The authors present a quantity termed charge–spin susceptibility, which measures the charge response to spin degrees of freedom in strongly correlated materials. This quantity is simple to evaluate using both standard density functional theory and many-body electronic structure techniques, enabling comparison between different levels of theory. A benchmark on 28 layered magnetic materials shows that large values of charge–spin susceptibility correlate with unconventional ground states such as disordered magnets and unconventional superconductivity.

The conventional paradigm of condensed matter physics involves the partitioning of electronic ground states into descriptions of spin and electronic degrees of freedom, potentially with small coupling terms between them. Many materials fall under this paradigm, such as antiferromagnetic insulators, ferromagnetic metals, and non-magnetic insulators. However, this weak coupling paradigm is insufficient to describe some materials, which often show unconventional ground states and excitations. Examples of unconventional behavior include the unconventional superconductors,1,2 unconventional magnetic states,3–5 spin liquids,6 and strong magnetodielectric effects.7 It is still an open question how to predict a priori whether a material breaks the weakly coupled spin/electron paradigm, even if such coupling has been previously studied, mostly in the context of model Hamiltonians.1,8–12

One of the most prominent examples of charge–spin interactions are the high-temperature unconventional superconductors such as copper oxides and iron-based pnictides and chalcogenides. As pointed out by Scalapino,1 there is vast experimental evidence such as the proximity between magnetic ordered phases and superconductivity, suggesting that in these materials, the coupling between orbital (charge) and magnetic (spins) degrees of freedom plays a crucial role in determining the physics.

Despite recent progress in the study of model Hamiltonians commonly associated with materials showing unconventional properties,13–18 the study of real materials with such properties remains challenging. There have been several attempts to computationally predict whether specific groups of materials show unconventional electronic phases, for example, unconventional superconductivity.19–21 However, to our knowledge, most of these similarity-based searches have attained limited success. In this manuscript, we adopt a somewhat different approach: we concentrate on exploring a new computational probe of charge–spin coupling in materials. We present and test the charge–spin susceptibility, which is computed from first principles in a simple way. This quantity measures the response of a material’s charge density to changes in its spin density.

We use density functional theory (DFT) and quantum Monte Carlo (QMC) calculations to estimate the charge–spin coupling in a set of 28 layered strongly correlated transition metal compounds including unconventional behavior such as unconventional superconductivity in the cuprates and iron-based superconductors,2,22 disordered magnetic ground states, and bad metallic behavior in materials such as BaCo2As2 and Sr2VO4.4,5,23 In order to assess the quality of the DFT results, for a small set of materials, we compare the DFT-derived charge–spin susceptibility predictions with those obtained from fixed-node diffusion Monte Carlo. We find that while DFT+U predicts a different charge–spin response from the QMC results, it is sufficient to distinguish large responses from small responses. Using DFT+U calculations on the entire set of materials, we find that materials with a large charge–spin susceptibility often present unconventional phases, while materials in the same class with smaller charge–spin susceptibility do not.

In this work, we concentrate on investigating a set of 28 magnetic layered materials containing transition metal atoms with magnetic moments arranged in diverse two dimensional structural motifs. Some of these materials are well known to present unconventional phases of matter as described in Sec. I. In Table I, we list the materials in our test set along with the low-temperature magnetic and electronic phases they show upon chemical doping or pressure.

TABLE I.

Materials in our test set ordered according to their average charge–spin susceptibility, χcs, obtained from DFT+U with U = 5 eV. Each material’s magnetic and electronic phases are shown in the two central columns. The arrows stand for phases achieved by chemical doping or pressure.

MaterialMagnetismCharge transportχcsU=5
BaCo2As224  D4  M4  0.45 
Sr2VO425  D5  I → M, B23  0.42 
T′-La2CuO426  N → S, D27  I → B, uSC27  0.41 
Sr2CoO428  F28  M → I, SC28–31  0.39 
o-BaFe2As232  S → D33,34 M → B, uSC34–36  0.37 
t-BaFe2As237  S → D33,34 M → B, uSC34–36  0.35 
FeTe38  S → D39  M → uSC, B39,40 0.32 
FeS41  D → S42  M → uSC41  0.30 
t-FeSe43  D → S44  uSC → B, M40,45 0.26 
Sr2FeO446  N46  I → M47  0.22 
o-FeSe43  D → S44  uSC → B, M40,45 0.20 
T-La2CuO448  N → S, D2  I → B, PG, uSC2  0.15 
CaCuO249  N → D50  I → uSC50  0.14 
SrCuO251  N → D52  I → uSC53  0.13 
K2CoF454  N55  I56  0.10 
TaS257  D58  I → M, SC59  0.09 
Sr2CrO460  N61  I60  0.06 
BaCr2As224  N62  M62  0.05 
La2NiO463  N → S64  I63  0.04 
BaMn2As265  N66  I → M67  0.03 
NiPSe368  N69  I69  0.03 
La2CoO470  N → S71  I72  0.02 
Sr2MnO473  N74  I74  0.02 
TeCuO375  N7  I7  0.02 
CrGeTe376  F76  I76  0.01 
K2CuF477  F78  I79  0.01 
K2NiF480  N81  I81  0.01 
SeCuO382  F7  I7  0.01 
 
MaterialMagnetismCharge transportχcsU=5
BaCo2As224  D4  M4  0.45 
Sr2VO425  D5  I → M, B23  0.42 
T′-La2CuO426  N → S, D27  I → B, uSC27  0.41 
Sr2CoO428  F28  M → I, SC28–31  0.39 
o-BaFe2As232  S → D33,34 M → B, uSC34–36  0.37 
t-BaFe2As237  S → D33,34 M → B, uSC34–36  0.35 
FeTe38  S → D39  M → uSC, B39,40 0.32 
FeS41  D → S42  M → uSC41  0.30 
t-FeSe43  D → S44  uSC → B, M40,45 0.26 
Sr2FeO446  N46  I → M47  0.22 
o-FeSe43  D → S44  uSC → B, M40,45 0.20 
T-La2CuO448  N → S, D2  I → B, PG, uSC2  0.15 
CaCuO249  N → D50  I → uSC50  0.14 
SrCuO251  N → D52  I → uSC53  0.13 
K2CoF454  N55  I56  0.10 
TaS257  D58  I → M, SC59  0.09 
Sr2CrO460  N61  I60  0.06 
BaCr2As224  N62  M62  0.05 
La2NiO463  N → S64  I63  0.04 
BaMn2As265  N66  I → M67  0.03 
NiPSe368  N69  I69  0.03 
La2CoO470  N → S71  I72  0.02 
Sr2MnO473  N74  I74  0.02 
TeCuO375  N7  I7  0.02 
CrGeTe376  F76  I76  0.01 
K2CuF477  F78  I79  0.01 
K2NiF480  N81  I81  0.01 
SeCuO382  F7  I7  0.01 
 
 M = metal 
N = Néel order B = bad metal 
S = stripe order PG = pseudogap 
F = ferromagnet SC = conventional superconductor 
D = disordered uSC = unconventional superconductor 
 I = insulator 
 M = metal 
N = Néel order B = bad metal 
S = stripe order PG = pseudogap 
F = ferromagnet SC = conventional superconductor 
D = disordered uSC = unconventional superconductor 
 I = insulator 

We shall start by clarifying the relation between the charge–spin susceptibility and the coupling between charge and spin degrees of freedom in model effective Hamiltonians. While our methodology does not depend on any particular effective Hamiltonian being applicable, we will illustrate this relation using a toy model. Consider the effective Hamiltonian,

H=Ho+HS+λHoS,
(1)

where Ho describes the orbital degrees of freedom, while HS describes spin states. The term HoS accounts for interactions between the latter two sets of degrees of freedom, which are controlled by the coupling λ.

In order to compute the relation between the charge–spin coupling λ and the charge–spin susceptibility in a system governed by Eq. (1), consider a small deformation of the electronic wave function away from the ground state. Assume that this deformation amounts to a change in the ground state’s magnetic order, which results in a change in the system’s spin density, Δsi(r) ≡ si(r) − s0(r), where s0(r) stands for the ground state spin density, while si(r) stands for the new/deformed state’s spin density. In such a case, one can show (see the  Appendix) that to first order in the deformation, the resulting change in the charge density, Δρi(r) ≡ ρi(r) − ρ0(r) (with ρ0 and ρi standing for the ground state and deformed state charge density), is proportional to the change in the spin density,

Δρi(r)λwXi(r)Δsi(r).
(2)

In the above expression, λ is the coupling constant connecting the orbital and the spin levels, while w is the energy scale associated with the orbital degrees of freedom, and Xi(r) is a numerical factor related to the type of spin deformation. Thus, the ratio Δρi(r)/Δsi(r) gives direct access to the magnitude of λ/w.

A simple way of estimating the magnitude of the coupling λ/w is to compute the average charge–spin susceptibility, χcs, defined as83 

χcs1Ni=1Nχi,
(3)

where N stands for the number of different magnetic orders considered for each material (see online data84). χi stands for the pairwise charge–spin susceptibility of each magnetic order with respect to the ground state. This is defined as

χiΔρiΔsi,
(4)

where Δρisi) stands for the spatial fluctuations in charge (spin) density relative to the lowest-energy magnetic state. The former are given by

Δρi=drρi(r)ρ0(r),
(5a)
Δsi=drsi(r)s0(r),
(5b)

where ρ0(r) and s0(r) are the charge and spin distributions of the lowest-energy magnetic state.

We calculate χcs as defined in Eqs. (3)–(5), by generating several low-energy magnetic textures for each material. As represented in Fig. 1, in order to obtain a new magnetic order, we optimize initial magnetic textures that differ from the ground state’s one by a few flipped transition metal atoms’ magnetic moments. Then, we compute the charge density and spin density differences between the ground and the new state obtaining Δρi and Δsi from Eq. (5). With several different magnetic orders, we compute the average charge–spin susceptibility χcs from Eqs. (3) and (4).

FIG. 1.

Schematic representation of the methodology used to compute the charge–spin susceptibility. Starting from the ground state magnetic texture of, e.g., FeSe (iron atoms with stripe order), we flip some of the magnetic moments to obtain a different magnetic order (e.g., Néel magnetic order). We optimize that initial texture to obtain a low-energy state with that magnetic order. With these two states, we compute the difference between charge densities, Δρir, and spin densities, Δsi(r), from which we obtain the charge–spin susceptibility of this pair of states, χi = ∫|Δρi(r)|dr/∫|Δsi(r)|dr. In the panels above, yellow, red, and blue identify isodensity surfaces in the ab-plane (of the iron atoms) for the charge density, positive, and negative spin density, respectively. The iso-charge density surfaces (yellow) correspond to an isolevel of 0.15 e/Å3, while all the other isodensity surfaces (red/blue) correspond to an isolevel of 0.01 e/Å3.

FIG. 1.

Schematic representation of the methodology used to compute the charge–spin susceptibility. Starting from the ground state magnetic texture of, e.g., FeSe (iron atoms with stripe order), we flip some of the magnetic moments to obtain a different magnetic order (e.g., Néel magnetic order). We optimize that initial texture to obtain a low-energy state with that magnetic order. With these two states, we compute the difference between charge densities, Δρir, and spin densities, Δsi(r), from which we obtain the charge–spin susceptibility of this pair of states, χi = ∫|Δρi(r)|dr/∫|Δsi(r)|dr. In the panels above, yellow, red, and blue identify isodensity surfaces in the ab-plane (of the iron atoms) for the charge density, positive, and negative spin density, respectively. The iso-charge density surfaces (yellow) correspond to an isolevel of 0.15 e/Å3, while all the other isodensity surfaces (red/blue) correspond to an isolevel of 0.01 e/Å3.

Close modal

Since our objective is to screen a large set of materials against the charge–spin susceptibility, we decided to base our search protocol on a low-cost but sufficiently accurate computational method. With that in mind, we chose Kohn–Sham density functional theory (KS–DFT).85 Most of the calculations presented in this work were performed using the KS–DFT approach,85 as implemented in the QUANTUM ESPRESSO code.86 The exchange–correlation energy was approximated by the generalized gradient approximation (GGA) using the Perdew–Burke–Ernzerhof (PBE) functional.87 To improve the description of the d orbitals, we used the DFT+U scheme of Cococcioni and de Gironcoli.88 Interactions between valence and core electrons were described by pseudopotentials in the accurate set of the Standard Solid-State Pseudopotentials library.89,90 The Kohn–Sham orbitals were expanded in a plane-wave basis with a cutoff energy Ec (Ry), while a cutoff of 4Ec was used for the charge density (see online data for Ec of each material84). The Ec of a given compound was chosen to be the largest Ec among those of its constituent chemical elements. The Ec of an atomic species was estimated from checking for convergence of the single-atom’s total energy against Ec. Convergence was assumed when total energy changed less than 0.01 Ry upon an increase of Ec by 10 Ry. The Brillouin zone (BZ) was sampled using a Γ-centered 6 × 6 × 6 grid following the scheme proposed by Monkhorst–Pack.91 Total energy convergence against the BZ grid density was tested by doing 7 × 7 × 7 grid calculations for unpolarized and ferromagnetic textures. The crystal structure for each material was set up with the information available on the ICSD database92—see online data84 for the CIF(s) used in the calculations of each material. A supercell was used whenever the material unit cell had less than four transition metal atoms per unit cell. This ensures that we can generate sufficient magnetic textures to properly estimate the charge–spin susceptibility.

For each material, we performed multiple DFT+U calculations (with U = 0 eV, 5 eV, 10 eV) in order to assess the uncertainty in the charge–spin susceptibility estimate. With the aim of converging different magnetic orders, we performed calculations in which the self-consistent cycle started from different magnetic states (see Fig. 1), i.e., different orderings and magnitudes for the magnetic moments on the transition metal atoms. In Ref. 84, the reader can find data specifying all the DFT+U calculations that were performed, including material name, crystallographic identifier (CIF), Hubbard U, Ec cutoff, supercell size, k-point mesh, starting magnetic state, final magnetic state, band gap estimate, and total energy.

To check the accuracy of the results obtained from the DFT+U calculations, in Subsection III B, we compare charge–spin susceptibilities obtained from DFT+U with those obtained from the PBE and PBE0 hybrid functionals93 (as implemented in the CRYSTAL17 code94), as well as those obtained from the highly accurate fixed-node diffusion Monte Carlo (DMC)95 (as implemented on the quantum Monte Carlo package QWALK96).

Fixed-node diffusion Monte Carlo (DMC) is a fully first-principles stochastic framework to solve the Schrödinger equation, which yields a variational upper bound to the ground state.95 We employed a Slater–Jastrow trial wavefunction, as implemented in the QWalk package.96 We constructed the Slater determinant with orbitals from DFT calculations using the Crystal code94 employing the PBE0 functional.93 Previous studies have shown that per comparison to other commonly used DFT functionals, PBE0 typically gives the best wave function nodes.97–99 The Brillouin zone (BZ) was also sampled using a Gamma-centered 6 × 6 × 6 Monkhorst–Pack grid.91 We used Dirac–Fock pseudopotentials and ECPs specially constructed for quantum Monte Carlo computations.100,101 We controlled finite-size errors by using 2 × 1 × 1 supercells and averaging over the sampled k-points. We used a time step of 0.005 Ha−1. This setup has been shown to give a good description of challenging materials such as the cuprates102 and FeSe.99 

For each material in Subsection III B, we considered several magnetic textures. Some of these have zero total spin projection Sz = 0 (i.e., antiferromagnetic-like orders), while others have Sz ≠ 0 (ferromagnetic or flip orders). The datasets provided online84 identify the magnetic textures considered for each material. There we have also included figures representing those magnetic textures.84 

In Fig. 2, we plot the change in charge and spin density between the lowest and second-lowest energy magnetic textures for K2NiF4, SrCuO2, FeSe, and BaCo2As2. Stoichiometric K2NiF4 is a typical Mott insulator showing Néel order.81 SrCuO2 is a Néel ordered magnetic insulator, while FeSe is a Hund’s metal without long-range magnetic order at atmospheric pressure. SrCuO2 and FeSe are representative of the families of cuprate and iron-based high-temperature unconventional superconductors, well known to support a wide range of uncommon phases, ranging from strange metallic behavior to non-trivial magnetic states and high-temperature unconventional superconductivity.2,44,45 BaCo2As2 is a disordered magnetic metal.4 In the bottom row of Fig. 2, we show the pairwise charge–spin susceptibility χi [see Eqs. (4) and (5)] resulting from those magnetic textures.

FIG. 2.

DFT+U (5 eV) charge (top row) and spin (bottom row) density differences in the ab-plane of the two lowest-energy magnetic textures for K2NiF4, SrCuO2, FeSe, and BaCo2As2. Atoms are represented by spheres as follows: Ni (yellow), F (cyan), and K (gray) for K2NiF4; Cu (brown), O (red), and Sr (blue) for SrCuO2; Fe (red) and Se (yellow) for FeSe; Co (green), Ba (gray), and As (cyan) for BaCo2As2. The red and blue isodensity surfaces stand for positive and negative values. The charge and spin density difference surfaces in the panels corresponding SrCuO2, FeSe, and BaCo2As2 were drawn with isolevel 0.01 e/Å3, while K2NiF4 is at 0.002 e/Å3 for charge density and 0.005 e/Å3 for spin density.

FIG. 2.

DFT+U (5 eV) charge (top row) and spin (bottom row) density differences in the ab-plane of the two lowest-energy magnetic textures for K2NiF4, SrCuO2, FeSe, and BaCo2As2. Atoms are represented by spheres as follows: Ni (yellow), F (cyan), and K (gray) for K2NiF4; Cu (brown), O (red), and Sr (blue) for SrCuO2; Fe (red) and Se (yellow) for FeSe; Co (green), Ba (gray), and As (cyan) for BaCo2As2. The red and blue isodensity surfaces stand for positive and negative values. The charge and spin density difference surfaces in the panels corresponding SrCuO2, FeSe, and BaCo2As2 were drawn with isolevel 0.01 e/Å3, while K2NiF4 is at 0.002 e/Å3 for charge density and 0.005 e/Å3 for spin density.

Close modal

In the leftmost column of Fig. 2, we see that the charge density of K2NiF4 is just slightly rearranged when the magnetic texture is modified. This weak charge density response to changes in magnetic order persists for other magnetic textures of K2NiF4, which indicates that charge and spin degrees of freedom are weakly coupled in this material.

In the remaining columns of Fig. 2, we can see that the charge density response in SrCuO2, FeSe, and BaCo2As2 is much stronger than that in K2NiF4 whose iso-charge density surfaces were fivefold magnified with respect to those of the other materials. However, the charge–spin response in SrCuO2, FeSe, and BaCo2As2 is rather different. Both the way in which charge rearranges and the magnitude of that rearrangement vary across these materials, as can be inferred from the shape and size of the isodensity-difference surfaces in Fig. 2. For instance, changing the magnetic order from Néel (ground state) to bicolinear order in SrCuO2 mostly results in electrons moving from oxygen 2px (2py) orbitals into copper 3dx2y2 ones. In FeSe, changing from the stripe (ground state) to the Néel order seems to largely transfer electrons from iron’s 3dxz, 3dyz, and 3dxy orbitals into its 3dz2r2. Similarly, the change from bicollinear to collinear magnetic order in BaCo2As2 mostly redistributes electrons among the 3d orbitals of cobalt.

The spin density differences for these four materials (see second row of Fig. 2) have similar magnitudes even if they are qualitatively different. This thus suggests that the charge–spin response in K2NiF4 should be much weaker than that in SrCuO2 and FeSe, which, in turn, seem to have a somewhat weaker response than BaCo2As2. Those observations are corroborated by the values of the pairwise charge–spin susceptibility χi calculated for these magnetic textures and shown in the third row of Fig. 2.

We now compare the charge–spin susceptibility computed from a few different methods: PBE+U87,88 with U = 0 eV, 5 eV, 10 eV (with the plane-wave code quantum espresso86); PBE087 (using the localized basis code CRYSTAL1794); and fixed-node diffusion Monte Carlo95 (using the quantum Monte Carlo package QWalk96). Due to the high computational cost of the DMC calculations, we performed this comparison for a set of four barium arsenides with a ThCr2Si2-like structure: BaM2As2 with M = Cr, Mn, Fe, and Mn. In Fig. 3, we compare the pairwise charge–spin susceptibility χi obtained with diffusion Monte Carlo (x-axis) with the χi obtained from density functional theory (y axis). Each panel in Fig. 3 makes this comparison for DFT calculations done with each of the four functionals mentioned above: PBE+U = 0, 5, 10, and PBE0.

FIG. 3.

Pairwise charge–spin susceptibility χi using different electronic structure methods. Each point corresponds to χi—see Eqs. (3) and (5)—for an excited magnetic texture with respect to the ground state. The point colors identify the material in the set BaM2As2 with M = Cr, Mn, Fe, and Co. The point’s horizontal position is set by the diffusion Monte Carlo χi, and the vertical one is set by the density functional theory χi. Each panel corresponds to a different DFT functional: PBE+U = 0, 5, 10, and PBE0 on the (a), (b), (c), and (d) panels, respectively. The table shows the root mean square deviation (RMSD) between the χi computed from a DFT method and those obtained with DMC.

FIG. 3.

Pairwise charge–spin susceptibility χi using different electronic structure methods. Each point corresponds to χi—see Eqs. (3) and (5)—for an excited magnetic texture with respect to the ground state. The point colors identify the material in the set BaM2As2 with M = Cr, Mn, Fe, and Co. The point’s horizontal position is set by the diffusion Monte Carlo χi, and the vertical one is set by the density functional theory χi. Each panel corresponds to a different DFT functional: PBE+U = 0, 5, 10, and PBE0 on the (a), (b), (c), and (d) panels, respectively. The table shows the root mean square deviation (RMSD) between the χi computed from a DFT method and those obtained with DMC.

Close modal

In this figure, the PBE+U derived pairwise charge–spin susceptibilities generally follow the trends of the more expensive PBE0 and DMC results. The ordering of these four materials according to their values of pairwise charge–spin susceptibility χi resulting from DFT calculations with the PBE0 and DFT+U = 5 eV, 10 eV functionals is the same as that resulting from DMC: χMnχCrχFeχCo.

The deviation between the DFT-derived pairwise charge–spin susceptibilities and those calculated from DMC is shown in the table of Fig. 3. PBE0 deviates the least from DMC, with a root mean square deviation of RMSDPBE0 = 0.04 considerably smaller than those resulting from the DFT+U calculations. Part of the discrepancy between the PBE+U and PBE0/DMC χi’s arises from the fact that some magnetic textures obtained with PBE+U are often quantitatively different from those obtained with PBE0/DMC. Among the PBE+U functionals, U = 5 eV is the one that better captures the response of charge to changes in the spin texture of these materials. The PBE+U = 5 eV χi root mean square deviation from DMC is RMSDU=5 = 0.11, while that of U = 0 eV and U = 10 eV is slightly larger: RMSDU=0 = 0.17 and RMSDU=10 = 0.14.

Even if the quantitative agreement between the DFT+U methods and DMC is not perfect, these still capture the qualitative trends in charge–spin susceptibility, enabling their use in charge–spin susceptibility screenings of large sets of materials. In what follows, we will show the results for DFT+U with U = 5 eV since this functional minimizes deviations from the DMC results—see table of Fig. 3.

In some materials, the charge–spin response is strongly dependent on the type of change in the magnetic texture. That can be seen in Fig. 3 where the pairwise charge–spin susceptibilities show a large spread for BaFe2As2 but a small one for BaCr2As2. This is reminiscent of electron–phonon coupling physics, in which some phonons are more strongly coupled to the material’s electronic degrees of freedom than others. For simplicity, we take the average [see Eq. (3)] but have checked that different strategies do not affect the results in Sec. III C.

In Fig. 4, we show the charge–spin susceptibility for the materials in the entire test set from Table I obtained using DFT+U with U = 5 eV. Each material is colored according to its family: copper oxides, barium arsenides, MPX3’s, iron chalcogenides, transition metal dichalcogenides, and 214 materials. Overall, we find that materials showing large charge–spin susceptibility in Fig. 4, typically present unconventional properties either in their stoichiometric form, or when put under pressure or chemically doped. This is shown in Fig. 5 where we color materials according to whether their pressure vs doping phase diagram shows unconventional ground states. Both these families are well separated: χconv = 0.08 ± 0.10 for conventional materials, while χunc = 0.28 ± 0.12 for unconventional ones. Using a two-sided t-test, the populations have different means with a p-value of 3 × 10−5. In this section, we discuss in some detail the materials in our test set showing large charge–spin susceptibility.

FIG. 4.

Average charge–spin susceptibility for the test set materials. Materials are positioned according to their low-temperature magnetic and transport properties (x axis) and their average charge–spin susceptibility (y axis) calculated with DFT+U with U = 5 eV. Each point corresponds to one material. Those showing large charge–spin susceptibility are labeled and tend to be materials with unconventional behavior.

FIG. 4.

Average charge–spin susceptibility for the test set materials. Materials are positioned according to their low-temperature magnetic and transport properties (x axis) and their average charge–spin susceptibility (y axis) calculated with DFT+U with U = 5 eV. Each point corresponds to one material. Those showing large charge–spin susceptibility are labeled and tend to be materials with unconventional behavior.

Close modal
FIG. 5.

Average charge–spin susceptibility for materials with conventional and unconventional ground states. The x-axis position of materials is set according to their undoped and unpressured ground state. A material’s average charge–spin susceptibility (from DFT+U with U = 5 eV) determines the y-axis position. Each point corresponds to one material. The green (blue) colored points identify materials with unconventional (conventional) ground states. A material has unconventional ground states if its doping-pressure phase diagram shows unconventional superconductivity, disordered magnetic states, or bad metal—see Table I. The right hand side panel shows the distribution of conventional (blue) and unconventional (green) materials in our test set according to their charge–spin susceptibility. Using a t-test, the populations are different with a p-value of 3 × 10−5.

FIG. 5.

Average charge–spin susceptibility for materials with conventional and unconventional ground states. The x-axis position of materials is set according to their undoped and unpressured ground state. A material’s average charge–spin susceptibility (from DFT+U with U = 5 eV) determines the y-axis position. Each point corresponds to one material. The green (blue) colored points identify materials with unconventional (conventional) ground states. A material has unconventional ground states if its doping-pressure phase diagram shows unconventional superconductivity, disordered magnetic states, or bad metal—see Table I. The right hand side panel shows the distribution of conventional (blue) and unconventional (green) materials in our test set according to their charge–spin susceptibility. Using a t-test, the populations are different with a p-value of 3 × 10−5.

Close modal

The results in Fig. 4 show four copper oxides with sizable values of charge–spin susceptibility: T-La2CuO4, CaCuO2, and SrCuO2 all have χcs ≈ 0.15, while T′-La2CuO4 shows χcs ≈ 0.40. All these four materials are Néel ordered2,27,50,52 insulators and well known to become unconventional superconductors under chemical doping.2,27,50,53 Several other uncommon phases have been observed in these materials, ranging from a pseudogap phase, to strange metallicity, short-range magnetic and charge order.2,27,50,52

Two other members of the copper oxide family, TeCuO3 and SeCuO3, show very small charge–spin susceptibility. TeCuO3 is a Néel ordered insulator, while SeCuO3 is a ferromagnetic insulator,7 both well known to show magnetodielectric properties.7 As opposed to the four cuprates with copper-oxide planes discussed above, unconventional phases have not been observed in either TeCuO3 or SeCuO3, consistent with the small value of susceptibility computed here.

The iron pnictides and chalcogenides o-FeSe, t-FeSe, FeS, FeTe, t-BaFe2As2, and o-BaFe2As2 all show χcs ≈ 0.30–0.37. These materials are all metallic magnets, some showing stripe magnetic textures (BaFe2As2 and FeTe34,39) ,while others only present short-range order (FeSe and FeS42,44). All these materials have unconventional superconducting phases induced by pressure or doping,36,40,41,45 as well as short-range magnetic order and bad metal phases.33,35,39,40,44

Sr2FeO4 has χcs = 0.22 and is an antiferromagnetic semiconductor. Chemically doping this compound weakens both its antiferromagnetic ordering and semiconducting character without completely suppressing the electronic gap.103–105 It has been shown that pressure induces a semiconductor-to-metal transition at P ≈ 18 GPa,47 but so far no unconventional phases have been observed on its phase diagram down to 5 K for pressures up to 30 GPa.

Sr2CoO4 is ferromagnetic and metallic at low temperatures.28 Upon chemical doping with Y,28 La,29 and Nd,30 the ferromagnetism weakens and semiconducting behavior arises. This material becomes a superconductor at around 5 K when doped with H2O,31 with very similar properties to the cuprate superconductors. We were not aware of this result prior to the study; the material was identified purely due to the charge–spin descriptor, χcs = 0.39.

Sr2VO4 is a multi-orbital Mott insulator with no long-range magnetic order5 that can be driven into a metallic state by hydrostatic pressure (≈20 GPa–24 GPa).23 An unconventional metal emerges at low temperatures in the vicinity of the pressure-driven transition.23 In contrast to Sr2VO4 thin films,106 attempts to chemically dope the bulk crystal did not succeed in making it metallic.107 Since our calculations suggest it has a strong charge–spin coupling, χcs = 0.42, a more comprehensive exploration of different ways of chemically doping this material may reveal novel phases.

BaCo2As2 has a rather large charge–spin susceptibility, χcs = 0.45. It is a disordered magnetic metal4 that seems to remain so upon both chemical doping with K4 and hydrostatic pressure (up to 8 GPa).108 It thus seems similar to Sr2VO4 in the sense that there is a disordered magnetic state. We note that these two systems are not ordinary magnetic materials, and both exhibit nonstandard ground states. Thus, the charge–spin descriptor succeeded in identifying unusual physics in these materials.

K2CoF4 has been classified as a 2D Ising magnet55 owing to its strongly anisotropic magnetic interactions. To our knowledge, this material’s behavior under pressure or chemical doping has been very sparsely studied,109 potentially due to the presence of fluorine. Its charge–spin susceptibility, χcs = 0.10, is slightly below that of some superconducting cuprates.

TaS2 is a Mott insulator associated with a charge-density wave (CDW) phase.59 Pressure induces a metal-to-insulator transition and a superconducting state below 5 K.59 Recent experiments58 suggest that the low-temperature CDW phase has short-range magnetic order, which supported proposals that this material might realize a quantum spin liquid state.110 Our calculations, done with the undistorted crystal structure, give χcs = 0.09 just below what we get for some cuprates.

The compounds with lower charge–spin susceptibility in Fig. 5 (listed on the bottom of Table I) comprise materials that show conventional phases, mostly insulators with Néel antiferromagnetic order and ferromagnetic order. Some of these materials are known to become metallic (e.g., BaMn2As2) or to acquire spin and/or charge stripe order (e.g., La2NiO4 and La2CoO4) upon doping or pressure, but none of them presents unconventional electronic phases. This indicates that, as suggested by our calculations, their charge and spin degrees of freedom are weakly coupled.

We have presented a new way of probing charge–spin coupling in materials. It is based on the charge–spin susceptibility, a quantity that estimates the magnitude of the coupling between charge and spin degrees of freedom in a material. This quantity is straightforward to compute in a high-throughput workflow, and when applied to a collection of layered materials containing transition metal atoms suggests that materials with high charge–spin coupling exhibit unconventional phenomena.

All the materials in our test set known to present unconventional phases (green colored in Fig. 5) show average charge–spin susceptibilities χcs ≥ 0.09. Among the 16 materials with charge–spin susceptibility above this value, only three have not been observed to show unconventional physics, Sr2CoO4, Sr2FeO4, and K2CoF4, which is a rather small rate of false positive identifications. Perhaps more importantly, the false negative rate was zero within our test set. This rate is certainly good enough to motivate experimental investigation into materials.

The computation of the charge–spin susceptibility for the materials in our test set was performed using DFT+U, a low-cost method which is sufficiently accurate to highlight the same qualitative trends found using the highly accurate many-body method fixed-node diffusion Monte Carlo. These calculations are inexpensive enough that they can be used in large-scale probes of the strength of the coupling between electronic and magnetic degrees of freedom in materials. The trends found here suggest that the charge–spin susceptibility is a valuable quantity for computational searches for new unconventional ground states, including unconventional superconductivity similar to iron-based and cuprate superconductors.

The data that support the findings of this study are openly available in Materials Data Facility111,112 at http://doi.org/10.18126/6oby-l2lp, Ref. 84.

This work was supported by the Center for Emergent Superconductivity, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Award No. DEAC0298CH1088. The authors thank Daniel Shoemaker for many illuminating discussions. The computational resources used in this work were provided by the University of Illinois Campus Cluster and the Blue Waters sustained-petascale computing project, which is supported by the National Science Foundation (Award Nos. OCI-0725070 and ACI-1238993) and the state of Illinois. Blue Waters is a joint effort of the University of Illinois at Urbana-Champaign and its National Center for Superconducting Applications. L.K.W. was supported by a grant from the Simons Foundation as part of the Simons Collaboration on the many-electron problem.

In the context of a system governed by the Hamiltonian in Eq. (1), assume that we fix the portion of the wave function associated with the spin degrees of freedom, ϕ(r), to a particular magnetic order. Then, the orbital degrees of freedom will be described by φ(r; ϕ). The charge and the spin density of such configuration will be given by

ρ(r)=σ|φσ(r)|2+|ϕσ(r)|2,
(A1a)
s(r)=σσ|φσ(r)|2+|ϕσ(r)|2,
(A1b)

where σ = ± identifies the spin projection, while φσ is a short-hand for φσ(r; ϕ).

Let us apply an infinitesimal deformation away from the ground state on the portion of the wave function describing the spin degrees of freedom, δϕσ(r) = ϕσ(r) − ϕ0σ(r) (where ϕσ is the deformed wave function and ϕ0σ is the ground state one). We can then write the differential of each component of the charge and spin densities as follows:

δ|ϕσ|2=2|ϕσ|δ|ϕσ|,
(A2a)
δ|φσ|2=2|φσ|αδ|φσ|δ|ϕα|δ|φα|,
(A2b)

where we used δ|φσ|=αδ|φσ|δ|ϕα|δ|ϕα| under the assumption that the wave function’s orbital degrees of freedom component only depend on the absolute value of the spin states component, φσ(r; ϕσ) ≃ φσ(r; |ϕσ|).

We can write the differential of the charge and spin densities as

δρ=2σ,α|φσ|δ|φσ|δ|ϕα|+|ϕσ|δσαδ|ϕα|,
(A3a)
δs=2σ,ασ|φσ|δ|φσ|δ|ϕα|+|ϕσ|δσαδ|ϕα|,
(A3b)

where δσα is the Kronecker delta.

Consider now that the small deformation is such that it only changes the spin states’ magnetic order, preserving their contribution to the charge density, i.e., δ∑σ|ϕσ|2 ≃ 0. This implies that σ2|ϕσ|δ|ϕσ| ≃ 0, which allows us to write

δ|ϕ|=|ϕ+||ϕ|δ|ϕ+|.
(A4)

Under this approximation, we can write δρ and δs as

δρ=2λwϒ+|ϕ+||ϕ|ϒδ|ϕ+|,
(A5a)
δs=2Ξ++|ϕ+||ϕ|Ξδ|ϕ+|,
(A5b)

where Υασ|φσ|fσα and Ξαλwϒα+|ϕα|. In these expressions, we used the definition δ|φσ|δ|ϕα|λwfσα, where according to the main text’s notation, λ stands for the coupling between the spin states and the orbital degrees of freedom, while w corresponds to a material-specific energy scale.

Using the above expressions, we can write δρ in terms of δs as

δρ(r)=λwϒ+|ϕ+||ϕ|ϒΞ++|ϕ+||ϕ|Ξδs(r),
(A6)

which using Eq. (A4) can be simplified into

δρ(r)=λwX(r)δs(r),
(A7)

where X(r) = Υ+(r)/|ϕ+(r)|.

Equation (A7) has the same form of main text’s Eq. (2), only that in the latter we use a perturbation theory notation: Δsi(r) = si(r) − s0(r) [instead of δs(r)] and Δρi(r) = ρi(r) − ρ0(r) [instead of δρ(r)], where ρi and si stand for the charge and spin densities of the i deformation away from the ground state’s charge and spin densities, ρ0(r) and s0(r).

1.
D. J.
Scalapino
,
Rev. Mod. Phys.
84
,
1383
(
2012
).
2.
C.
Proust
and
L.
Taillefer
,
Annu. Rev. Condens. Matter Phys.
10
,
409
(
2019
).
3.
A.
Fert
,
N.
Reyren
, and
V.
Cros
,
Nat. Rev. Mater.
2
,
17031
(
2017
).
4.
V. K.
Anand
,
D. G.
Quirinale
,
Y.
Lee
,
B. N.
Harmon
,
Y.
Furukawa
,
V. V.
Ogloblichev
,
A.
Huq
,
D. L.
Abernathy
,
P. W.
Stephens
,
R. J.
McQueeney
,
A.
Kreyssig
,
A. I.
Goldman
, and
D. C.
Johnston
,
Phys. Rev. B
90
,
064517
(
2014
).
5.
I.
Yamauchi
,
K.
Nawa
,
M.
Hiraishi
,
M.
Miyazaki
,
A.
Koda
,
K. M.
Kojima
,
R.
Kadono
,
H.
Nakao
,
R.
Kumai
,
Y.
Murakami
,
H.
Ueda
,
K.
Yoshimura
, and
M.
Takigawa
,
Phys. Rev. B
92
,
064408
(
2015
).
6.
H.
Takagi
,
T.
Takayama
,
G.
Jackeli
,
G.
Khaliullin
, and
S. E.
Nagler
,
Nat. Rev. Phys.
1
,
264
(
2019
).
7.
G.
Lawes
,
T.
Kimura
,
C. M.
Varma
,
M. A.
Subramanian
,
N.
Rogado
,
R. J.
Cava
, and
A. P.
Ramirez
,
Prog. Solid State Chem.
37
,
40
(
2009
).
8.
F.
Guinea
,
G.
Gómez-Santos
, and
D. P.
Arovas
,
Phys. Rev. B
62
,
391
(
2000
).
9.
G.
Jackeli
and
N. M.
Plakida
,
Physica B
259-261
,
723
(
1999
).
10.
S.
Pandey
,
S.
Das
,
B.
Kamble
,
S.
Ghosh
,
D.
Singh
,
R.
Ray
, and
A.
Singh
,
Phys. Rev. B
77
,
134447
(
2008
).
11.
M.
Johnson
and
J.
Byers
,
Phys. Rev. B
67
,
125112
(
2003
).
12.
X. G.
Zheng
,
C. N.
Xu
,
E.
Tanaka
,
Y.
Tomokiyo
,
M.
Suzuki
, and
E. S.
Otabe
,
Physica C
357-360
,
181
(
2001
).
13.
S.
Lederer
,
Y.
Schattner
,
E.
Berg
, and
S. A.
Kivelson
,
Proc. Natl. Acad. Sci. U. S. A.
114
,
4905
(
2017
).
14.
H.-C.
Jiang
,
Z.-Y.
Weng
, and
S. A.
Kivelson
,
Phys. Rev. B
98
,
140505
(
2018
).
15.
I.
Esterlis
,
B.
Nosarzewski
,
E. W.
Huang
,
B.
Moritz
,
T. P.
Devereaux
,
D. J.
Scalapino
, and
S. A.
Kivelson
,
Phys. Rev. B
97
,
140501
(
2018
).
16.
K.
Seki
,
Y.
Otsuka
,
S.
Yunoki
, and
S.
Sorella
,
Phys. Rev. B
99
,
125145
(
2019
).
17.
E.
Berg
,
S.
Lederer
,
Y.
Schattner
, and
S.
Trebst
,
Annu. Rev. Condens. Matter Phys.
10
,
63
(
2019
).
18.
N. C.
Costa
,
K.
Seki
,
S.
Yunoki
, and
S.
Sorella
,
Commun. Phys.
3
,
80
(
2020
).
19.
M.
Klintenberg
and
O.
Eriksson
,
Comput. Mater. Sci.
67
,
282
(
2013
).
20.
C.-H.
Yee
and
G.
Kotliar
,
Phys. Rev. B
89
,
094517
(
2014
).
21.
A. S.
Botana
,
V.
Pardo
, and
M. R.
Norman
,
Phys. Rev. Mater.
1
,
021801
(
2017
).
22.
Q.
Si
,
R.
Yu
, and
E.
Abrahams
,
Nat. Rev. Mater.
1
,
16017
(
2016
).
23.
S.
Karmakar
and
P. S.
Malavi
,
Phys. Rev. Lett.
114
,
166402
(
2015
).
24.
M.
Pfisterer
and
G.
Nagorsen
, “
Zur struktur ternärer übergangsmetallarsenide/on the structure of ternary arsenides
,”
Z. Naturforsch. B
35
,
703
(
1980
).
25.
M.
Cyrot
,
B.
Lambert-Andron
,
J. L.
Soubeyroux
,
M. J.
Rey
,
P.
Dehauht
,
F.
Cyrot-Lackmann
,
G.
Fourcaudot
,
J.
Beille
, and
J. L.
Tholence
,
J. Solid State Chem.
85
,
321
(
1990
).
26.
R.
Hord
,
H.
Luetkens
,
G.
Pascua
,
A.
Buckow
,
K.
Hofmann
,
Y.
Krockenberger
,
J.
Kurian
,
H.
Maeter
,
H.-H.
Klauss
,
V.
Pomjakushin
,
A.
Suter
,
B.
Albert
, and
L.
Alff
,
Phys. Rev. B
82
,
180508
(
2010
).
27.
K.
Jin
,
N. P.
Butch
,
K.
Kirshenbaum
,
J.
Paglione
, and
R. L.
Greene
,
Nature
476
,
73
(
2011
).
28.
X. L.
Wang
and
E.
Takayama-Muromachi
,
Phys. Rev. B
72
,
064401
(
2005
).
29.
Y.
Shimada
,
S.
Miyasaka
,
R.
Kumai
, and
Y.
Tokura
,
Phys. Rev. B
73
,
134424
(
2006
).
30.
Q.
Yao
,
H.
Kimura
,
X. L.
Wang
,
K.
Konstantinov
,
H.
Zhao
,
H.
Qiu
,
R.
Tanahashi
,
K.
Ozawa
, and
S. X.
Dou
,
J. Appl. Phys.
111
,
07D708
(
2012
).
31.
K.
Takada
,
H.
Sakurai
,
E.
Takayama-Muromachi
,
F.
Izumi
,
R. A.
Dilanian
, and
T.
Sasaki
,
Nature
422
,
53
(
2003
).
32.
R.
Mittal
,
S. K.
Mishra
,
S. L.
Chaplot
,
S. V.
Ovsyannikov
,
E.
Greenberg
,
D. M.
Trots
,
L.
Dubrovinsky
,
Y.
Su
,
T.
Brueckel
,
S.
Matsuishi
,
H.
Hosono
, and
G.
Garbarino
,
Phys. Rev. B
83
,
054503
(
2011
).
33.
H.
Luo
,
R.
Zhang
,
M.
Laver
,
Z.
Yamani
,
M.
Wang
,
X.
Lu
,
M.
Wang
,
Y.
Chen
,
S.
Li
,
S.
Chang
,
J. W.
Lynn
, and
P.
Dai
,
Phys. Rev. Lett.
108
,
247002
(
2012
).
34.
X. F.
Wang
,
T.
Wu
,
G.
Wu
,
H.
Chen
,
Y. L.
Xie
,
J. J.
Ying
,
Y. J.
Yan
,
R. H.
Liu
, and
X. H.
Chen
,
Phys. Rev. Lett.
102
,
117005
(
2009
).
35.
S.
Kasahara
,
T.
Shibauchi
,
K.
Hashimoto
,
K.
Ikada
,
S.
Tonegawa
,
R.
Okazaki
,
H.
Shishido
,
H.
Ikeda
,
H.
Takeya
,
K.
Hirata
,
T.
Terashima
, and
Y.
Matsuda
,
Phys. Rev. B
81
,
184519
(
2010
).
36.
M.
Rotter
,
M.
Tegel
, and
D.
Johrendt
,
Phys. Rev. Lett.
101
,
107006
(
2008
).
37.
J.-E.
Jørgensen
and
T. C.
Hansen
,
Eur. Phys. J. B
78
,
411
(
2010
).
38.
V. P. S.
Awana
,
A.
Pal
,
A.
Vajpayee
,
B.
Gahtori
, and
H.
Kishan
,
Physica C
471
,
77
(
2011
).
39.
A.
Martinelli
,
A.
Palenzona
,
M.
Tropeano
,
C.
Ferdeghini
,
M.
Putti
,
M. R.
Cimberle
,
T. D.
Nguyen
,
M.
Affronte
, and
C.
Ritter
,
Phys. Rev. B
81
,
094115
(
2010
).
40.
B. C.
Sales
,
A. S.
Sefat
,
M. A.
McGuire
,
R. Y.
Jin
,
D.
Mandrus
, and
Y.
Mozharivskyj
,
Phys. Rev. B
79
,
094521
(
2009
).
41.
X.
Lai
,
H.
Zhang
,
Y.
Wang
,
X.
Wang
,
X.
Zhang
,
J.
Lin
, and
F.
Huang
,
J. Am. Chem. Soc.
137
,
10148
(
2015
).
42.
S.
Holenstein
,
U.
Pachmayr
,
Z.
Guguchia
,
S.
Kamusella
,
R.
Khasanov
,
A.
Amato
,
C.
Baines
,
H.-H.
Klauss
,
E.
Morenzoni
,
D.
Johrendt
, and
H.
Luetkens
,
Phys. Rev. B
93
,
140506
(
2016
).
43.
J. N.
Millican
,
D.
Phelan
,
E. L.
Thomas
,
J. B.
Leão
, and
E.
Carpenter
,
Solid State Commun.
149
,
707
(
2009
).
44.
Q.
Wang
,
Y.
Shen
,
B.
Pan
,
X.
Zhang
,
K.
Ikeuchi
,
K.
Iida
,
A. D.
Christianson
,
H. C.
Walker
,
D. T.
Adroja
,
M.
Abdel-Hafiez
,
X.
Chen
,
D. A.
Chareev
,
A. N.
Vasiliev
, and
J.
Zhao
,
Nat. Commun.
7
,
12182
(
2016
).
45.
S.
Medvedev
,
T. M.
McQueen
,
I. A.
Troyan
,
T.
Palasyuk
,
M. I.
Eremets
,
R. J.
Cava
,
S.
Naghavi
,
F.
Casper
,
V.
Ksenofontov
,
G.
Wortmann
, and
C.
Felser
,
Nat. Mater.
8
,
630
(
2009
).
46.
S. E.
Dann
,
M. T.
Weller
,
D. B.
Currie
,
M. F.
Thomas
, and
A. D.
Al-Rawwas
,
J. Mater. Chem.
3
,
1231
(
1993
).
47.
G. K.
Rozenberg
,
A. P.
Milner
,
M. P.
Pasternak
,
G. R.
Hearne
, and
R. D.
Taylor
,
Phys. Rev. B
58
,
10283
(
1998
).
48.
C.
Rial
,
E.
Morán
,
M. A.
Alario-Franco
,
U.
Amador
, and
N. H.
Andersen
,
Physica C
278
,
122
(
1997
).
49.
H.
Schwer
and
J.
Karpinski
,
Stud. High Temp. Supercond.
24
,
49
(
1997
).
50.
A.
Lombardi
,
M.
Mali
,
J.
Roos
,
D.
Brinkmann
, and
I.
Mangelschots
,
Phys. Rev. B
54
,
93
(
1996
).
51.
G.
Er
,
Y.
Miyamoto
,
F.
Kanamaru
, and
S.
Kikkawa
,
Physica C
181
,
206
(
1991
).
52.
I. A.
Zaliznyak
,
H.
Woo
,
T. G.
Perring
,
C. L.
Broholm
,
C. D.
Frost
, and
H.
Takagi
,
Phys. Rev. Lett.
93
,
087202
(
2004
).
53.
M. G.
Smith
,
A.
Manthiram
,
J.
Zhou
,
J. B.
Goodenough
, and
J. T.
Markert
,
Nature
351
,
549
(
1991
).
54.
D.
Babel
and
E.
Herdtweck
,
Z. Anorg. Allg. Chem.
487
,
75
(
1982
).
55.
D. J.
Breed
,
K.
Gilijamse
, and
A. R.
Miedema
,
Physica
45
,
205
(
1969
).
56.
G.
Samoggia
,
F.
Parmigiani
, and
F.
Leccabue
,
Solid State Commun.
55
,
157
(
1985
).
57.
A.
Spijkerman
,
J. L.
de Boer
,
A.
Meetsma
,
G. A.
Wiegers
, and
S.
van Smaalen
,
Phys. Rev. B
56
,
13757
(
1997
).
58.
M.
Kratochvilova
,
A. D.
Hillier
,
A. R.
Wildes
,
L.
Wang
,
S.-W.
Cheong
, and
J.-G.
Park
,
npj Quantum Mater.
2
,
42
(
2017
).
59.
B.
Sipos
,
A. F.
Kusmartseva
,
A.
Akrap
,
H.
Berger
,
L.
Forró
, and
E.
Tutiš
,
Nat. Mater.
7
,
960
(
2008
).
60.
T.
Baikie
,
Z.
Ahmad
,
M.
Srinivasan
,
A.
Maignan
,
S. S.
Pramana
, and
T. J.
White
,
J. Solid State Chem.
180
,
1538
(
2007
).
61.
H.
Sakurai
,
J. Phys. Soc. Jpn.
83
,
123701
(
2014
).
62.
D. J.
Singh
,
A. S.
Sefat
,
M. A.
McGuire
,
B. C.
Sales
,
D.
Mandrus
,
L. H.
VanBebber
, and
V.
Keppens
,
Phys. Rev. B
79
,
094429
(
2009
).
63.
J.
Rodriguez-Carvajal
,
M. T.
Fernandez-Diaz
, and
J. L.
Martinez
,
J. Phys.: Condens. Matter
3
,
3215
(
1991
).
64.
P.
Wochner
,
J. M.
Tranquada
,
D. J.
Buttrey
, and
V.
Sachan
,
Phys. Rev. B
57
,
1066
(
1998
).
65.
E.
Brechtel
,
G.
Cordier
, and
H.
Schaefer
, “
Neue ternäre erdalkaliverbindungen des arsens mit mangan/new ternary alkaline earth compounds of arsenic with manganese
,”
Z. Naturforsch. B
33
,
820
(
1978
).
66.
J.
An
,
A. S.
Sefat
,
D. J.
Singh
, and
M.-H.
Du
,
Phys. Rev. B
79
,
075120
(
2009
).
67.
A.
Pandey
,
R. S.
Dhaka
,
J.
Lamsal
,
Y.
Lee
,
V. K.
Anand
,
A.
Kreyssig
,
T. W.
Heitmann
,
R. J.
McQueeney
,
A. I.
Goldman
,
B. N.
Harmon
,
A.
Kaminski
, and
D. C.
Johnston
,
Phys. Rev. Lett.
108
,
087005
(
2012
).
68.
R.
Brec
,
G.
Ouvrard
,
A.
Louisy
, and
J.
Rouxel
,
Ann. Chimie Sci. Matériaux
5
,
499
(
1980
).
69.
G. L.
Flem
,
R.
Brec
,
G.
Ouvard
,
A.
Louisy
, and
P.
Segransan
,
J. Phys. Chem. Solids
43
,
455
(
1982
).
70.
S. J.
Skinner
and
G.
Amow
,
J. Solid State Chem.
180
,
1977
(
2007
).
71.
P.
Babkevich
,
P. G.
Freeman
,
M.
Enderle
,
D.
Prabhakaran
, and
A. T.
Boothroyd
,
Nat. Commun.
7
,
11632
(
2016
).
72.
P.
Babkevich
,
D.
Prabhakaran
,
C. D.
Frost
, and
A. T.
Boothroyd
,
Phys. Rev. B
82
,
184425
(
2010
).
73.
K.
Tezuka
,
M.
Inamura
,
Y.
Hinatsu
,
Y.
Shimojo
, and
Y.
Morii
,
J. Solid State Chem.
145
,
705
(
1999
).
74.
T.-H.
Kao
,
H.
Sakurai
,
T.
Kolodiazhnyi
,
Y.
Suzuki
,
M.
Okabe
,
T.
Asaka
,
K.
Fukuda
,
S.
Okubo
,
S.
Ikeda
,
S.
Hara
,
T.
Sakurai
,
H.
Ohta
, and
H.-D.
Yang
,
J. Mater. Chem. C
3
,
3452
(
2015
).
75.
E.
Philippot
and
M.
Maurin
,
Rev. Chim. Miner.
13
,
162
(
1976
).
76.
V.
Carteaux
,
D.
Brunet
,
G.
Ouvrard
, and
G.
Andre
,
J. Phys.: Condens. Matter
7
,
69
(
1995
).
77.
E.
Herdtweck
and
D.
Babel
,
Z. Anorg. Allg. Chem.
474
,
113
(
1981
).
78.
I.
Yamada
,
J. Phys. Soc. Jpn.
33
,
979
(
1972
).
79.
W.
Kleemann
,
J.
Ferre
, and
F. J.
Schafer
,
J. Phys. C: Solid State Phys.
14
,
4463
(
1981
).
80.
S. K.
Yeh
,
S. Y.
Wu
,
C. S.
Lee
, and
Y.
Wang
,
Acta Crystallogr., Sect. B
49
,
806
(
1993
).
81.
R. J.
Birgeneau
,
H. J.
Guggenheim
, and
G.
Shirane
,
Phys. Rev. B
1
,
2211
(
1970
).
82.
R.
Escamilla
,
J.
Gallardo-Amores
,
E.
Morán
, and
M.
Alario-Franco
,
J. Solid State Chem.
168
,
149
(
2002
).
83.
A.
Narayan
,
B.
Busemeyer
, and
L. K.
Wagner
, arXiv:1705.01008 [cond-mat.supr-con] (
2017
).
84.
J. N. B.
Rodrigues
and
L. K.
Wagner
,
Mater. Data Facil.
(
2019
).
85.
W.
Kohn
and
L. J.
Sham
,
Phys. Rev.
140
,
A1133
(
1965
).
86.
P.
Giannozzi
,
S.
Baroni
,
N.
Bonini
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
G. L.
Chiarotti
,
M.
Cococcioni
,
I.
Dabo
,
A.
Dal Corso
,
S.
de Gironcoli
,
S.
Fabris
,
G.
Fratesi
,
R.
Gebauer
,
U.
Gerstmann
,
C.
Gougoussis
,
A.
Kokalj
,
M.
Lazzeri
,
L.
Martin-Samos
,
N.
Marzari
,
F.
Mauri
,
R.
Mazzarello
,
S.
Paolini
,
A.
Pasquarello
,
L.
Paulatto
,
C.
Sbraccia
,
S.
Scandolo
,
G.
Sclauzero
,
A. P.
Seitsonen
,
A.
Smogunov
,
P.
Umari
, and
R. M.
Wentzcovitch
,
J. Phys.: Condens. Matter
21
,
395502
(
2009
).
87.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
,
Phys. Rev. Lett.
77
,
3865
(
1996
).
88.
M.
Cococcioni
and
S.
de Gironcoli
,
Phys. Rev. B
71
,
035105
(
2005
).
89.
K.
Lejaeghere
,
G.
Bihlmayer
,
T.
Björkman
,
P.
Blaha
,
S.
Blügel
,
V.
Blum
,
D.
Caliste
,
I. E.
Castelli
,
S. J.
Clark
,
A.
Dal Corso
,
S.
de Gironcoli
,
T.
Deutsch
,
J. K.
Dewhurst
,
I.
Di Marco
,
C.
Draxl
,
M.
Dułak
,
O.
Eriksson
,
J. A.
Flores-Livas
,
K. F.
Garrity
,
L.
Genovese
,
P.
Giannozzi
,
M.
Giantomassi
,
S.
Goedecker
,
X.
Gonze
,
O.
Grånäs
,
E. K. U.
Gross
,
A.
Gulans
,
F.
Gygi
,
D. R.
Hamann
,
P. J.
Hasnip
,
N. A. W.
Holzwarth
,
D.
Iuşan
,
D. B.
Jochym
,
F.
Jollet
,
D.
Jones
,
G.
Kresse
,
K.
Koepernik
,
E.
Küçükbenli
,
Y. O.
Kvashnin
,
I. L. M.
Locht
,
S.
Lubeck
,
M.
Marsman
,
N.
Marzari
,
U.
Nitzsche
,
L.
Nordström
,
T.
Ozaki
,
L.
Paulatto
,
C. J.
Pickard
,
W.
Poelmans
,
M. I. J.
Probert
,
K.
Refson
,
M.
Richter
,
G.-M.
Rignanese
,
S.
Saha
,
M.
Scheffler
,
M.
Schlipf
,
K.
Schwarz
,
S.
Sharma
,
F.
Tavazza
,
P.
Thunström
,
A.
Tkatchenko
,
M.
Torrent
,
D.
Vanderbilt
,
M. J.
van Setten
,
V.
Van Speybroeck
,
J. M.
Wills
,
J. R.
Yates
,
G.-X.
Zhang
, and
S.
Cottenier
,
Science
351
,
aad3000
(
2016
).
90.
G.
Prandini
,
A.
Marrazzo
,
I. E.
Castelli
,
N.
Mounet
, and
N.
Marzari
,
npj Comput. Mater.
4
,
72
(
2018
).
91.
H. J.
Monkhorst
and
J. D.
Pack
,
Phys. Rev. B
13
,
5188
(
1976
).
92.
M.
Hellenbrandt
,
Crystallogr. Rev.
10
,
17
(
2004
).
93.
J. P.
Perdew
,
M.
Ernzerhof
, and
K.
Burke
,
J. Chem. Phys.
105
,
9982
(
1996
).
94.
R.
Dovesi
,
A.
Erba
,
R.
Orlando
,
C. M.
Zicovich-Wilson
,
B.
Civalleri
,
L.
Maschio
,
M.
Rérat
,
S.
Casassa
,
J.
Baima
,
S.
Salustro
, and
B.
Kirtman
,
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
8
,
e1360
(
2018
).
95.
W. M. C.
Foulkes
,
L.
Mitas
,
R. J.
Needs
, and
G.
Rajagopal
,
Rev. Mod. Phys.
73
,
33
(
2001
).
96.
L. K.
Wagner
,
M.
Bajdich
, and
L.
Mitas
,
J. Comput. Phys.
228
,
3390
(
2009
).
97.
J. c. v.
Kolorenč
,
S.
Hu
, and
L.
Mitas
,
Phys. Rev. B
82
,
115108
(
2010
).
98.
H.
Zheng
and
L. K.
Wagner
,
Phys. Rev. Lett.
114
,
176401
(
2015
).
99.
B.
Busemeyer
,
M.
Dagrada
,
S.
Sorella
,
M.
Casula
, and
L. K.
Wagner
,
Phys. Rev. B
94
,
035108
(
2016
).
100.
M.
Burkatzki
,
C.
Filippi
, and
M.
Dolg
,
J. Chem. Phys.
126
,
234105
(
2007
).
101.
M.
Burkatzki
,
C.
Filippi
, and
M.
Dolg
,
J. Chem. Phys.
129
,
164115
(
2008
).
102.
L. K.
Wagner
,
Phys. Rev. B
92
,
161116
(
2015
).
103.
T.
Omata
,
K.
Ueda
,
H.
Hosono
,
M.
Katada
,
N.
Ueda
, and
H.
Kawazoe
,
Phys. Rev. B
49
,
10194
(
1994
).
104.
A. J.
Jennings
,
C. K. M.
Shaw
, and
S. J.
Skinner
,
Mater. Chem. Phys.
89
,
354
(
2005
).
105.
Q.
Zhao
,
M. D.
Smith
, and
H.-C.
zur Loye
,
J. Chem. Crystallogr.
41
,
674
(
2011
).
106.
J.
Matsuno
,
Y.
Okimoto
,
M.
Kawasaki
, and
Y.
Tokura
,
Appl. Phys. Lett.
82
,
194
(
2003
).
107.
F.
Deslandes
,
A. I.
Nazzal
, and
J. B.
Torrance
,
Physica C
179
,
85
(
1991
).
108.
C.
Ganguli
,
K.
Matsubayashi
,
K.
Ohgushi
,
Y.
Uwatoko
,
M.
Kanagaraj
, and
S.
Arumugam
,
Mater. Res. Bull.
48
,
4329
(
2013
).
109.
D. J.
Breed
,
K.
Gilijamse
,
J. W. E.
Sterkenburg
, and
A. R.
Miedema
,
J. Appl. Phys.
41
,
1267
(
1970
).
110.
K. T.
Law
and
P. A.
Lee
,
Proc. Natl. Acad. Sci. U. S. A.
114
,
6996
(
2017
).
111.
B.
Blaiszik
,
K.
Chard
,
J.
Pruyne
,
R.
Ananthakrishnan
,
S.
Tuecke
, and
I.
Foster
,
JOM
68
,
2045
(
2016
).
112.
B.
Blaiszik
,
L.
Ward
,
M.
Schwarting
,
J.
Gaff
,
R.
Chard
,
D.
Pike
,
K.
Chard
, and
I.
Foster
,
MRS Commun.
9
,
1125
(
2019
).