Using ab initio density functional theory, we study the electronic and magnetic properties of the van der Waals chain material OsCl4. In the nonmagnetic state, a strongly anisotropic band structure was observed, in agreement with its anticipated one-dimensional crystal geometry. Based on Wannier functions, we found that the four electrons of the 5d Os atom form a low-spin S = 1 state, with a large crystal field between the dxz/yz and dxy orbitals, corresponding to a strong Jahn–Teller distortion (Q3<0). As a consequence, the magnetic properties are mainly contributed by the dxz/yz states. Furthermore, when a Mott gap develops after the introduction of the Hubbard U and Hund coupling J, we found that the staggered spin order is the most likely magnetic state, namely, spins arranged as (↑-↓-↑-↓) with π wavevector along the chain. In addition, the energy differences between various spin states are small, suggesting a weak magnetic exchange coupling along the chain. Our results provide guidance to experimentalists and theorists working on quasi-one-dimensional osmium halides chain materials.

Over several decades, one-dimensional (1D) systems have attracted considerable attention due to their interesting physical properties.1–3 In these 1D systems, remarkable physical phenomena have been found or predicted, such as high-Tc superconductivity in iron or copper chains and ladders,4–10 exotic insulating ferromagnetism in a 1D iron oxychalcogenide,11 ferroelectricity triggered by phonon modes or spin ordering,12–16 charge density wave or spin density wave states due to the partial or complete condensation of excitations of free carriers,2,3,17,18 spin block states in the orbital-selective Mott regime of iron 1D ladders or chains,15,19–21 and several others. In general, the spin–orbit coupling (SOC) λ is considered to be negligible for 1D systems with 3d transition-metal (TM) atoms, where their physical properties are primarily induced by electronic correlation couplings (i.e., Hubbard repulsion U and Hund coupling JH).

However, the strength of the SOC parameter λ is enhanced in 4d/5d TM atoms22 leading to comparable values between λ, U, and JH, resulting in several intriguing electronic phases. For example, if the intrahopping t is larger than the typical Hund coupling, an interesting orbital-selective Peierls phase was found in some 1D dimer systems with 4d2 or 5d2 electronic configurations, where the localized band is induced by a Peierls distortion.23,24

Because of a reduced JH, 4d/5d atoms often favor the low-spin configuration in compounds with more than half-filled t2g orbitals. Under a cubic crystal field, the five d orbitals split into the lower t2g and higher eg bands separated by a crystal-field splitting energy (10 Dq). Then, by introducing the SOC effect, the Jeff=1/2 and Jeff=3/2 states separate from each other [see Fig. 1(a)]. However, in these 4d5 or 5d5 systems, the Hubbard repulsion U can dramatically modify the electronic distribution and induce the localization of the spin–orbit coupled pseudospin degree of freedom, leading to half-occupied Jeff=1/2 “spin–orbit Mott” insulating states.25,26 For the d4 electronic configurations, such as Ru4+, Os4+, and Ir5+, all four electrons occupy three degenerate t2g orbitals under a cubic crystal field, leading to a total spin S = 1 and orbital moment L = 1. Returning to the case U = 0, including SOC with λ>0, this d4 system is expected to be a nonmagnetic (NM) insulator of local two-hole J = 0 singlets.27,28 By considering mobile spin–orbital excitons, their condensation may lead to a magnetically ordered state according to theoretical studies.27–32 However, some real materials with d4 electronic configuration have displayed magnetism at low temperature, instead of the nonmagnetic insulator of J = 0 singlets, such as the double perovskite iridates Sr2YIrO633 and Ba2YIrO6.34 One theoretical calculation suggests that band structure effects induce the breakdown of the J = 0 singlet state dominance because the noncubic crystal-field effect is quite small in these compounds.35 

FIG. 1.

(a) Schematic energy splitting of the d4 low-spin configuration with strong SOC under a cubic crystal field. (b) Schematic energy splitting of the d4 low-spin configuration under a Jahn–Teller distortion (Q3<0).

FIG. 1.

(a) Schematic energy splitting of the d4 low-spin configuration with strong SOC under a cubic crystal field. (b) Schematic energy splitting of the d4 low-spin configuration under a Jahn–Teller distortion (Q3<0).

Close modal

The Jahn–Teller (JT) Q3 distortion is a common phenomenon in real materials, which would also cause the energy splitting of t2g orbitals. As displayed in Fig. 1(b), the JT distortion could induce the splitting of the three degenerate t2g orbitals into a two upper degenerate dxz and dyz states, separated from dxy by the crystal-field splitting energy (Δ) when Q3<0. If λ is large enough, the strong SOC could fully suppress the JT distortion, leading again to a stable J = 0 singlet in the d4 configuration system.36 Previous studies have mainly focused on three-dimensional or layered materials for the 4d/5d systems with the d4 electronic configurations. Due to the reduced dimensional phase space, the JT Q3 distortion is expected to be large in some 1D or quasi-1D vdW systems.12,24 Hence, this naturally introduces a simple question: is there any real 1D material with d4 configurations having strong JT distortion and SOC effects?

In this Letter, we use first-principles density functional theory (DFT) method with the generalized gradient approximation (GGA)37–40 to investigate the 1D vdW chain system OsCl4. Our results are summarized as follows. First, the ab initio DFT calculations indicate a strongly anisotropic electronic band structure for OsCl4, in agreement with its anticipated 1D crystal geometry. Based on the Wannier functions resulting from first-principles calculations, we obtained the relevant hopping amplitudes and the crystal-field splitting energy of the t2g orbitals for the Os atoms. Because of the large crystal-field splitting energy between dxz/yz and dxy, OsCl4 with the d4 configuration is a spin-1 S = 1 system, instead of a J = 0 singlet ground state. Furthermore, the superexchange Hubbard interaction is dominant, leading to a Mott antiferromagnetic (AFM) state in the chain direction. In addition, our DFT calculations suggest that the staggered spin order with wavevector π (↑-↓-↑-↓) is the most likely magnetic ground state along the chain.

As shown in Fig. 2(a), OsCl4 has an orthorhombic crystal structure with space group Cmmm (No. 65),41 where the OsCl6 octahedra form edge-sharing vdW 1D chains along the c-axis. Under ambient conditions, based on experiments,41 the nearest neighbor (NN) Os–Os distances (dOsOs=3.56 Å) are identical along each chain. In those chains, the system has four identical Os–Cl bonds (2.378 Å) in the bc plane and two short Os–Cl bonds (2.261 Å) along the a axis, resulting in a Jahn–Teller distortion Q3<0. In this case, the three degenerate t2g orbitals split into two higher energy levels (dxz and dyz) and one lower energy level (dxy). Considering the location of Os in the periodic table, the SOC effect should be large in OsCl4. Due to the compression of the OsCl6 octahedra, this system forms an effective S = 1 low-energy state if the crystal-field splitting Δ between dxz/yz and dxy is sufficiently large.

FIG. 2.

(a) Schematic crystal structure of the conventional cell of OsCl4 drawn by VESTA42 with the convention: blue = Os and green = Cl. (b) The crystal structure showing the Os–Cl–Os chain direction. The angles of the Cl–Os–Cl and Os–Cl–Os bonds are 83.08° and 96.92°, respectively. (c) DOS near the Fermi level based on the nonmagnetic states for OsCl4. Gray: total; red: Os; cyan: Cl. (d) Projected band structures of OsCl4 in the nonmagnetic state without SOC. The dxy orbital is on the bc plane, with the x- or y-axis along the Os–Cl directions, while the z-axis is along the a-axis. The Fermi level is shown with dashed horizontal lines. Each osmium orbital is represented by lines of different colors. The coordinates of the high-symmetry points in the bulk Brillouin zone (BZ) are Γ = (0, 0, 0), X = (0.5, 0, 0), S = (0.5, 0.5, 0), Y = (0, 0.5, 0), Z = (0, 0, 0.5), U = (0.5, 0, 0.5), R = (0.5, 0.5, 0.5), and T = (0, 0.5, 0.5).

FIG. 2.

(a) Schematic crystal structure of the conventional cell of OsCl4 drawn by VESTA42 with the convention: blue = Os and green = Cl. (b) The crystal structure showing the Os–Cl–Os chain direction. The angles of the Cl–Os–Cl and Os–Cl–Os bonds are 83.08° and 96.92°, respectively. (c) DOS near the Fermi level based on the nonmagnetic states for OsCl4. Gray: total; red: Os; cyan: Cl. (d) Projected band structures of OsCl4 in the nonmagnetic state without SOC. The dxy orbital is on the bc plane, with the x- or y-axis along the Os–Cl directions, while the z-axis is along the a-axis. The Fermi level is shown with dashed horizontal lines. Each osmium orbital is represented by lines of different colors. The coordinates of the high-symmetry points in the bulk Brillouin zone (BZ) are Γ = (0, 0, 0), X = (0.5, 0, 0), S = (0.5, 0.5, 0), Y = (0, 0.5, 0), Z = (0, 0, 0.5), U = (0.5, 0, 0.5), R = (0.5, 0.5, 0.5), and T = (0, 0.5, 0.5).

Close modal

Let us now discuss the electronic structure of OsCl4 in the nonmagnetic (NM) phase without SOC. According to the calculated density of states (DOS), the states near the Fermi level are mainly contributed by the Os-5dt2g orbitals, hybridized with some Cl-3p orbitals [see Fig. 2(c)]. Most states related to the Cl-3p orbitals are away from the Fermi level, located in the energy range from −6 to −1.5 eV. In this case, the charge transfer gap Δ = εdεp is large, indicating that OsCl4 is a Mott–Hubbard system rather than a charge-transfer system.

As shown in Fig. 2(d), OsCl4 displays a strongly quasi-1D electronic behavior near the Fermi level along the c-axis corresponding to its dominant chain geometry, where the band structure is much more dispersive along that axis (Γ-Z path) than other directions. Based on the projected band structures of Os's 5d orbitals, the eg orbitals (dx2y2 and d3z2r2) are located at high-energy bands above the Fermi level and, thus, unoccupied, in agreement with the analysis of the low-spin configuration with Os4+ valence and d4 configuration. Furthermore, two 5d electrons of Os occupy the dxy state with flatband characteristics except the Γ–Z path because the dxy orbital is lying in the bc plane. The other two 5d electrons of Os occupy the dxz and dyz orbitals. Consequently, this system can be accurately regarded as having four electrons per site on three t2g orbitals. Based on the band structure information, the magnetic properties of OsCl4 are contributed by the two electrons on the dxz and dyz orbitals.

Figure 3(a) shows that the bands begin to split after introducing the SOC to OsCl4, opening an energy gap (0.16 eV) above the Fermi level at the Γ point for the dxz/yz bands. In addition, we also considered the electronic correlations (U = 2 eV and J = 0.4 eV) on the Os sites, in a screened Hartree–Fock-like manner, as in the local density approximation (LDA) + U method with Liechtenstein format within the double-counting item.43 The lower-energy bands of the fully occupied dxy orbital begin to shift away of the Fermi level under the influence of the electronic correlations. To better understand the low-energy bands of OsCl4, we constructed the disentangled Wannier function of those low-energy bands in the NM phase without SOC, based on the maximally localized Wannier method.44 As shown in Figs. 3(b)–3(d), those orbitals display clearly dxz, dyz, and dxy characteristics. Based on the Wannier function basis (dxz, dyz, dxy), here referred to as γ = (0, 1, 2), respectively, we also obtained the NN hopping matrix for the NN two Os sites in the chain direction,

tγγ=[0.0640.0290.0000.0290.0700.0000.0000.0000.134].
(1)
FIG. 3.

(a) Band structure of OsCl4 in the nonmagnetic state with SOC and with SOC + U + J (U = 2 eV, J = 0.4 eV). The Fermi level is the horizontal dashed line. (b)–(d) Wannier function of the three Os t2g orbitals, with lobes of opposite signs colored as blue and yellow: (b) dxz, (c) dyz, (d) dxy. Os and Cl atoms are in blue and green, respectively. (e) AFM superexchange path for two NN sites.

FIG. 3.

(a) Band structure of OsCl4 in the nonmagnetic state with SOC and with SOC + U + J (U = 2 eV, J = 0.4 eV). The Fermi level is the horizontal dashed line. (b)–(d) Wannier function of the three Os t2g orbitals, with lobes of opposite signs colored as blue and yellow: (b) dxz, (c) dyz, (d) dxy. Os and Cl atoms are in blue and green, respectively. (e) AFM superexchange path for two NN sites.

Close modal

All the hopping matrix elements are in eV units. The crystal-field splitting Δ between dxz/yz and dxy orbitals is about −0.721 eV, caused by the JT distortion Q3. (Additional Wannier results can be found in the supplementary material.) Note that the angle formed by Os–Cl–Os is not 90°, leading to the local y-axis to be not strictly pointing along the Os–Cl bond, resulting in a small difference of hopping values between the dxz and dyz orbitals.

Because Δ is larger than JH and λ, this system would form stable S = 1 states with one fully occupied dxy orbital and two half-occupied dxz/yz orbitals, instead of the expected J = 0 singlet state in the 5d4 configuration. Based on our previous second-order perturbation theory analysis and density matrix renormalization group (DMRG) calculations, large entanglements between doubly occupied and half-filled orbitals can play a key role in stabilizing FM order in a 1D model (three electrons in four orbitals).11 However, the hopping between the doubly occupied dxy and the half-filled dxz/yz orbitals is forbidden in OsCl4. Then, the magnetic properties of this system are mainly decided by the dxz/yz orbitals. Due to the Pauli principle, the two NN sites favor AFM exchange coupling in the chain direction, as shown in Fig. 3(e). Hence, the superexchange Hubbard interaction is dominant, leading to a robust staggered AFM state. In this case, intuitively, the most likely magnetic ordering is AFM, i.e., (↑-↓-↑-↓). Furthermore, considering the values of the hopping amplitudes of the dxz/yz orbitals, this system should be easily localized after introducing the Hubbard interaction U.

To confirm our intuitive analysis, we used the LDA + U method43 to compare different spin configurations along the chain of OsCl4 by changing the onsite Coulomb interaction U and onsite exchange interaction J. Here, we considered AFM1 (↑-↓), AFM2 (↑-↑-↓-↓), AFM3 (↑-↑-↑-↓-↓-↓), AFM4 (↑-↑-↑-↑-↓-↓-↓-↓), FM, and NM configurations along the chain direction, while the magnetic coupling between NN chain is regarded as FM. (Detail can be found in the supplementary material.)

Based on the experimental crystal structure,41 using periodic boundary conditions, with DFT, we constructed the phase diagram of the 1D chain present in OsCl4 with electronic density n = 4. As shown in Fig. 4(a), there is only one AFM magnetic state (specifically, AFM1) stable in our DFT phase diagram, supporting our intuitive analysis. Note that we compared the energies of the several different magnetic configurations mentioned before, with or without SOC, and obtained the same results. This DFT phase diagram is in agreement with our previous DMRG study on the 1D chain model with three electrons in four orbitals, where the diagonal intraorbital hopping is dominant.16 Furthermore, we also observed a metal–insulator phase transition at U = 1 eV. Because the hopping terms are not too large, the effects of the Hubbard U occur at this relatively small value. Note that the location of the boundary between the metal and the insulator should be considered only as a crude approximation. However, the NM–AFM1 and metallic–insulator phase transitions are clearly established, by considering the physical effects of hoppings and electronic correlations.

FIG. 4.

(a) Phase diagram based on the experimental lattice constants of OsCl4, employing the LDA + U technique with the electronic density n = 4. Small solid circles indicate the specific values that were investigated with DFT calculations. (b) and (c) Projected band structures of OsCl4 for the AFM1 state with SOC for different inter-chain couplings: (b) AFM1 (FM coupling between chains) and (c) AFM1–AFM (AFM coupling between chains). The Fermi level is the horizontal dashed line.

FIG. 4.

(a) Phase diagram based on the experimental lattice constants of OsCl4, employing the LDA + U technique with the electronic density n = 4. Small solid circles indicate the specific values that were investigated with DFT calculations. (b) and (c) Projected band structures of OsCl4 for the AFM1 state with SOC for different inter-chain couplings: (b) AFM1 (FM coupling between chains) and (c) AFM1–AFM (AFM coupling between chains). The Fermi level is the horizontal dashed line.

Close modal

Next, we also relaxed the crystal structures for the spin configurations mentioned before based on the LDA + U method with Liechtenstein format.43 Both the lattice constants and atomic positions were fully relaxed with vdW interactions within zero damping vdW-D3.45 Note that the vdW interactions mainly affect the lattice constants between chains but do not change the results of magnetic properties. Based on previous experimental and theoretical studies for 5d Ir or Os compounds,46–48U is in the range 1–3 eV and J in the range 0.3–0.5 eV. Consequently, here, we used the averages U = 2 eV and J = 0.4 eV as the appropriate values for the Os atoms.22,48

Then, we compared the energies of those different spin configurations based on the optimized structures with or without SOC effect, as listed in Table I. The AFM1 magnetic order always has the lowest energy among all tested candidates, with or without SOC. Furthermore, all the energy differences between different spin states are small, suggesting a weak magnetic exchange coupling along the chain direction, in agreement with our previous intuitive analysis from the results of Wannier functions. Based on the difference of energy between the AFM1 and FM states, we obtained that the NN Heisenberg magnetic exchange couplings are 9.85 and 5.85 meV for the GGA + U and GGA + U + SOC calculations, respectively, suggesting that the magnetic transition temperature must be low. For the AFM1 phase without SOC, the calculated local spin magnetic moment of Os is about 1.3 μB/Os, corresponding to the low-spin S = 1 configuration in Os4+. Upon turning on the SOC, the spin quantization axis pointed along the x-axis but with only a tiny difference in energy with respect to the y-axis, indicating the spin favors lying in the xy plane, corresponding to the bc crystal plane. The calculated orbital moment is about 0.1 μB/Os. As shown in Fig. 4(b), we also calculated the band structure of the AFM1 phase of OsCl4 with GGA + U + SOC (U = 2 eV, J = 0.4 eV). The half-occupied dxz/yz orbitals display Mott-insulating behavior with a gap 0.7 eV, while the dxy orbital has fully occupied behavior.

TABLE I.

The optimized lattice constants (Å), energy differences (meV/Os) with respect to the AFM1 configuration taken as the reference of energy, and magnetic moment (μB/Os), for the various magnetic configurations here used. The experimental values (Exp. for short) are also listed for comparison. E(GGA + U) and E(GGA + U + SOC) indicate the calculated energies for different magnetic configurations without or with the SOC effect, respectively. The spin and orbital moments are distinguished by the symbols MS and ML. Note that there are two/three/four Os atoms along the c-axis in their minimum unit cell for the AFM2, AFM3, and AFM4 states, respectively, leading to the different lattice constants of c in those spin configurations.

AFM1AFM2AFM3AFM4FMNMExp.41 
a 8.072 8.072 8.065 8.066 8.066 8.125 7.929 
b 8.391 8.428 8.441 8.442 8.449 8.055 8.326 
c 3.600 7.207 10.813 14.421 3.614 3.558 3.560 
E (GGA + U1.296 1.310 1.320 1.300 1.342 ⋯ 
E (GGA + U + SOC) 0.0 2.5 4.1 5.9 11.7 259.5 ⋯ 
MS (GGA + U + SOC) 1.267 1.278 1.285 1.291 1.300 ⋯ 
ML (GGA + U + SOC) 0.099 0.101 0.102 0.102 0.103 ⋯ 
AFM1AFM2AFM3AFM4FMNMExp.41 
a 8.072 8.072 8.065 8.066 8.066 8.125 7.929 
b 8.391 8.428 8.441 8.442 8.449 8.055 8.326 
c 3.600 7.207 10.813 14.421 3.614 3.558 3.560 
E (GGA + U1.296 1.310 1.320 1.300 1.342 ⋯ 
E (GGA + U + SOC) 0.0 2.5 4.1 5.9 11.7 259.5 ⋯ 
MS (GGA + U + SOC) 1.267 1.278 1.285 1.291 1.300 ⋯ 
ML (GGA + U + SOC) 0.099 0.101 0.102 0.102 0.103 ⋯ 

Finally, we calculated the AFM coupling between chains for the AFM1 configuration. We obtained the optimized crystal lattices (a = 8.070 Å, b = 8.428 Å, and c = 3.601 Å) for the case of AFM interchain order, which are close to the values for the AFM1 with FM interchain order. However, the energy of the AFM1-AFM state is lower by about 1 meV than the AFM1 state, suggesting that the coupling between chains is AFM. The calculated local spin magnetic moments of Os are about 1.31 μB/Os and 1.28 μB/Os for the GGA + U and GGA + U + SOC calculations. In addition, we also calculated the band structure of the AFM1-AFM phase with GGA + U + SOC, which is similar to the band structure of AFM1. More results can be found in the supplementary material.

In summary, we systematically studied the vdW 1D chain compound OsCl4 by using first-principles DFT calculations. A strongly anisotropic 1D electronic band structure near the Fermi surface was observed in the NM state, in agreement with its dominant chain geometry. Based on the Wannier functions calculated from DFT, we obtained the NN hopping amplitudes and on-site energies for the Os atoms. A spin S = 1 for the d4 electronic configuration was observed due to the large crystal-field splitting energy of the t2g orbitals. As a consequence, intuitively, the AFM1 state is expected to be the most likely ground state along the 1D chain direction (↑-↓-↑-↓). This staggered AFM order with π vector was found in our study in a robust portion of the DFT phase diagram at many values of U and J/U, based on the experimental crystal structure of OsCl4. In addition, we also relaxed the crystal structures with the different magnetic configurations using LDA + U (U = 2 eV and J = 0.4 eV). The AFM1 magnetic order with a Mott gap always has the lowest energy among all tested candidates, with or without SOC. The small energy differences between different magnetic configurations suggest a small Heisenberg magnetic exchange coupling along the chain direction and a low magnetic transition temperature. We believe our results could encourage additional detailed experimental studies of quasi-1D osmium halide chain materials. With the help of crystal growers, measurements of the temperature dependence of the susceptibility as well as neutron scattering experiments could confirm our predictions.

See the supplementary material for additional theoretical results corresponding to OsCl4, including calculation method, Wannier fitting, bandgaps, magnetic moments, and electronic structures.

The work of Y.Z., L.-F.L., A.M., and E.D. was supported by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences (BES), Materials Sciences, and Engineering Division.

The authors have no conflicts to disclose.

The data that support the findings reported in this study are available from the corresponding author upon request.

1.
E.
Dagotto
,
Rev. Mod. Phys.
66
,
763
(
1994
).
2.
G.
Grüner
,
Density Waves in Solids
(
Perseus
,
Cambridge, MA
,
2000
).
3.
P.
Monceau
,
Adv. Phys.
61
,
325
(
2012
).
4.
H.
Takahashi
,
A.
Sugimoto
,
Y.
Nambu
,
T.
Yamauchi
,
Y.
Hirata
,
T.
Kawakami
,
M.
Avdeev
,
K.
Matsubayashi
,
F.
Du
,
C.
Kawashima
,
H.
Soeda
,
S.
Nakano
,
Y.
Uwatoko
,
Y.
Ueda
,
T. J.
Sato
, and
K.
Ohgushi
,
Nat. Mater.
14
,
1008
(
2015
).
5.
Y.
Zhang
,
L. F.
Lin
,
J. J.
Zhang
,
E.
Dagotto
, and
S.
Dong
,
Phys. Rev. B
95
,
115154
(
2017
).
6.
J.-J.
Ying
,
H. C.
Lei
,
C.
Petrovic
,
Y.-M.
Xiao
, and
V.-V.
Struzhkin
,
Phys. Rev. B
95
,
241109(R)
(
2017
).
7.
Y.
Zhang
,
L. F.
Lin
,
J. J.
Zhang
,
E.
Dagotto
, and
S.
Dong
,
Phys. Rev. B
97
,
045119
(
2018
).
8.
Y.
Zhang
,
L. F.
Lin
,
A.
Moreo
,
S.
Dong
, and
E.
Dagotto
,
Phys. Rev. B
100
,
184419
(
2019
).
9.
E.
Dagotto
and
T. M.
Rice
,
Science
271
,
618
(
1996
).
10.
M.
Uehara
,
T.
Nagata
,
J.
Akimitsu
,
H.
Takahashi
,
N.
Mori
, and
K.
Kinoshita
,
J. Phys. Soc. Jpn.
65
,
2764
(
1996
).
11.
L. F.
Lin
,
Y.
Zhang
,
G.
Alvarez
,
A.
Moreo
, and
E.
Dagotto
,
Phys. Rev. Lett.
127
,
077204
(
2021
).
12.
L. F.
Lin
,
Y.
Zhang
,
A.
Moreo
,
E.
Dagotto
, and
S.
Dong
,
Phys. Rev. Mater.
3
,
111401(R)
(
2019
).
13.
L. F.
Lin
,
Y.
Zhang
,
A.
Moreo
,
E.
Dagotto
, and
S.
Dong
,
Phys. Rev. Lett.
123
,
067601
(
2019
).
14.
C.
Yang
,
M.
Chen
,
S.
Li
,
X.
Zhang
,
C.
Hua
,
C.
Xiao
,
S. A.
Yang
,
P.
He
,
Z.-A.
Xu
, and
Y.
Lu
,
ACS Appl. Mater. Interfaces
13
,
13517
(
2021
).
15.
Y.
Zhang
,
L. F.
Lin
,
A.
Moreo
,
S.
Dong
, and
E.
Dagotto
,
Phys. Rev. B
101
,
144417
(
2020
).
16.
Y.
Zhang
,
L. F.
Lin
,
A.
Moreo
,
G.
Alvarez
, and
E.
Dagotto
,
Phys. Rev. B
103
,
L121114
(
2021
).
17.
J.
Gooth
,
B.
Bradlyn
,
S.
Honnali
,
C.
Schindler
,
N.
Kumar
,
J.
Noky
,
Y.
Qi
,
C.
Shekhar
,
Y.
Sun
,
Z.
Wang
,
B. A.
Bernevig
, and
C.
Felser
,
Nature
575
,
315
(
2019
).
18.
Y.
Zhang
,
L.-F.
Lin
,
A.
Moreo
,
S.
Dong
, and
E.
Dagotto
,
Phys. Rev. B
101
,
174106
(
2020
).
19.
J.
Herbrych
,
J.
Heverhagen
,
N.
Patel
,
G.
Alvarez
,
M.
Daghofer
,
A.
Moreo
, and
E.
Dagotto
,
Phys. Rev. Lett.
123
,
027203
(
2019
).
20.
J.
Herbrych
,
J.
Heverhagen
,
N.
Patel
,
G.
Alvarez
,
M.
Daghofer
,
A.
Moreo
, and
E.
Dagotto
,
Proc. Natl. Acad. Sci. U. S. A.
117
,
16226
(
2020
).
21.
Y.
Zhang
,
L.-F.
Lin
,
G.
Alvarez
,
A.
Moreo
, and
E.
Dagotto
,
Phys. Rev. B
104
,
125122
(
2021
).
22.
W.
Witczak-Krempa
,
G.
Chen
,
Y. B.
Kim
, and
L.
Balents
,
Annu. Rev. Condens. Matter Phys.
5
,
57
(
2014
).
23.
S. V.
Streltsov
and
D. I.
Khomskii
,
Phys. Rev. B
89
,
161112(R)
(
2014
).
24.
Y.
Zhang
,
L. F.
Lin
,
A.
Moreo
, and
E.
Dagotto
,
Phys. Rev. B
104
,
L060102
(
2021
).
25.
B. J.
Kim
,
H.
Jin
,
S. J.
Moon
,
J.-Y.
Kim
,
B.-G.
Park
,
C. S.
Leem
,
J.
Yu
,
T. W.
Noh
,
C.
Kim
,
S.-J.
Oh
,
J.-H.
Park
,
V.
Durairaj
,
G.
Cao
, and
E.
Rotenberg
,
Phys. Rev. Lett.
101
,
076402
(
2008
).
26.
C.
Lu
and
J.-M.
Liu
,
Adv. Mater.
32
,
1904508
(
2020
).
27.
G.
Khaliullin
,
Phys. Rev. Lett.
111
,
197201
(
2013
).
28.
O. N.
Meetei
,
W. S.
Cole
,
M.
Randeria
, and
N.
Trivedi
,
Phys. Rev. B
91
,
054412
(
2015
).
29.
C.
Svoboda
,
M.
Randeria
, and
N.
Trivedi
,
Phys. Rev. B
95
,
014409
(
2017
).
30.
N.
Kaushal
,
J.
Herbrych
,
A.
Nocera
,
G.
Alvarez
,
A.
Moreo
,
F. A.
Reboredo
, and
E.
Dagotto
,
Phys. Rev. B
96
,
155111
(
2017
).
31.
A. J.
Kim
,
H. O.
Jeschke
,
P.
Werner
, and
R.
Valentí
,
Phys. Rev. Lett.
118
,
086401
(
2017
).
32.
N.
Kaushal
,
R.
Soni
,
A.
Nocera
,
G.
Alvarez
, and
E.
Dagotto
,
Phys. Rev. B
101
,
245147
(
2020
).
33.
G.
Cao
,
T. F.
Qi
,
L.
Li
,
J.
Terzic
,
S. J.
Yuan
,
L. E.
DeLong
,
G.
Murthy
, and
R. K.
Kaul
,
Phys. Rev. Lett.
112
,
056402
(
2014
).
34.
J.
Terzic
,
H.
Zheng
,
F.
Ye
,
H. D.
Zhao
,
P.
Schlottmann
,
L. E.
De Long
,
S. J.
Yuan
, and
G.
Cao
,
Phys. Rev. B
96
,
064436
(
2017
).
35.
S.
Bhowal
,
S.
Baidya
,
I.
Dasgupta
, and
T. S.
Dasgupta
,
Phys. Rev. B
92
,
121113
(
2015
).
36.
S. V.
Streltsov
and
D. I.
Khomskii
,
Phys. Rev. X
10
,
031043
(
2020
).
37.
G.
Kresse
and
J.
Hafner
,
Phys. Rev. B
47
,
558
(
1993
).
38.
G.
Kresse
and
J.
Furthmüller
,
Phys. Rev. B
54
,
11169
(
1996
).
39.
P. E.
Blöchl
,
Phys. Rev. B
50
,
17953
(
1994
).
40.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
,
Phys. Rev. Lett.
77
,
3865
(
1996
).
41.
F. A.
Cotton
and
C. E.
Rice
,
Inorg. Chem.
16
,
1865
(
1977
).
42.
K.
Momma
and
F.
Izumi
,
J. Appl. Crystallogr.
44
,
1272
(
2011
).
43.
A.
Liechtenstein
,
V.
Anisimov
, and
J.
Zaanen
,
Phys. Rev. B
52
,
R5467
(
1995
).
44.
A. A.
Mostofi
,
J. R.
Yates
,
Y. S.
Lee
,
I.
Souza
,
D.
Vanderbilt
, and
N.
Marzari
,
Comput. Phys. Commun.
178
,
685
(
2008
).
45.
S.
Grimme
,
J.
Antony
,
S.
Ehrlich
, and
S.
Krieg
,
J. Chem. Phys.
132
,
154104
(
2010
).
46.
B.
Yuan
,
J. P.
Clancy
,
A. M.
Cook
,
C. M.
Thompson
,
J.
Greedan
,
G.
Cao
,
B. C.
Jeon
,
T. W.
Noh
,
M. H.
Upton
,
D.
Casa
,
T.
Gog
,
A.
Paramekanti
, and
Y.-J.
Kim
,
Phys. Rev. B
95
,
235114
(
2017
).
47.
E.
Şaşioğlu
,
C.
Friedrich
, and
S.
Blügel
,
Phys. Rev. B
83
,
12101(R)
(
2011
).
48.
Y.
Du
,
X.
Wan
,
L.
Sheng
,
J.
Dong
, and
S. Y.
Savrasov
,
Phys. Rev. B
85
,
174424
(
2012
).

Supplementary Material