Introduction

The square-lattice Heisenberg model is the subject of intense numerical and experimental investigations. In spin-1/2 systems—such as cuprates1 and copper deuteroformate tetradeurate (CFTD)2—higher-order exchange interactions are inferred from observations of magnon dispersions along the magnetic zone boundary3,4. While a detailed magnon characterization is useful to understand quantum-fluctuation effects5, exchange incompatibility is typically avoided in these systems. Indeed, the antiferromagnetic (AF) nearest-neighbor (NN) exchange interaction (J1 > 0) and the ferromagnetic next-nearest-neighbor (NNN) interaction (J2 < 0) in these systems stabilize the classical AF Néel order. Instead, magnetic exchange incompatibility requires both J1 > 0 and J2 > 0. This regime of the J1J2 model is the subject of extensive computational investigations for both spin S = 1/26,7,8,9,10 and S = 111,12 systems. In a narrow range near \({J}_{2}/{J}_{1}\) ~ 1/2, magnetic frustration is found to dominate, and exotic quantum phases such as the spin-liquid state13 are predicted. Several calculations show that the Néel order is destroyed there and the ground state has a valence-bond character13,14,15, although its exact nature is still the subject of debate8,16. However, only very few square-lattice systems exhibit substantial magnetic frustration17,18, and even fewer display tunable magnetic interactions19. As a result, approaching the interesting parameter regime in real materials remains an ongoing issue.

In this article, we provide a high-resolution resonant inelastic X-ray scattering (RIXS) study of magnetic excitations in epitaxial thin films of the canonical S = 1 system La2NiO4, grown on different substrates. We discover a marked, upward dispersion along the AF zone boundary \(({\!\,}^{1}{/}_{2},0)\to ({\!\,}^{1}{/}_{4}{,}{\!\,}^{1}{/}_{4})\), which reveals the presence of AF NNN interactions that partly frustrate the NN ones. By employing ab initio calculations, we demonstrate that these results can only be explained by including the multi-orbital nature of 3d8-Ni systems. Furthermore, we observe a correlation between the relative strength of the magnetic interactions and the strain applied to the films. Our results demonstrate that 214-type nickelates are a promising class of materials for the study of the AF square-lattice Heisenberg model. Moreover, the use of thin films provides a clear route to tune the magnetic frustration and explore so far inaccessible regions of the magnetic-phase diagram.

Results

Our thin films of La2NiO4 (LNO) on SrTiO3 (STO), LaAlO3 (LAO), (LaAlO3)0.3(Sr2TaAlO6)0.7 (LSAT), and NdGaO3 (NGO) substrates are characterized by atomic force microscopy, X-ray diffraction, and X-ray absorption spectroscopy (see Fig. 1). The atomic force microscopy images display a step-like morphology indicating an excellent layer-by-layer growth. Diffraction patterns probing the (0, 0, ) reciprocal direction demonstrate good single crystallinity and allow us to extract the c-axis lattice parameters of the films. The epitaxial strain applied by the substrates is supported by the film c-axis and in-plane lattice-parameter dependence (Fig. 1b). X-ray absorption spectra recorded on the LNO/STO and LNO/NGO, shown in Fig. 1c, are consistent with observations on related nickelates20,21,22,23. The Ni L-edge features on the tail of the La M-edge.

Fig. 1: Characterization measurements on thin films of La2NiO4.
figure 1

a X-ray diffraction (at 300 K) probing the (0, 0, ) direction of 12-nm thin films of La2NiO4 on substrates as indicated. The inset represents the atomic force microscopy image showing the step-like morphology of the films (here for the NGO substrate). The color scale corresponds to film thickness. b c-axis lattice parameter versus in-plane epitaxial strain (at 300 K) calculated for La2NiO4 (purple dots) and La2CuO432 (gray dots) films grown on different substrates as a relative change of in-plane parameters in reference to bulk (diamonds) with a = 3.868 Å and c = 12.679 Å for an isomorphic La2NiO4 structure59 and a = 3.803 Å and c = 13.156 Å for La2CuO460. The errorbar correspond to the spread of in-plane parameters measured by X-ray diffraction. c X-ray absorption spectra around the Ni L-edge. The dominant peak corresponds to the La M-edge. b, c Dashed lines are guides to the eye.

The RIXS spectra of La2NiO4 films were measured at the Ni L3 edge (853 eV). These spectra exhibit key RIXS excitations, including high-energy dd-excitations (at ~0.5–3 eV), an elastic scattering contribution at 0 eV, as well as phonon and magnon excitations in between. The dd-excitations have a multi-peak structure, qualitatively similar to other 3d8 systems, such as NiO20,21,22,24. As shown in Fig. 2a, the relative intensities of the peaks are different in the three samples, due to the different crystal-fields acting on the Ni atoms (see Fig. S1, Supplementary Note 1). However, all our La2NiO4 films display the most intense dd-excitation around 1.1 eV and a second less intense excitation just below 1.6 eV. This is consistent with what is reported in bulk La2NiO425 (see Fig. S2, Supplementary Note 1). The subtraction of the elastic peak clearly highlights the presence of multiple low-energy features (see Fig. 2b, c).

Fig. 2: Resonant inelastic X-ray scattering spectra of La2NiO4.
figure 2

a Raw spectra recorded in La2NiO4 films with substrates as indicated. a (inset) Schematics of the photon-in-photon-out resonant inelastic X-ray scattering (RIXS) geometry with horizontally polarized light (π) and azimuthal sample rotation angle ϕ. b, c Low-energy part of the RIXS spectra with momentum transfer and film substrates indicated. The solid red line indicates a three-component fit with phonon, magnon (shaded), and multi-magnon (continuum) contributions. The elastic scattering channel is subtracted in (b, c).

To extract the dispersion of magnetic excitations, we assumed a two-mode model with the addition of a high-energy continuum “background" (Fig. 2b, c). Each of these components is represented by a Gaussian profile. This provides an effective fitting model of excitations for all measured film systems and momenta. Our interpretation of the proposed model is based on the hypothesis that the lower-energy mode (~40 meV) stems from an optical phonon, while the higher-energy mode (strongly dispersing between 60 and 120 meV) is a magnon. This assignment is supported by previous neutron scattering measurements that identified the phonon part via an out-of-plane oxygen buckling mode26,27. The interpretation of the higher-energy mode as a magnon is consistent with earlier RIXS25 and neutron studies28 of bulk La2NiO4. The resulting magnon dispersions are shown in Fig. 3.

Fig. 3: Magnon dispersion of La2NiO4 films.
figure 3

ac Magnon excitation energies (open dots) along high-symmetry directions for La2NiO4 on substrates as indicated. Solid lines represent the same spin-wave model evaluated for different exchange parameters within the confidence intervals of the fitted parameters. The curves corresponding to the best-fit values (marked in bold in the legend) are reported as thicker lines. The middle segment, X → Σ, is part of the antiferromagnetic zone boundary. The error bars are determined from the fitting uncertainty.

Due to lower-energy resolution, the previous RIXS study25 did not resolve any phonon excitations. The unresolved phonon excitation implied that the phonon and magnon spectral weights were merged. This, in turn influences the extraction of the magnon dispersion. Having access, in this work, to a higher-energy resolution, we can distinguish between the nearly momentum-independent phonon mode and the dispersive magnon branch along the three measured high-symmetry directions. In all the film systems explored, the magnon energy reaches its maximum at the AF zone boundary, at the Σ point \(({\!\,}^{1}{/}_{4},{\!\,}^{1}{/}_{4})\), referred to as EΣ, while it displays a local minimum at the X point \(({\!\,}^{1}{/}_{2},0)\), referred to as EX. This evidently anisotropic shape of magnon dispersion was not reported in earlier studies25,28, except for a recent inelastic neutron scattering experiment29. Furthermore, the energy EΣ is different for all three substrates. In particular, it increases as a function of compressive strain, with an enhancement of 18 ± 4 meV (~20%) from LNO/STO to LNO/LAO.

Discussion

By resolving both the phonon and magnon modes, we find that all samples exhibit a substantial dispersion of magnetic excitations along the AF zone boundary. This directly implies the presence of higher-order effective magnetic exchange interactions. In La2CuO4 and related Mott insulating cuprates, the zone-boundary dispersion has been interpreted in terms of a positive ring-exchange interaction that emerges naturally from a single-orbital Hubbard model30,31,32. There is, however, an important difference between the zone-boundary dispersion of La2CuO4 and La2NiO4: in contrast to La2CuO4, the zone- boundary dispersion of La2NiO4 has its maximum at the AF zone boundary Σ point rather than at the X point. As such, the magnon dispersion of La2NiO4 is (as could be expected) inconsistent with a single-band Hubbard model in the strong coupling limit (where the projection onto a Heisenberg spin Hamiltonian is viable).

As a first step, we parameterize the magnon dispersion of La2NiO4 using a phenomenological spin-wave model that includes effective NN and NNN exchange interactions, respectively, J1 and J2 (Fig. 3), plus an easy-plane anisotropy K, already reported by previous measurements28,29. As a starting point, we employ the AF structure of the bulk La2NiO4 determined by neutron diffraction33,34,35, with the spin direction parallel to the crystallographic a-axis. The model is solved in a linear spin-wave (large-S) limit, and the calculated dispersion is fitted to the measured one (see “Methods”). Fitting the experimental (exp) data yields an effective NN exchange interaction \({J}_{1}^{\exp } \sim 30\) meV consistent with previous neutron and RIXS results25,36. Due to the demonstrated finite zone-boundary dispersion, our spin-wave model fitting also yields a moderate NNN exchange interaction \({J}_{2}^{\exp }\). Importantly, \({J}_{2}^{\exp }\) is positive and enhanced by compressive strain3,4. In what follows, we wish to extract the frustration parameter \({{{{{{{\mathcal{G}}}}}}}}={J}_{2}^{\exp }/{J}_{1}^{\exp }\) with the highest precision. Within our spin-wave model, EX = 4SZc(J1 − 2J2) and EΣ = 4SZc(J1 − J2), where Zc is the quantum renormalization factor for spin-wave energies, which is taken as Zc = 1.0937. This gives \({{{{{{{{\mathcal{G}}}}}}}}}^{-1}=1{+}^{{E}_{\Sigma }}{/}_{({E}_{\Sigma }-{E}_{X})}\). The frustration parameter \({{{{{{{\mathcal{G}}}}}}}}\) is thus derived directly from the experimental data, with high precision (EΣ and EX are extracted with error lower than 5 meV) and plotted as a function of the c lattice parameter in Fig. 4 (see also Table S1, Supplementary Note 2). Due to the Poisson effect, the c lattice parameter undergoes a proportional shrinkage when the in-plane parameters expand. Our X-ray diffraction measurements confirm this relationship (Fig. 1b), indicating that the c-axis lattice parameter can serve as an indirect probe of the in-plane strain. Therefore, our findings demonstrate a nearly linear correlation between magnetic frustrations and epitaxial strain.

Fig. 4: Strain-dependent magnetic frustration.
figure 4

The frustration parameter \({{{{{{{\mathcal{G}}}}}}}}={J}_{2}^{\exp }/{J}_{1}^{\exp }\) derived from the experimental data (squares; left axis) is presented as a function of the c-axis lattice parameter. The error bars for the experimental data are calculated as a propagation of standard deviations extracted from the fits. The results for films are combined with data for bulk La2NiO4 from ref. 29. The experimental frustration \({{{{{{{\mathcal{G}}}}}}}}\) is compared to the ratio \({J}_{2}^{{{{{{{{\rm{cal}}}}}}}}}/{J}_{1}^{{{{{{{{\rm{cal,corr}}}}}}}}}\) derived from the DFT and cRPA calculations (diamonds; right axis). Note that the calculated \({J}_{2}^{{{{{{{{\rm{cal}}}}}}}}}\) only contains a contribution to the full next-nearest-neighbor coupling J2. Therefore, the comparison merely highlights a similar trend of the frustration under in-plane compression. The dashed line is a guide-to-the-eye.

We stress that, for interaction strengths and hop**s that are realistic for cuprates and nickelates, J2 > 0 is hard to reconcile with a single-band Hubbard model. A positive J2 implies an effective AF NNN exchange interaction, at odds with what is observed in cuprates4,30 and d9 infinite-layer nickelates38,39. Both systems have indeed been successfully described using a single d-orbital framework3,4,40,41. Therefore, we argue that the magnon zone-boundary dispersion in LNO signals physics beyond the single-orbital Hubbard model. We propose that the multi-orbital (\({d}_{{x}^{2}-{y}^{2}}\), \({d}_{{z}^{2}}\)) nature of nickelates42,43,44 must be explicitly considered. Already in La2CuO4, due to the short apical oxygen distance, a small but significant orbital hybridization between \({d}_{{z}^{2}}\) and \({d}_{{x}^{2}-{y}^{2}}\) has been reported45. In La2NiO4 the apical oxygen distance is even shorter, as exemplified by the reduced c lattice parameter (see Fig. 1b), and hence an even more pronounced hybridization is expected.

To rationalize the trend in the exchange interactions obtained from our spin-wave fits, we derive a two-orbital low-energy model for La2NiO4 on different substrates from first principles (see the Method section). For the Ni \({d}_{{z}^{2}}\) and \({d}_{{x}^{2}-{y}^{2}}\) orbitals (labeled α and β), we compute the (next) nearest-neighbor hop** parameters \({t}^{({\prime} )}\), the crystal-field splitting Δeg, local Coulomb (Hubbard) interaction U and Hund’s exchange JH using experimental lattice constants from Table 1. Noteworthy46, the hop** parameters and Coulomb interactions, listed in Table 1, hardly change under varying in-plane compression. This is different from calculations for the cuprate family, see ref. 32, and agrees with our experiments, which show substantially smaller changes in the magnon spectrum than for the cuprates. What is most affected by strain in Table 1 is the crystal-field splitting Δeg by which the Ni \({d}_{{x}^{2}-{y}^{2}}\) orbital is higher in energy than the \({d}_{{z}^{2}}\) orbital. When going from the STO to the LAO substrate, in-plane strain pushes the \({d}_{{x}^{2}-{y}^{2}}\) orbital further up in energy, as it is pointing towards the now closer in-plane oxygen sites that are charged negatively.

Table 1 Parameters of the two-orbital Hubbard model

This crystal-field splitting Δeg enters the calculated (cal) two-orbital superexchange as follows:

$${J}_{1}^{{{{{{{{\rm{cal}}}}}}}}}=\frac{{t}_{\alpha \beta }^{2}}{U+{J}_{H}-{\Delta }_{eg}}+\frac{{t}_{\alpha \beta }^{2}}{U+{J}_{H}+{\Delta }_{eg}}+\frac{{t}_{\alpha \alpha }^{2}+{t}_{\beta \beta }^{2}}{U+{J}_{H}},$$
(1)

where we extend the formula of ref. 47 to finite Δeg (see Supplementary Note 3). As Δeg appears once with a plus and once with a minus sign in the denominator, the crystal-field splitting enters \({J}_{1}^{{{{{{{{\rm{cal}}}}}}}}}\) in a higher-than-linear order.

The magnetic exchange couplings \({J}_{1}^{{{{{{{{\rm{cal}}}}}}}}}\) determined by Eq. (1) are displayed in Table 1. They show the same qualitative tendency as in our experiment, i.e., an increase of both \({J}_{1}^{\exp }\) (see Fig. 3) and \({J}_{1}^{{{{{{{{\rm{cal}}}}}}}}}\) with compressive strain. Quantitatively, the ab initio calculated \({J}_{1}^{{{{{{{{\rm{cal}}}}}}}}}\) is however too large. This has two major origins: (i) The cRPA interactions are, here, taken at zero frequency U = U(ω = 0). Additional renormalizations from the frequency dependence U(ω)48 are often mimicked through an empirical enhancement of U. Increasing U, we could easily obtain quantitative agreement with the experimental J1, but at the cost of a free fit parameter and most likely only an accidental agreement. (ii) Eq. (1) only includes terms to second-order perturbation theory in t. For the one-band Hubbard model, higher-order processes have been calculated and yield a correction from J1 = 4t2/U to a reduced J1 = 4(t2/U − 16t4/U3)40,49. Higher-order terms are expected to reduce J1 also in the two-orbital setting. Here, we estimate these corrections by the one-orbital prescription t2/U ⟶ t2/U − 16t4/U3. Then, e.g., for the NGO substrate, the leading contribution to \({J}_{1}^{{{{{{{{\rm{cal}}}}}}}}}\), i.e., \({t}_{\beta \beta }^{2}/({U}_{{{{{{{{\rm{eff}}}}}}}}})\) with Ueff = U + JH, reduces from 47 to 37 meV. Applying this substitution to Eq. (1) yields the corrected exchange couplings \({J}_{1}^{{{{{{{{\rm{cal,corr}}}}}}}}}\) listed in Table 1, which are in better agreement with the measured values (see Fig. 3).

Table 1 also lists the next-nearest-neighbor exchange \({J}_{2}^{{{{{{{{\rm{cal}}}}}}}}}\) that can be obtained with the same second-order formula Eq. (1), except now using the hop**s \({t}^{{\prime} }\) instead of t (see Supplementary Note 3 for details). Crucially, our calculations predict a positive J2, in agreement with the experiment. Moreover, \({J}_{2}^{{{{{{{{\rm{cal}}}}}}}}}\) shows the same qualitative tendency as in the experiment as a function of epitaxial strain. On a quantitative level, the calculated values are, however, a factor ~3–5 lower than the experimental results. The reason is that contributions to J2 from higher-order exchange processes, of order t4/U3 and \({t}^{{\prime} }{t}^{2}/{U}^{2}\), become (relatively) more important for \({J}_{2}^{{{{{{{{\rm{cal}}}}}}}}}\), as the second-order terms are now based on the much smaller \({t}^{{\prime} }\).

The key difference between the multi-orbital case of LNO and the one-orbital cuprates is the larger U (UcRPA ≈ 3.1 eV for two-orbitals while UcRPA ≈ 1.9 eV for a one Ni \({d}_{{x}^{2}-{y}^{2}}\) orbital setup) and the additional Hund’s JH ≈ 0.5 eV in the denominator of Eq. (1). As a consequence, the balance for the NNN exchange coupling shifts from a ferromagnetic ring exchange \({J}_{2} \sim -\!\!{t}^{4}/{U}_{eff}^{3} \, < \, 0\), that overpowers the AF second-order exchange \({J}_{2} \sim {{t}^{{\prime} }}^{2}/{U}_{eff} \, > \, 0\) in the one-orbital cuprates, toward dominance of the latter in the multi-orbital LNO. This change in hierarchy explains the main qualitative differences in magnon dispersion between LNO and cuprates: the opposite sign of the effective J2. For LNO, with a positive J2, the zone-boundary dispersion shows a notable minimum at \(({\!\,}^{1}{/}_{2},0)\), see Fig. 3, whereas a maximum occurs for the negative J2 in cuprates.

Conclusions

The ab initio calculations indicate that the magnetic frustration in La2NiO4 is caused by the multi-orbital nature of 3d8 nickelates. More importantly, our results demonstrate that the degree of frustration is amplified by compressive strain (see Fig. 4), with a pivotal role played by the crystal-field splitting. Indeed, with the substrates used, the magnetic frustration increases four-fold with respect to the bulk, bridging half the way toward the exotic realm anticipated for J2/J1 ~ 1/2. Thus, our study suggests an effective tool for tuning antiferromagnetic interactions within square-lattice systems. We speculate that the approach is applicable beyond La2NiO4 and may offer an experimental route to reach so far unexplored regions of the magnetic-phase diagram, potentially allowing to investigate exotic states induced by magnetic frustration.

Methods

Film growth and characterization

Thin films of La2NiO4 were grown by RHEED-equipped Radio-frequency off-axis magnetron sputtering50 on (001) STO, (001) LAO, (001) LSAT and (110) NGO substrates. These films were grown in an argon atmosphere at 700 °C. Their qualities were confirmed by atomic force microscopy and x-ray diffraction. Their insulating character was confirmed by resistivity measurements of the LNO/STO film (see Fig. S3, Supplementary Note 4).

RIXS experiments

Ni L-edge RIXS experiments for STO, LAO and NGO substrates were carried out at the I21 beamline51 at the DIAMOND Light Source. All spectra were collected in the grazing exit geometry using linear horizontal polarized incident light with the scattering angle fixed to 2θ = 154°. The energy resolution was estimated from the elastic scattering on amorphous carbon tape and was between 37 and 41 meV (full-width-at-half-maximum, FWHM). All films were measured at base temperature T = 16 K. We define the reciprocal space (qx, qy, qz) in reciprocal lattice units (h, k, ) = (qxa/2π, qyb/2π, qzc/2π) where a, b, and c are the pseudo-tetragonal lattice parameters. RIXS spectra were acquired along three in-plane paths: (0, 0) → (0, 1/2), \((0,0)\to ({\!\,}^{1}{/}_{4},{\!\,}^{1}{/}_{4})\) and \((0,{\!\,}^{1}{/}_{2})\to ({\!\,}^{1}{/}_{4},{\!\,}^{1}{/}_{4})\). Extraction of low-energy excitations around (0, 0) is limited by energy resolution. Due to kinematic constraints Γ points at higher zones cannot be reached, as well. RIXS intensities are normalized to the weight of the dd-excitations52. The data for the LSAT substrate were collected at the ID32 beamline at the European Synchrotron Radiation Facility (ESRF) (see description in Supplementary Note 5, Fig. S4).

Phenomenological spin-wave model

The effective superexchange parameters were extracted from the measured dispersion using a linear spin-wave model. We included effective couplings between the first and second nearest neighbors, plus an easy-plane anisotropy K, with the resulting Hamiltonian:

$${{{{{{{\mathcal{H}}}}}}}}={J}_{1}\mathop{\sum}_{\langle i,j\rangle }{{{{{{{{\bf{S}}}}}}}}}_{i}\cdot {{{{{{{{\bf{S}}}}}}}}}_{j}+{J}_{2}\mathop{\sum}_{\langle \langle i,j\rangle \rangle }{{{{{{{{\bf{S}}}}}}}}}_{i}\cdot {{{{{{{{\bf{S}}}}}}}}}_{j}+K{\sum}_{i}{\left({S}_{i}^{z}\right)}^{2}$$
(2)

where 〈i, j〉 and 〈〈i, j〉〉 denote pairs of first and second nearest neighbors Ni atoms, respectively. The fitting procedure has been carried out using the SpinW package53. As an input we have used the AF structure of the bulk La2NiO4 determined by neutron diffraction33,34,35, with the spin direction parallel to the crystallographic a-axis. The dispersion in the approximation of the linear spin-wave theory is represented by4,29:

$$\hslash \omega = \,{Z}_{c}\sqrt{\left({A}_{{{{{{{{\bf{q}}}}}}}}}^{2}-{B}_{{{{{{{{\bf{q}}}}}}}}}^{2}\right)}\\ {A}_{{{{{{{{\bf{q}}}}}}}}} = \,4S\left[\frac{K}{4}+{J}_{1}-{J}_{2}(1-{\nu }_{h}{\nu }_{k})\right]\\ {B}_{{{{{{{{\bf{q}}}}}}}}} = \,4S\left[{J}_{1}\,\frac{{\nu }_{h}+{\nu }_{k}}{2}-\frac{K}{4}\right]$$
(3)

where \({\nu }_{x}=\cos (2\pi x)\). The quantum renormalization factor for spin-wave velocity is fixed to Zc = 1.09, as usual for S = 1 systems37. The value of the easy-plane anisotropy K mostly controls the size of the magnon gap at the Γ point. Since this value is very hard to obtain from RIXS spectra, we have fixed K = 0.5 meV in agreement with previous inelastic neutron scattering data28. We have also neglected other interactions <10−1 meV, such as easy-axis anisotropy, inter-layer coupling, and Dzyaloshinskii–Moriya interactions29,34.

Ab initio calculations

Electronic structure calculations were performed with density functional theory in the local density approximation using a full-potential linearized muffin-tin orbital (FPLMTO) code54, after the structures were optimized with WIEN2k55 using the PBE functional. We mimicked the influence of the substrates by simulating bulk La2NiO4 using the experimental lattice constants of the thin films. The reference calculation for the bulk uses lattice constants from ref. 29. All calculations assume the space group I4/mmm and are paramagnetic. The resulting band structures are displayed in Fig. S5, Supplementary Note 6. The FPLMTO calculations were converged using 123 reducible k-points and include local orbitals for the Ni-3p and La-5p states. The internal atomic positions were relaxed with WIEN2k using 63 reducible k-points, a cutoff parameter RMTKMAX = 7 and partial waves inside the atomic spheres up to l = 5, until the forces were below 1 mRy per Bohr radius (for details of the relaxed structures, see Supplementary Note 7, Table S2). The tight-binding hop** and crystal-field parameters have been extracted from a projection onto maximally localized Wannier orbitals56,57 of Ni \(3{d}_{{x}^{2}-{y}^{2}}\) and \(3{d}_{{z}^{2}}\) character. Matrix elements of the static (ω = 0) and local screened Coulomb interaction (Hubbard U and Hund’s JH) have been estimated from calculations in the constrained random phase approximation (cRPA)48 for entangled band structures58 in the Wannier basis57 using 6 × 6 × 6 reducible momentum-points in the Brillouin zone. For the two-particle product basis, states are kept up to an angular cutoff of l = 4 and down to an overlap eigenvalue of 10−4.