Introduction

In the realm of quantum spin liquids (QSLs), quantum fluctuations prevent localized magnetic moments from establishing conventional magnetic order, wherein excited nonlocal quasiparticles hold promise for applications in decoherence-free topological quantum computing1,2,3,4,5,6,7. The Kitaev model, exhibiting exchange frustration between localized magnetic moments residing on a honeycomb lattice, is one viable model for the QSL since it is exactly solvable through the introduction of Majorana fermions3. The Hamiltonian is given by \({{{\mathcal{H}}}}_{{{\rm{Kitaev}}}}={\sum }_{\mu }{\sum }_{{\langle i,i^{\prime} \rangle }_{\mu }}K{S}_{i}^{\mu }{S}_{i^{\prime} }^{\mu }\), where K represents the coupling constant for the bond-dependent Ising-type interactions on three-type μ (x, y, and z) bonds on the honeycomb lattice, while \({S}_{i}^{\mu }\) (\({S}_{i^{\prime} }^{\mu }\)) signifies the μ component of the spin-1/2 operator at site i (its nearest-neighbor site \(i^{\prime}\) on the μ bond). The realization of Kitaev-type interactions has been achieved in spin-orbit coupled Mott insulators, where the interplay of electron correlation and spin-orbit coupling (SOC) is pivotal8,9. This is notably evident in the spin-orbital entangled Kramers doublet Γ7, described by jeff = 1/2 pseudospins, which typically originates from the low-spin d5 electron configuration under the octahedral crystal field (OCF). Indeed, the presence of dominant Kitaev interactions has been unveiled for several quasi-two-dimensional honeycomb compounds, such as A2IrO3 (A = Na, Li), α-RuCl3, and other related materials9,10,11,12,13,14,15,16,17,18,19,20,21. In these compounds, the Kitaev interactions stem from second-order perturbation processes with respect to the electron hopping mediated by ligands in edge-sharing MX6 octahedra, where M represents a transition metal cation, and X denotes a ligand ion (see Fig. 1a).

Fig. 1: Challenge in the derivation of effective exchange interactions by second-order perturbation calculations for 4f-electron systems.
figure 1

a Schematic of the Kitaev model realized in an edge-sharing network of RX6 octahedra. Three-type μ (x, y, and z) bonds on the honeycomb lattice are distinguished. The hopping paths R-X-R on a z bond are represented by the purple lines. b Schematic of the calculations. We successfully cover the second-order perturbation calculations on a massive scale by employing a parallelization scheme spanning all the possible intermediate states \(\left\vert 4{f}^{n-1}\right\rangle \otimes \left\vert 4{f}^{n+1}\right\rangle\), given the initial and final states \(\left\vert 4{f}^{n}\right\rangle \otimes \left\vert 4{f}^{n}\right\rangle\). The sequence of perturbation processes is schematically depicted for the case of n = 5: 1 represents the initial state \(\left\vert 4{f}^{5}\right\rangle \otimes \left\vert 4{f}^{5}\right\rangle\), 2 denotes an f-electron hopping from one \(\left\vert 4{f}^{5}\right\rangle\) to the other \(\left\vert 4{f}^{5}\right\rangle\), 3 represents an intermediate state \(\left\vert 4{f}^{4}\right\rangle \otimes \left\vert 4{f}^{6}\right\rangle\), 4 denotes an f-electron hopping from \(\left\vert 4{f}^{6}\right\rangle\) to \(\left\vert 4{f}^{4}\right\rangle\), and 5 represents the final state \(\left\vert 4{f}^{5}\right\rangle \otimes \left\vert 4{f}^{5}\right\rangle\) that is the same as the initial state. c The number of the intermediate states, Nint, for the 4fn-4fn states: Nint = 182 for n = 1 and n = 13, Nint = 182,182 for n = 3 and n = 11, and Nint = 6,012,006 for n = 5 and n = 9. The low-spin d5 case with Nint = 30 is shown for comparison. The initial/final and the intermediate states are schematically depicted for each electron configuration.

Besides the d-electron transition metal compounds, rare-earth materials with f electrons meet the requirements for the Kitaev interactions: cooperation of electron correlation and SOC. There exist rare-earth quasi-two-dimensional honeycomb materials, which would be deemed as intriguing platforms for realizing the Kitaev model, e.g, Na2PrO322,23, SmI324, DyCl325, ErX3 (X = Cl, Br, I)26,27, and YbCl328. Notably, the expectation is that antiferromagnetic (AFM) Kitaev interactions would manifest in the 4f1 electron configuration in A2PrO3 including Na2PrO3, in contrast to the ferromagnetic (FM) ones that typically dominate in d-electron systems20,29,30. Nonetheless, the design and discovery of f-electron materials with strong Kitaev-type interactions remain elusive. This is mainly because the formulation of low-energy effective models for 4f-electron systems remains a formidable challenge, as it requires significant computational efforts for second-order perturbation calculations (see Fig. 1b). Indeed, the number of the intermediate states in the perturbation processes becomes 182,182 and 6,012,006 for 4f3 (4f11) and 4f5 (4f9), respectively, in stark contrast to only 30 for the low-spin d5 electron configurations (see Fig. 1c).

To address this issue, we develop a highly parallel computational program capable of exhaustively performing second-order perturbation calculations on a massive scale. The program comprises three key steps. First, the eigenvectors \(\left\vert 4{f}^{n}\right\rangle\) and the eigenvalues \({E}_{4{f}^{n}}\) for all the many-electron states with 4fn electron configurations are prepared. In this step, the Coulomb interaction \({{{\mathcal{H}}}}_{{{\rm{int}}}}\) between 4f electrons is taken into account, along with the subsequent SOC \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\), based on the Russell-Saunders coupling scheme31. Second, upon using these many-electron states, the initial and final 4fn-4fn states for neighboring sites (the tensor products \(\left\vert 4{f}^{n}\right\rangle \otimes \left\vert 4{f}^{n}\right\rangle\) of 4fn state pairs) and all the possible 4fn−1-4fn+1 intermediate states (the tensor products \(\left\vert 4{f}^{n-1}\right\rangle \otimes \left\vert 4{f}^{n+1}\right\rangle\) of 4fn−1 and 4fn+1 state pairs) are automatically generated. Third, effective magnetic couplings are estimated based on the second-order perturbation expansion with respect to the 4f electron hopings \({{{\mathcal{H}}}}_{{{\rm{hop}}}}\). It is worth noting that the program can be flexibly extended beyond the second-order perturbation; it is capable of computing higher-order contributions, including multiple-spin interactions. We emphasize that even the second-order perturbation calculations are impracticable without efficient parallel computation (Fig. 1a) since the number of the intermediate states exceeds 6+ million for the 4f5 and 4f9 cases (Fig. 1b). This parallelization is achieved by implementing the Message Passing Interface in the C++ programming language.

In this study, we employ the program for the design of rare-earth Kitaev-type materials. For the 4fn-4fn states with n = 1, 3, 5, 9, 11, and 13, we assume a perfect OCF \({{{\mathcal{H}}}}_{{{\rm{OCF}}}}\) within the edge-sharing RX6 octahedra (R = rare-earth ions), along with \({{{\mathcal{H}}}}_{{{\rm{int}}}}\) and \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\). This results in the formation of spin-orbital entangled Kramers doublet for all n, depending on the crystal field parameters. In the perturbation, we take into account the indirect 4f-p-4f electron hoppings \({{{\mathcal{H}}}}_{{{\rm{hop}}}}\) via p orbitals of ligand X with the use of the Slater-Koster transfer integrals tpfπ and tpfσ32, and the p-4f energy difference Δp-f in the intermediate states. Our analysis reveals that in all cases the low-energy Hamiltonian can be effectively described by two predominant exchange interactions between the pseudospins for the Kramers doublet: the bond-independent isotropic Heisenberg interaction denoted as J [given in \({{{\mathcal{H}}}}_{{{\rm{Heisenberg}}}}={\sum }_{\langle i,i^{\prime} \rangle }J{{{\bf{S}}}}_{i}\cdot {{{\bf{S}}}}_{i^{\prime} }\), where \({{{\bf{S}}}}_{i}={({S}_{i}^{x},{S}_{i}^{y},{S}_{i}^{z})}^{{{\rm{T}}}}\)] and the bond-dependent anisotropic Kitaev interaction K. In most instances, both J and K exhibit AFM behavior. Notably, in the cases of 4f3 (as exemplified in Nd3+) and the electron-hole counterpart 4f11 (Er3+), we find that K largely dominates over J, which realizes situations close to the pure Kitaev model. This finding opens up unexpected opportunities for investigating the Kitaev QSLs in 4f-electron systems. Furthermore, beyond the scope of the Kitaev model, our computational program can also be applied to a wide range of 4f-electron magnets, which would contribute to future exploration of exotic rare-earth magnetism.

Results

Ground-state Kramers doublets

Let us begin with the analysis of the crystal field splitting of the ground-state multiplets given by the Russell-Saunders coupling scheme, focusing on the 4fn electron configurations with odd n20. In 4f1 electron configuration (Fig. 2a), the Coulomb interaction \({{{\mathcal{H}}}}_{{{\rm{int}}}}\) is irrelevant, leaving 14-fold 2F manifold. This is split by \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\) into the 2F5/2 sextet and the 2F7/2 octet. The ground-state 2F5/2 sextet is further split by \({{{\mathcal{H}}}}_{{{\rm{OCF}}}}\) into the Γ7 doublet and Γ8 quartet. Since the Γ7 doublet has lower energy than the Γ8 quartet, the 4f1 case gives the Γ7 Kramers doublet in the ground state. In the 4f3 electron configuration (Fig. 2b), \({{{\mathcal{H}}}}_{{{\rm{int}}}}\) gives 52-fold 4I manifold in the ground state, which is split by \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\) into four multiplets 4I9/2, 4I11/2, 4I13/2, and 4I15/2. The lowest-energy 4I9/2 dectet is further split by \({{{\mathcal{H}}}}_{{{\rm{OCF}}}}\) into the Γ6 doublet and two Γ8 quartets. The ground state depends on the crystal field parameters B40 and B60 (see Methods), and the Γ6 Kramers doublet is selected when B40 is predominant. In the 4f5 electron configuration (Fig. 2c), the lowest-energy 6H5/2 sextet selected by \({{{\mathcal{H}}}}_{{{\rm{int}}}}\) and \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\) is split by \({{{\mathcal{H}}}}_{{{\rm{OCF}}}}\) into the Γ7 doublet and the Γ8 quartet, and the ground state is given by the lower-energy Γ7 Kramers doublet, similar to the 4f1 case. In the 4f9 electron configuration (Fig. 2d), the lowest-energy 6H15/2 sexdectet selected by \({{{\mathcal{H}}}}_{{{\rm{int}}}}\) and \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\) is split by \({{{\mathcal{H}}}}_{{{\rm{OCF}}}}\) into the Γ6 doublet, the Γ7 doublet, and the three Γ8 quartets. The ground state is either the Γ6 doublet or the Γ7 doublet depending on the crystal field parameters. In the 4f11 electron configuration (Fig. 2e), the ground state is given by the Γ7 doublet when B60 is predominant. Finally, in the 4f13 electron configuration (Fig. 2f), the ground state is given by the Γ6 doublet, irrespective of the crystal field parameters. Thus, all 4fn cases considered here can offer the Kramers doublet in the ground state of an isolated ion.

Fig. 2: Schematic representation of multiplet splittings for 4fn electron configurations with odd integers n (except for n = 7).
figure 2

a 4f1, b 4f3, c 4f5, d 4f9, e 4f11, and f 4f13 cases are represented. In each configuration, the ground-state multiplet \({2S+1\atop }L\), initially determined by the Coulomb interaction \({{{\mathcal{H}}}}_{{{\rm{int}}}}\) (left), undergoes splitting by the spin-orbit coupling \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\), resulting in the ground-state multiplet \({2S+1\atop }{L}_{J}\) (middle). Subsequently, \({2S+1\atop }{L}_{J}\) is further split by the octahedral crystal field \({{{\mathcal{H}}}}_{{{\rm{OCF}}}}\), leading to the formation of ground-state Kramers doublets Γ7 (in red) or Γ6 (in blue) or quartets Γ6 (in thick green). In the cases of 4f3 in (b), 4f9 in (d), and 4f11 in (e), the ground state is contingent upon the parameters B40 and B60 governing \({{{\mathcal{H}}}}_{{{\rm{OCF}}}}\); we present two extreme cases of B40 = 0 (left) and B60 = 0 (right). The 4f7 case is not shown as the orbital is quenched. The corresponding wave functions for the Kramers doublets are also depicted, with red and blue denoting spin-up and spin-down density profiles, respectively.

Table 1 explicitly enumerates all the accessible ground-state Kramers doublets characterized by the pseudospin jeff = 1/2. In this table, each jeff = 1/2 state is described with j and jz representations; j (jz) is the (secondary) total angular momentum quantum number. For the pseudospin jeff = 1/2 degree of freedom, one can introduce the operator \({{\bf{S}}}={({S}^{x},{S}^{y},{S}^{z})}^{{{\rm{T}}}}\) defined by

$${S}^{\mu }={\mathbb{S}}\left(\begin{array}{cc}\left\langle +\right\vert {j}^{\mu }\left\vert +\right\rangle &\left\langle +\right\vert {j}^{\mu }\left\vert -\right\rangle \\ \left\langle -\right\vert {j}^{\mu }\left\vert +\right\rangle &\left\langle -\right\vert {j}^{\mu }\left\vert -\right\rangle \end{array}\right)=\frac{1}{2}{\sigma }^{\mu },$$
(1)

where j = (jxjyjz) and σ = (σxσyσz) are the vectors of the total angular momentum operators and the Pauli matrices, respectively, and \({\mathbb{S}}\) is a real scalar.

Table 1 Ground-state multiplets and possible Kramers doublets for 4fn electron configurations with odd integers n (except for n = 7)

Effective exchange couplings

Subsequently, the program proceeds to determine the second-quantized representations with multiple f-orbital bases for \(\left\vert 4{f}^{n}\right\rangle \otimes \left\vert 4{f}^{n}\right\rangle\) by the aforementioned Kramers doublets, which is commonly used for the initial and final states of the perturbation. Additionally, it constructs the representations for all conceivable intermediate states \(\left\vert 4{f}^{n-1}\right\rangle \otimes \left\vert 4{f}^{n+1}\right\rangle\). The energy difference between the initial/final states and the intermediate states is determined by two key parameters in the Hamiltonian \({{{\mathcal{H}}}}_{{{\rm{int}}}}\), namely, the onsite Coulomb interaction U and the Hund’s-rule coupling JH, as well as another in \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\), namely, the SOC coefficient λ. For obtaining the values of U and λ, the Herbst-Wilkins table33 and the Freeman-Watson table34 are consulted, respectively. The parameter JH is adjusted to achieve the alignment of energy differences between different multiplets according to the Dieke diagram35; see Supplementary Note 1.

Given the representations and the excitation energies described above, J and K are calculated by employing the parallelization scheme for perturbation calculations spanning the intermediate states. For the hopping parameters, we adopt the Slater-Koster transfer integrals32, while changing the ratio ∣tpfπ/tpfσ∣ between 0 and 1; we take tpfπ/tpfσ < 0. The results are summarized in Fig. 3 for all 4fn cases. In the 4f1Γ7 case (Fig. 3a), which includes 182 4f0-4f2 intermediate states, our results emphasize the dominance of the AFM K over compatibly the subdominant AFM J in the wide range of 0 < ∣tpfπ/tpfσ∣ ≤ 0.8. This behavior aligns with the findings based on the first-principles calculations in refs. 30,31. The magnitudes of K and J both monotonically increase with ∣tpfπ/tpfσ∣, which is a general trend seen also for most of the other cases below. In the 4f3Γ6 case with 182,182 4f2-4f4 intermediate states (Fig. 3b), the intriguing scenario arises in which the AFM K overwhelmingly outweighs non-negligible AFM J; this is particularly pronounced at ∣tpfπ/tpfσ∣ ≃ 1.0, where J almost vanishes. We also emphasize that K is one order of magnitude larger than that in the 4f1 case. Note that the coupling constants K and J are given in unit of \({t}_{pf\sigma }^{4}{\Delta }_{p-f}^{-2}\) eV−1; we will discuss the actual values later. In the 4f5Γ7 case with 6,012,006 4f4-4f6 intermediate states (Fig. 3c), the AFM K becomes predominant compared to the subdominant AFM J in the entire region of ∣tpfπ/tpfσ∣. In the 4f9 case (Fig. 3d), there are two cases, Γ6 and Γ7, depending on the crystal field parameters, both of which include 6,012,006 4f8-4f10 intermediate states. In the Γ6 case, K turns to be FM, while J remains AFM and predominant compared to K. Meanwhile, in the Γ7 case, both K and J are AFM, while J is again predominant. In the case of 4f11Γ7 (Fig. 3e), the trends mirror the electron-hole counterpart, the 4f3Γ6 case; the AFM K becomes far predominant compared to the AFM J, while J does not decrease for large ∣tpfπ/tpfσ∣. We note that the magnitude of K is also large comparable to the 4f3 case. Finally, in the 4f13Γ6 case (Fig. 3f), the result is quite different from the electron-hole counterpart, the 4f1 case; K is notably suppressed compared to the predominant AFM J. This is, however, consistent with the prior findings in ref. 36. It should be noted that the off-diagonal terms, referred to as Γ and Γ′, were found to be zero across all the cases since we omit the direct 4f-4f electron hoppings in the present analyses.

Fig. 3: Two coupling constants, isotropic Heisenberg interaction J and anisotropic Kitaev interaction K, derived by the second-order perturbation for 4fn-4fn electron configurations with odd integers n (except for n = 7).
figure 3

a 4f1Γ7, b 4f3Γ6, c 4f5Γ7, d 4f9Γ6 and Γ7, e 4f11Γ7, and f 4f13Γ6 cases are represented. The data are plotted for ∣tpfπ/tpfσ∣; we take tpfπ/tpfσ < 0.

The results in Fig. 3 highlight that, in the majority of instances, aside from the 4f9 and 4f13 cases, the AFM K prevails over the subdominant AFM J. This suggests a heightened propensity for robust Kitaev interactions within diverse 4f-electron systems. This is more clearly demonstrated by plotting the ratio of ∣K/J∣ in Fig. 4. Except for the 4f9 and 4f13 cases (and the 4f1 case for large ∣tpfπ/tpfσ∣), ∣K/J∣ is greater than 1, indicating the predominant Kitaev interactions. Interestingly, besides the 4f1 case, ∣K/J∣ consistently exhibits monotonic increases with ∣tpfπ/tpfσ∣. It is noteworthy that the substantial predominance of AFM K over AFM J is particularly viable, especially in the cases of 4f3Γ6 and 4f11Γ7. In both scenarios, it is observed that ∣K/J∣ > 4 for ∣tpfπ/tpfσ∣ ≳ 0.6, which includes the realistic range of the parameters37. In addition, the magnitude of K is considerably larger than in the other cases. To emphasize these prominent properties, we show the estimates of J, K, and ∣K/J∣ in Table 2, assuming the typical values of the parameters as tpfσ = 0.35 eV, tpfπ/tpfσ = −0.737, and Δp-f = 1 eV. Notably, for the 4f3Γ6 and 4f11Γ7 configurations, it is demonstrated that K = 1.21 meV and 1.27 meV, respectively, which are one order of magnitude larger than the other cases, and furthermore, K/J = 6.89 and 5.21, signifying the substantial AFM K prevalence over the AFM J.

Fig. 4: Ratio between the Kitaev and Heisenberg interactions, ∣K/J∣, for different 4fn cases.
figure 4

We exclude the data when ∣K∣ or ∣J∣ are extremely small: \(| K| \, < \, 1{0}^{-5}{t}_{pf\sigma }^{4}{\Delta }_{p{{-}}f}^{-2}\) eV−1 and \(| J| \, < \, 1{0}^{-5}{t}_{pf\sigma }^{4}{\Delta }_{p{{-}}f}^{-2}\) eV−1.

Table 2 Plausible estimates of the isotropic Heisenberg interaction J, anisotropic Kitaev interaction K, and their ratio ∣K/J∣ for various Kramers doublets in 4f-electron systems

Candidate materials

Let us finally discuss candidate materials for the 4f1, 4f3, 4f5, and 4f11 cases where K dominates J in our calculations. First, for 4f1, the authors and their collaborators previously identified A2PrO3 (A = alkali metals) as potential Kitaev-type magnets with the conventional assumption in the Russell-Saunders coupling scheme whereby the ordering of energy scales is given as \({{{\mathcal{H}}}}_{{{\rm{int}}}} \, > \, {{{\mathcal{H}}}}_{{{\rm{SOC}}}} \, \gg \, {{{\mathcal{H}}}}_{{{\rm{OCF}}}}\)29,30. However, the tetravalent Pr4+ ion is recently recognized to reside in the intermediate coupling regime \({{{\mathcal{H}}}}_{{{\rm{SOC}}}} \sim {{{\mathcal{H}}}}_{{{\rm{OCF}}}}\)38,39,40. We have verified that in this regime the AFM K is reduced to be subdominant, while the AFM J prevails41. Second, for 4f3, Nd3+-based materials are considered promising, although honeycomb lattice compounds with Nd3+ have not yet been identified to the best of our knowledge. This observation suggests avenues for additional materials design in the exploration of Nd3+-based Kitaev-type magnets. Third, for 4f5, SmI3 has recently undergone experimental scrutiny as a potential host for the Kitaev QSL, given its absence of long-range spin magnetic order down to 0.1 K24. Further experiments are awaited to identify the relevant magnetic interactions. Finally, for 4f11, Er3+-based van der Waals magnets ErX3 (X = Cl, Br, I) were studied26,27. These materials have similar lattice structures to that of the prime candidate for the Kitaev QSL, α-RuCl3, and were shown to exhibit noncollinear vortex-type magnetic orders. A recent experiment for ErBr3 discussed the relevance of long-range dipolar interactions42. However, note that similar vortex-like magnetic orders were also found in extensions of the Kitaev mdoel30,43,44. It would be intriguing to revisit the Er3+-based materials by using ab initio approaches.

Discussion

Our comprehensive approach, leveraging a parallel computational program capable of massive-scale second-order perturbation calculations, has provided insights into the nature of exchange interactions in rare-earth quasi-two-dimensional honeycomb lattices. The observed dominance of the anisotropic Kitaev interaction over the isotropic Heisenberg interaction in certain cases, particularly for 4f3 and 4f11 configurations, opens new avenues for investigating the Kitaev-type QSL. In particular, our results highlight Nd3+ and Er3+-based magnets as plausible candidates for the Kitaev QSL. The developed computational program extends its utility beyond the Kitaev model, which would address a wide range of exchange interactions in 4f-electron systems. Particularly when used in conjunction with ab initio calculations, which allow for an in-depth numerical analysis of parameters for crystal fields and electron hoppings, the program will more accurately describe spin systems in 4f-electron systems, including the effects of not only direct 4f-4f direct hopping but also charge-transfer processes and cyclic exchanges14,45, which will be explored in subsequent researches. We also note that, although we demonstrated the simplest cases of OCF in this study, the program is applicable to any perturbation problem for various types of crystal fields with lower symmetry. This work not only contributes to advancing our understanding of rare-earth Kitaev-type materials but also lays the groundwork for future exploration of exotic magnetism in this intriguing field of research.

We note that computational speed could be significantly enhanced by separating \({c}_{im\sigma }^{{\dagger} }\) and \({c}_{i^{\prime} m^{\prime} \sigma }\) in \({{{\mathcal{H}}}}_{{{\rm{hop}}}}\)46, given that the site indices i and \(i^{\prime}\) pertain exclusively to subspaces with 4fn+1 and 4fn−1 configurations, respectively [see equations (12) and (13)]. This strategy would reduce the size of the perturbation calculations from \(\scriptstyle2\left({14}\atop{n-1}\right)\left({14}\atop{n+1}\right)\) (considering \(\left\vert 4{f}^{n-1}\right\rangle \otimes \left\vert 4{f}^{n+1}\right\rangle\)) to \(\scriptstyle14\left(\left({14}\atop{n-1}\right)+\left({14}\atop{n+1}\right)\right)\) (considering \(\left\vert 4{f}^{n-1}\right\rangle\) and \(\left\vert 4{f}^{n+1}\right\rangle\) separately). For example, at n = 6, this approach would decrease the size from 6,012,066 to 76,076, demonstrating a substantial reduction that warrants testing in future studies.

Methods

Coulomb interactions

The Hamiltonian \({{{\mathcal{H}}}}_{{{\rm{int}}}}\) describing the Coulomb interactions between f electrons is given by

$$\begin{array}{l}{{{\mathcal{H}}}}_{{{\rm{int}}}}={\sum} _{i}\sum\limits_{{m}_{1},{m}_{2},{m}_{3},{m}_{4}}\sum\limits_{{\sigma }_{1},{\sigma }_{2}}{\delta }_{{m}_{1}+{m}_{2},{m}_{3}+{m}_{4}}\sum\limits_{k=0,2,4,6}\\ {F}^{k}{C}^{(k)}({m}_{1},{m}_{4}){C}^{(k)}({m}_{2},{m}_{3}){c}_{i{m}_{1}{\sigma }_{1}}^{{\dagger} }{c}_{i{m}_{2}{\sigma }_{2}}^{{\dagger} }{c}_{i{m}_{3}{\sigma }_{2}}{c}_{i{m}_{4}{\sigma }_{1}},\end{array}$$
(2)

where Fk and C(k) denote the Slater-Condon parameters and the Guant coefficients, respectively (k = 0, 2, 4, 6); δ is the Kronecker delta; \({c}_{im\sigma }^{{\dagger} }\) and cimσ represent creation and annihilation operators of an electron at site i in the spherical harmonics basis, respectively (m and σ = ±1 denote the magnetic and spin quantum numbers, respectively). Here, the Slater-Condon parameters are related with the onsite Coulomb interaction U and the Hund’s-rule coupling JH as47,48

$$U={F}^{0},$$
(3)
$${J}_{{{\rm{H}}}}=\frac{1}{6435}\left(286{F}^{2}+195{F}^{4}+250{F}^{6}\right).$$
(4)

\({{{\mathcal{H}}}}_{{{\rm{int}}}}\) is diagonalized using the lowering operators of orbital and spin angular momenta, L and S, respectively, both of which commute with \({{{\mathcal{H}}}}_{{{\rm{int}}}}\):

$${L}^{-}={\sum}_{i}{\sum}_{{m}_{1},{m}_{2}}{\sum}_{\sigma }{\delta }_{{m}_{1},{m}_{2}+1}\sqrt{(\ell +m+1)(\ell -m)}{c}_{i{m}_{1}\sigma }^{{\dagger} }{c}_{i{m}_{2}\sigma },$$
(5)

where is the orbital quantum number taken as  = 3 for the f-orbital manifold, and

$${S}^{-}={\sum}_{i}{\sum}_{m}{\sum}_{{\sigma }_{1},{\sigma }_{2}}{\delta }_{{\sigma }_{1},{\sigma }_{2}+1}\sqrt{\left(\frac{1}{2}+\frac{{\sigma }_{2}}{2}+1\right)\left(\frac{1}{2}-\frac{{\sigma }_{2}}{2}\right)}{c}_{im{\sigma }_{1}}^{{\dagger} }{c}_{im{\sigma }_{2}}.$$
(6)

A leading eigenvector of the ground-state multiplet \({}^{2{S}_{1}+1}{L}_{1}\) with L1 (S1) being the largest L (S) for 4fn electron configuration is given as \(\underbrace{{c}_{i,\ell ,+1}^{{\dagger} }\cdots {c}_{i,\ell -n+1,+1}^{{\dagger} }}_{n}\left\vert 0\right\rangle\), where \(\left\vert 0\right\rangle\) denotes the vacuum state. Given this leading eigenvector, all the eigenvectors expanded within the \({}^{2{S}_{1}+1}{L}_{1}\) subspace can be derived by successively applying either L or S to the leading eigenvector. Then, the leading eigenvector for the multiplet \({}^{2{S}_{1}+1}{L}_{2}\), where L2 is the second-largest L, is constructed as a vector orthogonal to the eigenvector within the \({}^{2{S}_{1}+1}{L}_{1}\) subspace with the expectation values of Lz and Sz being L2 and S1, respectively:

$${L}^{z}={\sum}_{i}{\sum}_{m}{\sum}_{\sigma }m{c}_{im\sigma }^{{\dagger} }{c}_{im\sigma },$$
(7)
$${S}^{z}={\sum}_{i}{\sum}_{m}{\sum}_{\sigma }\frac{\sigma }{2}{c}_{im\sigma }^{{\dagger} }{c}_{im\sigma }.$$
(8)

Similarly, all the eigenvectors within the \({}^{2{S}_{1}+1}{L}_{2}\) subspace can be derived by successively applying either L or S to the leading eigenvector. This process, involving the application of either L or S to the leading eigenvector with orthogonality, is repeated for the remaining subspaces as well. Finally, we assess the numerical validity of the derived eigenvectors by examining their orthogonality and verifying the absence of nonzero off-diagonal elements in \({{{\mathcal{H}}}}_{{{\rm{int}}}}\).

Spin-orbit coupling

The Hamiltonian \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\) describing the effect of the SOC is given by

$${{{\mathcal{H}}}}_{{{\rm{SOC}}}}={\sum}_{i}{{{\mathcal{H}}}}_{{{\rm{SOC}}},i},$$
(9)

where

$$\begin{array}{l}{{{\mathcal{H}}}}_{{{\rm{SOC}}},i}=\frac{\lambda }{2}{\sum}_{m=-\ell }^{\ell }\sum\limits_{\sigma }m\sigma {c}_{im\sigma }^{{\dagger} }{c}_{im\sigma }\\ +\frac{\lambda }{2}{\sum }_{m=-\ell }^{\ell -1}\sqrt{\ell +m+1}\sqrt{\ell -m}({c}_{im+1-}^{{\dagger} }{c}_{im+}+{c}_{im+}^{{\dagger} }{c}_{im+1-}),\end{array}$$
(10)

where λ > 0 is the SOC coefficient.

\({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\) is diagonalized using the lowering operator of total angular momentum, J, which commutes with \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\): J = L + S. A leading eigenvector of the \({2S+1\atop }{L}_{{J}_{1}}\) with J1 being the largest J (= L + S) is given as the eigenvector with the expectation value of Jz (= Lz + Sz) being J1. All the eigenvectors within the \({2S+1\atop }{L}_{{J}_{1}}\) subspace can be derived by successively applying either J to the leading eigenvector. Then, the leading eigenvector of the \({2S+1\atop }{L}_{{J}_{2}}\), where J2 is the second-largest J2 (= L + S − 1), is constructed as a vector orthogonal to the eigenvector within the \({2S+1\atop }{L}_{{J}_{2}}\) subspace with the expectation values of Jz being J2. This process, involving the application of J to the leading eigenvector with orthogonality, is repeated for the remaining subspaces as well. Finally, we assess the numerical validity of the derived eigenvectors by examining their orthogonality and verifying the absence of nonzero off-diagonal elements in \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\). We also confirm that this approach using L and S for \({{{\mathcal{H}}}}_{{{\rm{int}}}}\) and J for \({{{\mathcal{H}}}}_{{{\rm{SOC}}}}\) yields the same eigenvectors and eigenvalues for the 4f2 electron configuration as those obtained using the Wigner 3-j symbols30.

Octahedral crystal field

The Hamiltonian \({{{\mathcal{H}}}}_{{{\rm{OCF}}}}\) describing the octahedral crystal field is given by

$${{{\mathcal{H}}}}_{{{\rm{OCF}}}}={B}_{40}{O}_{4}+{B}_{60}{O}_{6},$$
(11)

where O4 = O40 + 5O44 and O6 = O60 − 21O64. Ors (s = −r, −r + 1, ⋯, r) are the rank-r Stevens operators49, and B40 and B60 are the coefficients.

Electron hopping

The Hamiltonian \({{{\mathcal{H}}}}_{{{\rm{hop}}}}\) describing the kinetic energy of electron hopping via indirect 4f-p-4f hopping processes is given by

$${{{\mathcal{H}}}}_{{{\rm{hop}}}}={\sum}_{\mu }{\sum}_{{\langle i,i^{\prime} \rangle }_{\mu }}{{{\mathcal{H}}}}_{{{\rm{hop}}},ii^{\prime} }^{\left(\mu \right)},$$
(12)

where \({{{\mathcal{H}}}}_{{{\rm{hop}}},ii^{\prime} }^{\left(\mu \right)}\) denotes the electron hopping between nearest-neighbor sites i and \(i^{\prime}\) on the μ bond (μ = x, y, and z) as

$${{{\mathcal{H}}}}_{{{\rm{hop}}},ii^{\prime} }^{\left(\mu \right)}=\sum\limits_{m,m^{\prime} }\sum\limits_{\sigma =\pm }\left({\sum}_{o,p}\frac{{t}_{im,op,\sigma }{t}_{i^{\prime} m^{\prime} ,op,\sigma }}{{\Delta }_{p{{-}}f}}{c}_{im\sigma }^{{\dagger} }{c}_{i^{\prime} m^{\prime} \sigma }+{{\rm{h.c.}}}\right).$$
(13)

tiu,op,σ is the transfer integral for spin σ between 4f orbital u at site i and p orbital p (= x, y, and z) at one of two ligand sites o (= 1 and 2) shared by two RX6 octahedra for the sites i and \(i^{\prime}\), and Δp-f is the energy difference between p and 4f orbitals. For tim,op,σ and \({t}_{i^{\prime} m^{\prime} ,op,\sigma }\), we refer to ref. 33.

Perturbation expansion

The effective Hamiltonian for a pair of jeff = 1/2 pseudospins for nearest-neighbor sites i and \(i^{\prime}\) on a μ bond is calculated by

$${h}_{ii^{\prime} }^{\left(\mu \right)}=\sum\limits_{a,b,c,d=\pm }{\sum}_{n}\frac{\left\langle c,d\right\vert {{{\mathcal{H}}}}_{{{\rm{hop}}},ii^{\prime} }^{(\mu )}\left\vert n\right\rangle \left\langle n\right\vert {{{\mathcal{H}}}}_{{{\rm{hop}}},ii^{\prime} }^{\left(\mu \right)}\left\vert a,b\right\rangle }{{E}_{0}-{E}_{n}}\left\vert c,d\right\rangle \left\langle a,b\right\vert .$$
(14)

where \(\left\vert a,b\right\rangle\) and \(\left\vert c,d\right\rangle\) are the initial and final two-site states with 4fn-4fn electron configurations described in Table 1 at each site, and \(\left\vert n\right\rangle\) is the intermediate states with 4fn+1-4fn−1 electron configurations; E0 is the energy for the initial and final states, while En is for the intermediate state \(\left\vert n\right\rangle\).