configurations: (i) homogeneous high-spin (HS) d4d^{4}; (ii) homogeneous low-spin (LS) d4d^{4}; and (iii) inhomogeneous HS d5d^{5} + d3d^{3}, also depicted by the schematic level diagrams in Fig. 2, and identify the ground state as function of interaction parameters UU and JJ. Details are described in App. A. Fig. 2 depicts the case without an explicit eg{e_{g}} site splitting, i.e., Δeg=0\Delta_{e_{g}}=0. Thick black lines separate the regions where the three different cases (i)-(iii) are the ground state configuration. The homogeneous LS configuration (ii) is generally favorable for small JJ (JΔegt2g/2.830.88J\lesssim\Delta_{e_{g}-t_{2g}}/2.83\approx 0.88 eV, except for the small part separating the LS and inhomogeneous HS regions, where the critical JJ is further decreased by decreasing UU). The inhomogeneous charge-disproportionated (CD) state (iii) is the ground state for larger JJ and U1.51JU\lesssim 1.51J (except, again, for the small part separating CD and homogeneous LS regions, where the critical UU is smaller). For large JJ and UU, the homogeneous HS state (i) is obtained. We note that in the Slater parametrization of the local interaction, a positive UU represents an average repulsive electron-electron interaction (independent of JJ), which is why we consider the whole region with U>0U>0 as potentially physically relevant.

A simple estimate for the boundary between metallic and (Mott-) insulating regions can be obtained by comparing the lowest inter-site charge transfer excitation energy with the average electronic kinetic energy. For the homogeneous phases, this kinetic energy can be approximated by the relevant band width (see, e.g., Ref. 4), resulting in the blue-shaded “Mott-insulating” region in Fig. 2 (see App. A for details). Since the t2g{t_{2g}} band width is smaller than the eg{e_{g}} band width, the Mott phase extends to smaller UU values in the LS region than in the HS region.

For the CD region, a good estimate for the corresponding effective band width/kinetic energy is more difficult to obtain. For the transition to the homogeneous metallic phase it turns out to be significantly smaller than WegW_{e_{g}} and essentially goes to zero when Δeg\Delta_{e_{g}} becomes large [8]. For simplicity, we therefore denote the whole region below the CD transition line as “CDI” in the atomic limit (red shaded area in Fig. 2). Considering only hopping among either nominal d5d^{5} or d3d^{3} sites defines limits for the CDI region towards small UU (see App. A). Simple band-width estimates for the kinetic energy result in the limits indicated by the red-striped regions in Fig. 2.

Introducing a nonzero eg{e_{g}} site splitting Δeg\Delta_{e_{g}} (corresponding to a nonzero R1+{R_{1}^{+}} distortion of the underlying crystal structure), extends the CD region slightly by rigidly shifting the phase boundaries towards the homogeneous HS and LS phases (see App. A) as indicated by the small arrows in Fig. 2. This also shifts the metal/insulator phase boundary within the homogeneous HS region towards higher UU. In contrast, reducing the eg{e_{g}} band width shifts the metal/insulator phase boundary in the homogeneous HS phase in the opposite direction, while not affecting any of the other main phase boundaries.

III.2 Phase diagram for equivalent sites

Refer to caption
Figure 3: The characteristic properties for the phase diagram in the basic TB model with equivalent sites: (a) average ege_{g} occupation negn_{e_{g}}, (b) occupation difference δn\delta n between the two sub-lattices, (c) averaged spectral weight A(0)A(0), and (d) quasiparticle weight ZZ, estimated from the slope of ImΣ(iω)\imaginary\Sigma(\mathrm{i}\omega). Atomic limit boundaries are indicated as solid gray lines. In (d), the empty circles indicate data points where the system is insulating and thus ZZ is ill-defined.

Next, we present the results of our full DMFT calculations for the TB model with Δeg=0\Delta_{e_{g}}=0, i.e., for the case without a site-dependent shift of the ege_{g} levels. Fig. 3 depicts various observables as function of interaction parameters UU and JJ.

The total ege_{g} occupation per site, negn_{e_{g}}, shown in Fig. 3(a), indicates a sharp transition from LS (neg=0n_{e_{g}}=0) to HS (neg=1n_{e_{g}}=1) for J1J\geq 1 eV, essentially independent of UU, in very good agreement with the expectation obtained from the atomic limit discussed in Sec. III.1. Only for J=0.75J=0.75 eV and U2U\leq 2 eV, we obtain a small intermediate-spin region where the ege_{g} occupation lies between 0 and 1.

The most important result is visible in the occupation difference between the two sublattices, δn\delta n, shown in Fig. 3(b). Here, in the region of small UU but large JJ, i.e., in the region where the atomic limit predicts the inhomogeneous CD ground state, we indeed obtain a non-zero value for δn\delta n, indicating that the system exhibits a spontaneously symmetry-broken state with higher and lower ege_{g} occupation on the two sublattices. Note that this CD region lies fully within the HS region where the average ege_{g} occupation is equal to 1. Inspection of the spectral weight at zero energy, A(0)A(0), shows that the charge disproportionation coincides with a transition to an insulating state, indicated by A(0)0A(0)\approx 0 (dark regions in Fig. 3(c)). This CD insulating region is separated from the Mott-insulating regime, where δn=0\delta n=0 and which occurs for large UU, by a metallic region. It can be seen that the boundary between the Mott-insulating and the metallic region (both in the HS and LS regime) also matches very well with the atomic limit prediction from Sec. III.1, indicated by the gray lines in Fig. 3. On the CD side, the MIT occurs together with the emerging charge disproportionation for δn1\delta n\gtrsim 1, somewhat below the homogeneous-inhomogeneous transition line predicted from the atomic limit. However, we note that the distance of the atomic-limit prediction and the actual charge disproportionation is significantly less than the ege_{g} band width.

In Fig. 3(c), there is a significant change in spectral weight, A(0)A(0), at the LS/HS transition in the metallic region. Fig. 4(a) shows corresponding spectral functions (for U=1.5U=1.5 eV), separated in t2gt_{2g} and ege_{g} contributions. It can be seen that for J=0.5J=0.5 eV the full spectral weight is generated by the t2gt_{2g} contribution, while the ege_{g} states are still unoccupied. Around J=0.75J=0.75 eV the transition occurs, where both ege_{g} and t2gt_{2g} states contribute to the spectral weight at zero energy. Above this LS/HS transition, at J=1J=1 eV, the t2gt_{2g} spectral function is gapped, and A(0)A(0) is entirely due to the ege_{g} states, i.e., a situation that corresponds to an orbitally selective Mott state. Here, the ege_{g} quasiparticle peak is lower and broader compared to the t2gt_{2g} peak for J=0.5J=0.5 eV, which results in the lower spectral weight in the phase diagram for the HS case.

Refer to caption
Figure 4: Evolution of the ege_{g}- and t2gt_{2g}-resolved spectral functions (a) and corresponding imaginary parts of the self energy on the Matsubara axis (b), for the transition from low- (top, J=0.5J=0.5 eV) to intermediate- (middle, J=0.75J=0.75 eV) to high-spin (bottom, J=1.0J=1.0 eV) in the metallic region of the phase diagram (for U=1.5U=1.5 eV). Also indicated are quasiparticle weights ZZ estimated by integrating A(ω)A(\omega) over the shaded areas in (a) or obtained from the slope of ImΣ(iω0)\imaginary\Sigma(\mathrm{i}\omega\rightarrow 0) (with the corresponding fourth order fits shown as solid lines) in (b).

Fig. 3(d) shows the quasiparticle weight ZZ in the metallic region, averaged over all orbitals with non-vanishing spectral weight. ZZ is largest along the line separating homogeneous/inhomogeneous ground states in the atomic limit, and decreases towards both insulating phases. This is consistent with analogous results for the three-orbital case in Ref. 9, where this behavior was related to the competition between the two different insulating phases, resulting in the so-called “Janus effect” [52; 4].

Fig. 4(b) depicts the imaginary part of the self-energy on the Matsubara axis across the LS/HS transition within the metallic regime. As described in Sec. II.4, the quasiparticle weight ZZ is obtained by fitting the low-frequency behavior of ImΣ(iω)\imaginary\Sigma(\mathrm{i}\omega). The corresponding fits are shown in Fig. 4(b). It can be seen that the self-energies exhibit rather strong deviations from simple Fermi liquid behavior. A linear dependence on iω\mathrm{i}\omega can only be observed in a very small range close to ω0\omega\rightarrow 0, and there is a rather large non-vanishing imaginary part for zero frequency. This could in part be due to the relatively high temperature (β=40\beta=40\,eV, corresponding to approximately room temperature) in our calculations. Thus, the interpretation of [1ImΣ/(iω)]1[1-\partial\imaginary\Sigma/\partial(\mathrm{i}\omega)]^{-1} as effective quasiparticle renormalization is somewhat approximate, here. This can also be inferred from the comparison with the “quasiparticle weights” extracted by integrating over the central peak of the corresponding spectral functions in Fig. 3. (Note that the latter of course also just represents a very rough and approximate measure of the quasiparticle renormalization.)

Refer to caption
Figure 5: Multiplet analysis for the model with equivalent sites, sampled along a rectangular path in the phase diagram, as described in the main text. For the different segments of this path, either UU or JJ are kept constant while the other parameter is varied as indicated on the xx-axis. For the multiplet probabilities pN=3p_{N=3}, pN=4p_{N=4}, pN=5p_{N=5}, the solid lines correspond to the sum over all spin quantum numbers |ms||m_{s}| and the dotted lines to the sum over only the HS multiplets, i.e., those with maximal spins |ms|=N/2|m_{s}|=N/2. All quantities are averaged over the two sublattices. Blue, red, and white background represents Mott-insulating, CDI, and metallic regions expected from the atomic limit.

To obtain further insight into the different phases, we also analyze the probabilities of the different atomic multiplet states, obtained from the DMFT Quantum-Monte-Carlo solver. These probabilities describe how likely it is for the corresponding state to be visited during the simulation of the effective impurity problem and allow to distinguish between different occupations and LS/HS configurations in a more quantitative way than in the phase diagrams shown in Fig. 3. To simplify the analysis, we sum up the probabilities for Hilbert states with the same number of electrons NN and the same magnetic spin quantum number |ms||m_{s}| 111We note that within the density-density approximation for the interaction Hamiltonian, Eq. (II.1), the magnetic quantum number msm_{s} and not the total spin SS is the relevant quantum number indicating the spin state of the system.. It turns out that only states with N=3N=3, 44, and 55 have significant probabilities; these together with the quasiparticle weight ZZ are shown along a rectangular path through the phase diagram in Fig. 5.

The first segment of this path corresponds to constant U=1.5U=1.5 eV and increasing JJ from 0.5 eV to 2.0 eV, i.e., starting from the LS metallic, then to the HS metallic, and then into the CDI phase. In the LS metallic phase for J=0.5J=0.5 eV, all multiplets with N=3N=3, 44, and 55 are visited, but the probabilities of the HS states (maximal |ms||m_{s}|) for N=4N=4 and 5 are close to zero. In the transition region between LS and HS metallic phases at J=0.75J=0.75 eV, the HS multiplets with |ms|=N/2|m_{\mathrm{s}}|=N/2 for N=4N=4 and N=5N=5 become more populated, while for J=1J=1 eV, the HS states amount to nearly the total probabilities.

For increasing JJ, the probability for N=4N=4 decreases linearly while the probabilities for N=3N=3 and N=5N=5 increase accordingly. When the system enters the CDI phase (J>1.5J>1.5 eV), the probabilities on the two inequivalent sites become different, but the averages over both sites, shown in Fig. 5, continue their linear trend. In particular, even deep inside the CDI phase, for J=2J=2 eV, the probability of N=4N=4 remains relatively large on both sites, indicating significant fluctuations between multiplets with different NN, even in the insulating state. We note that the presence of such fluctuations and the finite probability of the N=4N=4 multiplet is consistent with the gradual increase of the occupation difference δn\delta n within the CDI phase shown in Fig. 3.

Fig. 6 shows the site-resolved eg{e_{g}} spectral functions in this regime (U=1.5U=1.5 eV and J1.5J\geq 1.5 eV). The spectral functions for J>1.5J>1.5 eV exhibit a clear gap, consistent with the zero spectral weight in Fig. 3, despite the system not being in a pure local charge state, as follows from the multiplet probabilities. This differs from a recent study using a slave-boson method [9], where the Hund’s insulator in the two-orbital model was characterized by a complete charge disproportionation where only the nominal charge multiplets are populated. It also indicates that the transition to the CDI state cannot be understood purely in terms of atomic-limit considerations such as the ones in Sec. III.1.

Refer to caption
Figure 6: Site-resolved eg{e_{g}} spectral functions for the transition from HS metallic (top) to CDI phase (middle and bottom). A gap opens for J1.75J\geq 1.75 eV. Site 1 is the more occupied one. However, both sites exhibit spectral weight above as well as below the gap (albeit with clearly different strength).

Increasing UU from 1.5 eV to 6.0 eV for fixed J=1.5J=1.5 eV (second segment in Fig. 5) results in a transition from the CDI phase back to the HS metallic region and then moves the system towards the Mott insulating region. Here, the N=4N=4 probability increases again, while the N=3N=3 and N=5N=5 probabilities decrease accordingly, and become zero at the boundary to the Mott insulator. Also, the quasiparticle weight ZZ decreases continuously towards zero when approaching the Mott phase, while it remains quite large even very close to the CDI phase boundary. However, as discussed above, it is unclear whether [1ImΣ/(iω)]1[1-\partial\imaginary\Sigma/\partial(\mathrm{i}\omega)]^{-1} is really a good quantitative measure for the quasiparticle renormalization, here. Nevertheless, we note that a rather sharp drop of ZZ on approaching the Hund’s insulator (equivalent to our CDI phase) has also been observed in Ref. 9 for a three-orbital model.

Finally, when crossing the HS/LS transition within the Mott insulating region, the HS N=4N=4 multiplet vanishes abruptly, and then, when leaving the Mott phase towards the LS metallic region, the N=3N=3 and 44 multiplets appear again, together with a gradual increase of ZZ.

Thus, to preliminary summarize the results in this section, we emphasize that we establish the presence of a spontaneously charge-disproportionated insulating phase, for strong JJ and intermediate values of UU, in the cubic five-orbital multi-band Hubbard model with an average of four electrons per site, analogous to results obtained for similar two- and three-orbital models [8; 9]. The CDI phase is characterized by a gap opening at zero frequency, but still exhibits a strong local mixing of multiplets with different number of electrons NN. This is in contrast to the Mott insulating phase, which can be characterized by a unique local multiplet with N=4N=4. In addition to the metallic and two distinct insulating phases, a LS/HS transition occurs. This transition is very sharp inside the Mott-insulating regime but a small intermediate spin region can be observed for small UU in the metallic phase.

III.3 Introducing an intrinsic site splitting and effect of reduced band width

Refer to caption
Figure 7: (a) Occupation difference δn\delta n between inequivalent sublattices and (b) spectral weight A(0)A(0) for the case with an intrinsic site splitting Δeg\Delta_{e_{g}}, plotted as function of the interaction parameters UU and JJ.

Next, we investigate the effect of an intrinsic energy difference between the eg{e_{g}} levels on the two sublattices, which is introduced, for example, by the structural distortion that generally accompanies the charge disproportionation, i.e., the R1+{R_{1}^{+}} breathing mode of the oxygen octahedra surrounding the transition-metal sites in the perovskite structure. We use a splitting of Δeg=0.436\Delta_{e_{g}}=0.436 eV as obtained from our fit to the band structure of CaFeO3 described in Sec. II.3.

Fig. 7 shows the spectral weight A(0)A(0) and occupation difference δn\delta n for this case. It can be seen that the region corresponding to the CDI phase is significantly increased compared to the case with Δeg=0\Delta_{e_{g}}=0 in Fig. 3. The transition to the insulating state now occurs already very close to the homogeneous/inhomogeneous transition line estimated from the atomic limit. Similar behavior has also been observed for the two-band model [8]. Furthermore, due to the intrinsic energy difference, a small δn>0\delta n>0 already develops in the metallic phase. On the other hand, this small occupation difference is completely suppressed when approaching the Mott-insulating phase, and the corresponding MIT occurs again very close to the atomic limit predictions based on the ege_{g} band width. Note that the corresponding lines in Fig. 7 are shifted relative to those in Fig. 3 for Δeg=0\Delta_{e_{g}}=0 as indicated in the schematic phase diagram, Fig. 2.

Due to the large extent of the CDI phase, there is now a direct competition between the CDI and the homogeneous LS metallic phase for small JJ. Nevertheless, also in this case, the corresponding phase boundary appears to be rather well described by the atomic limit prediction (gray lines in Fig. 7). Similar to the case with Δeg=0\Delta_{e_{g}}=0, we obtain a small intermediate-spin region (without significant charge disproportionation) for U2U\leq 2 eV and J=0.75J=0.75 eV (data not shown in Fig. 7).

Finally, we also assess the effect of reducing the ege_{g} band width, while keeping the local crystal field and the inter-site eg{e_{g}} splitting constant. To simplify the corresponding analysis, we use the observation that within the HS region (J1J\geq 1 eV), the gradient of all quantities depicted in the UU-JJ phase diagrams in Fig. 3 and Fig. 7 appears to be perpendicular to the two transition lines bordering the metallic phase, i.e., perpendicular to lines with U1.51J=constantU-1.51J=\text{constant}. In Fig. 8, we therefore plot all available data points for the spectral weight A(0)A(0) within the HS region of the phase diagram, both for full and reduced band width, as a function of the “effective” interaction parameter U1.51JU-1.51J (which corresponds to the difference in the local interaction energies between d4+d4d^{4}+d^{4} and d3+d5d^{3}+d^{5}).

Refer to caption
Figure 8: Effect of an 80 % reduction of the ege_{g} band width on the spectral weight A(0)A(0) for the case with nonzero intrinsic site splitting. Only data for the HS region with neg>0.5n_{e_{g}}>0.5 is shown as function of the “effective” interaction U1.51JU-1.51J. The colored background and the vertical blue line indicate the atomic-limit predictions for the two distinct insulating phases, analogous to Fig. 2, with the white metallic region in between.

Indeed, all data points plotted in this way essentially condense to a single line. In agreement with the simple atomic-limit prediction (indicated in Fig. 8 as vertical line or shaded regions), the transition to the Mott-insulating phase is shifted towards smaller interaction parameters if the band width is reduced, while the transition to the CD phase is barely affected by this change. However, the occupation difference δn\delta n in the CD region (not shown) is slightly increased for the case with reduced band width. Furthermore, one can also see that spectral weight disappears only very gradually towards the CDI state, while it disappears rather abruptly at the Mott transition.

We note that, even though the reduction of the ege_{g} band width does not significantly affect the CDI phase boundary in our calculations, the band width could nevertheless affect the coupling to the structural distortion and can thus be an important control parameter of the system. In Ref. 20, it was shown that in the two-orbital model the band width controls the relevant electronic susceptibility and is therefore crucial for stabilizing the CD through coupling to the structural distortion.

III.4 Comparison with an effective two-orbital model for equivalent sites

After establishing and analyzing the phase diagram of the full five-orbital model, we now compare this model to the simplified effective two-orbital description introduced in Sec. II.2. Since by construction the effective two-orbital model assumes a HS configuration on each site, we only compare the HS part of the phase diagram. Our objective is to test how qualitatively and quantitatively accurate the computationally less demanding simplified model is. To ensure a meaningful comparison with the full five-orbital model, the interaction parameters of the two-orbital model are always chosen to be consistent with the corresponding parameters of the full model, as described in Sec. II.2. Furthermore, we only present the case with Δeg=0\Delta_{e_{g}}=0, i.e., without an intrinsic ege_{g} site splitting between the two sublattices. Our test calculations (not shown) indicate that a site splitting does not alter the quality of the comparison between the full five-orbital model and the effective two-orbital case.

Refer to caption
Figure 9: Comparison between the simplified effective two-orbital model and the full five-orbital description in the HS region for the case without an intrinsic site splitting, i.e., Δeg=0\Delta_{e_{g}}=0. (a) The occupation difference δn\delta n between inequivalent sublattices and (b) the spectral weight A(0)A(0) are shown as function of the effective interaction parameter U1.51JU-1.51J. Shaded regions indicate the CD and Mott insulating regions predicted from the atomic limit (Sec. III.1), white background indicates the metallic region.

Fig. 9 depicts the occupation difference δn\delta n and spectral weight A(0)A(0) as function of the effective interaction U1.51JU-1.51J, both for the effective two-orbital model and the full five-orbital model. Similar to the case presented in Fig. 8, all points for the same data set, obtained for different combinations of UU and JJ, essentially reduce to a single line, demonstrating that U1.51JU-1.51J is indeed the relevant interaction parameter characterizing the behavior of the model in the HS regime.

Both models give very similar results, with only two small quantitative differences. First, the transition to the CDI phase occurs for slightly smaller effective interaction in the simplified two-orbital model. Due to the large increase of δn\delta n at the transition, this can lead to quantitative differences in the charge disproportionation of up to 0.5 electrons, but only in a very small region close to the phase transition. Second, the spectral weight at zero energy in the metallic phase is slightly reduced in the effective two-orbital model.

Refer to caption
Figure 10: Comparison of kk-resolved spectral functions between the full five-orbital model (a and c) and the simplified effective two-orbital model (b and d) in the HS metallic state close to the HS/LS transition (a and b) and in the CD insulating state (c and d). The blue lines in (b) are the bands obtained from the non-interacting part of the effective model, after taking into account the local spin splitting as well as the renormalized and spin-mixed hopping matrices.

Thus, it appears that the basic observables, such as occupations and the resulting phase boundaries, as well as the gap opening at the Mott transition, obtained from the simplified two-orbital model agree well with the corresponding results for the full five-orbital model. However, this is not true for more complex spectral properties, as can be seen from the kk-resolved spectral functions shown in Fig. 10 (obtained through analytic continuation of the self-energy).

For the full five-orbital model, the occupied and unoccupied t2gt_{2g} bands appear as rather sharp features. For the case with U=2.0U=2.0 eV and J=1.0J=1.0 eV these bands are situated at around -3 eV and 1 eV. The ege_{g} spectral weight is much more incoherent, with significant lifetime broadening due to many-body correlations. In the metallic state obtained for J=1.0J=1.0 eV, some dispersion is still visible between ω=1 eV\omega=$-1\text{\,}\mathrm{e}\mathrm{V}$ and 1 eV1\text{\,}\mathrm{e}\mathrm{V}. At higher energies, a broad incoherent spectral background is present at around 2 eV.

In contrast, the kk-resolved spectral functions obtained from the effective two-orbital model exhibit much weaker many-body effects. In the metallic regime for J=1.0J=1.0 eV, a sharp band structure is visible, in particular around ω=0\omega=0. The band dispersion is essentially identical to the one following from the renormalized and spin-mixed hopping amplitudes, also shown as blue lines for comparison in Fig. 10(b). The Coulomb interaction between the ege_{g} electrons only leads to a weak lifetime broadening of the individual bands and mostly increases the splitting between local majority and minority spin channels. Close to the CDI transition at J=2J=2 eV, the local minority spin bands are shifted to even higher energies and a gap starts opening at zero frequency, consistent with the full five-orbital model. However, in the two-orbital model, the bands around the gap seem to remain rather well defined, while the band edges at approximately ±1\pm 1 eV become rather fuzzy, which does not happen in the five-orbital case.

The fact that the effective two-orbital model exhibits much weaker many-body correlations than the full five-orbital model is to be expected, since the interaction with the t2gt_{2g} core spins merely enters via the renormalized hopping amplitudes and in the form of “classical” Zeeman fields. However, the local self-energies also seem to exhibit some non-trivial differences, leading to pronounced lifetime broadening in different energy regions. Nevertheless, these differences do not seem to critically affect integrated quantities such as occupations or the spectral weight around zero energy (see Fig. 9).

IV Conclusions and Summary

In summary, we have established the existence of a charge-disproportionated insulating phase (or Hund’s insulator) within a generic cubic five-orbital full dd-shell model and an average filling of four electrons per site. Our results are consistent with previous studies of two- and three-orbital models [8; 9], which all show the same overall phase diagram with two distinct insulating phases, the standard homogeneous Mott insulator for large UU and a spontaneously charge disproportionated insulator for large JJ and moderate UU, separated by a metallic region.

For the present case of the cubic five-orbital model, additional complexity arises from the presence of the LS/HS transition. In particular for U<2U<2 eV and J1J\lesssim 1 eV, the metallic LS state competes with the CDI state. We note that this parameter regime is realistic for materials such as ferrates, where the Coulomb repulsion between the dd electrons can be strongly screened due to the (potentially negative) charge-transfer character and the resulting strong hybridization with the oxygen pp bands. Furthermore, it can be expected that the incorporation of intersite Coulomb interactions, neglected in the present study to focus on the effect of the local Hund’s coupling, will also favor the charge disproportionated state and thus further reduce the minimal JJ required to obtain the CDI phase for a given UU [54].

Our results also show that the CDI phase is strongly affected by a small energy splitting of the ege_{g} levels between the charge-disproportionated sublattices, introduced here to mimic the effect of an additional structural distortion that accompanies the CD. Again, this is analogous to previous results obtained for the two-orbital case [8], and shows that the CDI couples very strongly to such a structural distortion. Thus, for quantitative studies of specific materials, it is crucial to account for this coupling and treat the structural distortion as a free variable [20].

An interesting aspect is also the unconventional nature of the CDI state. While for the case of the homogeneous Mott insulator, the critical UU and JJ values for the MIT can be estimated with rather good accuracy from the atomic limit, this is not the case for the transition to the (spontaneous) CDI state. Our results show that a gap in the spectral function opens even though there are still significant fluctuations between different local multiplets. This is also consistent with the gradual increase of the charge disproportionation. Nevertheless, for the case with a nonzero site splitting Δeg\Delta_{e_{g}}, the homogeneous/CD phase boundary is well approximated by the corresponding atomic limit.

Finally, we have also compared the full five-orbital model with a simplified effective two-orbital description applicable to the HS limit. The simplified model gives results very similar to the full model in the HS region, even though the transition to the CDI state occurs at slightly lower effective interaction U1.51JU-1.51J than in the full model. In spite of clear differences in the kk-resolved spectral functions, integrated quantities such as occupations or the zero-energy spectral weight A(ω=0)A(\omega=0) agree very well between the two models. However, assuming a realistic parameter regime for perovskite transition-metal oxides such as CaFeO3 around J1J\lesssim 1 eV and U=U= 1-2 eV, our calculated phase diagrams indicates that these materials could be rather close to the LS/HS transition, and therefore might require a treatment within the full model.

Our work shows that the formation of a charge-disproportionated insulator, in a regime where the Hund’s coupling outweighs the strongly screened Coulomb repulsion, is a general feature of multi-band Hubbard models at specific filling levels. Based on this observation, our rather generic cubic five-orbital model represents a good starting point and can serve as reference for future studies of specific materials including, e.g., ferrates and manganites.

*

Acknowledgements.
We thank Sophie Beck and Alexander Hampel for their help regarding the use of soliDMFT and TRIQS/DFTTools as well as for many enlightening discussions, and Alexander Gillmann, who performed initial calculations related to this project for his BSc thesis. This work was supported by ETH Zurich and the Swiss National Science Foundation through NCCR-MARVEL. Calculations have been performed on the cluster “Piz Daint” hosted by the Swiss National Supercomputing Centre and supported under project IDs s889 (User Lab) and mr26 (MARVEL).

Appendix A Atomic-limit calculations

The local interaction energies for all low-energy configurations (both LS and HS) with 2N62\leq N\leq 6 electrons per site, corresponding to the interaction Hamiltonian in Eq. (II.1) are listed in Table 2. The corresponding energies can be expressed exactly in terms of the Slater integrals F0F_{0}, F2F_{2}, and F4F_{4} (see, e.g., Ref. 35), while for the expression in terms of UU and JJ we use the definitions described in Sec. II.1. The full local energies EN(LS/HS)E_{N}^{\text{(LS/HS)}} in the atomic limit are then easily obtained by adding multiples of the crystal-field energy Δegt2g±Δeg/2\Delta_{{e_{g}}-{t_{2g}}}\pm\Delta_{{e_{g}}}/2 according to the corresponding ege_{g} occupation and the relevant sublattice.

Table 2: Interaction energies of the lowest HS/LS NN electron configuration (for 2N62\leq N\leq 6), corresponding to the Hamiltonian in Eq. (II.1), expressed both in terms of the Slater integrals F0F_{0}, F2F_{2}, F4F_{4}, and in terms of the interaction parameters UU and JJ. Note that for N=2N=2 and N=3N=3 there is no ambiguity between HS and LS.
N=2N=2 F05/49F224/441F4F_{0}-5/49F_{2}-24/441F_{4} U1.171JU-1.171J
N=3N=3 3F015/49F272/441F43F_{0}-15/49F_{2}-72/441F_{4} 3U3.513J3U-3.513J
N=4N=4, LS 6F015/49F244/441F46F_{0}-15/49F_{2}-44/441F_{4} 6U3.169J6U-3.169J
N=4N=4, HS 6F021/49F2189/441F46F_{0}-21/49F_{2}-189/441F_{4} 6U6J6U-6J
N=5N=5, LS 10F020/49F240/441F410F_{0}-20/49F_{2}-40/441F_{4} 10U3.997J10U-3.997J
N=5N=5, HS 10F035/49F2315/441F410F_{0}-35/49F_{2}-315/441F_{4} 10U10J10U-10J
N=6N=6, LS 15F030/49F260/441F415F_{0}-30/49F_{2}-60/441F_{4} 15U5.995J15U-5.995J
N=6N=6, HS 15F035/49F2315/441F415F_{0}-35/49F_{2}-315/441F_{4} 15U10J15U-10J

The transition boundary between the homogeneous d4d^{4} HS and LS states is then obtained by comparing the corresponding energies:

0\displaystyle 0 =E4(LS)E4(HS)\displaystyle=E_{4}^{\mathrm{(LS)}}-E_{4}^{\mathrm{(HS)}}
=649F2+145441F4Δegt2g\displaystyle=\frac{6}{49}F_{2}+\frac{145}{441}F_{4}-\Delta_{{e_{g}}-{t_{2g}}}
2.83JΔegt2g\displaystyle\approx 2.83J-\Delta_{{e_{g}}-{t_{2g}}} (6)

For the transition between the inhomogeneous CD region and the homogeneous HS and LS regions, respectively, we compare the corresponding energies per two sites:

0\displaystyle 0 =(E5(HS)+E3)2E4(HS)\displaystyle=\left(E_{5}^{\mathrm{(HS)}}+E_{3}\right)-2E_{4}^{\mathrm{(HS)}}
=F0849F2149F4Δeg\displaystyle=F_{0}-\frac{8}{49}F_{2}-\frac{1}{49}F_{4}-\Delta_{e_{g}}
U1.51JΔeg,\displaystyle\approx U-1.51J-\Delta_{e_{g}}\quad, (7)

and:

0\displaystyle 0 =(E5(HS)+E3)2E4(LS)\displaystyle=\left(E_{5}^{\mathrm{(HS)}}+E_{3}\right)-2E_{4}^{\mathrm{(LS)}}
=F02049F2299441F4+2(Δegt2gΔeg/2)\displaystyle=F_{0}-\frac{20}{49}F_{2}-\frac{299}{441}F_{4}+2(\Delta_{{e_{g}}-{t_{2g}}}-\Delta_{e_{g}}/2)
U7.17J+2Δegt2gΔeg.\displaystyle\approx U-7.17J+2\Delta_{{e_{g}}-{t_{2g}}}-\Delta_{e_{g}}\quad. (8)

Here, of course, we assume that in the presence of an eg{e_{g}} site splitting, the HS d5d^{5} state is formed on the site with the lower eg{e_{g}} energy.

The border of the Mott insulating regions within the homogeneous HS and LS phases are estimated by comparing the energy of the lowest inter-site charge transfer excitation (2d4d5+d32d^{4}\rightarrow d^{5}+d^{3}) with the relevant band width. Note that for the homogeneous HS region, this results in the same energy terms as in Eq. (7), E5(HS)+E32E4(HS)E_{5}^{\mathrm{(HS)}}+E_{3}-2E_{4}^{\mathrm{(HS)}}, but now this is set equal to the eg{e_{g}} band width WegW_{e_{g}}. The same argument is also applied to identify the metal-insulator boundary within the homogeneous LS phase, where the electrons are restricted to states within the t2g{t_{2g}} sub-manifold and therefore the inter-site charge transfer energy is compared to the t2g{t_{2g}} band width Wt2gW_{t_{2g}}:

Wt2g\displaystyle W_{t_{2g}} =(E5(LS)+E3)2E4(LS)\displaystyle=\left(E_{5}^{\mathrm{(LS)}}+E_{3}\right)-2E_{4}^{\mathrm{(LS)}}
=F0549F224441F4\displaystyle=F_{0}-\frac{5}{49}F_{2}-\frac{24}{441}F_{4}
U1.17J.\displaystyle\approx U-1.17J\quad. (9)

Note that if the right side of Eq. (A) is set to zero, this also defines a potential transition to an inhomogeneous LS phase, which is indeed more favorable than the HS CD state for J<Δegt2g/3J<\Delta_{{e_{g}}-{t_{2g}}}/3. However, with Δegt2g2.5\Delta_{{e_{g}}-{t_{2g}}}\approx 2.5 eV, this requires J0.83J\lesssim 0.83 eV and simultaneously U1.17J0.98U\lesssim 1.17J\lesssim 0.98 eV. Since we do not consider this parameter range in our DMFT calculations, we neglect the corresponding phase also in the schematic phase diagram in Fig. 2.

The estimation of the metal-insulator boundary within the inhomogeneous CD region is less straightforward. As mentioned in Sec. III.1, the effective kinetic energy required to destabilize the CDI state towards the homogeneous metal cannot be estimated from simple band-width considerations (at least for Δeg=0\Delta_{e_{g}}=0). In addition, the CDI state can be destabilized towards low UU values through an effective kinetic hopping within only one specific sublattice. As pointed out in Ref. 8, this involves indirect hopping processes, which again complicates a simple band-width estimation of the corresponding kinetic energy.

For the case with hopping only among the nominal d5d^{5} sites (2d5d4+d62d^{5}\rightarrow d^{4}+d^{6}), one obtains:

(Wt2g+Weg)/2\displaystyle(W_{t_{2g}}+W_{e_{g}})/2 (E6(HS)+E4(HS))2E5(HS)\displaystyle\leq\left(E_{6}^{\mathrm{(HS)}}+E_{4}^{\mathrm{(HS)}}\right)-2E_{5}^{\mathrm{(HS)}}
=F0+1449F2+126441F4\displaystyle=F_{0}+\frac{14}{49}F_{2}+\frac{126}{441}F_{4}
(Δegt2gΔeg/2)\displaystyle\quad-(\Delta_{{e_{g}}-{t_{2g}}}-\Delta_{e_{g}}/2)
=U+4JΔegt2g+Δeg/2\displaystyle=U+4J-\Delta_{{e_{g}}-{t_{2g}}}+\Delta_{e_{g}}/2 (10)

Note that here we have used the average band width as an upper limit for the effective kinetic energy, which corresponds to the red-hatched area around J1J\approx 1 eV in Fig. 2. The true effective band width corresponds to an indirect hopping between d5d^{5} sites and is, therefore, expected to be significantly smaller, and thus this transition barely plays a role in our considerations. We can also identify (Wt2g+Weg)/2+Δegt2gWfulld(W_{t_{2g}}+W_{e_{g}})/2+\Delta_{{e_{g}}-{t_{2g}}}\equiv W_{\mathrm{full-d}} leading to

WfulldΔeg/2U+4J\displaystyle W_{\mathrm{full-d}}-\Delta_{e_{g}}/2\leq U+4J (11)

For the case with hopping only between nominal d3d^{3} sites, and assuming that the involved d4d^{4} state is LS, one obtains (again using the full t2g{t_{2g}} band width as upper limit):

Wt2g\displaystyle W_{t_{2g}} (E4(LS)+E2)2E3\displaystyle\leq\left(E_{4}^{\mathrm{(LS)}}+E_{2}\right)-2E_{3}
=F0+1049F2+76441F4\displaystyle=F_{0}+\frac{10}{49}F_{2}+\frac{76}{441}F_{4}
U+2.69J\displaystyle\approx U+2.69J (12)

This only predicts a metallic state in the region where the hopping between d5d^{5} sites is already active, and is therefore not relevant. However, if one considers also hopping through a d4d^{4} HS state, and using an average band width as in Eq. (A), one obtains:

(Wt2g+Weg)/2\displaystyle(W_{t_{2g}}+W_{e_{g}})/2 (E4(HS)+E2)2E3\displaystyle\leq\left(E_{4}^{\mathrm{(HS)}}+E_{2}\right)-2E_{3}
=F0+449F269441F4\displaystyle=F_{0}+\frac{4}{49}F_{2}-\frac{69}{441}F_{4}
+(Δegt2g+Δeg/2)\displaystyle\quad+(\Delta_{{e_{g}}-{t_{2g}}}+\Delta_{e_{g}}/2)
U0.15J+Δegt2g+Δeg/2\displaystyle\approx U-0.15J+\Delta_{{e_{g}}-{t_{2g}}}+\Delta_{e_{g}}/2 (13)

This can in principle destabilize the CDI state for very low UU and very high JJ and is also indicated by the red-hatched area in Fig. 2.

Finally, we note that none of the multiplet energies listed above would be affected by the spin-flip and pair-hopping terms that we ignore within the density-density approximation, Eq. (II.1). These additional terms would only affect higher energy multiplets, which are not crucial for a correct description of the ground state phase diagram, as they do not affect the phase boundaries obtained from this atomic limit consideration. This provides a good a priori justification for our use of the density-density approximation.

References