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

Magnetic Couplings in Edge-Sharing High-Spin d7d^{7} Compounds

Stephen M. Winter Department of Physics and Center for Functional Materials, Wake Forest University, NC 27109, USA
Abstract

High-spin d7d^{7} Co(II) compounds have recently been identified as possible platforms for realising highly anisotropic and bond-dependent couplings featured in quantum-compass models such as the celebrated Kitaev model. In order to evaluate this potential, we consider all symmetry-allowed contributions to the magnetic exchange for ideal edge-sharing bonds. Though a combination of ab-initio and cluster many-body calculations we conclude that bond-dependent couplings are generally suppressed in favor of Heisenberg exchange for real materials. Consequences for several prominent materials including Na2Co2TeO6 and BaCo2(AsO4)2 are discussed.

I Introduction

Pursuit of strongly anisotropic dd-block magnets has been motivated by the possibility of material realization of quantum compass modelsNussinov and Van Den Brink (2015), such as Kitaev’s celebrated honeycomb modelKitaev (2006). In these materials, competition between different bond-dependent magnetic interactions produces an extensive classical degeneracy conducive to quantum spin-liquid ground statesHermanns et al. (2018); Broholm et al. (2020); Zhou et al. (2017). Realising these conditions in real materials requires precise tuning and suppression of the usual isotropic magnetic exchange. This can be accomplished, in principle, in edge-sharing compounds with d5d^{5} filling and strong spin-orbital coupling. Remarkably, for ideal considerations, the specific spin-orbital composition of the local moments suppresses all couplings except those bond-dependent Ising couplings precisely prescribed by the Kitaev modelJackeli and Khaliullin (2009). This revelation initiated a number of studiesWinter et al. (2017a); Trebst (2017) in 5d55d^{5} Ir(IV) compounds such as A2IrO3 (A = Na, Li)Singh and Gegenwart (2010) and the 4d54d^{5} Ru(III) compoundPlumb et al. (2014); Banerjee et al. (2016, 2017) α\alpha-RuCl3. These studies have revealed clear evidence of dominant bond-dependent anisotropic couplings in these compoundsHwan Chun et al. (2015); Suzuki et al. (2021), leading to a variety of anomalous behaviors from the breakdown of conventional magnon excitationsWinter et al. (2017b) to the possibility of a field-induced spin liquidBanerjee et al. (2018); Kasahara et al. (2018); Yokoi et al. (2021). However, while Kitaev couplings are thought to be the largest interaction, other couplings of similar magnitude always lift the classical degeneracy leading to magnetic order at zero field.

In this context, the seminal work of Liu et al.Liu and Khaliullin (2018); Liu (2021); Liu et al. (2020) and Sano et al. Sano et al. (2018) renewed hope for realizing Kitaev’s spin liquid, by showing that the magnetic exchange in high-spin 3d73d^{7} Co(II) ions may also produce dominant Kitaev interactions for ideal considerations. In particular, these studies assumed the dominant hopping between transition metal ions occurs via hybridization with ligand ions, which suppresses other couplings. While this condition is satisfied for 5d55d^{5} Ir(IV) compounds such as A2IrO3, the presence of significant direct metal-metal hopping in 4d54d^{5} α\alpha-RuCl3 is the primary source of non-Kitaev interactionsRau et al. (2014); Winter et al. (2016). It is not clear that these assumptions are satisfied in 3d3d systems.

The possibility of strong bond-dependent Kitaev interactions also challenges the conventional view that Co(II) compounds typically have bond-independent XXZ anisotropy largely driven by the effects of local crystal field distortions on the j1/2j_{1/2} doublets Lines (1963); Oguchi (1965). For example, CoNb2O6 (CNO) is considered to be a prototypical 1D Ising ferromagnetScharf et al. (1979); Maartense et al. (1977); Kobayashi et al. (1999), and has been studied in the context of transverse-field Ising criticality Lee et al. (2010); Coldea et al. (2010); Morris et al. (2014). The structure consists of zigzag chains of edge-sharing CoO6 octahedra. While this bonding geometry might be expected to produce large bond-dependent couplings, the dominant nearest neighbor interaction is known to have an Ising form SiαSjαS_{i}^{\alpha}S_{j}^{\alpha} with a common α\alpha-axis for all bonds. Recent studies have highlighted the importance of small deviationsFava et al. (2020); Morris et al. (2021), but it is nonetheless evident that the Kitaev coupling is not dominant.

More recently, the pursuit of 2D honeycomb materials with large bond-dependent couplings has drawn attention to Na3Co2SbO6 (NCSO), and Na2Co2TeO6 (NCTO). Both materials show zigzag antiferromagnetic orderLefrançois et al. (2016); Bera et al. (2017); Wong et al. (2016); Bera et al. (2017); Chen et al. (2021). This ground state is natural for strong bond-dependent couplingsChaloupka et al. (2013); Rau et al. (2014), although longer-range Heisenberg J2J_{2} and J3J_{3} may also be invokedFouet et al. (2001); Kimchi and You (2011) Indeed, analysis of inelastic neutron scattering has led to a wide variety of proposed models for the couplingsChen et al. (2021); Songvilay et al. (2020); Kim et al. (2021); Lin et al. (2021), which span the entire range from dominant Heisenberg to dominant Kitaev. Overall, the relative role of nearest neighbor bond-dependent coupling vs. longer range Heisenberg exchange remains unclear.

Two more honeycomb materials of recent interest are BaCo2(AsO4)2 (BCAO) and BaCo2(PO4)2 (BCPO). Of these, BCPO displays only short-range incommensurate correlations, suggesting strong frustrationNair et al. (2018). BCAO orders in a state intermediary between zigzag antiferromagnetic and ferromagnetic states with unconventional magnon dispersionRegnault et al. (2006, 2018), which has been discussed as an incommensurate helimagnetRegnault et al. (1977) or double stripe zigzagRegnault et al. (2018). Under applied field in-plane, BCAO undergoes a series of phase transitionsRegnault et al. (1977) between magnetization plateaus, and was proposed to host a field-induced spin liquidZhong et al. (2020); Zhang et al. (2021). However, this was recently called into question due to the appearance of sharp magnon modes in each of the phasesShi et al. (2021). As with NCTO, the relative role of different couplings is a subject of much discussion; the first ab-initio studiesDas et al. (2021); Maksimov et al. (2022) favored a nearly XXZ model, in contrast with the assumption of large Kiteav interactions.

All of these findings call for a reinvestigation of the magnetic couplings in edge-sharing Co(II) materials. In this work, we find, in contrast to the assumptions of the initial theoretical analysis, that ligand-mediated hopping is not large in these compounds. For this reason the character of the magnetic couplings is significantly altered from the expected Kitaev form. In particular, ferromagnetic Heisenberg JJ typically dominates the nearest neighbor couplings, while a variety of smaller anisotropic couplings may appear depending on the specific details of the hopping and crystal field distortions, which are discussed in this work.

The paper is organized as follows: We first define the electronic Hamiltonian, and review the single-ion ground state, and the effect of crystal field distortions on the spin-orbital composition of the j1/2j_{1/2} moments. We then analyze the full set of relevant symmetry-allowed hoppings in edge-sharing bonds, and present calculations of such hoppings for real materials to establish their physically relevant ranges. On the basis of these hoppings, we then compute the resulting magnetic couplings. Finally, we summarize the results, and discuss them in the context of materials of recent interest.

II Electronic Hamiltonian

II.1 General Form

We first define the electronic Hamiltonian relevant for the edge-sharing bond depicted in Fig. 1(a), which is conventionally called a Z-bond to indicate that the global cubic zz-axis is perpendicular to the shared edge. There are generally two different orbital bases that can be employed in the description of magnetic couplings. One may consider explicitly either (i) the full basis of metal dd-orbitals and ligand pp-orbitals, or (ii) the dd-orbitals only, for which the effects of the pp-orbitals are downfolded into the dd-orbital Wannier function basis. The latter Wannier functions are anti-bonding combinations of dd- and pp-orbitals depicted schematically in Fig. 1(b). In the following, we mostly consider the downfolded basis, but discuss the deficiencies of this approach below.

Refer to caption
Figure 1: (a) Edge-sharing bond with cubic coordinates defined. (b) Cartoon of the dd-orbital Wannier functions in the downfolded dd-only basis.

In the downfolded dd-only basis, the total Hamiltonian is given by:

=hop+ii\displaystyle\mathcal{H}=\mathcal{H}_{\rm hop}+\sum_{i}\mathcal{H}_{i} (1)

where hop\mathcal{H}_{\rm hop} denotes the hopping between Co sites. i\mathcal{H}_{i} is the local Hamiltonian at each Co site ii, which is a sum, respectively, of Coulomb interactions, crystal-field splitting, and spin-orbit coupling:

i=U+CFS+SOC\displaystyle\mathcal{H}_{i}=\mathcal{H}_{U}+\mathcal{H}_{\rm CFS}+\mathcal{H}_{\rm SOC} (2)

The SOC term is written:

SOC=λ𝐋i𝐒i\displaystyle\mathcal{H}_{\rm SOC}=\lambda\mathbf{L}_{i}\cdot\mathbf{S}_{i} (3)

where the atomic SOC constant for Co is roughly λCo60\lambda_{\rm Co}\approx 60 meV. The Coulomb interactions are most generally written:

U=α,β,δ,γσ,σUαβγδci,α,σci,β,σci,γ,σci,δ,σ\displaystyle\mathcal{H}_{U}=\sum_{\alpha,\beta,\delta,\gamma}\sum_{\sigma,\sigma^{\prime}}U_{\alpha\beta\gamma\delta}\ c_{i,\alpha,\sigma}^{\dagger}c_{i,\beta,\sigma^{\prime}}^{\dagger}c_{i,\gamma,\sigma^{\prime}}c_{i,\delta,\sigma} (4)

where α,β,γ,δ\alpha,\beta,\gamma,\delta label different dd-orbitals. 111The coefficients UαβγδU_{\alpha\beta\gamma\delta} may be grouped according to the number of unique orbital indices, from one to four. For example, the intra-orbital Hubbard terms ni,α,ni,α,n_{i,\alpha,\uparrow}n_{i,\alpha,\downarrow} have one unique index α\alpha, while the inter-orbital Hubbard terms ni,α,σni,β,σn_{i,\alpha,\sigma}n_{i,\beta,\sigma^{\prime}} have two unique indices α,β\alpha,\beta. In the spherically symmetric approximation Sugano (2012), the Coulomb coefficients with three and four indices vanish unless at least one of the orbitals is an ege_{g} orbital. For this reason, t2gt_{2g}-only (and ege_{g}-only) models reduce to the familiar Kanamori formGeorges et al. (2013); Pavarini (2014), which includes only Hubbard density-density repulsion, Hund’s exchange, and pair-hopping contributions. However, when both ege_{g} and t2gt_{2g} orbitals are considered together, it is important to include the full rotationally symmetric Coulomb terms. This is particularly true when computing anisotropic magnetic exchange, because any approximations to the Coulomb Hamiltonian are likely to explicitly break rotational symmetry, leading to erroneous sources of anisotropy. In the spherically symmetric approximation Sugano (2012), the coefficients UαβγδU_{\alpha\beta\gamma\delta} are all related to the three Slater parameters F0,F2,F4F_{0},F_{2},F_{4}. In terms of these, the familiar t2gt_{2g} Kanamori parameters are, for example:

Ut2g=F0+449(F2+F4)\displaystyle U_{t2g}=F_{0}+\frac{4}{49}\left(F_{2}+F_{4}\right) (5)
Jt2g=349F2+20441F4\displaystyle J_{t2g}=\frac{3}{49}F_{2}+\frac{20}{441}F_{4} (6)

Unless otherwise stated, we use Ut2g=3.25U_{t2g}=3.25 eV, and Jt2g=0.7J_{t2g}=0.7 eV to model Co(II) compounds, following Ref. Das et al., 2021. We also take the approximate ratio F4/F2=5/8F_{4}/F_{2}=5/8, which is appropriate for 3d3d elementsPavarini (2014).

For the crystal-field Hamiltonian, we consider an ideal trigonal distortion within D3dD_{3d} site symmetry. The Hamiltonian can be written:

CFS=σ𝐜i,σ𝔻𝐜i,σ\displaystyle\mathcal{H}_{\rm CFS}=\sum_{\sigma}\mathbf{c}_{i,\sigma}^{\dagger}\ \mathbb{D}\ \mathbf{c}_{i,\sigma} (7)

where:

𝐜i,σ=(ci,yz,σci,xz,σci,xy,σci,z2,σci,x2y2,σ)\displaystyle\mathbf{c}_{i,\sigma}^{\dagger}=\left(c_{i,yz,\sigma}^{\dagger}\ c_{i,xz,\sigma}^{\dagger}\ c_{i,xy,\sigma}^{\dagger}\ c_{i,z^{2},\sigma}^{\dagger}\ c_{i,x^{2}-y^{2},\sigma}^{\dagger}\right) (8)

creates an electron in one of the dd-orbital (Wannier) functions at site ii. In terms of the global cubic (xyz)(xyz) coordinates defined in Fig. 2, the CFS matrix can be written:

𝔻=(0Δ2Δ200Δ20Δ200Δ2Δ2000000Δ100000Δ1)\displaystyle\mathbb{D}=\left(\begin{array}[]{ccccc}0&\Delta_{2}&\Delta_{2}&0&0\\ \Delta_{2}&0&\Delta_{2}&0&0\\ \Delta_{2}&\Delta_{2}&0&0&0\\ 0&0&0&\Delta_{1}&0\\ 0&0&0&0&\Delta_{1}\end{array}\right) (14)

where Δ1\Delta_{1} is the t2gt_{2g}-ege_{g} splitting, and Δ2\Delta_{2} is the trigonal term. Trigonal distortions are relevant for BCAO, BCPO, NCTO, and NCSO. Generally, Δ2>0\Delta_{2}>0 corresponds to trigonal elongation, as shown in Fig. 2, although the actual sign is further influenced by the details of the ligand environments and longer ranged Coulomb potentials. Without SOC, the t2gt_{2g} levels are split into a doubly degenerate ee pair and a singly degenerate aa level, with EaEe=3Δ2E_{a}-E_{e}=3\Delta_{2}. As discussed below, the trigonal splitting has a strong impact on the nature of the local moments.

Refer to caption
Figure 2: Energy level diagram showing the splitting of the local single-particle electronic orbitals in the absence of spin-orbit coupling. Occupied orbitals in the high-spin d7d^{7} configuration are indicated.

The effective dd-dd hopping between transition metal sites is described by:

hop=ij,σ𝐜i,σ𝕋ij𝐜j,σ\displaystyle\mathcal{H}_{\rm hop}=\sum_{ij,\sigma}\mathbf{c}_{i,\sigma}^{\dagger}\ \mathbb{T}_{ij}\ \mathbf{c}_{j,\sigma} (15)

For an ideal edge-sharing bond with 9090^{\circ} metal-ligand-metal bond angles, C2vC_{2v} symmetry restricts the form of the hopping matrices. In terms of the global (x,y,z)(x,y,z) coordinates defined in Fig. 3, the matrices are constrained to take the following form, for the Z-bond:

𝕋Z=(t1t2000t2t100000t3t6000t6t400000t5)\displaystyle\mathbb{T}_{Z}=\left(\begin{array}[]{ccccc}t_{1}&t_{2}&0&0&0\\ t_{2}&t_{1}&0&0&0\\ 0&0&t_{3}&t_{6}&0\\ 0&0&t_{6}&t_{4}&0\\ 0&0&0&0&t_{5}\end{array}\right) (21)

Of these, t1,t3,t4t_{1},t_{3},t_{4}, and t5t_{5} are primarily direct hopping between the dd-orbitals on the metal atoms, as shown in Fig. 3. By contrast, t2t_{2} and t6t_{6} have significant contributions from both direct hopping and ligand-assisted hopping, which arises from the hybridization between the metal and ligand orbitals. Realistic ranges for these hoppings are discussed in section II.3.

Refer to caption
Figure 3: Summary of symmetry allowed hoppings for ideal Z-bonds with C2vC_{2v} symmetry and 9090^{\circ} metal-ligand-metal bond angles. t1,t3,t4t_{1},t_{3},t_{4}, and t5t_{5} arise from direct metal-metal hopping, while t2t_{2} and t6t_{6} have contributions from both direct and ligand-assisted processes. The global (xyz)(xyz) and local (e^1e^2e^3)(\hat{e}_{1}\hat{e}_{2}\hat{e}_{3}) coordinates are shown. The pp-orbital contributions to the Wannier functions have been omitted except in cases where ligand-assisted hopping is depicted.

Previous worksLiu and Khaliullin (2018); Liu (2021); Liu et al. (2020); Sano et al. (2018) on the possibility of dominant Kitaev magnetic exchange in high-spin 3d73d^{7} Co(II) ions have highlighted the ideal regime is found for small trigonal splitting (Δ2λ\Delta_{2}\ll\lambda) and dominant ligand-assisted hopping (t2t1,t3,t4,t5,t6t_{2}\gg t_{1},t_{3},t_{4},t_{5},t_{6}). In principle, if either of these conditions is violated, the interactions may deviate significantly from the Kitaev form. We consider these possible deviations in the following sections.

II.2 Composition of Local Moments

For the “high-spin” d7d^{7} case, the ground state has nominal configuration (t2g)5(eg)2(t_{2g})^{5}(e_{g})^{2}, with three unpaired electrons (S=3/2S=3/2), as shown in Fig. 2. For 3d3d transition metals, the constants in the Hamiltonian i\mathcal{H}_{i} typically satisfy JH,Δ1λ,Δ2J_{H},\Delta_{1}\gg\lambda,\Delta_{2}. This ensures that configurations belonging precisely to the (t2g)5(eg)2(t_{2g})^{5}(e_{g})^{2}, S=3/2S=3/2 manifold carry the dominant weight in the low energy multiplets, even when SOC and trigonal splitting are considered. For this reason, the low-energy single-ion states may be conveniently described by projecting i\mathcal{H}_{i} into this manifold, and diagonalizing. This procedure was employed in Ref. Lines, 1963, and is repeated in this section for the purpose of discussion. More detailed discussions can be found in Ref. Lines, 1963; Liu, 2021.

In the absence of trigonal splitting (Δ2=0\Delta_{2}=0), there is a three-fold orbital degeneracy associated with the t2gt_{2g} levels, leading to an effective orbital momentum Leff=1L_{\rm eff}=1. Spin-orbit coupling SOC=λ𝐋𝐒\mathcal{H}_{\rm SOC}=\lambda\mathbf{L}\cdot\mathbf{S} splits the low energy multiplets into Jeff=1/2J_{\rm eff}=1/2, 3/2, and 5/25/2 states. For Δ2=0\Delta_{2}=0, the multiplet energies satisfy:

E3/2E1/2=12λ\displaystyle E_{3/2}-E_{1/2}=\frac{1}{2}\lambda (22)
E5/2E1/2=43λ\displaystyle E_{5/2}-E_{1/2}=\frac{4}{3}\lambda (23)

Given λCo60\lambda_{\rm Co}\approx 60 meV, the j1/2j3/2j_{1/2}\to j_{3/2} excitation is expected to appear in the range of 30\sim 30 meV, as has been seen experimentally in numerous compoundsSarte et al. (2018); Ross et al. (2017); Songvilay et al. (2020); Kim et al. (2021).

For finite Δ2\Delta_{2}, the evolution of the multiplet energies is shown in Fig. 4(a). For all values of trigonal splitting, the single-ion ground state is a doublet; it is smoothly connected to the pure Jeff=1/2J_{\rm eff}=1/2 doublet as Δ2\Delta_{2} vanishes. The ground state doublet can be written in terms of |mL,mS|m_{L},m_{S}\rangle states asLines (1963):

|j1/2,+12=\displaystyle\left|j_{1/2},+\frac{1}{2}\right\rangle= c1|1,32+c2|0,12+c3|1,12\displaystyle\ c_{1}\left|-1,\frac{3}{2}\right\rangle+c_{2}\left|0,\frac{1}{2}\right\rangle+c_{3}\left|1,-\frac{1}{2}\right\rangle (24)
|j1/2,12=\displaystyle\left|j_{1/2},-\frac{1}{2}\right\rangle= c1|1,32+c2|0,12+c3|1,12\displaystyle\ c_{1}\left|1,-\frac{3}{2}\right\rangle+c_{2}\left|0,-\frac{1}{2}\right\rangle+c_{3}\left|-1,\frac{1}{2}\right\rangle (25)

where the coefficients cnc_{n} vary with Δ2,λ\Delta_{2},\lambda as shown in Fig. 4(b). More detailed expressions for the |mL,mS|m_{L},m_{S}\rangle states in terms of the occupation of the single-particle dd-orbitals is given in Appendix A. Analytical expression for cnc_{n} can be found in Ref. Lines, 1963; for Δ2=0\Delta_{2}=0, the coefficients are c1=1/2c_{1}=1/\sqrt{2}, c2=1/3c_{2}=1/\sqrt{3}, c3=1/6c_{3}=1/\sqrt{6}. From Fig. 4(b), it is evident that the composition of the lowest doublet changes dramatically with finite Δ2\Delta_{2}.

Refer to caption
Figure 4: Evolution of the single-ion properties with trigonal CFS Δ2\Delta_{2} following Ref. Lines, 1963. (a) Energy spectrum. (b) Wavefunction coefficients cnc_{n} for the ground state j1/2j_{1/2} doublet. (c) gg-tensor components for the ground state doublet. g||g_{||} refers to the direction parallel to the trigonal distortion axis, x^+y^+z^\hat{x}+\hat{y}+\hat{z}.

In the limit of large trigonal elongation Δ2>0\Delta_{2}>0, the wavefunction coefficients converge to c21c_{2}\to 1 and c1,c30c_{1},c_{3}\to 0. This reflects the fact that the unpaired hole in the t2gt_{2g} levels would occupy the singly degenerate aa level, thus quenching the orbital moment completely (see Fig. 2). As a result, spin-orbit coupling effects are suppressed, and the fourfold degeneracy of the nearly pure S=3/2S=3/2 moment is restored, such that the energetic splitting between the lowest two doublets becomes small (Fig. 4(a)). In this limit, the ms=±1/2m_{s}=\pm 1/2 states lie slightly below the ms=±3/2m_{s}=\pm 3/2 states due to the residual effects of SOC, and thus the splitting of the lowest two doublets may be considered as a residual easy-plane single-ion anisotropy. As such, the gg-tensor (Fig. 4(c)) for the lowest doublet satisfies g>g||g_{\perp}>g_{||}, where g||g_{||} refers to the component along the trigonal (111)(111) axis.

For the opposite case of trigonal compression Δ2<0\Delta_{2}<0, the unpaired hole in the t2gt_{2g} levels occupies the doubly degenerate ee levels (see Fig. 2), thus retaining some orbital degeneracy consistent with Leff=1/2L_{\rm eff}=1/2. As depicted in Fig. 4(a), the effect of SOC is then to split the S=3/2S=3/2, Leff=1/2L_{\rm eff}=1/2 manifold into four doublets. The lowest doublet remains energetically separated from the higher states. For increasing trigonal distortion, the wavefunction coefficients for the lowest doublet approach c11c_{1}\to 1, c2,c30c_{2},c_{3}\to 0, implying that it is composed of pure ms=±3/2m_{s}=\pm 3/2 states, according to Eq. (10, 11). Consistently, the gg-tensor satisfies g||gg_{||}\gg g_{\perp} in this limit (Fig. 4(c)).

Finally, it should be emphasized that an effective J=12J=\frac{1}{2} model that incorporates only the lowest doublet remains valid only as long as the energetic splitting between the lowest doublet and first excited state remains large compared to the intersite magnetic exchange. For large Δ2<0\Delta_{2}<0, the gap between the lowest doublets converges to λ/320\lambda/3\sim 20 meV, which should typically exceed the intersite magnetic coupling. For this reason, a model incorporating only the lowest doublet may remain valid even for large Δ2<0\Delta_{2}<0. For Δ2>0\Delta_{2}>0, which is expected to be relevant for BCAO, BCPO, NCTO, and NCSO, the range of validity is roughly constrained to Δ2<λ/2\Delta_{2}<\lambda/2. For larger values of Δ2\Delta_{2}, it would be more appropriate to consider an S=3/2S=3/2 model with easy-plane single-ion anisotropy. A detailed analysis of the experimental values of Δ2\Delta_{2} for various Co(II) compounds may be found in Ref. Liu, 2021. All of these materials are expected to be well described by a J=12J=\frac{1}{2} model, which can be verified experimentally by the observation of the well-defined local j1/2j3/2j_{1/2}\to j_{3/2} excitations in an energy range larger than the dispersion of magnetic excitations Sarte et al. (2018); Ross et al. (2017); Songvilay et al. (2020); Kim et al. (2021). For the materials of interests, the trigonal splitting is therefore not sufficiently large to invalidate the j1/2j_{1/2} picture, but may nonetheless have a considerable effect on the magnetic couplings, as discussed in the following sections.

II.3 Survey of Hopping in Real Materials

As noted in previous sections, the key question for the realization of dominant Kitaev coupling in high-spin d7d^{7} compounds is the relative importance of direct metal-metal vs. ligand-assisted hopping processes. In terms of the usual Slater-Koster hoppings, the downfolded dd-only hoppings are given approximately by:

t1\displaystyle t_{1}\approx 12tddπ\displaystyle\ \frac{1}{2}t_{dd}^{\pi} (26)
t2\displaystyle t_{2}\approx 12tddπ+(tpdπ)2Δpd\displaystyle\ -\frac{1}{2}t_{dd}^{\pi}+\frac{(t_{pd}^{\pi})^{2}}{\Delta_{pd}} (27)
t3\displaystyle t_{3}\approx 34tddσ\displaystyle\ \frac{3}{4}t_{dd}^{\sigma} (28)
t4\displaystyle t_{4}\approx 12tddσ\displaystyle\ -\frac{1}{2}t_{dd}^{\sigma} (29)
t5\displaystyle t_{5}\approx tddπ\displaystyle\ t_{dd}^{\pi} (30)
t6\displaystyle t_{6}\approx 34tddσtpdπtpdσΔpd\displaystyle\ \frac{\sqrt{3}}{4}t_{dd}^{\sigma}-\frac{t_{pd}^{\pi}t_{pd}^{\sigma}}{\Delta_{pd}} (31)

where Δpd\Delta_{pd} is the charge-transfer energy from Co dd to the ligand pp orbitals. Here, we have ignored contributions from tddδt_{dd}^{\delta}. For t2t_{2} and t6t_{6}, the second terms represent ligand-assisted hopping. There are two main factors that can affect these: (i) the Co-Co bond lengths (which primarily affect the direct overlap of metal dd-orbitals on adjacent metal sites) and (ii) the degree of hybridization with the ligands (as quantified by the relative weight of the pp-orbitals in the dd-orbital Wannier functions, which scales as tpd/Δpd\propto t_{pd}/\Delta_{pd}).

In order to investigate the hopping trends with bond-length dependence, we first considered hypothetical cubic CoO (NaCl type; Fm3¯mFm\bar{3}m) structures with a symmetrically stretched unit cells, allowing us to investigate a continuous range of geometries. This construction maintains 90 Co-O-Co bond angles, which deviates slightly from real materials, but allows us to consider ideal trends. In order to estimate the hoppings, we employed fully relativistic density functional theory calculations performed with FPLOKoepernik and Eschrig (1999); Opahle et al. (1999) at the GGA (PBEPerdew et al. (1996)) level. Hopping integrals were extracted by formulating Wannier orbitals via projection onto atomic pp and/or dd-orbitals. For example, the Slater-Koster hoppings were extracted from the CoO calculations by fitting with an 8-band (3d+2p)(3d+2p) model including explicitly the O orbitals. Results are shown in Fig. 5(a) as a function of Co-Co distance.

Refer to caption
Figure 5: Evolution of the relevant hoppings as a function of Co-Co distance. (a) Hoppings in the full (3d+2p)(3d+2p) scheme for hypothetical symmetrically stretched structures of CoO (see text). (b) Hoppings in the downfolded dd-only scheme. Solid lines correspond to hypothetical stretched cubic CoO. Points correspond to real materials; BCAO = BaCo2(AsO4)2, NCTO = Na2Co2TeO6, NCSO = Na3Co2SbO6, CNO = CoNb2O6.

As expected, the largest magnitude corresponds to tpdσt_{pd}^{\sigma}, which quantifies the hopping between ligand pp and metal ege_{g} orbitals. This is followed in magnitude by tpdπt_{pd}^{\pi}, which quantifies the hopping between ligand pp and metal t2gt_{2g} orbitals. The difference in magnitude between tpdσt_{pd}^{\sigma} and tpdπt_{pd}^{\pi} results in a larger degree of metal-ligand hybridization for the Wannier ege_{g} orbitals than the t2gt_{2g} orbitals. This is, of course, the origin of octahedral crystal field splitting Δ1\Delta_{1}, as described by standard ligand field theory. From the signs of the Slater-Koster hoppings, one can see that the contribution to t2t_{2} and t6t_{6} from ligand-assisted hopping is always positive, while the direct hopping contribution is always negative. The balance of this competition is therefore governed by the charge-transfer energy Δpd\Delta_{pd}. For 3d3d transition metal compounds, this quantity is typically large, which reduces ligand-metal hybridization relative to their 4d4d and 5d5d counterparts. It is precisely this effect that leads to smaller t2gegt_{2g}-e_{g} splitting Δ1\Delta_{1} for 3d3d transition metal compounds Gray (1965), which is a key requirement for stability of the high-spin state in Co 3d73d^{7} compounds. For this reason, ligand-assisted hopping is expected to be suppressed overall. This leads to strong competition between hopping processes and overall suppression of t2t_{2} and t6t_{6}. In order to demonstrate this effect, we show in Fig. 5(b) (solid lines) the downfolded hoppings extracted for the hypothetical CoO structures by Wannier fitting with a 5-band (3d-only) basis. The results are compatible with (but not exactly given by) Eq’ns (26) - (31) with Δpd4.5\Delta_{pd}\sim 4.5 eV. For all lattice constants, we find that t3t_{3} is the dominant hopping rather than the ligand-assisted t2t_{2} and t6t_{6}.

While we leave complete discussion of individual materials for later work, it is useful to confirm realistic ranges of hoppings. In order to do so, we performed additional calculations on several prominent oxide materials based on literature structures: CoNb2O6 (Ref. Sarvezuk et al., 2011), BaCo2(AsO4)2 (Ref. Dordević, 2008), Na3Co2SbO6 (Ref. Songvilay et al., 2020), and Na2Co2TeO6 (Ref. Xiao et al., 2019)222The Na2Co2TeO6 structure contains disorder in the Na position, in which each Na position has occupancy 2/3. To perform calculations, we artificially increased the occupancy to 1, which corresponds to Na3Co2TeO6. It is expected this change in the filling should have minimal impact on the computed hoppings., as well as triangular lattice compounds CoCl2, CoBr2 and CoI2 according to crystal structures from Ref. Wilkinson et al., 1959; Wyckoff and Wyckoff, 1963. These edge-sharing Co(II) compounds span a wide range Co-Co distances, e.g. from 2.9\sim 2.9 Å in BaCo2(AsO4)2Dordević (2008) to 3.9\sim 3.9 Å in CoI2Wyckoff and Wyckoff (1963). For each case, we employed fully relativistic density functional theory calculations performed with FPLOKoepernik and Eschrig (1999); Opahle et al. (1999) at the GGA (PBEPerdew et al. (1996)) level (as above). Hoppings in the downfolded dd-only basis were estimated by projection onto dd-orbitals, for which the cubic projection coordinates were defined to be orthogonal but minimize the difference with the corresponding Co-X (X = O, Cl, Br, I) bond vectors in the (distorted) octahedra.

Computed hoppings for the real materials are shown in Fig. 5(b). In general, t1t_{1}, t4t_{4} and t5t_{5} are small, and are thus omitted from the figure. From these results, it is evident that the physical region corresponds to large direct hopping t3t_{3} and subdominant ligand-mediated hopping t6t_{6}. Similarly, the ligand-mediated t2t_{2} is significantly suppressed, such that |t2||t1|,|t4|,|t5|0.05|t_{2}|\sim|t_{1}|,|t_{4}|,|t_{5}|\lesssim 0.05 eV. We find that this applies equally to Co oxides and halides. A similar situationWellm et al. (2021) was recently proposed for Na2BaCo(PO4)2. On this basis, we conclude: for edge-sharing materials with Co-Co bond lengths \sim 3.0 Å, direct hopping almost certainly dominates. This differs from the previous theoretical works predicting large Kitaev couplingsLiu and Khaliullin (2018); Liu (2021); Liu et al. (2020); Sano et al. (2018), which considered ligand-mediated hopping t2t_{2} and t6t_{6} to be the largest. This discrepancy calls for a reexamination of the magnetic couplings.

III Magnetic Couplings

III.1 General Form

For ideal edge-sharing bonds with C2vC_{2v} symmetry, the magnetic couplings may be written in the familiar formRau et al. (2014):

ij=\displaystyle\mathcal{H}_{ij}= J𝐒i𝐒j+KSiγSjγ+Γ(SiαSjβ+SiβSjα)\displaystyle\ J\ \mathbf{S}_{i}\cdot\mathbf{S}_{j}+K\ S_{i}^{\gamma}S_{j}^{\gamma}+\Gamma\left(S_{i}^{\alpha}S_{j}^{\beta}+S_{i}^{\beta}S_{j}^{\alpha}\right)
+Γ(SiαSjγ+SiγSjα+SiβSjγ+SiγSjβ)\displaystyle+\Gamma^{\prime}\left(S_{i}^{\alpha}S_{j}^{\gamma}+S_{i}^{\gamma}S_{j}^{\alpha}+S_{i}^{\beta}S_{j}^{\gamma}+S_{i}^{\gamma}S_{j}^{\beta}\right) (32)

where (α,β,γ)=(x,y,z)(\alpha,\beta,\gamma)=(x,y,z) for the Z-bonds, (y,z,x)(y,z,x) for the X-bonds, and (z,x,y)(z,x,y) for the Y-bonds, in terms of the global xyzxyz coordinates. This definition turns out to be convenient for the case of trigonal splitting Δ2=0\Delta_{2}=0. For finite Δ2\Delta_{2}, following Ref. Lines, 1963, the alterations to the nature of the local moments are expected to induce significant uniaxial anisotropy along the trigonal axis. This axial anisotropy is more apparent in alternative local XXZ coordinates shown in Fig. 3. In particular, for each bond we define local coordinates: e^1\hat{e}_{1} is parallel to the bond and e^3=(x^+y^+z^)/3\hat{e}_{3}=(\hat{x}+\hat{y}+\hat{z})/\sqrt{3} is along the global trigonal axis. Thus, the couplings may be written Ross et al. (2011); Maksimov et al. (2019):

ij=\displaystyle\mathcal{H}_{ij}= Jxy(Si1Sj1+Si2Sj2)+JzSi3Sj3\displaystyle\ J_{xy}\left(S_{i}^{1}S_{j}^{1}+S_{i}^{2}S_{j}^{2}\right)+J_{z}S_{i}^{3}S_{j}^{3} (33)
+2J±±(Si1Sj1Si2Sj2)+Jz±(Si3Sj2+Si2Sj3)\displaystyle\ +2J_{\pm\pm}\left(S_{i}^{1}S_{j}^{1}-S_{i}^{2}S_{j}^{2}\right)+J_{z\pm}\left(S_{i}^{3}S_{j}^{2}+S_{i}^{2}S_{j}^{3}\right)

where the superscript numbers refer to the local directions. The two parameterizations may be relatedMaksimov et al. (2019) via:

Jxy=\displaystyle J_{xy}= J+13(KΓ2Γ)\displaystyle\ J+\frac{1}{3}\left(K-\Gamma-2\Gamma^{\prime}\right) (34)
Jz=\displaystyle J_{z}= J+13(K+2Γ+4Γ)\displaystyle\ J+\frac{1}{3}\left(K+2\Gamma+4\Gamma^{\prime}\right) (35)
J±±=\displaystyle J_{\pm\pm}= 16(K+2Γ2Γ)\displaystyle\ -\frac{1}{6}\left(K+2\Gamma-2\Gamma^{\prime}\right) (36)
Jz±=\displaystyle J_{z\pm}= 23(KΓ+Γ)\displaystyle\ -\frac{\sqrt{2}}{3}\left(K-\Gamma+\Gamma^{\prime}\right) (37)

A similar parameterization was also employed in Ref. Liu, 2021; Liu et al., 2020. Both parameterizations are employed below, where appropriate.

III.2 Method and Downfolding Schemes

Refer to caption
Figure 6: Example hopping paths for holes contributing to magnetic exchange. Terms resulting from (a) and (b) are captured in the downfolded dd-only scheme. Terms resulting from (c) are not captured, and must be estimated from perturbative expressions (see text).

In previous works, the magnetic couplings have been discussed in terms of perturbation theory in both the full d+pd+p basis, and the downfolded dd-only basis. It is useful to discuss how terms in each scheme are related.

As an example of this downfolding, we may consider the ligand-mediated exchange process shown in Fig. 6(a), in which a hole hops from a metal dd-orbital to a ligand pp-orbital, and subsequently to a dd-orbital on an adjacent metal site, thus yielding an excited (d8,d6)(d^{8},d^{6}) configuration. From this excited configuration, the reverse process returns the system to the ground (d7,d7)(d^{7},d^{7}) configuration. Such processes contribute terms of order 𝒪[(tpd2/Δpd)2/Udd]\mathcal{O}[\ (t_{pd}^{2}/\Delta_{pd})^{2}/U_{dd}\ ] to the magnetic exchange, where Δpd\Delta_{pd} is the charge transfer energy, and UddU_{dd} is the Coulomb repulsion between dd-electrons. The same process can also be captured within a dd-only picture, by identifying teff=tpd2/Δpdt_{\rm eff}=t_{pd}^{2}/\Delta_{pd} as an effective ligand-mediated ddd-d hopping, as discussed above. In this case, the magnetic couplings arise at order 𝒪[teff2/Udd]\mathcal{O}[\ t_{\rm eff}^{2}/U_{dd}\ ]. Such terms are naturally captured in the downfolded dd-only scheme.

We next consider the “cyclic” exchange process depicted in Fig. 6(b). In this case, a hole first hops from one metal to a ligand, and then a second hole hops from the second metal to another ligand. The system is returned to the ground (d7,d7)(d^{7},d^{7}) configuration by two additional hops of the holes from the ligands to opposite metal sites from where they started. Such processes contribute to the magnetic exchange at order 𝒪[(tpd2/Δpd)2/(2Δpd)]\mathcal{O}[\ (t_{pd}^{2}/\Delta_{pd})^{2}/(2\Delta_{pd})\ ]. In the downfolded dd-only basis, these processes are not distinguished from those in Fig. 6(a). Rather, they reflect the fact that the Coulomb repulsion within the dd-orbital Wannier functions is renormalized to U~1Udd1+(2Δpd)1\tilde{U}^{-1}\sim U_{dd}^{-1}+(2\Delta_{pd})^{-1} via hybridization of the dd- and pp-orbitals. Thus both the “antiferromagnetic” superexchange and cyclic exchange processes are captured, in principle, by exchange contributions at order 𝒪[teff2/U~]\mathcal{O}[\ t_{\rm eff}^{2}/\tilde{U}\ ], assuming the hoppings and Coulomb repulsion terms are suitably renormalized.

Finally, we consider the process depicted in Fig. 6(c). In this case, two holes hop from their parent metal sites to the same ligand, but occupy different orbitals. They interact with eachother via Hund’s coupling at the ligand site, and then hop back to their parent metal sites. Such processes constitute the usual Goodenough-Kanamori ferromagnetic superexchange. When downfolded into the dd-only basis, these processes reflect the fact that the tails of the Wannier orbitals extend onto the ligands, which results in an enhancement of the effective intersite Hund’s coupling. Such terms are not explicitly included in our dd-only model Eq’n (1). We therefore estimate the effects of the additional contributions using previously derived perturbative expressions. From Ref. Liu and Khaliullin, 2018, there is a correction to both JJ and KK given approximately by:

δJ\displaystyle\delta J\approx γΔpd2(52(tpdσ)4+32(tpdσ)2(tpdπ)2+(tpdπ)4)\displaystyle\ -\frac{\gamma}{\Delta_{pd}^{2}}\left(\frac{5}{2}(t_{pd}^{\sigma})^{4}+\frac{3}{2}(t_{pd}^{\sigma})^{2}(t_{pd}^{\pi})^{2}+(t_{pd}^{\pi})^{4}\right) (38)
δK\displaystyle\delta K\approx γΔpd2(12(tpdσ)2(tpdπ)212(tpdπ)4)\displaystyle\ -\frac{\gamma}{\Delta_{pd}^{2}}\left(\frac{1}{2}(t_{pd}^{\sigma})^{2}(t_{pd}^{\pi})^{2}-\frac{1}{2}(t_{pd}^{\pi})^{4}\right) (39)
γ=\displaystyle\gamma= 40JHp81(Δpd+Up/2)2\displaystyle\ \frac{40J_{H}^{p}}{81(\Delta_{pd}+U_{p}/2)^{2}} (40)

where JHpJ_{H}^{p} is the Hund’s coupling at the ligand, UpU_{p} is the excess Coulomb repulsion at the ligand ion, Δpd4.5\Delta_{pd}\approx 4.5 eV is the charge-transfer gap. We take the same approximations as Ref. Liu and Khaliullin, 2018 (Up=0.7Ut2g,JHp=0.3UpU_{p}=0.7\ U_{t2g},\ J_{H}^{p}=0.3\ U_{p}), and consider tpdσ1t_{pd}^{\sigma}\approx 1 eV, tpdπ0.5t_{pd}^{\pi}\approx-0.5 eV, according to Fig. 5(a). From this, we estimate the correction to the Kitaev coupling to be negligible δK0.1\delta K\sim 0.1 meV, while the corrections to the Heisenberg coupling may be typically in the range δJ2\delta J\sim-2 to 6-6 meV. The reason for the discrepancy in magnitudes between δJ\delta J and δK\delta K is that the dominant contribution to δJ\delta J comes from hopping between the ege_{g} and pp orbitals (depicted in Fig. 6(c)), as parameterized by the large hopping tpdσt_{pd}^{\sigma}. This process does not contribute to KK. The small correction δK\delta K is henceforth neglected.

In the following, we compute all exchange contributions that are captured within the downfolded dd-only basis using exact diagonalization of the model Eq’n (1) for two Co sites. The couplings are extracted by projecting onto the ideal j1/2j_{1/2} doublets defined in eq’n (24, 25). This procedure is described in more detail in Ref. Winter et al., 2016, and is guaranteed to yield couplings that converge to the results of perturbation theory with respect to hop\mathcal{H}_{\rm hop} for small values of tn/Ut_{n}/U (with all other contributions to the Hamiltonian, which appear in i\mathcal{H}_{i}, being treated exactly). The reason for adopting this method is that analytical perturbation theory is generally challenging for finite Δ2\Delta_{2} and the full set of allowed hoppings t16t_{1-6}. The expressions below are intended to be used with hopping integrals extracted from DFT, which are assumed to be suitably renormalized. The correction δJ\delta J must be added to the computed couplings to estimate the omitted ligand exchange processes depicted in Fig. 6(c).

Refer to caption
Figure 7: (a-c) Evolution of the magnetic couplings in the dd-only model for ideal edge-sharing bond with no trigonal distortion (Δ2=0\Delta_{2}=0) and Δ1=1.1\Delta_{1}=1.1 eV, U=3.25U=3.25 eV, JH=0.7J_{H}=0.7 eV, t1=|t3|/4,t4=t5=|t3|/4,t6=+0.1t_{1}=|t_{3}|/4,t_{4}=t_{5}=-|t_{3}|/4,t_{6}=+0.1 eV. The ferromagnetic correction δJ\delta J due to ligand exchange processes is not included (see text). (d) Computed couplings along the path indicated by a dashed line in (a-c), interpolating between the direct and ligand-assisted hopping regimes. A correction δJ=2\delta J=-2 meV has been added. The small correction δK\delta K has been neglected.

III.3 General Hopping Dependence

In this section, we first consider the magnetic couplings for the case Δ2=0\Delta_{2}=0, for which Γ=0\Gamma^{\prime}=0 strictly. Up to second order in ddd-d hopping (and fourth order in dpd-p hopping), the couplings may be written:

J=\displaystyle J= 𝐭𝕄J𝐭T+δJ\displaystyle\ \mathbf{t}\cdot\mathbb{M}_{J}\cdot\mathbf{t}^{T}+\delta J (41)
K=\displaystyle K= 𝐭𝕄K𝐭T\displaystyle\ \mathbf{t}\cdot\mathbb{M}_{K}\cdot\mathbf{t}^{T} (42)
Γ=\displaystyle\Gamma= 𝐭𝕄Γ𝐭T\displaystyle\ \mathbf{t}\cdot\mathbb{M}_{\Gamma}\cdot\mathbf{t}^{T} (43)

where:

𝐭=(t1t2t3t4t5t6)\displaystyle\mathbf{t}=\left(t_{1}\ t_{2}\ t_{3}\ t_{4}\ t_{5}\ t_{6}\right) (44)

and 𝕄\mathbb{M} parameterizes the contributions that are captured by (downfolded) ddd-d hopping. The matrices are a function of Fn,λ,ΔnF_{n},\lambda,\Delta_{n}. To estimate 𝕄\mathbb{M}, we computed the magnetic couplings for a grid of hoppings 0.05<tn<+0.05-0.05<t_{n}<+0.05 eV using exact diagonalization of the full dd-orbital model U+CFS+SOC+hop\mathcal{H}_{U}+\mathcal{H}_{\rm CFS}+\mathcal{H}_{\rm SOC}+\mathcal{H}_{\rm hop} for two metal sites, and fit the resulting couplings. For this purpose, we use Δ1=1.1\Delta_{1}=1.1 eV and λ=0.06\lambda=0.06 eV, which is consistent with estimates from DFT in the previous sections and Ut2g=3.25U_{t2g}=3.25 eV, and Jt2g=0.7J_{t2g}=0.7 eV, following Ref. Das et al., 2021. This provides an estimate of the couplings in the perturbative regime:

𝕄J=\displaystyle\mathbb{M}_{J}= (t1t2t3t4t5t6t15501431020t207600077t30033220t400026000t500002590t600000165)\displaystyle\ \left(\begin{array}[]{c|cccccc}&t_{1}&t_{2}&t_{3}&t_{4}&t_{5}&t_{6}\\ \hline\cr t_{1}&-55&0&143&10&2&0\\ t_{2}&0&-76&0&0&0&-77\\ t_{3}&0&0&-33&2&2&0\\ t_{4}&0&0&0&260&0&0\\ t_{5}&0&0&0&0&259&0\\ t_{6}&0&0&0&0&0&165\end{array}\right) (52)
𝕄K=\displaystyle\mathbb{M}_{K}= (t1t2t3t4t5t6t112801199350t2010800086t3008250t4000400t5000010t600000147)\displaystyle\ \left(\begin{array}[]{c|cccccc}&t_{1}&t_{2}&t_{3}&t_{4}&t_{5}&t_{6}\\ \hline\cr t_{1}&128&0&-119&-9&35&0\\ t_{2}&0&-108&0&0&0&86\\ t_{3}&0&0&-8&-2&5&0\\ t_{4}&0&0&0&-4&0&0\\ t_{5}&0&0&0&0&1&0\\ t_{6}&0&0&0&0&0&-147\end{array}\right) (60)
𝕄Γ=\displaystyle\mathbb{M}_{\Gamma}= (t1t2t3t4t5t6t103400049t200116210t30000031t40000067t5000000t6000000)\displaystyle\ \left(\begin{array}[]{c|cccccc}&t_{1}&t_{2}&t_{3}&t_{4}&t_{5}&t_{6}\\ \hline\cr t_{1}&0&-34&0&0&0&49\\ t_{2}&0&0&-116&-2&1&0\\ t_{3}&0&0&0&0&0&-31\\ t_{4}&0&0&0&0&0&-67\\ t_{5}&0&0&0&0&0&0\\ t_{6}&0&0&0&0&0&0\end{array}\right) (68)

in units of 10310^{-3}/eV. Recall, for real materials we generally anticipate |t3|>|t6|>|t1||t2||t4||t5||t_{3}|>|t_{6}|>|t_{1}|\sim|t_{2}|\sim|t_{4}|\sim|t_{5}|. Furthermore, t1>0t_{1}>0, t3<0t_{3}<0, t4<0t_{4}<0, t6>0t_{6}>0 (see section II.3). These results highlight several key aspects:

Heisenberg JJ: For JJ, there are various contributions of different sign. Those arising from hopping between t2gt_{2g} orbitals (t1,t2,t3t_{1},t_{2},t_{3}) are exclusively ferromagnetic. Processes involving hopping between ege_{g} orbitals (t4,t5t_{4},t_{5}) are exclusively antiferromagnetic. The terms related to ege_{g}-t2gt_{2g} hopping tend to be antiferromagnetic t62\propto t_{6}^{2} and t2t6t_{2}t_{6} given that ab-initio tends to yield t2<0t_{2}<0 and t6>0t_{6}>0.

Kitaev KK: There are also different contributions to KK of varying sign. Hopping between ege_{g} orbitals (t4,t5t_{4},t_{5}) makes little contribution to the anisotropic couplings overall. The sign of the contribution from t2gt_{2g}-t2gt_{2g} hopping depends on the balance of transfer integrals: terms t22\propto t_{2}^{2} are ferromagnetic, while terms t12\propto t_{1}^{2} and t1t3t_{1}t_{3} are antiferromagnetic. Contributions related to t2gt_{2g}-ege_{g} hopping may take both signs: terms t62\propto t_{6}^{2} are ferromagnetic, while terms t2t6\propto t_{2}t_{6} depend on the sign of t2t_{2}.

Off-diagonal Γ\Gamma: For the off-diagonal couplings, the primary contribution arises at order t2t3t_{2}t_{3}, and as a result sign(Γ)\text{sign}(\Gamma) is typically the same as sign(t2t3)-\text{sign}(t_{2}t_{3}). There are no contributions that are diagonal with respect to the hopping pairs. A similar result appears in Ref. Liu and Khaliullin, 2018.

The appearance of hopping combinations such as t2t6t_{2}t_{6}, which do not conserve t2gt_{2g} and ege_{g} occupancy, may appear surprising at first. If the ground state doublets have approximate configuration (t2g)5(eg)2(t_{2g})^{5}(e_{g})^{2}, one might expect terms mixing the occupancy to be forbidden at low orders, because they do not connect ground states. However, in reality, neither occupancy is preserved by either spin-orbit coupling or the full Coulomb terms, which are treated exactly (not perturbatively) in this approach.

For general parameters, we expect all three couplings to be finite. The computed hopping-dependence of K,J,ΓK,J,\Gamma are shown in Fig. 7 for the choice t1=|t3|/4,t4=t5=|t3|/4,t6=+0.1t_{1}=|t_{3}|/4,t_{4}=t_{5}=-|t_{3}|/4,t_{6}=+0.1 eV. These ratios of hoppings are compatible with the ab-initio estimates in section II.3. With this choice, we interpolate between the limits of dominant ligand-mediated vs. direct hopping.

In the hypothetical regime of pure ligand-mediated hopping (finite t2,t6>0t_{2},t_{6}>0 ), we find Γ=0\Gamma=0. If we ignore the ferromagnetic correction δJ\delta J, then J>0J>0 is antiferromagnetic and K<0K<0 is ferromagnetic. These findings verify expectations from perturbation theory for this limitLiu and Khaliullin (2018); Sano et al. (2018). The Kitaev coupling is the largest, with values |K/J|110|K/J|\sim 1-10 depending on the precise balance of hoppings. If we consider also the ferromagnetic correction δJ2\delta J\sim-2 to 6-6 meV discussed in the previous section, the overall sign of JJ should reverse, and the magnitude may be suppressed, such that dominant Kitaev coupling is possible with some tuning.

By contrast, for the physically relevant region of large t3t_{3} and finite values of all hoppings, we anticipate that ferromagnetic J<0J<0 is the dominant coupling, particularly due to contributions t1t3\propto t_{1}t_{3} and the correction δJ\delta J. In fact, δJ\delta J (which is just the regular ferromagnetic super-exchange for 90 bondsGoodenough (1963)) is the largest contribution. All possible combinations of signs of KK and Γ\Gamma are possible depending on the balance of hoppings, but their magnitude is suppressed relative to JJ. For very short Co-Co bond lengths, where the direct hopping contribution to t2t_{2} is the largest (t2<0t_{2}<0), the tendency is for K,Γ<0K,\Gamma<0. For longer bond lengths, where ligand-mediated contributions are the largest (t2>0t_{2}>0), then K,Γ>0K,\Gamma>0.

III.4 Effect of Trigonal Distortion

We next consider the effects of trigonal distortion. Given the relatively small value of the atomic SOC constant λCo60\lambda_{\rm Co}\approx 60 meV, small distortions may be relevant for Co(II) compounds. This make clarifying the size and sign of Δ2\Delta_{2} important for modelling such materials.

In general, for Δ2<0\Delta_{2}<0, as the moments become more axial, components of the exchange along the e^3\hat{e}_{3} direction are expected to be enhanced compared to the e^1\hat{e}_{1} and e^2\hat{e}_{2} directionsLines (1963). As a result JxyJ_{xy} and J±±J_{\pm\pm} should be suppressed relative to JzJ_{z} and Jz±J_{z\pm}. Trigonal elongation Δ2>0\Delta_{2}>0 should have the opposite effect. As the moments become more planar, JxyJ_{xy} and J±±J_{\pm\pm} should be relatively enhanced.

Refer to caption
Figure 8: Anisotropic ferromagnetic corrections as a function of trigonal distortion, for δJ0=2\delta J_{0}=-2 meV. (a) Local XXZ scheme. (b) Global J,K,Γ,ΓJ,K,\Gamma,\Gamma^{\prime} scheme.
Refer to caption
Figure 9: Magnetic couplings for ideal edge-sharing bond with finite trigonal distortion. The parameters are otherwise the same as Fig. 7. (a-f): Δ2/λ=0.5\Delta_{2}/\lambda=-0.5, (g-l): Δ2/λ=+0.5\Delta_{2}/\lambda=+0.5. The ferromagnetic corrections δJ,δΓ,δΓ\delta J,\delta\Gamma,\delta\Gamma^{\prime} due to ligand exchange processes are not included (see text). (e,k): Couplings along the path interpolating between direct and ligand-assisted hoppings depicted in (a,g) in the J,K,Γ,ΓJ,K,\Gamma,\Gamma^{\prime} scheme. (f,l): Couplings along the path in the Jz,Jxy,Jz±,J±±J_{z},J_{xy},J_{z\pm},J_{\pm\pm} scheme. For (e,f,k,l), solid lines indicate the results of dd-dd exchange only, and dashed lines indicate corrected values J+δJJ+\delta J, according to δJ0=2\delta J_{0}=-2 meV.

Following Ref. Lines, 1963; Liu et al., 2020, the ferromagnetic corrections to JJ resulting from ligand exchange processes are rendered anisotropic, with:

δJxy\displaystyle\delta J_{xy}\approx uxy2δJ0\displaystyle\ u_{xy}^{2}\ \delta J_{0} (69)
δJz\displaystyle\delta J_{z}\approx uz2δJ0\displaystyle\ u_{z}^{2}\ \delta J_{0} (70)
uxy=\displaystyle u_{xy}= 35(23c1c3+2c22)\displaystyle\ \frac{3}{5}\left(2\sqrt{3}c_{1}c_{3}+2c_{2}^{2}\right) (71)
uz=\displaystyle u_{z}= 35(1+2(c12c32))\displaystyle\ \frac{3}{5}\left(1+2(c_{1}^{2}-c_{3}^{2})\right) (72)

with cnc_{n} given in Fig. 4(b) and δJ0\delta J_{0} defined according to eq’n (38). Thus, in the J,K,Γ,ΓJ,K,\Gamma,\Gamma^{\prime} parameterisation, these corrections correspond to:

δJ=\displaystyle\delta J= 13(uz2+2uxy2)δJ0\displaystyle\ \frac{1}{3}\left(u_{z}^{2}+2u_{xy}^{2}\right)\delta J_{0} (73)
δΓ=\displaystyle\delta\Gamma= δΓ=13(uz2uxy2)δJ0\displaystyle\ \delta\Gamma^{\prime}=\frac{1}{3}\left(u_{z}^{2}-u_{xy}^{2}\right)\delta J_{0} (74)

Here, we have continued to neglect the small corrections δK0\delta K_{0}. As with the undistorted case, the corrections to the anisotropic couplings J±±J_{\pm\pm} and Jz±J_{z\pm} are predicted to be small. They have also been neglected in the following. The evolution of the ferromagnetic corrections with distortion are shown in Fig. 8. For Δ2<0\Delta_{2}<0, the ratio δJz/δJxy>1\delta J_{z}/\delta J_{xy}>1 is, in principle, unbounded (and should increase continuously with trigonal distortion). For Δ2>0\Delta_{2}>0 the degree of anisotropy is restricted, because the distortion-induced effects are bounded 1/4Jz/Jxy<11/4\lesssim J_{z}/J_{xy}<1. The lower bound is reached for large Δ2\Delta_{2}, where the orbital moment is quenched, thus restoring the full degeneracy of the S=3/2S=3/2 states. As discussed above, a low-energy model including only the lowest doublet would no longer be sufficient, so this limit does not represent a physically sensible model. In terms of global cubic coordinates [Fig. 8(b)], the trigonal distortion primarily introduces off-diagonal couplings, where Δ2<0\Delta_{2}<0 tends to be associated with δΓ,δΓ<0\delta\Gamma,\delta\Gamma^{\prime}<0, and vice versa.

To explore the exchange contributions from (downfolded) dd-dd hopping, we recomputed the couplings using exact diagonalization with significant distortion Δ2/λ=±0.5\Delta_{2}/\lambda=\pm 0.5 to emphasize the effects. Results are shown in Fig. 9 for the choice t1=|t3|/4,t4=t5=|t3|/4,t6=+0.1t_{1}=|t_{3}|/4,t_{4}=t_{5}=-|t_{3}|/4,t_{6}=+0.1 eV, which is compatible with the ab-initio estimates. In Fig. 9(e,f) and (k,l), we also show the effect of corrections δJ\delta J. The results are as follow:

Trigonal compression: For Δ2<0\Delta_{2}<0, as shown in Fig. 9(a-e), we find all four of the couplings J,K,Γ,ΓJ,K,\Gamma,\Gamma^{\prime} may be of similar magnitude. This is particularly true in the region of large ligand-assisted hopping. For the physically relevant region of large direct hopping (t3t2t_{3}\gg t_{2}), we find that KK is still relatively suppressed (same as for Δ2=0\Delta_{2}=0), but large Γ,Γ\Gamma,\Gamma^{\prime} are induced, with sign(Γ,Γ)\text{sign}(\Gamma,\Gamma^{\prime}) typically being the same as sign(J)\text{sign}(J). These results are more easily interpreted in the alternative XXZ parameterization shown in Fig. 9(f). In particular, as the local moments become more axial with larger trigonal distortion, the coupling becomes dominated by a ferromagnetic Ising exchange JzJ_{z}. Overall, the estimated ferromagnetic correction δJz\delta J_{z} is quite large compared to the regular dd-dd contributions. For the physically relevant region, we anticipate Jxy=2J_{xy}=-2 to 0 meV, Jz=3J_{z}=-3 to 10-10 meV, J±±=0.5J_{\pm\pm}=-0.5 to +0.5+0.5 meV, and Jz±=0.5J_{z\pm}=-0.5 to +1.5+1.5 meV for significant trigonal distortion of Δ=λ/2\Delta=-\lambda/2. . As a result, we expect such materials to be described mostly by Ising couplings with a common axis for every bond.

Trigonal elongation: For Δ2>0\Delta_{2}>0, we find that KK is less suppressed. The distortions induce off-diagonal couplings following roughly sign(Γ,Γ)\text{sign}(\Gamma,\Gamma^{\prime}) typically the same as sign(J)-\text{sign}(J). In the XXZ parameterization, this corresponds to an enhancement of JxyJ_{xy}. In the hypothetical ligand-assisted hopping region, we find that JzJ_{z} may be almost completely suppressed due to different values of the ferromagnetic shifts δJz\delta J_{z} and δJxy\delta J_{xy}. While JxyJ_{xy} appears to be the largest coupling in this limit, the bond-dependent couplings Jz±J_{z\pm} and J±±J_{\pm\pm} may also remain significant. For the physically relevant region, we find that JxyJ_{xy} is typically the dominant coupling, with Jxy/Jz4J_{xy}/J_{z}\sim 4, which is the hypothetical limit. Overall, we anticipate Jxy=2J_{xy}=-2 to 10-10 meV, Jz=0.5J_{z}=-0.5 to 4-4 meV, J±±=2J_{\pm\pm}=-2 to +1+1 meV, and Jz±=0J_{z\pm}=0 to +1+1 meV for significant trigonal distortion of Δ=+λ/2\Delta=+\lambda/2.

III.5 Longer Range Couplings

While we have discussed above that t2gt_{2g}-ligand hybridization should generally be small in 3d3d transition metal oxides (as reflected by small tpdπt_{pd}^{\pi}), the ege_{g}-ligand hybridization may still play a significant role in the magnetic exchange due to the large tpdσt_{pd}^{\sigma}. This fact is what leads to significant ferromagnetic corrections δJ\delta J for nearest neighbor bonds. For longer range bonds, such as third neighbors in edge-sharing honeycomb lattices, it gives rise to a large hopping between dx2y2d_{x^{2}-y^{2}} orbitals shown in Fig. 10 at order (tpdσ)2tppσ/Δpd20.05(t_{pd}^{\sigma})^{2}t_{pp}^{\sigma}/\Delta_{pd}^{2}\sim 0.05 to 0.1 eV. This is equivalent to a 3rd neighbor t5t_{5}, which allows the associated coupling to be readily estimated from the matrices 𝕄\mathbb{M}. In particular, we estimate (for Δ2=0\Delta_{2}=0):

J3+0.5 to+2.5 meV\displaystyle J_{3}\approx+0.5\text{ to}+2.5\text{ meV} (75)
K3Γ30\displaystyle K_{3}\approx\Gamma_{3}\approx 0 (76)

This is the only major third neighbor hopping pathway, so there are no additional terms to compete, and a relatively large antiferromagnetic J3J_{3} should be expected for all honeycomb materials with partially occupied ege_{g} orbitals. For finite Δ2\Delta_{2}, the interactions are rendered anisotropic, with XXZ anisotropy reflecting the composition of the local moments. These estimates may be confirmed by considering the triangular lattice compound Ba3CoSb2O9, for which the closest Co2+ neighbors have the same geometry as depicted in Fig. 10. In this case, it is well establishedMa et al. (2016); Ito et al. (2017); Ghioldi et al. (2018); Kamiya et al. (2018) that the interactions are described by Jxy=+1.7J_{xy}=+1.7 meV, Jz=+1.5J_{z}=+1.5 meV. Third neighbor interactions of similar magnitude should be expected for all honeycomb materials BCAO, BCPO, NCSO, and NCTO.

Refer to caption
Figure 10: 3rd neighbor hopping relevant to J3J_{3}.

IV Discussion

In this work, we have considered the magnetic couplings in edge-sharing d7d^{7} compounds. On this basis, we make several observations:

(1) All of the edge-sharing Co(II) oxides considered in this work appear to fall outside the regime of primary focus in previous theoretical studiesLiu and Khaliullin (2018); Liu (2021); Liu et al. (2020); Sano et al. (2018). In particular, direct hopping likely dominates over ligand-assisted hopping (t3t2t_{3}\gg t_{2}). In the realistic regime, we find that KK is generally suppressed compared to JJ, which calls into question models with dominant KK proposed for these materials.

(2) The presence of the ege_{g} spins also opens additional exchange pathways. The ege_{g} orbitals mix relatively strongly with the ligand pp-orbitals in contrast to the t2gt_{2g} orbitals. This enhances the importance of certain ligand exchange processes, which are responsible for ferromagnetic couplings in materials with 90 bond angles in the Goodenough-Kanamori description Goodenough (1963). Analogous contributions are typically an order of magnitude smaller in low-spin d5d^{5} compounds, such as A2IrO3 and α\alpha-RuCl3, where the ege_{g} orbitals are empty. The ege_{g} spins also participate in longer range antiferromagnetic couplings, corresponding to third neighbors for honeycomb lattice materials. It may be commonly expected that first and third neighbor interactions have similar magnitude in materials such as BCAO, BCPO, NCTO, and NCSO. Thus, while it was initially hoped that the smaller spatial extent of the 3d3d orbitals would suppress longer range coupling relative to those found in 4d54d^{5} and 5d55d^{5} compounds, the presence of ege_{g} spins likely enhances long-range interactions.

(3) Compared to heavy d5d^{5} Kitaev materials such as iridates A2IrO3 and α\alpha-RuCl3, the weak spin-orbit coupling of Co increases the relative importance of local distortions. There are also many separate contributions to the magnetic couplings whose balance depends sensitively on local parameters such as JH,UJ_{H},U, and Δ1\Delta_{1}. This makes anticipating the magnetic Hamiltonian somewhat challenging. For Co(II) oxides, fortuitous fine-tuning may result in a different balance of couplings, but we anticipate that ferromagnetic JJ (or equivalently Jz,JxyJ_{z},J_{xy}) is likely always the largest coupling. The signs and magnitudes of the other couplings K,Γ,ΓK,\Gamma,\Gamma^{\prime} are influenced by the crystal field splitting and specific details of the hoppings. We find regions with all possible signs and relative magnitudes.

(4) Real materials with small trigonal distortions (discussed in Section III.3) are likely described by |K/J|0.2|K/J|\sim 0.2 to 0.5, and KΓK\approx\Gamma; specifically: J8J\sim-8 to 2-2 meV, K2K\sim-2 to +2+2 meV, and Γ1\Gamma\sim-1 to +3+3 meV. These magnitudes are compatible with the overall scale of interactions reported in the experimental literatureRegnault et al. (1977); Nair et al. (2018); Regnault et al. (2018); Fava et al. (2020). Γ\Gamma^{\prime} is predicted to be small unless there are significant departures from ideal symmetry of the bonds (i.e. Δ2=0\Delta_{2}=0, C2hC_{2h} bond symmetry, 90 M-L-M bond angles). It is not clear that a uniquely dominant KK is possible even for weak trigonal distortions.

(5) For systems with significant crystal field distortions (section III.4), our findings are compatible with the historical description of Co(II) magnetic couplings in terms of XXZ models by M. E. Lines (Ref. Lines, 1963). This is true particularly because these considerations apply to the ferromagnetic ligand exchange processes, which we estimate are at least as important as processes involving dd-dd hopping. In this case, the considerations discussed in Ref. Liu et al., 2020 become equivalent to the classic results of Lines. For trigonal crystal fields with Δ2<0\Delta_{2}<0 (corresponding to g||>gg_{||}>g_{\perp}), the Ising anisotropy induced by the crystal field may be very large (Jz/Jxy1J_{z}/J_{xy}\gg 1), such that the couplings are dominated by a ferromagnetic JzJ_{z} with a common Ising axis for every bond. For positive crystal field Δ2>0\Delta_{2}>0 (corresponding to g>g||g_{\perp}>g_{||}), XXZ anisotropy is more limited (Jxy/Jz4J_{xy}/J_{z}\leq 4), but may still be quite significant. Such materials are generally more desirable for realising strongly bond-dependent couplings, because the XXZ anisotropy is not as dominant.

(6) From the perspective of chemistry, it is unclear how to access the desirable ligand-assisted hopping regime where Kitaev coupling is largest. It is necessary to increase the metal-ligand hybridization relative to direct hopping between transition metal ions. This may typically be achieved by matching the electronegativity of the metals and ligands, such that Δpd\Delta_{pd} is small. As a general trend, electronegativity of transition metals increases for heavier atoms, while the opposite is true for pp-block ligand ions. The combination of heavy metals with heavy ligands results in the most covalent metal-ligand bonds. However, with increased covalency comes increased t2gt_{2g}-ege_{g} splitting, and heavier metals tend to have reduced Coulomb terms JHJ_{H}. As a result, the high-spin (t2g)5(eg)2(t_{2g})^{5}(e_{g})^{2} state is only typically achievable in Co(II) compounds. By contrast, Rh(II) and Ni(III) tend to adopt low-spin (t2g)6(eg)1(t_{2g})^{6}(e_{g})^{1} ground states. The most promising avenue then appears to be the combination of Co(II) with heavier ligands. With this in mind, we computed the hopping integrals for the triangular lattice compounds CoCl2, CoBr2 and CoI2. For these compounds, we still estimate |t3/t2|4|t_{3}/t_{2}|\sim 4. Thus, it is not clear that Kitaev-dominant exchange can be achieved without extreme fine tuning.

(7) Regarding NCSO and NCTO: Experimental estimatesKim et al. (2021); Liu et al. (2020); Liu (2021) for NCSO and NCTO suggest Δ2+4\Delta_{2}\sim+4 to +13+13 meV, which corresponds to a significant easy-plane XXZ anisotropy with δΓ,δΓ0.20.5\delta\Gamma,\delta\Gamma^{\prime}\sim 0.2-0.5 meV. Large departures from XXZ symmetry are also unlikely, given the small magnitude of the magnon gap (1\sim 1 meV) observed in experiment at both the Γ\Gamma-pointLin et al. (2021); Chen et al. (2021) and the ordering wavevectorSongvilay et al. (2020); Kim et al. (2021); Chen et al. (2021); Lin et al. (2021). It may further be remarked that the field-evolution of the ESR modesLin et al. (2021) follow expectations for moderate easy-plane XXZ anisotropy. Thus, we may anticipate J1<0J_{1}<0, and Γ,Γ>0\Gamma,\Gamma^{\prime}>0, with the sign of KK being dependent on the microscopic details.

It may be possible to place some constraints on the interactions on the basis of the ordered moment directions in the zigzag state. Related discussions appear in Ref. Sanders et al., 2021. For NCTO, it was initially reported that the ordered moments lie nearly in the honeycomb plane, oriented along the direction of the ferromagnetic chainsLefrançois et al. (2016). This orientation was later disputed, with a significant tilting of the ordered moments out of the plane being reported in Ref. Samarakoon et al., 2021. We may consider each case separately.

If we consider the zigzag domain with magnetic wavevector parallel to the Z-bond, then moments oriented strictly in-plane would point along e^2Z=(1,1,2)/6\hat{e}_{2}^{Z}=(1,1,-2)/\sqrt{6} in cubic coordinates. The magnetic state would be left invariant under a 180 rotation around (111) followed by time reversal; it is thus reasonable to assume the couplings possess the same symmetry. This orientation generally requiresChaloupka and Khaliullin (2016) K,Γ>0K,\Gamma>0, and places the constraint on the couplings Γ+δΓ=K+Γ+δΓ\Gamma+\delta\Gamma=K+\Gamma^{\prime}+\delta\Gamma^{\prime} (i.e. Jz±=0J_{z\pm}=0). Taken together, we suggest J1=3,J3=+2.5,K=Γ=+0.25,Γ=+0.5J_{1}=-3,J_{3}=+2.5,K=\Gamma^{\prime}=+0.25,\Gamma=+0.5 meV as an appropriate starting point for analysis. These considerations favor the model proposed in Ref. Lin et al., 2021, and correspond to Jxy=3.25,Jz=2.25,J±±=0.125,Jz±=0J_{xy}=-3.25,J_{z}=-2.25,J_{\pm\pm}=-0.125,J_{z\pm}=0 meV. The antiferromagnetic sign of KK is possible with large t3t_{3} and small t2>0t_{2}>0.

In contrast, if the moments are tilted out of plane, then fewer constraints may be placed on the couplings a-priori. However, such tilting likely requires K<0K<0. In this case, we suggest J1=3,J3=+2.5,K=0.25,Γ=+0.25,Γ=+0.5J_{1}=-3,J_{3}=+2.5,K=-0.25,\Gamma^{\prime}=+0.25,\Gamma=+0.5 meV as an appropriate starting point for analysis. The ferromagnetic sign of KK is possible with large t3t_{3} and small t2<0t_{2}<0.

(8) Regarding BCAO: The breadth of experimental data on BCAO, in terms of the progression of field-induced phasesHuyan et al. (2022) and inelastic neutron data, provide a number of clues towards the magnetic model. While we leave full elaboration for future studyMaksimov et al. (2022), some comments can be made. A recent reinvestigationRegnault et al. (2018) of the zero-field structure suggested it might better be described by a double stripe \uparrow\uparrow\downarrow\downarrow analogue of the zigzag antiferromagnet, with moments oriented nearly along the in-plane e^2Z\hat{e}_{2}^{Z} direction, as with NCTO. This orientation points to K,Γ>0K,\Gamma>0. The large discrepancy between in-plane critical fields (0.2, 0.5 T) and the out-of-plane critical field (4T)Zhang et al. (2021) suggests significant anisotropy. Indeed, the gg-tensor appears to satisfyRegnault et al. (2018) gz0.5gxyg_{z}\sim 0.5\ g_{xy}, and within an XXZ model, Jz0.4JxyJ_{z}\sim 0.4\ J_{xy}. These findings point to significant crystal field effects, with Δ20.2\Delta_{2}\sim 0.2 to 0.25λ0.25\ \lambda, i.e. Δ215\Delta_{2}\sim 15 meV, implying significant Γ\Gamma and Γ\Gamma^{\prime} in the global coordinate scheme. In the XXZ scheme, the nearly in-plane moments suggest small Jz±J_{z\pm}, while an apparently small anisotropy between in-plane field directionsRegnault et al. (2018) may place restrictions on J±±J_{\pm\pm}. In contrast, the authors of Ref. Zhang et al., 2021 have advocated for small JJ, large K<0K<0, and small average off-diagonal coupling Γ¯=(Γ+2Γ)/3\bar{\Gamma}=(\Gamma+2\Gamma^{\prime})/3 on the basis of THz spectroscopy experiments. It should be emphasized that these conditions are not mutually compatible: small Jz±J_{z\pm} and J±±J_{\pm\pm} implies small KK, and large anisotropy between JxyJ_{xy} and JzJ_{z} implies large Γ+2Γ>0\Gamma+2\Gamma^{\prime}>0. If we consider BCAO to be in the physical regime of hoppings, our findings contradict the Kitaev-dominant model. We propose a model similar to NCTO with in-plane moments: KK is small and likely antiferromagnetic, J<0J<0 is the dominant coupling, and Γ,Γ>0\Gamma,\Gamma^{\prime}>0 reflect a planar XY-anisotropy |Jxy|>|Jz||J_{xy}|>|J_{z}|. The anomalous aspects of the ground state are then understood as a competition between J1,J2J_{1},J_{2}, and J3J_{3}, as previously suggestedRegnault et al. (1977); Nair et al. (2018); Regnault et al. (2018). These suggestions are compatible with the recent ab-initio estimatesDas et al. (2021); Maksimov et al. (2022).

V Conclusions

In high-spin Co(II) 3d73d^{7}, the spin-orbital nature of the jeff=1/2j_{\rm eff}=1/2 ground state results in complex bond-dependent magnetic couplings in edge-sharing compounds. The theoretical description of these interactions is complicated by the existence of a wide number of different contributions, the balance of which is difficult to predict without detailed knowledge of the microscopic hopping, Coulomb, and crystal field parameters. Based on DFT studies of a range of Co oxides and halides, we find particular discrepancies with previous assumptions regarding these microscopic parameters. Specifically, direct metal-metal hopping is more prominent than ligand-assisted hopping, which results in significant suppression, and possible sign-reversal of the Kitaev interactions. Ferromagnetic exchange resulting from ege_{g} spins likely strongly dominates the nearest neighbor exchange. Off-diagonal couplings are likely enhanced by crystal field effects, resulting in significant XXZ anisotropy that is consistent with previous description of materials like the Ising ferromagnet CoNb2O6. Further neighbor exchange is also promoted by the presence of the ege_{g} spins. Taken together, Co(II) materials will continue to represent interesting platforms for highly complex anisotropic magnetism. However, the realisation of Kitaev magnetism specifically appears to be unrealistic from the microscopic perspective.

VI Acknowledgements

We acknowledge useful discussions with D. Smirnov, Y. Jiang, S. Streltsov, P. Maksimov, R. Valenti, and E. Gati. We also thank P. Dai for bringing the problem to our attention. This work was supported by a pilot grant from the Center for Functional Materials. Computations were performed on the Wake Forest University DEAC Cluster, a centrally managed resource with support provided in part by Wake Forest University.

References

  • Nussinov and Van Den Brink (2015) Z. Nussinov and J. Van Den Brink, Rev. Mod. Phys. 87, 1 (2015).
  • Kitaev (2006) A. Kitaev, Ann. Phys. 321, 2 (2006).
  • Hermanns et al. (2018) M. Hermanns, I. Kimchi,  and J. Knolle, Annu. Rev. Condens. Matter Phys. 9, 17 (2018).
  • Broholm et al. (2020) C. Broholm, R. Cava, S. Kivelson, D. Nocera, M. Norman,  and T. Senthil, Science 367, eaay0668 (2020).
  • Zhou et al. (2017) Y. Zhou, K. Kanoda,  and T.-K. Ng, Rev. Mod. Phys. 89, 025003 (2017).
  • Jackeli and Khaliullin (2009) G. Jackeli and G. Khaliullin, Phys. Rev. Lett. 102, 017205 (2009).
  • Winter et al. (2017a) S. M. Winter, A. A. Tsirlin, M. Daghofer, J. van den Brink, Y. Singh, P. Gegenwart,  and R. Valenti, J. Phys. Condens. Matter 29, 493002 (2017a).
  • Trebst (2017) S. Trebst, arXiv preprint arXiv:1701.07056  (2017).
  • Singh and Gegenwart (2010) Y. Singh and P. Gegenwart, Phys. Rev. B 82, 064412 (2010).
  • Plumb et al. (2014) K. Plumb, J. Clancy, L. Sandilands, V. V. Shankar, Y. Hu, K. Burch, H.-Y. Kee,  and Y.-J. Kim, Phys. Rev. B 90, 041112 (2014).
  • Banerjee et al. (2016) A. Banerjee, C. Bridges, J.-Q. Yan, A. Aczel, L. Li, M. Stone, G. Granroth, M. Lumsden, Y. Yiu, J. Knolle, S. Bhattacharjee, D. L. Kovrizhin, R. Moessner, D. A. Tennant, D. G. Mandrus,  and S. E. Nagler, Nat. Mater. 15, 733 (2016).
  • Banerjee et al. (2017) A. Banerjee, J. Yan, J. Knolle, C. A. Bridges, M. B. Stone, M. D. Lumsden, D. G. Mandrus, D. A. Tennant, R. Moessner,  and S. E. Nagler, Science 356, 1055 (2017).
  • Hwan Chun et al. (2015) S. Hwan Chun, J.-W. Kim, J. Kim, H. Zheng, C. C. Stoumpos, C. Malliakas, J. Mitchell, K. Mehlawat, Y. Singh, Y. Choi, T. Gog, A. Al-Zein, M. M. Sala, M. Krisch, J. Chaloupka, G. Jackeli, G. Khaliullin,  and B. J. Kim, Nat. Phys. 11, 462 (2015).
  • Suzuki et al. (2021) H. Suzuki, H. Liu, J. Bertinshaw, K. Ueda, H. Kim, S. Laha, D. Weber, Z. Yang, L. Wang, H. Takahashi, K. Fürsich, M. Minola, B. V. Lotsch, B. J. Kim, H. Yavas, M. Daghofer, J. Chaloupka, G. Khaliullin, H. Gretarsson,  and B. Keimer, Nat. Commun. 12, 1 (2021).
  • Winter et al. (2017b) S. M. Winter, K. Riedl, P. A. Maksimov, A. L. Chernyshev, A. Honecker,  and R. Valentí, Nat. Commun. 8, 1 (2017b).
  • Banerjee et al. (2018) A. Banerjee, P. Lampen-Kelley, J. Knolle, C. Balz, A. A. Aczel, B. Winn, Y. Liu, D. Pajerowski, J. Yan, C. A. Bridges, A. T. Savici, B. C. Chakoumakos, M. D. Lumsden, D. A. Tennant, R. Moessner, D. G. Mandrus,  and S. E. Nagler, npj Quantum Mater. 3, 1 (2018).
  • Kasahara et al. (2018) Y. Kasahara, T. Ohnishi, Y. Mizukami, O. Tanaka, S. Ma, K. Sugii, N. Kurita, H. Tanaka, J. Nasu, Y. Motome, T. Shibauchi,  and Y. Matsuda, Nature 559, 227 (2018).
  • Yokoi et al. (2021) T. Yokoi, S. Ma, Y. Kasahara, S. Kasahara, T. Shibauchi, N. Kurita, H. Tanaka, J. Nasu, Y. Motome, C. Hickey, S. Trebst,  and Y. Matsuda, Science 373, 568 (2021).
  • Liu and Khaliullin (2018) H. Liu and G. Khaliullin, Phys. Rev. B 97, 014407 (2018).
  • Liu (2021) H. Liu, Int. J. Mod. Phys. B 35, 2130006 (2021).
  • Liu et al. (2020) H. Liu, J. Chaloupka,  and G. Khaliullin, Phys. Rev. Lett. 125, 047201 (2020).
  • Sano et al. (2018) R. Sano, Y. Kato,  and Y. Motome, Phys. Rev. B 97, 014408 (2018).
  • Rau et al. (2014) J. G. Rau, E. K.-H. Lee,  and H.-Y. Kee, Phys. Rev. Lett. 112, 077204 (2014).
  • Winter et al. (2016) S. M. Winter, Y. Li, H. O. Jeschke,  and R. Valentí, Phys. Rev. B 93, 214431 (2016).
  • Lines (1963) M. Lines, Phys. Rev. 131, 546 (1963).
  • Oguchi (1965) T. Oguchi, J. Phys. Soc. Japan 20, 2236 (1965).
  • Scharf et al. (1979) W. Scharf, H. Weitzel, I. Yaeger, I. Maartense,  and B. Wanklyn, J. Magn. Magn. Mater. 13, 121 (1979).
  • Maartense et al. (1977) I. Maartense, I. Yaeger,  and B. Wanklyn, Solid State Commun. 21, 93 (1977).
  • Kobayashi et al. (1999) S. Kobayashi, S. Mitsuda, M. Ishikawa, K. Miyatani,  and K. Kohn, Phys. Rev. B 60, 3331 (1999).
  • Lee et al. (2010) S. Lee, R. K. Kaul,  and L. Balents, Nat. Phys. 6, 702 (2010).
  • Coldea et al. (2010) R. Coldea, D. Tennant, E. Wheeler, E. Wawrzynska, D. Prabhakaran, M. Telling, K. Habicht, P. Smeibidl,  and K. Kiefer, Science 327, 177 (2010).
  • Morris et al. (2014) C. Morris, R. V. Aguilar, A. Ghosh, S. Koohpayeh, J. Krizan, R. Cava, O. Tchernyshyov, T. McQueen,  and N. Armitage, Phys. Rev. Lett. 112, 137403 (2014).
  • Fava et al. (2020) M. Fava, R. Coldea,  and S. Parameswaran, Proc. Natl. Acad. Sci. 117, 25219 (2020).
  • Morris et al. (2021) C. Morris, N. Desai, J. Viirok, D. Hüvonen, U. Nagel, T. Room, J. Krizan, R. Cava, T. McQueen, S. Koohpayeh, R. K. Kaul,  and N. P. Armitage, Nat. Phys. 17, 832 (2021).
  • Lefrançois et al. (2016) E. Lefrançois, M. Songvilay, J. Robert, G. Nataf, E. Jordan, L. Chaix, C. V. Colin, P. Lejay, A. Hadj-Azzem, R. Ballou,  and V. Simonet, Phys. Rev. B 94, 214416 (2016).
  • Bera et al. (2017) A. K. Bera, S. M. Yusuf, A. Kumar,  and C. Ritter, Phys. Rev. B 95, 094424 (2017).
  • Wong et al. (2016) C. Wong, M. Avdeev,  and C. D. Ling, J. Solid State Chem. 243, 18 (2016).
  • Chen et al. (2021) W. Chen, X. Li, Z. Hu, Z. Hu, L. Yue, R. Sutarto, F. He, K. Iida, K. Kamazawa, W. Yu, X. Lin,  and Y. Li, Phys. Rev. B 103, L180404 (2021).
  • Chaloupka et al. (2013) J. Chaloupka, G. Jackeli,  and G. Khaliullin, Phys. Rev. Lett. 110, 097204 (2013).
  • Fouet et al. (2001) J. Fouet, P. Sindzingre,  and C. Lhuillier, Eur. Phys. J. B 20, 241 (2001).
  • Kimchi and You (2011) I. Kimchi and Y.-Z. You, Phys. Rev. B 84, 180407 (2011).
  • Songvilay et al. (2020) M. Songvilay, J. Robert, S. Petit, J. A. Rodriguez-Rivera, W. D. Ratcliff, F. Damay, V. Balédent, M. Jiménez-Ruiz, P. Lejay, E. Pachoud, A. Hadj-Azzem, V. Simonet,  and C. Stock, Phys. Rev. B 102, 224429 (2020).
  • Kim et al. (2021) C. Kim, J. Jeong, G. Lin, P. Park, T. Masuda, S. Asai, S. Itoh, H.-S. Kim, H. Zhou, J. Ma,  and J.-G. Park, J. Phys. Condens. Matter 34, 045802 (2021).
  • Lin et al. (2021) G. Lin, J. Jeong, C. Kim, Y. Wang, Q. Huang, T. Masuda, S. Asai, S. Itoh, G. Günther, M. Russina, Z. Lu, J. Sheng, L. Wang, J. Wang, G. Wang, Q. Ren, C. Xi, W. Tong, L. Ling, Z. Liu, J. Mei, Z. Qu, H. Zhou, J.-G. Park, Y. Wan,  and J. Ma, Nat. Commun. 12, 1 (2021).
  • Nair et al. (2018) H. S. Nair, J. Brown, E. Coldren, G. Hester, M. Gelfand, A. Podlesnyak, Q. Huang,  and K. Ross, Phys. Rev. B 97, 134409 (2018).
  • Regnault et al. (2006) L. Regnault, C. Boullier,  and J. Henry, Physica B Condens. Matter 385, 425 (2006).
  • Regnault et al. (2018) L.-P. Regnault, C. Boullier,  and J. Lorenzo, Heliyon 4, e00507 (2018).
  • Regnault et al. (1977) L. Regnault, P. Burlet,  and J. Rossat-Mignod, Physica B Condens. Matter 86, 660 (1977).
  • Zhong et al. (2020) R. Zhong, T. Gao, N. P. Ong,  and R. J. Cava, Sci. Adv. 6, eaay6953 (2020).
  • Zhang et al. (2021) X. Zhang, Y. Xu, R. Zhong, R. Cava, N. Drichko,  and N. Armitage, arXiv preprint arXiv:2106.13418  (2021).
  • Shi et al. (2021) L. Shi, X. Wang, R. Zhong, Z. Wang, T. Hu, S. Zhang, Q. Liu, T. Dong, F. Wang,  and N. Wang, Phys. Rev. B 104, 144408 (2021).
  • Das et al. (2021) S. Das, S. Voleti, T. Saha-Dasgupta,  and A. Paramekanti, Phys. Rev. B 104, 134425 (2021).
  • Maksimov et al. (2022) P. A. Maksimov, A. V. Ushakov, Z. V. Pchelkina, Y. Li, S. M. Winter,  and S. V. Streltsov, arXiv preprint arXiv:2204.09695  (2022).
  • Note (1) The coefficients UαβγδU_{\alpha\beta\gamma\delta} may be grouped according to the number of unique orbital indices, from one to four. For example, the intra-orbital Hubbard terms ni,α,ni,α,n_{i,\alpha,\uparrow}n_{i,\alpha,\downarrow} have one unique index α\alpha, while the inter-orbital Hubbard terms ni,α,σni,β,σn_{i,\alpha,\sigma}n_{i,\beta,\sigma^{\prime}} have two unique indices α,β\alpha,\beta. In the spherically symmetric approximation Sugano (2012), the Coulomb coefficients with three and four indices vanish unless at least one of the orbitals is an ege_{g} orbital. For this reason, t2gt_{2g}-only (and ege_{g}-only) models reduce to the familiar Kanamori formGeorges et al. (2013); Pavarini (2014), which includes only Hubbard density-density repulsion, Hund’s exchange, and pair-hopping contributions. However, when both ege_{g} and t2gt_{2g} orbitals are considered together, it is important to include the full rotationally symmetric Coulomb terms. This is particularly true when computing anisotropic magnetic exchange, because any approximations to the Coulomb Hamiltonian are likely to explicitly break rotational symmetry, leading to erroneous sources of anisotropy.
  • Sugano (2012) S. Sugano, Multiplets of transition-metal ions in crystals (Elsevier, 2012).
  • Pavarini (2014) E. Pavarini, in Many-Electron Approaches in Physics, Chemistry and Mathematics (Springer, 2014) pp. 321–341.
  • Sarte et al. (2018) P. M. Sarte, R. A. Cowley, E. E. Rodriguez, E. Pachoud, D. Le, V. García-Sakai, J. W. Taylor, C. D. Frost, D. Prabhakaran, C. MacEwen, A. Kitada, A. J. Browne, M. Songvilay, Z. Yamani, W. J. L. Buyers, J. P. Attfield,  and C. Stock, Phys. Rev. B 98, 024415 (2018).
  • Ross et al. (2017) K. A. Ross, J. Brown, R. Cava, J. Krizan, S. E. Nagler, J. Rodriguez-Rivera,  and M. B. Stone, Phys. Rev. B 95, 144414 (2017).
  • Koepernik and Eschrig (1999) K. Koepernik and H. Eschrig, Physical Review B 59, 1743 (1999).
  • Opahle et al. (1999) I. Opahle, K. Koepernik,  and H. Eschrig, Physical Review B 60, 14035 (1999).
  • Perdew et al. (1996) J. P. Perdew, K. Burke,  and M. Ernzerhof, Physical review letters 77, 3865 (1996).
  • Gray (1965) H. B. Gray, Electrons and chemical bonding (WA Benjamin, Inc., 1965).
  • Sarvezuk et al. (2011) P. W. C. Sarvezuk, E. J. Kinast, C. Colin, M. Gusmão, J. Da Cunha,  and O. Isnard, Int. J. Appl. Phys. 109, 07E160 (2011).
  • Dordević (2008) T. Dordević, Acta Crystallogr. E 64, i58 (2008).
  • Xiao et al. (2019) G. Xiao, Z. Xia, W. Zhang, X. Yue, S. Huang, X. Zhang, F. Yang, Y. Song, M. Wei, H. Deng,  and D. Jiang, Crystal Growth & Design 19, 2658 (2019).
  • Note (2) The Na2Co2TeO6 structure contains disorder in the Na position, in which each Na position has occupancy 2/3. To perform calculations, we artificially increased the occupancy to 1, which corresponds to Na3Co2TeO6. It is expected this change in the filling should have minimal impact on the computed hoppings.
  • Wilkinson et al. (1959) M. Wilkinson, J. Cable, E. Wollan,  and W. Koehler, Phys. Rev. 113, 497 (1959).
  • Wyckoff and Wyckoff (1963) R. W. G. Wyckoff and R. W. Wyckoff, Crystal structures, Vol. 1 (Interscience publishers New York, 1963).
  • Wellm et al. (2021) C. Wellm, W. Roscher, J. Zeisner, A. Alfonsov, R. Zhong, R. J. Cava, A. Savoyant, R. Hayn, J. van den Brink, B. Büchner, O. Janson,  and V. Kataev, Phys. Rev. B 104, L100420 (2021).
  • Ross et al. (2011) K. A. Ross, L. Savary, B. D. Gaulin,  and L. Balents, Phys. Rev. X 1, 021002 (2011).
  • Maksimov et al. (2019) P. Maksimov, Z. Zhu, S. R. White,  and A. Chernyshev, Phys. Rev. X 9, 021017 (2019).
  • Goodenough (1963) J. B. Goodenough, Magnetism and the chemical bond, Vol. 1 (Interscience publishers, 1963).
  • Ma et al. (2016) J. Ma, Y. Kamiya, T. Hong, H. Cao, G. Ehlers, W. Tian, C. Batista, Z. Dun, H. Zhou,  and M. Matsuda, Physical review letters 116, 087201 (2016).
  • Ito et al. (2017) S. Ito, N. Kurita, H. Tanaka, S. Ohira-Kawamura, K. Nakajima, S. Itoh, K. Kuwahara,  and K. Kakurai, Nature communications 8, 1 (2017).
  • Ghioldi et al. (2018) E. A. Ghioldi, M. G. Gonzalez, S.-S. Zhang, Y. Kamiya, L. O. Manuel, A. E. Trumper,  and C. D. Batista, Physical Review B 98, 184403 (2018).
  • Kamiya et al. (2018) Y. Kamiya, L. Ge, T. Hong, Y. Qiu, D. Quintero-Castro, Z. Lu, H. Cao, M. Matsuda, E. Choi, C. Batista, ,  and M. Mourigal, Nature communications 9, 1 (2018).
  • Sanders et al. (2021) A. L. Sanders, R. A. Mole, J. Liu, A. J. Brown, D. Yu, C. D. Ling,  and S. Rachel, arXiv preprint arXiv:2112.12254  (2021).
  • Samarakoon et al. (2021) A. M. Samarakoon, Q. Chen, H. Zhou,  and V. O. Garlea, Physical Review B 104, 184415 (2021).
  • Chaloupka and Khaliullin (2016) J. Chaloupka and G. Khaliullin, Phys. Rev. B 94, 064435 (2016).
  • Huyan et al. (2022) S. Huyan, J. Schmidt, E. Gati, R. Zhong, R. J. Cava, P. C. Canfield,  and S. L. Bud’ko, arXiv preprint arXiv:2201.12233  (2022).
  • Georges et al. (2013) A. Georges, L. d. Medici,  and J. Mravlje, Annu. Rev. Condens. Matter Phys. 4, 137 (2013).

Appendix A Low-Energy Multiplet Composition

The pure L,SL,S multiplets |mL,mS|m_{L},m_{S}\rangle can be conveniently expressed in terms of the single-particle levels with precise orbital momentum:

|ea,σ=\displaystyle|e_{a,\sigma}\rangle= |dz2,σ\displaystyle\ |d_{z^{2},\sigma}\rangle (77)
|eb,σ=\displaystyle|e_{b,\sigma}\rangle= |dx2y2,σ\displaystyle\ |d_{x^{2}-y^{2},\sigma}\rangle (78)
|t+,σ=\displaystyle|t_{+,\sigma}\rangle= 12(|dyz,σ+i|dxz,σ)\displaystyle\ -\frac{1}{\sqrt{2}}\left(|d_{yz,\sigma}\rangle+i|d_{xz,\sigma}\rangle\right) (79)
|t0,σ=\displaystyle|t_{0,\sigma}\rangle= |dxy,σ\displaystyle\ |d_{xy,\sigma}\rangle (80)
|t,σ=\displaystyle|t_{-,\sigma}\rangle= 12(|dyz,σi|dxz,σ)\displaystyle\ \frac{1}{\sqrt{2}}\left(|d_{yz,\sigma}\rangle-i|d_{xz,\sigma}\rangle\right) (81)

This leads to:

|1,32=\displaystyle\left|-1,\frac{3}{2}\right\rangle= |ea,eb,t+,t0,t0,t,t,\displaystyle\ \left|e_{a,\uparrow}e_{b,\uparrow}t_{+,\uparrow}t_{0,\uparrow}t_{0,\downarrow}t_{-,\uparrow}t_{-,\downarrow}\right\rangle (82)
|0,12=\displaystyle\left|0,\frac{1}{2}\right\rangle= 13|ea,eb,t+,t+,t0,t,t,\displaystyle\ \frac{1}{\sqrt{3}}\left|e_{a,\uparrow}e_{b,\uparrow}t_{+,\uparrow}t_{+,\downarrow}t_{0,\downarrow}t_{-,\uparrow}t_{-,\downarrow}\right\rangle (83)
+13|ea,eb,t+,t+,t0,t,t,\displaystyle\ +\frac{1}{\sqrt{3}}\left|e_{a,\uparrow}e_{b,\downarrow}t_{+,\uparrow}t_{+,\downarrow}t_{0,\uparrow}t_{-,\uparrow}t_{-,\downarrow}\right\rangle
+13|ea,eb,t+,t+,t0,t,t,\displaystyle\ +\frac{1}{\sqrt{3}}\left|e_{a,\downarrow}e_{b,\uparrow}t_{+,\uparrow}t_{+,\downarrow}t_{0,\uparrow}t_{-,\uparrow}t_{-,\downarrow}\right\rangle
|1,12=\displaystyle\left|1,-\frac{1}{2}\right\rangle= 13|ea,eb,t+,t+,t0,t0,t,\displaystyle\ \frac{1}{\sqrt{3}}\left|e_{a,\uparrow}e_{b,\downarrow}t_{+,\uparrow}t_{+,\downarrow}t_{0,\uparrow}t_{0,\downarrow}t_{-,\downarrow}\right\rangle (84)
+13|ea,eb,t+,t+,t0,t0,t,\displaystyle\ +\frac{1}{\sqrt{3}}\left|e_{a,\downarrow}e_{b,\uparrow}t_{+,\uparrow}t_{+,\downarrow}t_{0,\uparrow}t_{0,\downarrow}t_{-,\downarrow}\right\rangle
+13|ea,eb,t+,t+,t0,t0,t,\displaystyle\ +\frac{1}{\sqrt{3}}\left|e_{a,\downarrow}e_{b,\downarrow}t_{+,\uparrow}t_{+,\downarrow}t_{0,\uparrow}t_{0,\downarrow}t_{-,\uparrow}\right\rangle

The time-reversed partners can be similarly obtained.