This paper deals with a significant family of compounds predicted by simplistic electronic structure theory to be metals but are, in fact, insulators. This false metallic state has been traditionally attributed in the literature to reflect the absence of proper treatment of electron-electron correlation (“Mott insulators”) whereas, in fact, even mean-field like density functional theory describes the insulating phase correctly if the restrictions posed on the simplistic theory are avoided. Such unwarranted restrictions included different forms of disallowing symmetry breaking described in this article. As the science and technology of conductors have transitioned from studying simple elemental metals such as Al or Cu to compound conductors such as binary or ternary oxides and pnictides, a special class of degenerate but gapped metals has been noticed. Their presumed electronic configurations show the Fermi level inside the conduction band or valence band, yet there is an “internal band gap” between the principal band edges. The significance of this electronic configuration is that it might be unstable toward the formation of states inside the internal band gap when the formation of such states costs less energy than the energy gained by transferring carriers from the conduction band to these lower energy acceptor states, changing the original (false) metal to an insulator. The analogous process also exists for degenerate but gapped metals with the Fermi level inside the valence band, where the energy gain is defined by transfer of electrons from the donor level to the unoccupied part of the valence band. We focus here on the fact that numerous electronic structure methodologies have overlooked some physical factors that could stabilize the insulating alternative, predicting instead false metals that do not really exist (note that this is in general not a physical phase transition, but a correction of a previous error in theory that led to a false prediction of a metal). Such errors include: (i) ignoring spin symmetry breaking, such as disallowing magnetic spin ordering in CuBi2O4 or disallowing the formation of polymorphous spin networks in paramagnetic LaTiO3 and YTiO3; (ii) ignoring structural symmetry breaking, e.g., not enabling energy-lowering bond disproportionation (Li-doped TiO2, SrBiO3, and rare-earth nickelates), or not exploring pseudo-Jahn–Teller-like distortions in LaMnO3, or disallowing spontaneous formation of ordered vacancy compounds in Ba4As3 and Ag3Al22O34; and (iii) ignoring spin–orbit coupling forcing false metallic states in CaIrO3 and Sr2IrO4. The distinction between false metals vs real insulators is important because (a) predicting theoretically that a given compound is metal even though it is found to be an insulator often creates the temptation to invoke high order novel physical effects (such as correlation in d-electron Mott insulators) to explain what was in effect caused by a more mundane artifact in a lower-level mean-field band theory, (b) recent prediction of exotic physical effects such as topological semimetals were unfortunately based on the above compounds that were misconstrued by theory to be metal, but are now recognized to be stable insulators not hosting exotic effects, and (c) practical technological applications based on stable degenerate but gapped metals such as transparent conductors or electrides for catalysis must rely on the systematically correct and reliable theoretical classification of metals vs insulators.

One of the most fundamental descriptors of solids is their designation as metals or insulators. Indeed, this distinction frames much of the discussion of their electronic, transport, superconducting, or topological characteristics.1–4 In standard theoretical descriptions, this distinction is represented via the construct of an energy vs wavevector (E vs k) band structure, where metals (respectively, insulators) have their Fermi level (EF) inside the continuous part of the band structure (respectively, band gap) energy region. Figure 1(a) shows schematically the density of states of an n-type metal, where the Fermi level resides inside the conduction band (CB). This is often the case when the composition weighted formal oxidation state (FOS) is positive, as is the case for Sr2+V5+O32−.5 A p-type metal would correspond to the case where the Fermi level resides in the valence band (VB), as is often the case when the composition weighted FOS is negative, e.g., Tl1+Cu21+Se22−.6 Unlike the band structure of elemental metals such as Al or Cu, where the next occupied band below the metallic band is generally an isolated, electronically inactive core electron band, recent interest in compound metals (such as the familiar groups ABX2, A2BX4, and ABX3 ternary chalcogenides and pnictides some being metallic) has focused attention on the special electronic configuration illustrated in Fig. 1(a), namely, where a principal occupied valence band is located just a couple of eV below the metallic conduction band. We will refer to a metallic system having its Fermi energy inside the conduction or valence band, yet there is a large internal band gap [Egint shown in Fig. 1(a)] between the principal valence band maximum and conduction band minimum as a “degenerate but gapped metal.”

FIG. 1.

Schematic illustration of (a) degenerate gapped metal having electrons in the principal conduction band and, in addition, an “internal” band gap Egint between the principal conduction and valence bands; (b) schematic illustration of a real insulator originating from a degenerate gapped metal. Blue: occupied states; white: unoccupied states. Here, the system invests energy Erec to create in the gap a new state via reconstruction, magnetization, or defect formation. In turn, this state accepts q electrons from the conduction band, contributing to energy lowering qΔE, where ΔE is the difference in band energies for electrons in the conduction band and localized occupied state. This creates a gap between the newly formed state and principal conduction band. Reproduced with permission from Malyi and Zunger, Phys. Rev. B 101, 235202 (2020). Copyright 2020 by the American Physical Society.

FIG. 1.

Schematic illustration of (a) degenerate gapped metal having electrons in the principal conduction band and, in addition, an “internal” band gap Egint between the principal conduction and valence bands; (b) schematic illustration of a real insulator originating from a degenerate gapped metal. Blue: occupied states; white: unoccupied states. Here, the system invests energy Erec to create in the gap a new state via reconstruction, magnetization, or defect formation. In turn, this state accepts q electrons from the conduction band, contributing to energy lowering qΔE, where ΔE is the difference in band energies for electrons in the conduction band and localized occupied state. This creates a gap between the newly formed state and principal conduction band. Reproduced with permission from Malyi and Zunger, Phys. Rev. B 101, 235202 (2020). Copyright 2020 by the American Physical Society.

Close modal

These prototypical electronic configurations are well known in dilute doped inorganic and organic insulators.7–11 Interestingly, there are also many pristine (undoped) stoichiometric solids, where the electronic structure corresponds to the configuration of degenerate gapped metal portrayed in Fig. 1(a). The most famous examples of such compounds are SrVO3,5 CaVO3,5 BaNbO3,12 and Ca6Al7O16,13 which exhibit metallic behavior based on the analysis of temperature behavior of resistivity5,13,14 and angle-resolved photoemission spectroscopy.15–17 The successful design of the novel degenerate gapped metals lies at the heart of the development of transparent conductors,18 electrides,19 and prediction of Dirac semimetals.20–22 For instance, it has been recently proposed that intrinsic degenerate gapped metals can be used as transparent conductors.23,24 Similarly, many electride13 compounds correspond to bulk solids that have degenerate gapped electronic structures where the electrons form a two-dimensional (2D) or 1D or 0D cloud resembling an intrinsic electron gas.

Spontaneous instabilities: Free carriers introduced via doping to insulators are the source of “doping bottlenecks,”25–27 whereby doping leads to the self-regulating formation of charged structural defects that compensate the intentional doping. But such free carrier instabilities can also occur in pristine compounds having analogous electronic configurations. Figure 1(b) illustrates the fact that such an electronic structure configuration might be unstable toward the formation of an “acceptor state” inside the principal band gap if its formation costs a “reconstruction (rec) energy” |Erec| that is smaller than the energy gained by transferring q electrons from the conduction band to these lower energy acceptor states |qΔE|, where ΔE is the band energy difference for electrons in the conduction/valence band and localized states. In general, spontaneously formed acceptor states may be structural defects (e.g., cation vacancies) as in BaNbO3, Ca6Al7O16, and Ag3Al22O34 and lead to (i) the observed off-stoichiometry even at low temperatures,23,28–31 (ii) reduced metallicity and, at the extreme, even to (iii) the emptying of the conduction band and thus a metal to insulator transition. But such acceptor states may also be electronic defects, rather than structural defects such as polaron that can sweep conduction electrons into its “acceptor” state.32–34 On the other hand, when |Erec| > |qΔE| as shown in Fig. 1(a), the metallic configuration can be stable against reconstructions, and the gapped metal configuration is real. The analogous process illustrated in Fig. 1(b) for n-type solids exists for degenerate but gapped metals that have their Fermi level inside the valence band.

The false metal syndrome: It is our observation that the current literature sometimes reflects confusion between false metals, real insulators, and degenerate gapped metals. We will discuss in the current paper when predictions of metallic configuration, such as those shown in Fig. 1(a) are valid, and when they are false, meaning that the actual electronic structure is that illustrated in Fig. 1(b). The distinction between false metals vs real metals vs real insulators is important because of a number of reasons. Indeed, predicting theoretically that a given compound is metal even though it is found to be an insulator often creates the temptation to invoke a high order novel physical effects (such as correlation in d-electron compounds35–37) to explain the misassignment, rather than searching for a more mundane artifact in a lower-level mean-field band theory. Examples include some Mott insulators such as 3d oxides, where historically naïve band theory35 using nonmagnetic (NM) spin configuration and a minimal unit cell forced, for an odd number of electrons per cell, a false metallic state, in contrast with the experiment. This sharp contradiction with the experiment prompted that the long-lasting tradition of a Hubbard Hamiltonian description of false metals would be Mott insulators that identify many-body correlation effects as the gapping mechanism. However, a number of examples came to light, illustrating that ordinary density functional theory (DFT) free from the restrictions to the nonmagnetic description or its commitment to minimal unit cell sizes already correctly gives an insulating ground state.38–42 This raises the question of what type of physics is responsible for the removal of false metal designation in favor of real insulators. This will be discussed in the present paper.

While recently developed open access databases20,43–45 of electronic band structures constitute an important contribution to materials science, false metals abound such databases (see a few examples in Fig. 2) and literature citing them are discussed below. Indeed, it was recently argued that the prediction of novel Dirac metal (BiO2)46 is unlikely to be valid, as a more proper description31 of the same compound shows that it spontaneously reconstructs to significantly lower energy, creating a trivial insulator structure. Hence, herein, we will show that the leading cause of the “false metal syndrome” is the incomplete application of electronic structure theory, omitting a degree of freedom in the calculation that, if enabled, will convert the false metal into a real insulator, as illustrated schematically in Fig. 1(b).

FIG. 2.

Examples of computational data for false metals available in Materials Project,43 Open Quantum Materials Database (OQMD),45 Automatic-FLOW for Materials Discovery (AFLOW),44 and Topological Materials Database20 with comparison to experimental data. The summary demonstrates that theoretically predicted electronic structures are often in disagreement with the corresponding experimental data. The reasons for the false metal predictions are not absence of electron-electron correlation but generally disallowing some modes of symmetry breaking in DFT. The cases where the insulating nature of compounds has been predicted in the respective literature from calculations of density of states (not from band structure calculations) are shown in italic. Since results in the databases are changing with time, we note that the data quoted here was retrieved on May 28, 2020.

FIG. 2.

Examples of computational data for false metals available in Materials Project,43 Open Quantum Materials Database (OQMD),45 Automatic-FLOW for Materials Discovery (AFLOW),44 and Topological Materials Database20 with comparison to experimental data. The summary demonstrates that theoretically predicted electronic structures are often in disagreement with the corresponding experimental data. The reasons for the false metal predictions are not absence of electron-electron correlation but generally disallowing some modes of symmetry breaking in DFT. The cases where the insulating nature of compounds has been predicted in the respective literature from calculations of density of states (not from band structure calculations) are shown in italic. Since results in the databases are changing with time, we note that the data quoted here was retrieved on May 28, 2020.

Close modal

Figure 3 summarizes the modalities that can lead to the prediction of false metals and will be discussed in the current article. The first category involves oversimplified computational assumptions such as restriction of the unit cell representation to cells that cannot geometrically accommodate low symmetry structures or use of exchange–correlation (XC) functionals that produce rather delocalized orbitals that are unable to take advantage of energy lowering broken symmetries. The second category involves oversimplified physical assumptions that restrict the formation of structural, magnetic, or defect breaking modes that could otherwise transform the configuration in Fig. 1(a) to that in Fig. 1(b). In this case, the initial configuration shown in Fig. 1(a) exists only under hypothetical, theoretical approximations that fail to take advantage of this energy lowering reconstruction.

FIG. 3.

Different causes of false metals and mechanisms of the corresponding band gap opening with examples. Compounds that are observed to be metals at high temperature are marked by a star.

FIG. 3.

Different causes of false metals and mechanisms of the corresponding band gap opening with examples. Compounds that are observed to be metals at high temperature are marked by a star.

Close modal

The paper is organized as follows. Section II provides basic information on the different computational approximations for describing electronic structures of compounds, which can result in the prediction of false metals. Section III gives details on symmetry breaking motifs that should be accounted for during the search of potential degenerate gapped metals, providing detailed examples for each case. In Sec. IV, we illustrate the cases when false metals can create in-gap polaron-like states and their importance. In Sec. V, we discuss the progress on the theoretical and experimental predictions of real degenerate gapped metals. Finally, we provide a short outlook and perspective of the field in Sec. VI.

Different styles of electronic structure theories have different vulnerabilities toward predicting false states of conductivity. Fixed-Hamiltonian band structure methodologies (see textbooks1–4) lack a feedback mechanism that allows the electronic structure and the crystal or spin structure to affect each other. This is the case for standard tight binding,47 k⋅p,48,49 or empirical pseudopotential method,50,51 that selected fixed atomic compositions, crystal structure with specific internal symmetry, unit cell size, and a fixed spin configuration that are not allowed to vary as the electronic structure does. These fixed features end up deciding uniquely the outcome electronic characteristics—whether true or false.

Modern electronic structure modalities based on DFT, on the other hand, allow the electronic structure to respond to the different occupation numbers, atomic positions, spin configurations, unit cell size, and symmetries, variables that are being explored during the calculation in search of a minimum energy self-consistent solution. This provides a feedback loop between the dynamic variables of the calculationI—{occupation numbers, atomic positions, spin configuration, cell symmetries}all able to change during the solution process, and the ensuing electronic structure II (stable atomic and magnetic structure, energies of states, and their localization). The agents mediating the effects of the variables of the calculation I on the ensuing electronic structure II are the well-known heuristic constructs of traditional solid-state chemistry such as bonding, charge transfer, and hybridization. An important manifestation of this feedback response, for example, is that electron addition (to empty states) and electron removal (from occupied states) can change the underlying difference in their energies, i.e., the band gap, transforming, for example, a (false) metal to a (real) insulator. We will explain in this article that this crucial nonlinear feedback between I and II is a property already present in properly formulated mean-field DFT (but not of unresponsive fixed-Hamiltonian band structure methodologies noted above), rather than the exclusive property of many-body dynamic correlation effects, an opinion as very often echoed in the literature. Thus, compounds such as NiO or cuprate superconductors that were depicted as (false) metals in the literature using naïve fixed Hamiltonian band structure models became real insulators using properly executed (symmetry-broken) DFT. The present article will illustrate numerous mechanisms (Fig. 3) where the above noted feedback loop underlying mean-field theories depicts correct symmetry broken insulation without recourse to strongly correlated symmetry preserving treatment.

Figure 4 illustrates the basic makeup of such electronic structure methods. For a given compound defined by its Atomic identities, Composition, and Structure type (ACS for short52), there are two types of inputs: one that is fixed (the inner circle) and one that changes iteratively during the calculation (outer circle). The fixed input contains the definition of the compound explored, i.e., ACS. This is encoded in the atomic numbers, pseudopotential, and the crystalline phase of interest. These quantities appear in the crystal electron–ion potential Vext(r), which is the changing part of the input. In addition, the fixed part also contains the type of screening Vscr(r) allowed for the external potential. In the context of DFT, this refers to the XC potential.53The changing part consists of parameters that are affected by the evolving electronic structure and the need to be optimized as the calculation goes forward. This includes energy-minimizing structural parameters consistent with the physical phase ACS being explored. Such parameters include the self-consistent charge and spin densities, atomic position parameters obtained via minimization of atomic forces, the size of cell (primitive or supercell), and type of spin configuration. These quantities can change during the calculation in order to find the lowest total energy. Different levels of computational sophistication do this, i.e., either automatic variations seeking lower total energy54–57 or “manually,” exploring discrete geometries. This framework allows coupling (i) the atomic structure and spin configuration to the (ii) charge density and potential, and hence can change the ensuing band structure dramatically as the (i) and (ii) feedback unfolds iteratively during self-consistency and ionic relaxation. Herein, the calculations are performed using a plane-wave pseudopotential density functional method as implemented by the Vienna Ab Initio Simulation Package (VASP)58–60 with the results visualized using Vesta.61 As we will see in this article, this strong nonlinear interrelation between structure vs band structure brings new possibilities of classification and, indeed, misclassification of compounds as metals or as insulators, depending on how this coupling is described.

FIG. 4.

Schematic illustration of electronic structure calculations using two types of input to establish the single-particle Schrödinger equation. The inner circle is the self-consistency of the charge density for a given Vext(r) and a given spin configuration, whereas the outer circle involves changing the geometry [i.e., Vext(r)] and the spin configuration in search of lower total energy.

FIG. 4.

Schematic illustration of electronic structure calculations using two types of input to establish the single-particle Schrödinger equation. The inner circle is the self-consistency of the charge density for a given Vext(r) and a given spin configuration, whereas the outer circle involves changing the geometry [i.e., Vext(r)] and the spin configuration in search of lower total energy.

Close modal

Given the computational framework described in Fig. 4 that permits validation, we can now return to Fig. 3 and examine the type of approximations beyond what Fig. 4 affords, checking which of these additional approximations limits the full extent of coupling between the electronic structure and atomic structure/spin configuration, thus leading to false metal designations.

From the computational DFT perspective, the description of the electronic structures of compounds requires the utilization of XC functional depicted in Fig. 4 as Vscr. There are, by now, a large number of candidate XC functionals,62 fitting different aspects. But recently, the examination of the minimal physics needed to correctly describe metal vs insulator states in both correlated d-electron and s-p electron systems38,39,63 suggests two major conditions that need to be satisfied: an ideal XC functional should (i) be able to distinguish occupied from unoccupied states and (ii) have reduced the self-interaction error (SIE).64,65 Indeed, the wide range of modern XC functional [e.g., local density approximation (LDA),64 Perdew–Burke–Ernzerhof (PBE),66 PBEsol67] suffer from the SIE arising from the spurious interaction of an electron with itself. Since the repulsive self-Coulomb interaction exceeds the corresponding attractive self-exchange-correlation, the net SIE is usually positive (repulsive toward electrons), resulting in excessively delocalized wave functions and very high (overly unbound) orbital energies. The tendency to delocalization due to excessive SIE in a given XC potential can be monitored by computing the degree to which the total energy deviates from linearity (the generalized Koopman's condition65) as a function of non-integer occupation number. This is a valid “shopping criteria” for selecting an XC functional that can produce spatially compact orbitals with reduced SIE, illustrated in Ref. 32. If this function bows down significantly below the linear Koopman's result, we refer to such XC functional as “soft,” else it is harder. Meeting the generalized Koopman's condition enforces the minimization of SIE; the SIE can lead to the underestimation of the band gap energy and even artificially stabilize the non-symmetry-broken configurations. This problem can be fixed by utilizing a self-interaction correction (SIC) discussed by Zunger and Freeman68,69 and by Perdew and Zunger,64 or XC functionals that have reduced self-interactions such as SCAN,70 DFT+U,71–74 or Heyd–Scuseria–Ernzerhof (HSE)75,76 functional. For instance, the SIC applied within DFT for CrO, FeO, CoO, and CuO has been shown to result in band gap opening.77 A similar tendency has also been demonstrated by DFT+U38,39 and hybrid functional calculations78 for 3d oxides. For instance, Fig. 5(a) illustrates here a case of antiferromagnetic (AFM) LaVO3, where the choice of a soft XC potential that deviates too much from the linearity condition can produce a false metal, whereas an XC functional with more reduced SIC gives an insulator [Fig. 5(b)].

FIG. 5.

(a) Using soft XC PBEsol functional (no U), one predicts a (false) metal; (b) utilization of hard XC PBEsol+U functional (U = 3.5 eV applied on V-d) in band gap opening for the experimentally observed AFM-C LaVO3 structure. In both calculations, the lattice vectors and atomic positions are fixed at equilibrium PBEsol values. The occupied states are shadowed in light blue. The band gap is shown in gray. SG denotes the space group number. We use the term “soft XC” for functionals that have a pronounced downward bowing total energy with respect to occupation number, while the term “hard XC” represents more self-interaction corrected functionals that more closely approach the linearity condition.

FIG. 5.

(a) Using soft XC PBEsol functional (no U), one predicts a (false) metal; (b) utilization of hard XC PBEsol+U functional (U = 3.5 eV applied on V-d) in band gap opening for the experimentally observed AFM-C LaVO3 structure. In both calculations, the lattice vectors and atomic positions are fixed at equilibrium PBEsol values. The occupied states are shadowed in light blue. The band gap is shown in gray. SG denotes the space group number. We use the term “soft XC” for functionals that have a pronounced downward bowing total energy with respect to occupation number, while the term “hard XC” represents more self-interaction corrected functionals that more closely approach the linearity condition.

Close modal

One of the computational inputs in performing periodic band calculations is the lattice vectors and internal atomic positions defining the unit cell. Tradition has it that one would use the smallest unit cell, generally as obtained from x-ray diffraction plus Rietveld refinement as, for instance, listed in Inorganic Crystal Structure Database (ICSD).79 The latter technique, however, has a large coherence length and thus averages over large volumes, sometimes omitting symmetry-lowering details of the local environment, as seen by experimental techniques with shorter coherence length (x-ray absorption fine structure, pair distribution function, and Raman scattering, to mention a few). Electronic structure methods such as those illustrated in Fig. 4 may also be sensitive to local symmetries and magnetic order, but the use of a unit cell with an overly restricted size leading to a high average symmetry may not allow this. This problem is well known in alloy theory, where a virtual crystal approximation to a substitutional A1−xBx alloy considers an artificial average structure where each atom is replaced by a virtual ⟨AB⟩ atom, thus artificially raising the symmetry and consequently inhibiting degeneracy removal and local relaxation that are physical effects observed when a larger unit cell is explored.80 Indeed, there is no general theorem or explanation arguing that the minimal (highest symmetry) unit cell is somehow physically valid.

A simple, common-sense computational test to validate the choice of a unit cell size for a given symmetry is to compute the total energy per atom for a given global cell symmetry but possibly different cell-internal atomic relaxation and spin structures, as a function of (super)cell sizes, and observing if the energy per atom is constant or decreasing. In the latter case, one might find a polymorphous network,81 whereby certain local structural features (such as atomic displacements and octahedral tilting) or spin configurations show a distribution of such local motifs, rather than a single sharp value. To computationally afford the opportunity of examining if symmetry breaking lowers the total energy, it is important to “nudge” the atomic positions so as to dislodge atoms from possible local minima and to avoid the practice of wave function symmetrization by equally occupying partner states of a degenerate level [i.e., do not use (0.5;0.5) for an e-level occupied by one electron, but rather (1;0)].

It turns out that each of these symmetry-breaking modalities can result in energy lowering, leading to local symmetry breaking. The electronic structure can react to the existence of such distributions, even changing from a false metal to real insulator, as will be shown below. Indeed, if a system has different local environments (structural or spin) {Si;i = 1, N}, then its physical property P (band gap, moments, and others) cannot be approximated as the properties ⟨P⟩ = P(S0) of the macroscopically averaged monomorphous structure S0, instead of the correct average Pobs = ΣP(Si) of the properties {P(Si)} of the individual, low symmetry microscopic configurations.38–40,81,82

Using large supercells and performing minimization with respect to cell internal atomic displacements establishes local displacements. There are two types of displacements in this discussion:

(a) Local intrinsic displacements: These arise from the intrinsic preference of chemical bonding (like the bent H–O–H bond angle in water,83 or electronically mandated static Jahn–Teller distortion,84 or due to a steric preference such as tilted octahedra85). This type of displacements exists even at the lowest temperatures wherein the question exists. Intrinsic displacements can be predicted quantum mechanically from the minimization of enthalpy without entropy contributions. They reflect symmetry breaking, such as symmetry lowering off-center atoms. These displacements do not spatially average to zero. As long as there is some length scale of ordering, one will get something finite in the bulk limit.

(b) Local displacements induced by thermal motion: Such displacement represents movements about the low T equilibrium geometry and can be simulated by stochastic movements (i.e., molecular dynamics and Monte-Carlo), causing Urbach tails in absorption. However, if this disorder is truly random, uncorrelated with no form of short or long-range order, then for an infinitely large unit cell, the average of such displacements gets zero weight scattering intensity.

The relevance of this discussion on false metals is twofold:

First, we will see that degenerate gapped metals tend to become real insulators when intrinsic symmetry breaking removes band degeneracies. Conversely, such systems can stay as metals due to the absence of symmetry breaking. Such absence can occur in two ways: (i) at low-temperature “intrinsic metals” such as SrVO3, where the chemical bonding does not require atomic displacements (as seen in ABO3 compounds with a Goldschmidt tolerance factor86 near one, that are stable in an ideal, undisplaced cubic phase). This gives real metals even at low T. (ii) Absence of symmetry breaking can also occur at rather high temperatures where strong thermal motions erase the intrinsic displacements. This can give metallic states at high T as in the cases of BiBaO387,88 and SmNiO3.89,90 This illustrates once again the relationship between structural symmetry breaking as a cause of insulating behavior and its absence as the illustration of metallic behavior. Calculations that attempt to simulate the low-temperature phases by ignoring structural symmetry breaking predict false metals for such phases.

Second, if one insists on no symmetry breaking, then the XC functional should have an explicitly discontinuous (non-differentiable) dependence on the density or density matrix that accounts for spin degeneracies and might predict band gaps. Such discontinuity is missing from all current practical XC approximations. Thus, using any of the current XC functionals without symmetry breaking polymorphous representation often does not open band gaps. This paper will illustrate in Sec. III such modalities of symmetry breaking that transform false metals into real insulators.

This section discusses illustrative examples of cases where the omission of local spin motives (Secs. III A and III B), or local positional motives (Secs. III C–III E), or spin–orbit coupling (SOC) (Sec. III F) leads to the designation of a compound as a false metal, whereas removal of such artificial constraints reveals they are insulators.

In simplified calculations, compounds are described using NM configurations. However, forcing the NM solution can create artificial stabilization of the false metal case, as illustrated schematically in Fig. 6(a). For example, a nonmagnetic scenario for CuBi2O4 was used in Refs. 91 and 92 to find a metallic phase having a partially occupied intermediate band (IB) in the principal gap, showing an eightfold band degeneracy at EF [Fig. 6(b)]. Analogously, the nonmagnetic scenario for NiO [Fig. 6(c)] obtained by assuming a small unit cell of 1 formula unit (f.u.) gives a p-type false metal.20 Such result led historically N. Mott to deduce that 3d oxides with an odd number of valence electrons must be metallic in standard band theory.35 However, experimentally both compounds are insulators. While the above studies are only simple illustrations of considering nonmagnetic systems, the nonmagnetic calculations are not rare. For instance, despite groundbreaking studies, recently developed topological databases20–22 are limited to the NM calculations. In practice, one does not need to guess if a material is NM or magnetic; in most cases, one can use the fact that DFT comes with its intrinsic total energy expression93 and compare the total energies of magnetic vs NM solutions and pickup the lowest energy one.

FIG. 6.

Assuming a nonmagnetic scenario, [(a)–(c)] one predicts a (false) metal; allowing spin-order [(d)–(f)] results in the formation of an insulator. (a) Schematic illustration of the electronic structure of a degenerate gapped metal assuming no spin order. (b) Actual calculation for a nonmagnetic CuBi2O4 band structure (using PBE+U with U = 6 eV for Cu-d states) gives a false n-type gapped metal. (c) Actual calculation of nonmagnetic NiO (using SCAN) in a primitive cell containing one formula unit predicting a false p-type gapped metal. (d) Schematic illustration of gap opening due to spin order resulting in moving electrons from the conduction band to lower unoccupied orbitals shown by the arrow. (e) and (f) Actual calculations of CuBi2O4 and NiO allowing AFM magnetic order, showing both systems are insulators. Occupied states are shadowed in light blue. The band gap is shown in gray. SG denotes the space group number. The figures for CuBi2O4 are redrawn using data from Ref. 31.

FIG. 6.

Assuming a nonmagnetic scenario, [(a)–(c)] one predicts a (false) metal; allowing spin-order [(d)–(f)] results in the formation of an insulator. (a) Schematic illustration of the electronic structure of a degenerate gapped metal assuming no spin order. (b) Actual calculation for a nonmagnetic CuBi2O4 band structure (using PBE+U with U = 6 eV for Cu-d states) gives a false n-type gapped metal. (c) Actual calculation of nonmagnetic NiO (using SCAN) in a primitive cell containing one formula unit predicting a false p-type gapped metal. (d) Schematic illustration of gap opening due to spin order resulting in moving electrons from the conduction band to lower unoccupied orbitals shown by the arrow. (e) and (f) Actual calculations of CuBi2O4 and NiO allowing AFM magnetic order, showing both systems are insulators. Occupied states are shadowed in light blue. The band gap is shown in gray. SG denotes the space group number. The figures for CuBi2O4 are redrawn using data from Ref. 31.

Close modal

The physical effect that will stabilize an insulating state: magnetic order:Fig. 6(d) shows schematically how magnetic order can result in band gap opening by creating the empty seats for conducting electrons resulting in the insulator. Such behavior is well-known for AFM compounds where, e.g., doubling the unit cell and allowing magnetic moment formation lead to the band gap opening even in simple band theory.94 This is indeed the case for both NiO and CuBi2O4 [Figs. 6(e) and 6(f)]. Thus, despite the eightfold band degeneracy at EF of the hypothetical nonmagnetic CuBi2O4,91,92 the lowest total energy magnetically ordered configuration is an insulator with a large band gap energy of 1.64 eV according to the PBE+U calculations.31 Similarly, NiO is an insulator with a band gap energy of 2.42 eV according to magnetic SCAN calculations. Since the nonmagnetic and AFM structures are identical in atomic positions, the band gap opening in both compounds originates from magnetization only. It should be noted that AFM solutions for both compounds are significantly lower in energy compared to nonmagnetic false metal (i.e., by 0.5 eV/f.u. and by 1.33 eV/f.u. for CuBi2O431 and NiO,63 respectively); hence, it is clear that metallic electronic structures are not likely to be realized experimentally in these compounds.

The experimental situation is indeed that NiO is the AFM insulator, which has the optical band gap energy of 3.68 eV.95 This system is often discussed as a Mott insulator where the gapping is caused by electron–electron repulsion.96 However, as discussed by Zhang et al.,63 the properties of the compound can be described with non-empirical XC density-functional (i.e., SCAN) without an on-site interelectronic repulsion, i.e., U =0 eV. CuBi2O4 is also a wide band gap magnetic insulator, which recently attracted significant attention for catalysis,97 thus suggesting that eight-band fermions near the Fermi level found in hypothetically nonmagnetic CuBi2O4 are not likely to be realized.

Paramagnetic (PM) compounds have non-zero local but zero total magnetic moments. Until recently, the properties of such systems have been explored as properties of globally average nonmagnetic structures [Fig. 7(a)],14,92,98–101 leading invariably to metallic prediction in contrast with the known insulating properties of many if not most PM ABO3 phases. Because of this, there has been a long-term belief that many properties of PM systems cannot be described within DFT methodology, and higher-order methods [e.g., dynamical mean-field theory (DMFT)37,102] should be applied to get the right result of the insulating phase. However, as discussed in Sec. II B, the properties of a globally average structure should not necessarily be the same as the properties of the system with non-zero local but zero total magnetic moments. To illustrate the limitation of such a naïve PM model equating it with a NM phase, we consider paramagnetic LaTiO3 (SG: 62) and YTiO3 (SG: 62) systems. As shown in Fig. 7(b), assuming a NM scenario, both compounds are n-type degenerate gapped metal with 1 e/f.u. in the conduction band and large separation between the principal valence and conduction bands, as reported in Refs. 20–22,103, and 104. However, experimentally, LaTiO3 and YTiO3 are insulators.105–108 

FIG. 7.

When a globally average (nonmagnetic) structure is used with a small unit cell, [(a) and (b)] one predicts a (false) metal; allowing a large unit cell with the spin disorder [(c) and (d)] results in the formation of an insulator. (a) Schematic depiction of naïve approximation of the paramagnetic compound using the nonmagnetic configuration resulting in the false metal state. (b) Actual calculations of the density of states for LaTiO3 and YTiO3 within the naïve approximation. (c) Schematic depiction of a spin-disordered large unit cell model of a paramagnetic compound resulting in band gap opening due to moving electrons from the conduction band to lower unoccupied orbitals as shown by the arrow. (d) Actual calculations of density of states for LaTiO3 and YTiO3 computed using for a 160-atom spin-disordered supercell. All results are presented for PBEsol+U calculations with a U value of 2.5 eV applied on Ti-d states. Occupied states are shadowed in light blue. The band gap is shown in gray.

FIG. 7.

When a globally average (nonmagnetic) structure is used with a small unit cell, [(a) and (b)] one predicts a (false) metal; allowing a large unit cell with the spin disorder [(c) and (d)] results in the formation of an insulator. (a) Schematic depiction of naïve approximation of the paramagnetic compound using the nonmagnetic configuration resulting in the false metal state. (b) Actual calculations of the density of states for LaTiO3 and YTiO3 within the naïve approximation. (c) Schematic depiction of a spin-disordered large unit cell model of a paramagnetic compound resulting in band gap opening due to moving electrons from the conduction band to lower unoccupied orbitals as shown by the arrow. (d) Actual calculations of density of states for LaTiO3 and YTiO3 computed using for a 160-atom spin-disordered supercell. All results are presented for PBEsol+U calculations with a U value of 2.5 eV applied on Ti-d states. Occupied states are shadowed in light blue. The band gap is shown in gray.

Close modal

The physical effect that will stabilize an insulating state: formation of a spin-disordered supercell: Allowing the atom to have a non-zero magnetic moment on each site can result in energy lowering and band gap opening [Fig. 7(c)] due to moving the conduction electrons to the lower energy level. It has been shown that modeling the paramagnetic compound as a spin-disordered system (each atom has non-zero spin, but the spins in the system are disordered) can be used to effectively reproduce experimental properties of binary38,63 and ternary39,40 paramagnetic compounds within the DFT calculations. These studies thus demonstrated that it is possible to model paramagnetic systems with DFT calculations despite the counter, long-term beliefs. Application of this model to LaTiO3(SG: 62) and YTiO3(SG: 62) results in insulators [Fig. 7(d)] with PBEsol+U band gap energies of 0.89 and 1.24 eV, respectively. As will be discussed in Sec. IV, the formation of an insulator is accompanied by electron localization on Ti atoms, resulting in e-trapped states.

The experimental situation: Paramagnetic LaTiO3(SG: 62) and YTiO3(SG: 62) are insulators as reported by multiple studies and confirmed by both temperature dependence of resistivity and photoemission data.105–108 These results thus prove that prediction of metals for both LaTiO3 and YTiO3 is the artifact of the nonmagnetic PM model, while the spin-disordered approximation can successfully represent the main experimental predictions for the PM phases, which is in line with that reported by Varignon et al.39,40

In which type of compounds would this insulator stabilization by a spin-disordered supercell occur? Since the model of spin-disordered systems is relatively new, the set of compounds where the band gap opening has been observed is currently limited to binary MnO, FeO, CoO, and NiO38,63 and a set of 3d ABO3 oxides.39,40 Specifically, it has been demonstrated that the consistent way to calculate the properties of PM phase as the polymorphous statistical average over the ensemble of microscopic configurations and not the properties of macroscopically average structure (see Sec. II B).

We close this subsection by discussing the manner in which our description of a PM phase constitutes a generalization of the well-known low-temperature spin ordered phase. The low-temperature AFM phase generally inherits some of the properties of the PM phase. The simple illustration of such behavior is the comparison of electronic properties of AFM and PM phases of YNiO3(SG: 14), showing that both phases have similar band gap energies of 0.59 and 0.49 eV [Fig. 8], respectively, according to PBEsol+U calculations. In contrast, the NM YNiO3(SG: 62) phase is a degenerate gapped conductor with a partially occupied intermediate gap state, which has higher total energy, about 0.13 eV/f.u. above both AFM and spin-disordered PM phase. Indeed, the low-temperature AFM phase can have spin and positional local motifs that are a subset of those in the high-temperature phase. For example, the AFM phase can have a single spin motif—each spin-up is coordinated by all spin-down sites, whereas the PM phase represents a generalization, having a distribution of local motifs where each spin-up is locally coordinated by m spin-up and N-m spin-down, where N is the coordination number and 0 < m < N. Similarly, the high-temperature phase can have a distribution of local displacements resembling geometrically the global structural motif that is responsible for a long-range order in the low-temperature ground state. An example is the high-temperature trigonal phase of FeSe having locally orthorhombic distortions mimicking the globally orthorhombic stable low-temperature phase.82 In general, the transition from high-temperature PM to the low-temperature ordered phase involves ordering vectors that select from the PM phase certain spin and structural motifs that are stabilized.

FIG. 8.

Comparison of electronic structures for (a) naïve nonmagnetic, (b) antiferromagnetic, and (c) 160-atom spin-disordered paramagnetic phases of YNiO3. The results demonstrate that electronic structure of the spin-disordered paramagnetic phase of YNiO3 is closer to antiferromagnetic order than to the naïve nonmagnetic approximation. Results are presented for the PBEsol+U exchange correlation functional with a U value of 2 eV applied to Ni-d states. Occupied states are shadowed in light blue. The band gap is shown in gray.

FIG. 8.

Comparison of electronic structures for (a) naïve nonmagnetic, (b) antiferromagnetic, and (c) 160-atom spin-disordered paramagnetic phases of YNiO3. The results demonstrate that electronic structure of the spin-disordered paramagnetic phase of YNiO3 is closer to antiferromagnetic order than to the naïve nonmagnetic approximation. Results are presented for the PBEsol+U exchange correlation functional with a U value of 2 eV applied to Ni-d states. Occupied states are shadowed in light blue. The band gap is shown in gray.

Close modal

In traditional first-principles calculations, compounds are usually described with the smallest possible primitive cells where each species is often represented via a so-called single local environment (SLE), as shown schematically in Fig. 9(a). While this approach allows a proper description of the properties of some compounds, it does not guaranty that the formation of structurally inequivalent Wyckoff positions does not lower the energy of such structures. This can lead, for instance, to a double local environment (DLE), where the chemical bonding can mandate a larger unit cell. Examples of the consequence of assuming SLE behavior are shown in Fig. 9 for monoclinic TiO2 [space group 12, a structure often denoted as “(B)”109] doped with Li interstitial and for cubic SrBiO3 (SG: 221). The consequences of this assumption are that the former case has the EF in the conduction band with 1 e/f.u., whereas the latter case has the EF in the valence band with 1 h/f.u., as shown in DFT calculations in Figs. 9(b) and 9(c), respectively. However, experimentally, both systems are insulators.110,111 There are a number of cases when using such approximation invalidates the proposed new functionality, e.g., Nanda et al.112,113 suggested cubic SLE SrBiO3 as a topological compound, but this structure is not the stable phase.

FIG. 9.

Assuming a single local environment (SLE), [(a)–(c)] one predicts a (false) metal; allowing double local environment (DLE) [(b)–(f)] results in the formation of an insulator. (a) Schematic illustration of degenerate gapped metal having a single local environment. (b) HSE calculation of density of states for a nonmagnetic SLE 16 f.u. supercell of monoclinic TiO2 (SG: 12) containing a Li interstitial atom demonstrating that it is an n-type degenerate gapped false conductor with Fermi level in the conduction band. (c) PBE+SOC calculation for SLE cubic SrBiO3 (SG: 221) showing that it is a p-type degenerate gapped false conductor with Fermi level in the valence band. (d) Schematic illustration of degenerate gapped metal having a double local environment resulting in band gap opening due to moving electrons from the conduction band to lower unoccupied orbitals as shown by the arrow. (e) HSE density of states for 16 f.u. supercell of monoclinic TiO2 (SG: 12) containing a single Li interstitial showing that the system is an insulator with e-trapped intermediate band caused by localization of electron on the part of Ti sublattice—formation of DLE. Occupied states are shadowed in light blue. The figure is redrawn using data from Ref. 114. (f) PBE+SOC band structure for DLE monoclinic SrBiO3 (SG: 14) showing that the compound is an insulator with band gap energy (shown in gray) of 0.26 eV. SG denotes the space group number.

FIG. 9.

Assuming a single local environment (SLE), [(a)–(c)] one predicts a (false) metal; allowing double local environment (DLE) [(b)–(f)] results in the formation of an insulator. (a) Schematic illustration of degenerate gapped metal having a single local environment. (b) HSE calculation of density of states for a nonmagnetic SLE 16 f.u. supercell of monoclinic TiO2 (SG: 12) containing a Li interstitial atom demonstrating that it is an n-type degenerate gapped false conductor with Fermi level in the conduction band. (c) PBE+SOC calculation for SLE cubic SrBiO3 (SG: 221) showing that it is a p-type degenerate gapped false conductor with Fermi level in the valence band. (d) Schematic illustration of degenerate gapped metal having a double local environment resulting in band gap opening due to moving electrons from the conduction band to lower unoccupied orbitals as shown by the arrow. (e) HSE density of states for 16 f.u. supercell of monoclinic TiO2 (SG: 12) containing a single Li interstitial showing that the system is an insulator with e-trapped intermediate band caused by localization of electron on the part of Ti sublattice—formation of DLE. Occupied states are shadowed in light blue. The figure is redrawn using data from Ref. 114. (f) PBE+SOC band structure for DLE monoclinic SrBiO3 (SG: 14) showing that the compound is an insulator with band gap energy (shown in gray) of 0.26 eV. SG denotes the space group number.

Close modal

The physical effect that will stabilize an insulating state is creation of different local environments:Fig. 9(d) illustrates how the formation of different local environments in a degenerate gapped metal can result in band gap opening and total energy lowering. This turns to be the case for both TiO2:Li and cubic SrBiO3, which both spontaneously disproportionate to lower energy structures when symmetry breaking is allowed. The false metal Li-doped monoclinic TiO2 reconstructs to lower energy insulator (with a HSE band gap energy of 1.12 eV) having an e-trapped intermediate band [Fig. 9(e)]. As will be discussed in Sec. IV, the formation of the e-trapped state and the band gap opening are due to the ability for Ti atoms to change the formal oxidation state from Ti4+ to Ti3+. This is evidenced by concomitant structural changes where clearly district local environments are formed—the average Ti4+–O bond length is 2.02 Å, while the corresponding value for the Ti3+–O bond is 2.09 Å. The same tendency is observed in the SrBiO3(SG: 221) supercell that spontaneously disproportionates to monoclinic SrBiO3(SG: 14),111 where Bi has a double local environment, lowering the system energy by 0.14 eV/atom. The resulting system is an insulator with the PBE+SOC band gap energy of 0.26 eV having a h-trapped intermediate band [Fig. 9(f)]. This reaction is caused by the ability of Bi to disproportionate to Bi3+ and Bi5+,115–117 which is confirmed by structural analysis; monoclinic SrBiO3 (SG: 14) contains structurally different Bi atoms with the average Bi–O bond lengths of 2.17 and 2.34 Å. One should note, however, that herein FOS is used as a label without assigning any physical meaning. Indeed, FOS does not describe the charge density distribution in solids due to “self-regulating response”118—a change of cation charge is usually counteracted by opposing change on the ligand.

The experimental situation is that the formation of e-trapped states originated from reduced Ti atom in TiO2:Li systems are well-known phenomena discussed by different groups110,114,119–121 as the cause for low electronic conductivity of TiO2:Li systems formed during lithiation of the TiO2-based electrode. Experimental investigations also confirm that despite numerous studies reported on the cubic-like structure of SrBiO3 (SG: 221),111,116 it always exists in a monoclinic structure (SG: 14). Moreover, as indicated above, the compound is always found to be an insulator, and the formation of structurally different Bi atoms has been often referenced as Bi3+ and Bi5+.116 These results thus prove that cubic SrBiO3 is an example of false metal caused by forcing the SLE in the compound that exhibits spontaneous energy lowering by bond disproportionation and formation of the DLE phase.

In which type of compounds would this insulator stabilization occur? The DLE behavior for compounds being a degenerate gapped n-type metal in the precursor SLE state is common for systems having cations which can exist in different oxidation states (e.g., Ti4+/Ti3+, Ce4+/Ce3+), where the splitting between the orbital energies of the two FOS is large. TiO2−x, CeO2−x, and V2O5−x are well-known examples of insulators having different local environments and e-trapped in-gap states localized on the reduced metal ions.120,122,123 This set of examples can be further extended to a wide family of ternary early transition metal oxides where the sum of FOS differs from zero (e.g., TiO2:Li).110,114,119 The DLE behavior for compounds being a degenerate gapped p-type metal in the precursor SLE state can be found either in (i) nonmagnetic insulators that include high-Z elements (e.g., BaBiO3,87,115,124 SrBiO3,117 CsTlF3,115 CsAuCl3,115 and CsTe2O6115) or in (ii) compounds having band gap opening as a result of the superposition of disproportionation and magnetism. Classic examples of such materials are rare-earth nickelates (ANiO3 with A = Sm, Eu, Y, and Lu39,115,125) and CaFeO3115. It is interesting that disproportionation in d-electron compounds such as SmNiO3 was initially described as an effect enabled specifically by electron correlation.126 However, the disproportionated structure was, in fact, also obtained in a straight DFT calculation based on broken symmetry,125 reflecting the physics of broken symmetry as indicated here.

The degeneracy-removing electronically-induced Jahn–Teller Q2 distortion as well as the structurally induced (e.g., by steric effects) pseudo-Jahn–Teller Q+2 distortions are known to exist in numerous perovskite compounds,127 but were often ignored when in describing metallic or insulating behavior. Simple approximations often assume high symmetry cubic structures that cannot geometrically accommodate such local symmetry lowering distortions. This assumed cubic structure is especially popular in the description of the Mott insulators, where it has been widely used for DMFT Hamiltonian mapping128 in the potential degenerate gapped metals at least until recently. For the case of potential degenerate gapped metals, forcing the cubic symmetry can artificially move EF to CB or VB [Fig. 10(a)]. LaMnO3 is an example of such behavior: it is a metal with EF in the CB and 1 e/f.u. free carriers if its pseudo-Jahn–Teller distortion is ignored [Fig. 10(b)]. These results conflict with available experimental data showing that the compound is an AFM insulator with a wide band gap at low temperature.129 

FIG. 10.

Ignoring the pseudo-Jahn–Teller Q+2 distortion, [(a) and (b)] one predicts for LaMnO3 a metal; [(c) and (d)] allowing energy lowering Q+2 distortion results in the formation of an insulator. (a) Schematic illustration of ideal cubic ABO3 structure having all octahedra identical with equivalent Mn-O bond length for each metal with Fermi level in the conduction band. (b) Actual calculation (using the SCAN XC) for cubic AFM LaMnO3 demonstrating that it is an n-type degenerate gapped false conductor with Fermi level in the conduction band. (c) Schematic illustration of Q+2 symmetry breaking in cubic ABO3 structure resulting in inequivalent B–O bond lengths and band gap opening due to moving electrons from the conduction band to lower unoccupied orbitals as shown by the arrow. (d) Actual calculation (using the SCAN XC) of band gap opening in AFM LaMnO3 as a result of Q+2 distortion. The band gap is shown in gray. SG denotes the space group number. The figure is redrawn using data from Ref. 130.

FIG. 10.

Ignoring the pseudo-Jahn–Teller Q+2 distortion, [(a) and (b)] one predicts for LaMnO3 a metal; [(c) and (d)] allowing energy lowering Q+2 distortion results in the formation of an insulator. (a) Schematic illustration of ideal cubic ABO3 structure having all octahedra identical with equivalent Mn-O bond length for each metal with Fermi level in the conduction band. (b) Actual calculation (using the SCAN XC) for cubic AFM LaMnO3 demonstrating that it is an n-type degenerate gapped false conductor with Fermi level in the conduction band. (c) Schematic illustration of Q+2 symmetry breaking in cubic ABO3 structure resulting in inequivalent B–O bond lengths and band gap opening due to moving electrons from the conduction band to lower unoccupied orbitals as shown by the arrow. (d) Actual calculation (using the SCAN XC) of band gap opening in AFM LaMnO3 as a result of Q+2 distortion. The band gap is shown in gray. SG denotes the space group number. The figure is redrawn using data from Ref. 130.

Close modal

The physical effect that will stabilize an insulating state is Q+2 distortions: The deviation from 1 of the Goldschmidt tolerance factor can lead to different energy lowering reconstructions resulting in a change of electronic properties. For the case of LaMnO3, the compound exhibits spontaneous Q+2 distortions127,130,131 resulting in band gap opening [see Figs. 10(c) and 10(d)]. Specifically, while cubic LaMnO3 has zero rotation angle and all Mn–O bonds are equivalent to each other, in the distorted orthorhombic structure, there is octahedra rotation and non-equivalent Mn–O bonds—in a cubic structure, the Mn–O bond length is 1.94 Å, while for the orthorhombic one, Mn–O bond lengths are 1.91, 1.97, and 2.21 Å. According to the SCAN calculations, LaMnO3 is an insulator having an e-trapped intermediate band occupied by 1 e/f.u. and a band gap energy of 0.81 eV. Importantly, this band gap opening is observed without using any electron–electron repulsion U value, suggesting that the true origin of the band gap opening is energy lowering structural symmetry breaking. The detailed analysis of such symmetry breaking and its effect on the electronic structure has been recently documented by Wang et al.,130 who also noted a significant change of effective masses in the compound as the result of symmetry lowering.

The experimental situation: Orthorhombic LaMnO3 is an AFM-A insulator having clear distortions with respect to the ideal cubic form.129,132 According to previous experimental data,129,132 the compound is expected to be cubic at high temperature, which indeed should be a metallic system with one of the highest known concentration of free carriers for degenerate gapped metals [e.g., Fig. 10(b)]. However, it has been shown that cubic LaMnO3 is unrealizable as despite having a cubic-like structure at high temperature, the MnO6 octahedra are still noticeably distorted locally as demonstrated by Qiu et al.133 It should be noted that at high temperature, the LaMnO3 can have the metallic-like behavior based on the analysis of resistivity vs temperature.134 This phenomenon is still not completely understood yet. One should note that while LaMnO3 discussed herein is the example of Q+2 distortion resulting in band gap opening, there is the range of other possible octahedra distortion mechanisms discussed in the literature as the cause of band gap opening. For instance, LaTiO3,135 KCrF3,136 and KCuF3137 exhibit the different Q2 distortions.

The formation of nonstoichiometric compounds is usually attributed to a growth effect rather than to a thermodynamically mandated specific instability. Hence, spontaneously defected compounds are usually neglected in many theoretical and experimental studies. This can lead to an incorrect prediction of a degenerate gapped metal. Figures 11(a) and (b) show the calculations for stoichiometric defect-free Ba4As3 and Ag3Al22O34 having a large internal gap and EF in principal valence and conduction bands, respectively. However, these compounds have never been realized experimentally under the stoichiometric conditions. All attempts to synthesize metallic Ba4As3 result in the formation of nonstoichiometric Ba4As3−x having insulating properties.138 The similar tendency is also found in attempts to synthesize Ag3Al22O34.23 Although both these compounds have not been widely studied theoretically, they attracted some attention. Specifically, stoichiometric Ag3Al22O34 has been predicted as a potential intrinsic transparent conductor,24 while stoichiometric Ba4As3 is shown in the Materials Project database43 as a potential gapped metal. Moreover, Ba4As3 represents a wide family of potential false metals (e.g., Ba4Bi3, Sr4Bi3, and Yb4Sb3) recently predicted to have topological properties in the Ba4As3-like structure.20–22 

FIG. 11.

When defect formation is disallowed, [(a) and (b)] one predicts a (false) metal; allowing vacancy formation [(c) and (d)] leads to spontaneous nonstoichiometry and formation of the real insulator. (a) Density of states for Ba4As3 showing that it is potential p-type degenerate gapped metal according to PBE calculations. Occupied states are shadowed in light blue. (b) Band structure of Ag3Al22O34 showing that it is potential n-type degenerate gapped metal as computed using PBE+U with a U value of 5 eV applied for Ag-d states. [(c) and (d)] Schematic illustrations of the formation of donor/acceptor vacancy in p-/n-type degenerate gapped metals resulting in electron–hole recombination reducing the vacancy formation energy. Donor and acceptor vacancies are shown as light blue and beige colors, respectively.

FIG. 11.

When defect formation is disallowed, [(a) and (b)] one predicts a (false) metal; allowing vacancy formation [(c) and (d)] leads to spontaneous nonstoichiometry and formation of the real insulator. (a) Density of states for Ba4As3 showing that it is potential p-type degenerate gapped metal according to PBE calculations. Occupied states are shadowed in light blue. (b) Band structure of Ag3Al22O34 showing that it is potential n-type degenerate gapped metal as computed using PBE+U with a U value of 5 eV applied for Ag-d states. [(c) and (d)] Schematic illustrations of the formation of donor/acceptor vacancy in p-/n-type degenerate gapped metals resulting in electron–hole recombination reducing the vacancy formation energy. Donor and acceptor vacancies are shown as light blue and beige colors, respectively.

Close modal

The physical effect that stabilizes an insulating state: spontaneous vacancy formation: For some degenerate gapped metals, the formation of point defects (e.g., intrinsic vacancies) can be spontaneous due to an electronic instability.23,31 The point is that while in traditional insulators, vacancy formation requires endothermic breaking the bonds, for degenerate gapped conductors, the presence of carriers in the conduction or valence band can change the defect physics, leading to spontaneous defect formation. For instance, for p-type degenerate gapped conductors [Fig. 11(c)], the formation of donor vacancy results in the moving electrons from the donor level to the hole states in the valence band, which can restore the part of the energy needed to form the vacancy. Similarly, for n-type degenerate gapped metal [Fig. 11(d)], the formation of acceptor vacancy can result in decay of conducting electrons to the acceptor level restoring part of the energy needed to create the vacancy. Such electron–hole recombination can result in spontaneous vacancy formation, which can induce significant deviation from stoichiometry at low-temperatures.23,31,139,140 To examine the possibility of the instability of stoichiometric Ba4As3 and Ag3Al22O34, we study the formation of As vacancy (donor) in the Ba4As3 and Ag vacancy (acceptor) in Ag3Al22O34. Taking into account all experimentally known stoichiometric phases in Ba–As and Ag–Al–O phases, the range of chemical potentials for stability of Ba4As3 and Ag3Al22O34 is predicted by calculations of an energy convex hull141 which defines the ACS having energy lower than any linear combination of any competing phases at the corresponding compositions. Thus, we find that Ba4As3 is on the convex hull [Fig. 12(a)] while Ag3Al22O34 is slightly above the convex hull [Fig. 12(c)]—i.e., there are no chemical potentials at which it is thermodynamically stable.

FIG. 12.

Spontaneous formation of As vacancy in Ba4As3 and Ag vacancy in Ag3Al22O34. (a) Ba-As energy convex hull computed using all experimentally known stoichiometric Ba-As compounds. The compounds above the convex hull are shown by the empty gray circles. (b) Defect formation energy for As vacancy in Ba4As3 as the function of As chemical potential. (c) Stability triangular for Ag3Al22O34 demonstrating that the compound is unstable with respect to decomposition to comparing phases when known experimental Ag–Al–O compounds are taken into account. (d) Defect formation energy for Ag vacancy in Ag3Al22O34 as a function of Ag chemical potential. The figure for Ag3Al22O34 is redrawn using data from Ref. 23. The results for Ag3Al22O34 are given for PBE+U calculations with a U value of 5 eV applied on Ag-d states. For Ba4As3, the results are given for PBE calculations. Fitted elemental-phase reference energies (FERE)143 is used to correct the elemental chemical potentials.

FIG. 12.

Spontaneous formation of As vacancy in Ba4As3 and Ag vacancy in Ag3Al22O34. (a) Ba-As energy convex hull computed using all experimentally known stoichiometric Ba-As compounds. The compounds above the convex hull are shown by the empty gray circles. (b) Defect formation energy for As vacancy in Ba4As3 as the function of As chemical potential. (c) Stability triangular for Ag3Al22O34 demonstrating that the compound is unstable with respect to decomposition to comparing phases when known experimental Ag–Al–O compounds are taken into account. (d) Defect formation energy for Ag vacancy in Ag3Al22O34 as a function of Ag chemical potential. The figure for Ag3Al22O34 is redrawn using data from Ref. 23. The results for Ag3Al22O34 are given for PBE+U calculations with a U value of 5 eV applied on Ag-d states. For Ba4As3, the results are given for PBE calculations. Fitted elemental-phase reference energies (FERE)143 is used to correct the elemental chemical potentials.

Close modal

The computed vacancy formation energies [Figs. 12(b) and 12(d)] demonstrate that stoichiometric Ba4As3 and Ag3Al22O34 are unstable with respect to the spontaneous formation of donor and acceptor vacancies, respectively. Moreover, the vacancy formation results in the reduction of carrier density due to electron–hole compensation—each As vacancy in Ba4As3 removes 3h from the valence band, and each Ag vacancy in Ag3Al22O34 depletes 1e from the conduction band. These results suggest that both compounds are unstable with respect to vacancy formation and hence unlikely to exist in stoichiometric forms (a far larger concentration of defects than considered here might exist).

The experimental situation: The perfectly stoichiometric Ba4As3(SG: 220) has never been synthesized. Stoichiometric Ba4As3 is a compound inspired by experimentally reported nonstoichiometric Ba4As3−x(SG: 220), which is indeed a simple insulator.138 Ba4As3−x(SG: 220) and Ba4As3(SG: 220) structures are similar, with the only difference being that in Ba4As3-x each As site has partial occupancy. While Ag3Al22O34 structure has been suggested experimentally,142 all attempts to synthesize the compound lead to close stoichiometry Ag3−xAl22O34+y having a similar structure and insulating properties.23 Taking into account the above defect calculations, we conclude that the experimentally observed nonstoichiometry of the compounds is caused not by the growth effect, but by the Fermi level instability-the energy lowering caused by electron–hole recombination as the result of vacancy formation is larger than the energy gain needed to break the chemical bonds.

A more specialized effect, pertaining to compounds containing a high-Z atom is the effect of gapping a false metal due to spin–orbit coupling. In traditional first-principles studies,43–45 the investigation of materials properties is usually limited to the non-SOC calculations. This computational approximation works sufficiently well for light elements, and the large set of material properties can be described well within this simplified approximation.144 In general, the SOC effect on electronic structure is usually discussed in terms of spin–orbit splitting. For degenerate gapped metals, SOC can result in the splitting of a conduction or valence band, giving a true insulator. An example of such behavior is CaIrO3 (SG: 63)—a compound which without SOC is predicted to be a p-type conductor145 with the Fermi level in the principal valence band [Fig. 13(a)] even when a hard XC functional (e.g., HSE) is used. However, this result contradicts experimental data suggesting that CaIrO3 is an insulator based on the measurement of electronic resistance vs temperature.146 When SOC is applied, the system becomes an insulator [see Fig. 13(b)],145 which is due to the splitting of the valence band. A similar example of band gap opening has been reported for Sr2IrO4.147 These explorations thus demonstrate the important contribution of the SOC in the false metal—real insulator transition, suggesting that SOC should be taken into account for the prediction of real degenerate gapped metals. Importantly, the band gap opening is only observed with the XC functional satisfying Sec. II A XC conditions—application of SOC on top of the PBE functional often does not result in band gap opening. It should be noted that recently developed topological databases20–22 provided the results for electronic structures of many compounds with included SOC. However, since all calculations were performed with soft XC functionals, it is likely that databases have a number of false predictions. For instance, both Sr2IrO4 and CaIrO3 are predicted to be topological metals in the databases.20–22 

FIG. 13.

When SOC is ignored, (a) one predicts a (false) metal; adding SOC (b) results in the formation of an insulator. Band structure projected onto dxy, dyz, and dzx orbitals for CaIrO3 computed with (a) HSE and (b) HSE+SOC. The highest occupied level is at 0 eV. The radii of semicircles and circles are proportional to the weights of the orbitals. SG denotes the space group number. Reproduced with permission from Kim et al., Phys. Rev. Lett. 115, 096401 (2015). Copyright 2015 by the American Physical Society.

FIG. 13.

When SOC is ignored, (a) one predicts a (false) metal; adding SOC (b) results in the formation of an insulator. Band structure projected onto dxy, dyz, and dzx orbitals for CaIrO3 computed with (a) HSE and (b) HSE+SOC. The highest occupied level is at 0 eV. The radii of semicircles and circles are proportional to the weights of the orbitals. SG denotes the space group number. Reproduced with permission from Kim et al., Phys. Rev. Lett. 115, 096401 (2015). Copyright 2015 by the American Physical Society.

Close modal

We have seen that the symmetry-breaking modes discussed herein can (i) result in band gap opening without the formation of any in-gap states, as demonstrated in Fig. 6(f) for AFM NiO or (ii) cause splitting of a subband from the continuum [Fig. 1(a)], forming an intermediate band within the principle band gap [Fig. 1(b)]. Inspection of the wave function of such split-off bands created via symmetry breaking can reveal the nature of in-gap states and its correlation to local structural distortions. The orbital character of the split-off intermediate band can either (a) mimic that of the nearest band edge from which it was split, corresponding to a normal situation of degeneracy removal with the possibly trapped carrier (because the intermediate band is not connected via dispersion to any other band), or (b) the split-off band can show localization of certain sublattices (see below TiO2:Li) behaving as a defect level, except that the defect here is electronic. This situation is common when the localizing sublattice atom can exist in more than one FOS, such as Ti3+ and Ti4+. In the dilute defect limit, this latter situation is often discussed as polaron—a quasiparticle originating from the interactions of electrons/holes with a lattice ion, often causing local distortions.148,149 Inspection of the wavefunctions (Fig. 14) of the broken symmetry cases described in Sec. III reveals the following instances of intermediate bands:

FIG. 14.

Wavefunction amplitudes (yellow isosurfaces) computed for electron (e) or hole (h) trapped intermediate bands (IBs) in TiO2:Li and ABO3 compounds demonstrating carrier-trapping. For all cases, the isosurface is set at 0.01 e/Bohr3. A, B, and oxygen atoms are shown as green, pink, and gray, respectively. SG denotes the space group number.

FIG. 14.

Wavefunction amplitudes (yellow isosurfaces) computed for electron (e) or hole (h) trapped intermediate bands (IBs) in TiO2:Li and ABO3 compounds demonstrating carrier-trapping. For all cases, the isosurface is set at 0.01 e/Bohr3. A, B, and oxygen atoms are shown as green, pink, and gray, respectively. SG denotes the space group number.

Close modal

Polaron-like states in TiO2:Li: Doping of ordinary insulators can result in different compensation mechanisms that remove free carriers, including the formation of electronic defects being polarons. A polaron can form without doping or with doping. An example of the latter is the widely discussed e-doped TiO2 systems114,122,150–152 and is demonstrated on the example of density of states for Li-doped TiO2 shown in Fig. 9(e). Indeed, Li doping of TiO2(SG: 12) results in disproportionation and carrier localization—one Li converts one Ti4+ to Ti3+. This is in agreement with the fact that the calculated charge density corresponding to the in-gap states shows localization of the e-trapped intermediate band on a single Ti atom [see Fig. 14(a)]. In fact, the existence of Li induced Ti3+ states has been used to quantify the Li content coming from surface reaction for TiO2-based electrode in Li-ion batteries.110 Among the considered compounds, TiO2:Li has the narrowest band of in-gap states, which is mainly due to a low concentration of reduced Ti3+ weakly interacting in the system.

Conventional split-off bands due to symmetry breaking can be divided into electron traps or hole traps:

Formation of electron-trapped valence band maxima: DFT depiction of early transition metal oxides (i.e., LaVO3, YTiO3, LaTiO3, and LaMnO3) exhibits the formation of an occupied “upper Hubbard Bands” that contain trapped electrons [see Figs. 5(b), 7(d), and 10(d)] with roughly the same wavefunction amplitude for the trapped carriers [see Figs. 14(b)–14(d)], suggesting that the compounds can be considered as La3+V3+O32−, Y3+Ti3+O32−, La3+Ti3+O32− and La3+Mn3+O32−. The e-trapped states in ABO3 compounds are reported in photoemission studies showing the existence of in-gap occupied states located about 1.3–1.5 eV below the principal conduction band (or the Fermi level).105–108 Moreover, for YTiO3, the in-gap state was attributed to the reduced Ti3+, which is in line with the above discussion.

Formation of hole-trapped conduction band minima: DFT depiction of YNiO3, SrBiO3, and CuBi2O4 shows clearly distinct formation of h-trapped states in Figs. 6(e), 8(b), and 9(f). Here, YNiO3 and SrBiO3 are the compounds having disproportionation in the low-temperature phase, where one can expect different charge density distributions on structurally different B atoms. For AFM monoclinic YNiO3, there are two structurally different Ni sublattices with one magnetic and one nonmagnetic. These Ni sublattices have different contributions to electronic structures. This is shown in Fig. 14(e), demonstrating that the h-trapped state is not only localized on O but also has contribution from Ni atoms. Here, due to DLE structure, one type of Ni atoms contributes significantly more than another one. This picture is similar to that for nonmagnetic SrBiO3 shown in Fig. 14(f).

The formation of hole-trapped conduction band minima and electron-trapped valence band maxima is often characteristic of existence of electronically distinct metal ions capable of carrying different valences. Such “electronic defects” are analogous to “atomic impurities/defects.” The latter traditional defect physics described by Koster and Slater153 and extended to 3D semiconductors by Hjalmarson et al.154 depict the formation of in-gap states in the language of the perturbation caused by impurity I substituting a host atom H, i.e., ΔVimp = V(I)-V(H). If the magnitude of this perturbation exceeds some threshold value (that depends on the bandwidth), in gap split-off states can form. Electronic defects are an analogous problem to atomic defects in that the perturbation can be defined as the difference of potential for a real insulator having in-gap states and corresponding false metal, for instance, as ΔVelec = V(Ti3+)-V(Ti4+).

New doping physics (“antidoping”) in real insulators originated from false metals: While the above examples are pristine false metals, recent discussion has focused on doping them.120,155,156 Specifically, it has been discovered that n-type doping of compounds having h-trapped IBs (e.g., YNiO3120 and SmNiO3155,156) can result in a shift of intensity of the intermediate band toward the valence band leading to reduction of conductivity and band gap opening. The similar band gap opening mechanisms can also be observed for hole doping of the compounds having e-trapped intermediate bands,120 where adding holes results in band shift toward the conduction band with following band gap opening. Such a unique doping response has been recently confirmed experimentally.157–159 Importantly, accounting for symmetry breaking, the formation of point defects, and using valid computational setups can be used to identify not only false metals but new compounds showing antidoping behavior.

We emphasize again that all of the behaviors seen above: split-off bands, polaron formation, and antidoping are found by mean-field DFT and do not call for special effects such as strong correlation.

The above examples of false metals demonstrate that real degenerate gapped metals—be that transparent conductors, or electrides, or Dirac semimetals—might be rare because they can be vulnerable to instabilities converting them to insulators by any of the mechanisms discussed in Sec. III. However, this does not mean that such compounds do not exist and cannot be found. To illustrate it better, let us consider the case of cubic SrVO3 as one of the most widely studied real degenerate gapped metals illustrating why this compound remains metal despite the above-discussed mechanisms. The compound contains light elements, and hence SOC does not affect its electronic structure noticeably. The internal band gap in this material is relatively small, which is expected to result in rather small energy lowering due to electron-hole recombination caused by the formation of acceptor defects—indeed, no significant deviation from the ideal composition is observed in SrVO3 under optimized conditions.160 SrVO3 has closer-to-one tolerance factor minimizing the need for tilting/rotation and atomic displacements in the compound. When SrVO3 is artificially distorted via tilting/rotation, theory predicts that band gap opening can be archived.39 However, since such tilting is not energetically favorable, the system is a degenerate gapped metal under normal conditions. These results demonstrate how using the classic Goldschmidt tolerance factor can be used as one of the inverse design principles for the search of potential degenerate gapped ABO3 conductors.

The family of real degenerate gapped metals is expanding owing to better understanding resulting from close theory and experiment collaboration. Significant progress has been made in artificially doped degenerate gapped compounds, including the classical transparent conducting oxides (e.g., ZnO:Al161 or In2O3:Sn162) which can accumulate a large concentration of free carriers, on the order of 1020 cm−3. From the theoretical side, the set of heavily doped insulators is expanding with novel prediction of potential n- and p-type transparent conductors163 with some of the most important systems already experimentally validated (i.e., n-type doped Ga2O3164 and BaSnO3165). The key steps here remain the understanding of doping limits25–27 of insulators and identifying main compensation mechanisms. While the intrinsic (not doped) degenerate gapped metals have not been the subject of intensive research until recently, the discovery of inorganic electrides13 and intrinsic transparent conductors24 has raised interest in the field. For instance, a number of new intrinsic degenerate gapped metals have been experimentally synthesized, including 2D Ca2N,166 1D Sr3CrN3,167 and 0D YH2168 electrides. Moreover, many potential gapped metals [e.g., LaH2, KRb3, K4Al3(SiO4)3, Ca5Bi3, Cs3O, Ba3CrN3, Ba3FeN3, Ba2NaO, and Li12Mg3Si4] found in high-throughput calculations of electrides169,170 and thermoelectrics171 are awaiting potential laboratory testing: are they stable metals or will they transform spontaneously to insulators, suggesting that they were false metals, not real metals in the first place?

An interesting property of degenerate gapped metals is the ability to balance the generally conflicting properties of (i) transparency (reflecting the internal gap plus the occupied portion of the CB, as well as plasma reflection due to free carriers), (ii) conductivity (reflecting the carriers inside the CB), and (iii) stability (with respect to the metal to insulator transition mechanisms discussed above). These three factors (i)–(iii) can be used as effective knobs for controlling the properties or real degenerate gapped conductors. For instance, it has been demonstrated that while BaNbO3 and Ca6Al7O16 are real gapped metals and stable with respect to decompositions to competing phases (i.e., there is a set of experimental conditions corresponding to specific chemical potentials where BaNbO3 and Ca6Al7O16 can exist), tuning of experimental conditions can result in stabilization of ordered vacancy compounds which have totally different optoelectronic properties.23 This eventually can open the possibility for tailoring optoelectronic properties by controllable nonstoichiometry to design new functional degenerate gapped metals.

The work analyzed and reviewed here points to three relevant possible “call to action” by the community:

  • Let us not jump into conclusions that the false metal syndrome—incorrect theoretical predictions of degenerate gapped metals based on naïve applications of DFT, using the least number of possible magnetic, orbital/structural degrees of freedom—must imply that strong electronelectron correlation is at play and has been omitted unjustifiably in these calculations. Indeed, before leapfrogging from N-DFT to highly correlated approaches, it is best to examine whether symmetry-broken DFT, being a bona fide mean-field theory, contains sufficient physics to explain the absence of false metals. In many cases, the realization that some prediction of metallic states has prematurely motivated the introduction of a dynamic electron–electron correlation in a symmetry-preserving picture, as the essential, must-have ingredient.37,172–174 Such DMFT papers on many ABO3 compounds have explicitly claimed failure of DFT in explaining disproportionation, Jahn–Teller displacements, orbital order, mass enhancement, and gapping in PM phases of many 3d compounds,126,172,174–177 whereas what actually failed was a naïve version of DFT. We pointed out here that there are avenues for removing the constraints on N-DFT theory other than disposing of DFT altogether, leading to a systematic understanding of which compounds are metallic and which are insulating band structures. Such a detailed exploration involves considering larger unit cells that support different local structural and magnetic short-range orders. This points to the possible scenario that mean-field like energy lowering via spin and structural symmetry breaking is the crucial, minimal mechanism at work for systematically explaining metal vs insulating band structures, whereas the consideration the extra physics of dynamic correlation is not needed to describe the actual effects of disproportionation, Jahn–Teller displacements, orbital order, mass enhancement, and gapping in 3d oxides. Conversely, the false metals syndrome needs to be better scrutinized by the DFT community, correcting the naïve approximations in N-DFT that led to it and propagated into databases (Fig. 2).

  • The community should beware of the scenario of predictions of exotic properties in compounds that, in fact, are false metals, i.e., do not exist.31,178 The results of N-DFT calculations have been often used as a platform to suggest new exotic physics such as topology, quantum confinement, and superconductivity. Herein, we call for caution regarding proposing compounds with new physical effects on such an uncertain platform (i.e., the compound does not exist). For instance, it has been recently demonstrated that certain topological properties predicted to “live” in compounds that are not realizable due to different energy structural/magnetic symmetry lowering modes.31,178 As an illustration, recent screening of topological compounds suggested that NiO, CuBi2O4, YTiO3, Sr2IrO4, CaIrO3, and LaTiO3 are topological semimetals,20–22 while as demonstrated in Sec. III, all the above compounds are false metals becoming real insulators when non-N-DFT is used.

  • Computational band structure databases and high-throughput calculations that recommend compounds as metals but use N-DFT must be scrutinized (viz., few examples in Fig. 2). This review also revealed the role of computational materials databases20,43–45 and high-throughput calculations179–184 to identify metallic vs insulating band structures. Open-access materials databases revolutionized the materials science by providing details on new applications of already well-known compounds and a number of potential compounds, which have not been synthesized before. However, the electronic properties of many degenerate gapped metals are still not described correctly within the existing studies (e.g., see examples in Fig. 2). The problem is not easy to handle at present because the databases are in constant change (i.e., hundreds/thousands of new compounds and their properties are added to each database every year), and compounds declared metals one day disappear and are replaced on another day by insulating entries (and this is not a phase transition, just correction of an error) without clear or accessible explanation, leading to confusion due to often weak documentation on the updates. Although it is only a matter of time until the accurate computational setups are commonly used, currently available data for electronic structures for degenerate gapped conductors should be used with great caution and be verified with band gap mechanisms discussed above. The common-sense steps which should be undertaken are: (i) use XC functionals that are able to distinguish occupied from unoccupied states and have reduced SIE; (ii) account SOC especially for high-Z elements; (iii) allow different structural/magnetic symmetry-breaking modalities that can result in energy lowering (e.g., magnetic order, spin-disorder, Jahn–Teller distortion, disproportionation); and (iv) verify the stability of compounds with respect to other competing phases and defect formation. All these steps should account for structural nudging and use the cell size not as the fixed input, but as the convergence parameter. In these ways, we can demonstrate not only the set of true degenerate gapped metals, but also identify the true mechanisms resulting in a change of electronic structures.

The work on electronic structure calculations was supported by the National Science Foundation, Division of Materials Research, Condensed Matter and Materials Theory (CMMT) within DMR-1724791. The study on doping of quantum materials was supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences, and Engineering Division within DE-SC0010467. The authors acknowledge the use of computational resources located at the National Renewable Energy Laboratory and sponsored by the Department of Energy's Office of Energy Efficiency and Renewable Energy. This work also utilized the Extreme Science and Engineering Discovery Environment (XSEDE) supercomputer resources, which are supported by the National Science Foundation Grant No. ACI-1548562. The authors thank Dr. Zhi Wang for the useful discussions.

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

1.
N. W.
Ashcroft
and
N. D.
Mermin
,
Solid State Physics
(
Holt, Rinehart and Winston
,
New York
,
1976
).
2.
C.
Kittel
,
Introduction to Solid State Physics
(
Wiley
,
New York
,
1976
).
3.
W. A.
Harrison
,
Electronic Structure and the Properties of Solids: The Physics of the Chemical Bond
(
Courier Corporation
,
2012
).
4.
J. M.
Ziman
,
Principles of the Theory of Solids
(
Cambridge University Press
,
Cambridge
,
1999
).
5.
B. L.
Chamberland
and
P. S.
Danielson
, “
Alkaline-earth vanadium (IV) oxides having the AVO3 composition
,”
J. Solid State Chem.
3
,
243
(
1971
).
6.
R.
Berger
and
C. F.
Van Bruggen
, “
TlCu2Se2: A p-type metal with a layer structure
,”
J. Less-Common Met.
99
,
113
(
1984
).
7.
E. F.
Schubert
,
Doping in III-V Semiconductors
(
Cambridge University Press
,
1993
).
8.
E. F.
Schubert
,
Delta-Doping of Semiconductors
(
Cambridge University Press
,
1996
).
9.
P. A.
Lee
,
N.
Nagaosa
, and
X.-G.
Wen
, “
Doping a Mott insulator: Physics of high-temperature superconductivity
,”
Rev. Mod. Phys.
78
,
17
(
2006
).
10.
B.
Lüssem
,
M.
Riede
, and
K.
Leo
, “
Doping of organic semiconductors
,”
Phys. Status Solidi A
210
,
9
(
2013
).
11.
R.
Kalish
, “
Doping of diamond
,”
Carbon
37
,
781
(
1999
).
12.
M. T.
Casais
,
J. A.
Alonso
,
I.
Rasines
, and
M. A.
Hidalgo
, “
Preparation, neutron structural study and characterization of BaNbO3: A Pauli-like metallic perovskite
,”
Mater. Res. Bull.
30
,
201
(
1995
).
13.
S.
Matsuishi
,
Y.
Toda
,
M.
Miyakawa
,
K.
Hayashi
,
T.
Kamiya
,
M.
Hirano
,
I.
Tanaka
, and
H.
Hosono
, “
High-density electron anions in a nanoporous single crystal: [Ca24Al28O64]4+(4e-)
,”
Science
301
,
626
(
2003
).
14.
L.
Zhang
,
Y. J.
Zhou
,
L.
Guo
,
W. W.
Zhao
,
A.
Barnes
,
H. T.
Zhang
,
C.
Eaton
,
Y. X.
Zheng
,
M.
Brahlek
,
H. F.
Haneef
,
N. J.
Podraza
,
M. H. W.
Chan
,
V.
Gopalan
,
K. M.
Rabe
, and
R.
Engel-Herbert
, “
Correlated metals as transparent conductors
,”
Nat. Mater.
15
,
204
(
2016
).
15.
T.
Yoshida
,
M.
Kobayashi
,
K.
Yoshimatsu
,
H.
Kumigashira
, and
A.
Fujimori
, “
Correlated electronic states of SrVO3 revealed by angle-resolved photoemission spectroscopy
,”
J. Electron. Spectrosc. Relat. Phenom.
208
,
11
(
2016
).
16.
M.
Takizawa
,
M.
Minohara
,
H.
Kumigashira
,
D.
Toyota
,
M.
Oshima
,
H.
Wadati
,
T.
Yoshida
,
A.
Fujimori
,
M.
Lippmaa
,
M.
Kawasaki
,
H.
Koinuma
,
G.
Sordi
, and
M.
Rozenberg
, “
Coherent and incoherent d band dispersions in SrVO3
,”
Phys. Rev. B
80
,
235104
(
2009
).
17.
T.
Yoshida
,
M.
Hashimoto
,
T.
Takizawa
,
A.
Fujimori
,
M.
Kubota
,
K.
Ono
, and
H.
Eisaki
, “
Mass renormalization in the bandwidth-controlled Mott-Hubbard systems SrVO3 and CaVO3 studied by angle-resolved photoemission spectroscopy
,”
Phys. Rev. B
82
,
085119
(
2010
).
18.
H.
Hosono
and
D. C.
Paine
,
Handbook of Transparent Conductors
(
Springer Science and Business Media
,
2010
).
19.
J. L.
Dye
, “
Electrides: Early examples of quantum confinement
,”
Acc. Chem. Res.
42
,
1564
(
2009
).
20.
M. G.
Vergniory
,
L.
Elcoro
,
C.
Felser
,
N.
Regnault
,
B. A.
Bernevig
, and
Z.
Wang
, “
A complete catalogue of high-quality topological materials
,”
Nature
566
,
480
(
2019
).
21.
T.
Zhang
,
Y.
Jiang
,
Z.
Song
,
H.
Huang
,
Y.
He
,
Z.
Fang
,
H.
Weng
, and
C.
Fang
, “
Catalogue of topological electronic materials
,”
Nature
566
,
475
(
2019
).
22.
F.
Tang
,
H. C.
Po
,
A.
Vishwanath
, and
X.
Wan
, “
Comprehensive search for topological materials using symmetry indicators
,”
Nature
566
,
486
(
2019
).
23.
O. I.
Malyi
,
M. T.
Yeung
,
K. R.
Poeppelmeier
,
C.
Persson
, and
A.
Zunger
, “
Spontaneous non-stoichiometry and ordering in degenerate but gapped transparent conductors
,”
Matter
1
,
280
(
2019
).
24.
X.
Zhang
,
L.
Zhang
,
J. D.
Perkins
, and
A.
Zunger
, “
Intrinsic transparent conductors without doping
,”
Phys. Rev. Lett.
115
,
176602
(
2015
).
25.
S. B.
Zhang
,
S.-H.
Wei
, and
A.
Zunger
, “
A phenomenological model for systematization and prediction of doping limits in II–VI and I–III–VI2 compounds
,”
J. Appl. Phys.
83
,
3192
(
1998
).
26.
A.
Zunger
, “
Practical doping principles
,”
Appl. Phys. Lett.
83
,
57
(
2003
).
27.
S. B.
Zhang
,
S. H.
Wei
, and
A.
Zunger
, “
Microscopic origin of the phenomenological equilibrium “Doping Limit Rule” in n-type III-V semiconductors
,”
Phys. Rev. Lett.
84
,
1232
(
2000
).
28.
E.
Özdaş
,
A. R.
Kortan
,
N.
Kopylov
,
A. P.
Ramirez
,
T.
Siegrist
,
K. M.
Rabe
,
H. E.
Bair
,
S.
Schuppler
, and
P. H.
Citrin
, “
Superconductivity and cation-vacancy ordering in the rare-earth fulleride Yb2.75C60
,”
Nature
375
,
126
(
1995
).
29.
D.
Schmid
,
M.
Ruckh
,
F.
Grunwald
, and
H. W.
Schock
, “
Chalcopyrite/defect chalcopyrite heterojunctions on the basis of CuInSe2
,”
J. Appl. Phys.
73
,
2902
(
1993
).
30.
X.
Xu
,
C.
Randorn
,
P.
Efstathiou
, and
J. T. S.
Irvine
, “
A red metallic oxide photocatalyst
,”
Nat. Mater.
11
,
595
(
2012
).
31.
O. I.
Malyi
,
G. M.
Dalpian
,
X.-G.
Zhao
,
Z.
Wang
, and
A.
Zunger
, “
Realization of predicted exotic materials: The burden of proof
,”
Mater. Today
32
,
35
(
2020
).
32.
S.
Lany
and
A.
Zunger
, “
Polaronic hole localization and multiple hole binding of acceptors in oxide wide-gap semiconductors
,”
Phys. Rev. B
80
,
085202
(
2009
).
33.
C.
Freysoldt
,
B.
Grabowski
,
T.
Hickel
,
J.
Neugebauer
,
G.
Kresse
,
A.
Janotti
, and
C. G.
Van de Walle
, “
First-principles calculations for point defects in solids
,”
Rev. Mod. Phys.
86
,
253
(
2014
).
34.
S.
Lany
, “
Predicting polaronic defect states by means of generalized Koopmans density functional calculations
,”
Phys. Status Solidi B
248
,
1052
(
2011
).
35.
N.
Mott
,
Metal-Insulator Transitions
(
CRC Press
,
London
,
1990
).
36.
M.
Imada
,
A.
Fujimori
, and
Y.
Tokura
, “
Metal-insulator transitions
,”
Rev. Mod. Phys.
70
,
1039
(
1998
).
37.
A.
Georges
,
G.
Kotliar
,
W.
Krauth
, and
M. J.
Rozenberg
, “
Dynamical mean-field theory of strongly correlated fermion systems and the limit of infinite dimensions
,”
Rev. Mod. Phys.
68
,
13
(
1996
).
38.
G.
Trimarchi
,
Z.
Wang
, and
A.
Zunger
, “
Polymorphous band structure model of gapping in the antiferromagnetic and paramagnetic phases of the Mott insulators MnO, FeO, CoO, and NiO
,”
Phys. Rev. B
97
,
035107
(
2018
).
39.
J.
Varignon
,
M.
Bibes
, and
A.
Zunger
, “
Origin of band gaps in 3d perovskite oxides
,”
Nat. Commun.
10
,
1658
(
2019
).
40.
J.
Varignon
,
M.
Bibes
, and
A.
Zunger
, “
Mott gapping in 3d ABO3 perovskites without Mott-Hubbard interelectronic repulsion energy U
,”
Phys. Rev. B
100
,
035119
(
2019
).
41.
R. M.
Wentzcovitch
,
W. W.
Schulz
, and
P. B.
Allen
, “
VO2: Peierls or Mott-Hubbard? A view from band theory
,”
Phys. Rev. Lett.
72
,
3389
(
1994
).
42.
S.
Xu
,
X.
Shen
,
K. A.
Hallman
,
R. F.
Haglund
, and
S. T.
Pantelides
, “
Unified band-theoretic description of structural, electronic, and magnetic properties of vanadium dioxide phases
,”
Phys. Rev. B
95
,
125105
(
2017
).
43.
A.
Jain
,
S. P.
Ong
,
G.
Hautier
,
W.
Chen
,
W. D.
Richards
,
S.
Dacek
,
S.
Cholia
,
D.
Gunter
,
D.
Skinner
,
G.
Ceder
, and
K. A.
Persson
, “
Commentary: The materials Project: A materials genome approach to accelerating Materials innovation
,”
APL Mater.
1
,
011002
(
2013
).
44.
S.
Curtarolo
,
W.
Setyawan
,
S.
Wang
,
J.
Xue
,
K.
Yang
,
R. H.
Taylor
,
L. J.
Nelson
,
G. L. W.
Hart
,
S.
Sanvito
,
M.
Buongiorno-Nardelli
,
N.
Mingo
, and
O.
Levy
, “
AFLOWLIB.ORG: A distributed materials properties repository from high-throughput ab initio calculations
,”
Comput. Mater. Sci.
58
,
227
(
2012
).
45.
J. E.
Saal
,
S.
Kirklin
,
M.
Aykol
,
B.
Meredig
, and
C.
Wolverton
, “
Materials design and discovery with high-throughput density functional theory: The Open Quantum Materials Database (OQMD)
,”
JOM
65
,
1501
(
2013
).
46.
S. M.
Young
,
S.
Zaheer
,
J. C. Y.
Teo
,
C. L.
Kane
,
E. J.
Mele
, and
A. M.
Rappe
, “
Dirac semimetal in three dimensions
,”
Phys. Rev. Lett.
108
,
140405
(
2012
).
47.
J. C.
Slater
and
G. F.
Koster
, “
Simplified LCAO method for the periodic potential problem
,”
Phys. Rev.
94
,
1498
(
1954
).
48.
J. M.
Luttinger
and
W.
Kohn
, “
Motion of electrons and holes in perturbed periodic fields
,”
Phys. Rev.
97
,
869
(
1955
).
49.
E. O.
Kane
, “
Band structure of indium antimonide
,”
J. Phys. Chem. Solids
1
,
249
(
1957
).
50.
D.
Brust
,
J. C.
Phillips
, and
F.
Bassani
, “
Critical points and ultraviolet reflectivity of semiconductors
,”
Phys. Rev. Lett.
9
,
94
(
1962
).
51.
M. L.
Cohen
and
T. K.
Bergstresser
, “
Band structures and pseudopotential form factors for fourteen semiconductors of the diamond and zinc-blende structures
,”
Phys. Rev.
141
,
789
(
1966
).
52.
A.
Zunger
, “
Inverse design in search of materials with target functionalities
,”
Nat. Rev. Chem.
2
,
0121
(
2018
).
53.
W.
Kohn
and
L. J.
Sham
, “
Self-consistent equations including exchange and correlation effects
,”
Phys. Rev.
140
,
A1133
(
1965
).
54.
G.
Trimarchi
and
A.
Zunger
, “
Global space-group optimization problem: Finding the stablest crystal structure without constraints
,”
Phys. Rev. B
75
,
104113
(
2007
).
55.
A. R.
Oganov
and
C. W.
Glass
, “
Crystal structure prediction using ab initio evolutionary techniques: Principles and applications
,”
J. Chem. Phys.
124
,
244704
(
2006
).
56.
N. L.
Abraham
and
M. I. J.
Probert
, “
A periodic genetic algorithm with real-space representation for crystal structure and polymorph prediction
,”
Phys. Rev. B
73
,
224104
(
2006
).
57.
S.
Goedecker
, “
Minima hopping: An efficient search method for the global minimum of the potential energy surface of complex molecular systems
,”
J. Chem. Phys.
120
,
9911
(
2004
).
58.
G.
Kresse
and
J.
Hafner
, “
Ab initio molecular dynamics for liquid metals
,”
Phys. Rev. B
47
,
558
(
1993
).
59.
G.
Kresse
and
J.
Furthmüller
, “
Efficiency of ab initio total energy calculations for metals and semiconductors using a plane-wave basis set
,”
Comput. Mater. Sci.
6
,
15
(
1996
).
60.
G.
Kresse
and
J.
Furthmüller
, “
Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set
,”
Phys. Rev. B
54
,
11169
(
1996
).
61.
K.
Momma
and
F.
Izumi
, “
VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data
,”
J. Appl. Crystallogr.
44
,
1272
(
2011
).
62.
M. G.
Medvedev
,
I. S.
Bushmarinov
,
J.
Sun
,
J. P.
Perdew
, and
K. A.
Lyssenko
, “
Density functional theory is straying from the path toward the exact functional
,”
Science
355
,
49
(
2017
).
63.
Y.
Zhang
,
J.
Furnes
,
R.
Zhang
,
Z.
Wang
,
A.
Zunger
, and
J.
Sun
, “
Symmetry-breaking polymorphous descriptions for correlated materials without interelectronic U
,”
Phys. Rev. B
102
,
045112
(
2020
).
64.
J. P.
Perdew
and
A.
Zunger
, “
Self-interaction correction to density-functional approximations for many-electron systems
,”
Phys. Rev. B
23
,
5048
(
1981
).
65.
A. J.
Cohen
,
P.
Mori-Sánchez
, and
W.
Yang
, “
Insights into current limitations of density functional theory
,”
Science
321
,
792
(
2008
).
66.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
, “
Generalized gradient approximation made simple
,”
Phys. Rev. Lett.
77
,
3865
(
1996
).
67.
J. P.
Perdew
,
A.
Ruzsinszky
,
G. I.
Csonka
,
O. A.
Vydrov
,
G. E.
Scuseria
,
L. A.
Constantin
,
X.
Zhou
, and
K.
Burke
, “
Restoring the density-gradient expansion for exchange in solids and surfaces
,”
Phys. Rev. Lett.
100
,
136406
(
2008
).
68.
A.
Zunger
and
A. J.
Freeman
, “
Self-consistent numerical-basis-set linear-combination-of-atomic-orbitals model for the study of solids in the local density formalism
,”
Phys. Rev. B
15
,
4716
(
1977
).
69.
A.
Zunger
and
A. J.
Freeman
, “
Ground-state electronic properties of diamond in the local-density formalism
,”
Phys. Rev. B
15
,
5049
(
1977
).
70.
J.
Sun
,
A.
Ruzsinszky
, and
J. P.
Perdew
, “
Strongly constrained and appropriately normed semilocal density functional
,”
Phys. Rev. Lett.
115
,
036402
(
2015
).
71.
A.
Liechtenstein
,
V.
Anisimov
, and
J.
Zaanen
, “
Density-functional theory and strong interactions: Orbital ordering in Mott-Hubbard insulators
,”
Phys. Rev. B
52
,
R5467
(
1995
).
72.
M.
Cococcioni
and
S.
De Gironcoli
, “
Linear response approach to the calculation of the effective interaction parameters in the LDA+U method
,”
Phys. Rev. B
71
,
035105
(
2005
).
73.
V. I.
Anisimov
,
J.
Zaanen
, and
O. K.
Andersen
, “
Band theory and Mott insulators: Hubbard U instead of Stoner I
,”
Phys. Rev. B
44
,
943
(
1991
).
74.
S. L.
Dudarev
,
G. A.
Botton
,
S. Y.
Savrasov
,
C. J.
Humphreys
, and
A. P.
Sutton
, “
Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+U study
,”
Phys. Rev. B
57
,
1505
(
1998
).
75.
J.
Heyd
,
G. E.
Scuseria
, and
M.
Ernzerhof
, “
Hybrid functionals based on a screened Coulomb potential
,”
J. Chem. Phys.
118
,
8207
(
2003
).
76.
A. V.
Krukau
,
O. A.
Vydrov
,
A. F.
Izmaylov
, and
G. E.
Scuseria
, “
Influence of the exchange screening parameter on the performance of screened hybrid functionals
,”
J. Chem. Phys.
125
,
224106
(
2006
).
77.
A.
Svane
and
O.
Gunnarsson
, “
Transition-metal oxides in the self-interaction–corrected density-functional formalism
,”
Phys. Rev. Lett.
65
,
1148
(
1990
).
78.
J.
He
and
C.
Franchini
, “
Screened hybrid functional applied to 3d0→3d8 transition-metal perovskites LaMO3 (M= Sc–Cu): Influence of the exchange mixing parameter on the structural, electronic, and magnetic properties
,”
Phys. Rev. B
86
,
235117
(
2012
).
79.
A.
Belsky
,
M.
Hellenbrandt
,
V. L.
Karen
, and
P.
Luksch
, “
New developments in the Inorganic Crystal Structure Database (ICSD): Accessibility in support of materials research and design
,”
Acta Crystallogr., Sect. B: Struct. Sci.
58
,
364
(
2002
).
80.
Z.
Wang
,
Q.
Liu
,
J.-W.
Luo
, and
A.
Zunger
, “
Digging for topological property in disordered alloys: The emergence of Weyl semimetal phase and sequential band inversions in PbSe–SnSe alloys
,”
Mater. Horiz.
6
,
2124
(
2019
).
81.
X.-G.
Zhao
,
G. M.
Dalpian
,
Z.
Wang
, and
A.
Zunger
, “
The polymorphous nature of cubic halide perovskites
,”
Phys. Rev. B
101
,
155137
(
2020
).
82.
Z.
Wang
,
X.
Zhao
,
R.
Koch
,
S. J. L.
Billinge
, and
A.
Zunger
, “
Understanding electronic peculiarities in tetragonal FeSe as local structural symmetry breaking
,” arXiv:1911.02670, [cond-mat] (
2020
).
83.
Y.
Marcus
, “
Effect of ions on the structure of water: Structure making and breaking
,”
Chem. Rev.
109
,
1346
(
2009
).
84.
H. A.
Jahn
,
E.
Teller
, and
F. G.
Donnan
, “
Stability of polyatomic molecules in degenerate electronic states - I—Orbital degeneracy
,”
Proc. R. Soc. London, Ser. A
161
,
220
(
1937
).
85.
A.
Glazer
, “
The classification of tilted octahedra in perovskites
,”
Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem.
28
,
3384
(
1972
).
86.
V. M.
Goldschmidt
, “
Die Gesetze der Krystallochemie
,”
Naturwissenschaften
14
,
477
(
1926
).
87.
B. J.
Kennedy
,
C. J.
Howard
,
K. S.
Knight
,
Z.
Zhang
, and
Q.
Zhou
, “
Structures and phase transitions in the ordered double perovskites Ba2BiIIIBiVO6 and Ba2BiIIISbVO6
,”
Acta Crystallogr., Sect. B
62
,
537
(
2006
).
88.
B. H.
Yan
,
M.
Jansen
, and
C.
Felser
, “
A large-energy-gap oxide topological insulator based on the superconductor BaBiO3
,”
Nat. Phys.
9
,
709
(
2013
).
89.
M. L.
Medarde
, “
Structural, magnetic and electronic properties of perovskites (R = rare earth)
,”
J. Phys.: Condens. Matter
9
,
1679
(
1997
).
90.
J. B.
Torrance
,
P.
Lacorre
,
A. I.
Nazzal
,
E. J.
Ansaldo
, and
C.
Niedermayer
, “
Systematic study of insulator-metal transitions in perovskites RNiO3 (R = Pr,Nd,Sm,Eu) due to closing of charge-transfer gap
,”
Phys. Rev. B
45
,
8209
(
1992
).
91.
B.
Bradlyn
,
J.
Cano
,
Z.
Wang
,
M. G.
Vergniory
,
C.
Felser
,
R. J.
Cava
, and
B. A.
Bernevig
, “
Beyond Dirac and Weyl fermions: Unconventional quasiparticles in conventional crystals
,”
Science
353
,
aaf5037
(
2016
).
92.
D.
Di Sante
,
A.
Hausoel
,
P.
Barone
,
J. M.
Tomczak
,
G.
Sangiovanni
, and
R.
Thomale
, “
Realizing double Dirac particles in the presence of electronic interactions
,”
Phys. Rev. B
96
,
121106
(
2017
).
93.
J.
Ihm
,
A.
Zunger
, and
M. L.
Cohen
, “
Momentum-space formalism for the total energy of solids
,”
J. Phys. C
12
,
4409
(
1979
).
94.
J.
Slater
, “
Magnetic effects and the Hartree-Fock equation
,”
Phys. Rev.
82
,
538
(
1951
).
95.
Z.
Zhang
,
Y.
Zhao
, and
M.
Zhu
, “
NiO films consisting of vertically aligned cone-shaped NiO rods
,”
Appl. Phys. Lett.
88
,
033101
(
2006
).
96.
N. F.
Mott
and
Z.
Zinamon
, “
The metal-nonmetal transition
,”
Rep. Prog. Phys.
33
,
881
(
1970
).
97.
G.
Sharma
,
Z.
Zhao
,
P.
Sarker
,
B. A.
Nail
,
J.
Wang
,
M. N.
Huda
, and
F. E.
Osterloh
, “
Electronic structure, photovoltage, and photocatalytic hydrogen evolution with p-CuBi2O4 nanocrystals
,”
J. Mater. Chem. A
4
,
2936
(
2016
).
98.
X.
Ren
,
I.
Leonov
,
G.
Keller
,
M.
Kollar
,
I.
Nekrasov
, and
D.
Vollhardt
, “
LDA+DMFT computation of the electronic spectrum of NiO
,”
Phys. Rev. B
74
,
195114
(
2006
).
99.
Y.
Shinohara
,
S.
Sharma
,
S.
Shallcross
,
N.
Lathiotakis
, and
E.
Gross
, “
Spectrum for nonmagnetic Mott insulators from power functional within reduced density matrix functional theory
,”
J. Chem. Theory Comput.
11
,
4895
(
2015
).
100.
J.
Kuneš
,
V.
Anisimov
,
A.
Lukoyanov
, and
D.
Vollhardt
, “
Local correlations and hole doping in NiO: A dynamical mean-field study
,”
Phys. Rev. B
75
,
165115
(
2007
).
101.
S.
Biermann
,
A.
Poteryaev
,
A.
Lichtenstein
, and
A.
Georges
, “
Dynamical singlets and correlation-assisted Peierls transition in VO2
,”
Phys. Rev. Lett.
94
,
026404
(
2005
).
102.
G.
Kotliar
,
S. Y.
Savrasov
,
K.
Haule
,
V. S.
Oudovenko
,
O.
Parcollet
, and
C. A.
Marianetti
, “
Electronic structure calculations with dynamical mean-field theory
,”
Rev. Mod. Phys.
78
,
865
(
2006
).
103.
H. T.
Dang
,
X.
Ai
,
A. J.
Millis
, and
C. A.
Marianetti
, “
Density functional plus dynamical mean-field theory of the metal-insulator transition in early transition-metal oxides
,”
Phys. Rev. B
90
,
125114
(
2014
).
104.
V. I.
Anisimov
,
A. I.
Poteryaev
,
M. A.
Korotin
,
A. O.
Anokhin
, and
G.
Kotliar
, “
First-principles calculations of the electronic structure and spectra of strongly correlated systems: Dynamical mean-field theory
,”
J. Phys.: Condens. Matter
9
,
7359
(
1997
).
105.
T.
Yoshida
,
A.
Ino
,
T.
Mizokawa
,
A.
Fujimori
,
Y.
Taguchi
,
T.
Katsufuji
, and
Y.
Tokura
, “
Photoemission spectral weight transfer and mass renormalization in the Fermi-liquid system La1−xSrxTiO3+y/2
,”
Europhys. Lett.
59
,
258
(
2002
).
106.
A.
Fujimori
,
I.
Hase
,
M.
Nakamura
,
H.
Namatame
,
Y.
Fujishima
,
Y.
Tokura
,
M.
Abbate
,
F. M. F.
de Groot
,
M. T.
Czyzyk
,
J. C.
Fuggle
,
O.
Strebel
,
F.
Lopez
,
M.
Domke
, and
G.
Kaindl
, “
Doping-induced changes in the electronic structure of LaxSr1−xTiO3: Limitation of the one-electron rigid-band model and the Hubbard model
,”
Phys. Rev. B
46
,
9841
(
1992
).
107.
A.
Fujimori
,
I.
Hase
,
H.
Namatame
,
Y.
Fujishima
,
Y.
Tokura
,
H.
Eisaki
,
S.
Uchida
,
K.
Takegahara
, and
F. M. F.
de Groot
, “
Evolution of the spectral function in Mott-Hubbard systems with d1 configuration
,”
Phys. Rev. Lett.
69
,
1796
(
1992
).
108.
K.
Morikawa
,
T.
Mizokawa
,
A.
Fujimori
,
Y.
Taguchi
, and
Y.
Tokura
, “
Photoemission spectral weight distribution in Y1-xCaxTiO3
,”
Phys. Rev. B
54
,
8446
(
1996
).
109.
R.
Marchand
,
L.
Brohan
, and
M.
Tournoux
, “
TiO2(B) a new form of titanium dioxide and the potassium octatitanate K2Ti8O17
,”
Mater. Res. Bull.
15
,
1129
(
1980
).
110.
Y.
Tang
,
Y.
Zhang
,
O. I.
Malyi
,
N.
Bucher
,
H.
Xia
,
S.
Xi
,
Z.
Zhu
,
Z.
Lv
,
W.
Li
,
J.
Wei
,
M.
Srinivasan
,
A.
Borgna
,
M.
Antonietti
,
Y.
Du
, and
X.
Chen
, “
Identifying the origin and contribution of surface storage in TiO2(B) nanotube electrode by in situ dynamic valence state monitoring
,”
Adv. Mater.
30
,
1802200
(
2018
).
111.
S. M.
Kazakov
,
C.
Chaillout
,
P.
Bordet
,
J. J.
Capponi
,
M.
NunezRegueiro
,
A.
Rysak
,
J. L.
Tholence
,
P. G.
Radaelli
,
S. N.
Putilin
, and
E. V.
Antipov
, “
Discovery of a second family of bismuth-oxide-based superconductors
,”
Nature
390
,
148
(
1997
).
112.
B.
Khamari
,
R.
Kashikar
, and
B. R. K.
Nanda
, “
Topologically invariant double Dirac states in bismuth-based perovskites: Consequence of ambivalent charge states and covalent bonding
,”
Phys. Rev. B
97
,
045149
(
2018
).
113.
R.
Kashikar
,
B.
Khamari
, and
B. R. K.
Nanda
, “
Second-neighbor electron hopping and pressure induced topological quantum phase transition in insulating cubic perovskites
,”
Phys. Rev. Mater.
2
,
124204
(
2018
).
114.
Y.
Zhang
,
O. I.
Malyi
,
Y.
Tang
,
J.
Wei
,
Z.
Zhu
,
H.
Xia
,
W.
Li
,
J.
Guo
,
X.
Zhou
,
Z.
Chen
,
C.
Persson
, and
X.
Chen
, “
Reducing the charge carrier transport barrier in functionally layer-graded electrodes
,”
Angew. Chem. Int. Ed.
56
,
14847
(
2017
).
115.
G. M.
Dalpian
,
Q.
Liu
,
J.
Varignon
,
M.
Bibes
, and
A.
Zunger
, “
Bond disproportionation, charge self-regulation, and ligand holes in ABX3 perovskites by density functional theory
,”
Phys. Rev. B
98
,
075135
(
2018
).
116.
F.
Li
,
B. A.
Davidson
,
R.
Sutarto
,
H.
Shin
,
C.
Liu
,
I.
Elfimov
,
K.
Foyevtsova
,
F.
He
,
G. A.
Sawatzky
, and
K.
Zou
, “
Epitaxial growth of perovskite SrBiO3 film on SrTiO3 by oxide molecular beam epitaxy
,”
Phys. Rev. Mater.
3
,
100802
(
2019
).
117.
K.
Foyevtsova
,
A.
Khazraie
,
I.
Elfimov
, and
G. A.
Sawatzky
, “
Hybridization effects and bond disproportionation in the bismuth perovskites
,”
Phys. Rev. B
91
,
121114
(
2015
).
118.
H.
Raebiger
,
S.
Lany
, and
A.
Zunger
, “
Charge self-regulation upon changing the oxidation state of transition metals in insulators
,”
Nature
453
,
763
(
2008
).
119.
Y.
Zhang
,
Y.
Tang
,
J.
Deng
,
W. R.
Leow
,
H.
Xia
,
Z.
Zhu
,
Z.
Lv
,
J.
Wei
,
W.
Li
,
C.
Persson
,
O. I.
Malyi
,
M.
Antonietti
, and
X.
Chen
, “
Correlating the Peukert's constant with phase composition of electrode materials in fast lithiation processes
,”
ACS Mater. Lett.
1
,
519
(
2019
).
120.
O. I.
Malyi
and
A.
Zunger
, “
Hole antidoping of oxides
,”
Phys. Rev. B
101
,
235202
(
2020
).
121.
B. J.
Morgan
and
G. W.
Watson
, “
GGA+U description of lithium intercalation into anatase TiO2
,”
Phys. Rev. B
82
,
144119
(
2010
).
122.
A. C. M.
Padilha
,
J. M.
Osorio-Guillén
,
A. R.
Rocha
, and
G. M.
Dalpian
, “
TinO2n-1 Magnéli phases studied using density functional theory
,”
Phys. Rev. B
90
,
035213
(
2014
).
123.
C.
Loschen
,
J.
Carrasco
,
K. M.
Neyman
, and
F.
Illas
, “
First-principles LDA+U and GGA+U study of cerium oxides: Dependence on the effective U parameter
,”
Phys. Rev. B
75
,
035115
(
2007
).
124.
S.
Balandeh
,
R. J.
Green
,
K.
Foyevtsova
,
S.
Chi
,
O.
Foyevtsov
,
F.
Li
, and
G. A.
Sawatzky
, “
Experimental and theoretical study of the electronic structure of single-crystal BaBiO3
,”
Phys. Rev. B
96
,
165127
(
2017
).
125.
J.
Varignon
,
M. N.
Grisolia
,
J.
Íñiguez
,
A.
Barthélémy
, and
M.
Bibes
, “
Complete phase diagram of rare-earth nickelates from first-principles
,”
npj Quantum Mater.
2
,
21
(
2017
).
126.
H.
Park
,
A. J.
Millis
, and
C. A.
Marianetti
, “
Site-selective Mott transition in rare-earth-element nickelates
,”
Phys. Rev. Lett.
109
,
156402
(
2012
).
127.
J.
Varignon
,
M.
Bibes
, and
A.
Zunger
, “
Origins versus fingerprints of the Jahn-Teller effect in d-electron ABX3 perovskites
,”
Phys. Rev. Res.
1
,
033131
(
2019
).
128.
L.
Craco
,
M. S.
Laad
,
S.
Leoni
, and
E.
Müller-Hartmann
, “
Insulator-metal transition in the doped 3d1 transition metal oxide LaTiO3
,”
Phys. Rev. B
70
,
195116
(
2004
).
129.
M.
Baldini
,
V. V.
Struzhkin
,
A. F.
Goncharov
,
P.
Postorino
, and
W. L.
Mao
, “
Persistence of Jahn-Teller distortion up to the insulator to metal transition in LaMnO3
,”
Phys. Rev. Lett.
106
,
066402
(
2011
).
130.
Z.
Wang
,
O. I.
Malyi
,
X.
Zhao
, and
A.
Zunger
, “
Mass enhancement in 3d and s-p perovskites from symmetry breaking
,” arXiv:2006.10099 (
2020
).
131.
M. M.
Schmitt
,
Y.
Zhang
,
A.
Mercy
, and
P.
Ghosez
, “
Electron-lattice interplay in LaMnO3 from canonical Jahn-Teller distortion notations
,”
Phys. Rev. B
101
,
214304
(
2020
).
132.
W. C.
Koehler
and
E. O.
Wollan
, “
Neutron-diffraction study of the magnetic properties of perovskite-like compounds LaBO3
,”
J. Phys. Chem. Solids
2
,
100
(
1957
).
133.
X.
Qiu
,
T.
Proffen
,
J. F.
Mitchell
, and
S. J. L.
Billinge
, “
Orbital correlations in the pseudocubic O and rhombohedral R phases of LaMnO3
,”
Phys. Rev. Lett.
94
,
177203
(
2005
).
134.
J. S.
Zhou
and
J. B.
Goodenough
, “
Paramagnetic phase in single-crystal LaMnO3
,”
Phys. Rev. B
60
,
R15002
(
1999
).
135.
M.
Cwik
,
T.
Lorenz
,
J.
Baier
,
R.
Müller
,
G.
André
,
F.
Bourée
,
F.
Lichtenberg
,
A.
Freimuth
,
R.
Schmitz
,
E.
Müller-Hartmann
, and
M.
Braden
, “
Crystal and magnetic structure of LaTiO3: Evidence for nondegenerate t2g orbitals
,”
Phys. Rev. B
68
,
060401
(
2003
).
136.
S.
Margadonna
and
G.
Karotsis
, “
Cooperative Jahn-Teller distortion, phase transitions, and weak ferromagnetism in the KCrF3 perovskite
,”
J. Am. Chem. Soc.
128
,
16436
(
2006
).
137.
A.
Okazaki
and
Y.
Suemune
, “
The crystal structure of KCuF3
,”
J. Phys. Soc. Jpn.
16
,
176
(
1961
).
138.
B.
Li
,
A.-V.
Mudring
, and
J. D.
Corbett
, “
Valence compounds versus metals. Synthesis, characterization, and electronic structures of cubic Ae4Pn3 phases in the systems Ae = Ca, Sr.,, Ba, Eu; Pn = As, Sb, Bi
,”
Inorg. Chem.
42
,
6940
(
2003
).
139.
G. L. W.
Hart
and
A.
Zunger
, “
Origins of nonstoichiometry and vacancy ordering in Sc1-xxS
,”
Phys. Rev. Lett.
87
,
275508
(
2001
).
140.
J. E.
Bernard
and
A.
Zunger
, “
Ordered-vacancy-compound semiconductors: Pseudocubic CdIn2Se4
,”
Phys. Rev. B
37
,
6835
(
1988
).
141.
F.
Ducastelle
,
Order and Phase Stability in Alloys
(
North-Holland
,
1991
).
142.
W. A.
England
,
A. J.
Jacobson
, and
B. C.
Tofield
, “
Structural studies of highly non-stoichiometric polycrystalline sodium and silver beta-aluminas
,”
Solid State Ionics
6
,
21
(
1982
).
143.
V.
Stevanović
,
S.
Lany
,
X.
Zhang
, and
A.
Zunger
, “
Correcting density functional theory for accurate predictions of compound enthalpies of formation: Fitted elemental-phase reference energies
,”
Phys. Rev. B
85
,
115104
(
2012
).
144.
W. P.
Huhn
and
V.
Blum
, “
One-hundred-three compound band-structure benchmark of post-self-consistent spin-orbit coupling treatments in density functional theory
,”
Phys. Rev. Mater.
1
,
033803
(
2017
).
145.
S.-W.
Kim
,
C.
Liu
,
H.-J.
Kim
,
J.-H.
Lee
,
Y.
Yao
,
K.-M.
Ho
, and
J.-H.
Cho
, “
Nature of the insulating ground state of the 5d postperovskite CaIrO3
,”
Phys. Rev. Lett.
115
,
096401
(
2015
).
146.
K.
Ohgushi
,
H.
Gotou
,
T.
Yagi
,
Y.
Kiuchi
,
F.
Sakai
, and
Y.
Ueda
, “
Metal-insulator transition in Ca1−xNaxIrO3 with post-perovskite structure
,”
Phys. Rev. B
74
,
241104
(
2006
).
147.
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
, “
Novel Jeff=1/2 Mott state induced by relativistic spin-orbit coupling in Sr2IrO4
,”
Phys. Rev. Lett.
101
,
076402
(
2008
).
148.
M.
Reticcioli
,
U.
Diebold
,
G.
Kresse
, and
C.
Franchini
, in
Handbook of Materials Modeling: Applications: Current and Emerging Materials
, edited by
W.
Andreoni
and
S.
Yip
(
Springer International Publishing
,
Cham
,
2019
), p.
1
.
149.
T.
Holstein
, “
Studies of polaron motion: Part I. The molecular-crystal model
,”
Ann. Phys.
8
,
325
(
1959
).
150.
A. R.
Elmaslmane
,
M. B.
Watkins
, and
K. P.
McKenna
, “
First-principles modeling of polaron formation in TiO2 polymorphs
,”
J. Chem. Theory Comput.
14
,
3740
(
2018
).
151.
A. C. M.
Padilha
,
H.
Raebiger
,
A. R.
Rocha
, and
G. M.
Dalpian
, “
Charge storage in oxygen deficient phases of TiO2: Defect Physics without defects
,”
Sci. Rep.
6
,
28871
(
2016
).
152.
J. P.
Allen
and
G. W.
Watson
, “
Occupation matrix control of d- and f-electron localisations using DFT + U
,”
Phys. Chem. Chem. Phys.
16
,
21016
(
2014
).
153.
G. F.
Koster
and
J. C.
Slater
, “
Simplified impurity calculation
,”
Phys. Rev.
96
,
1208
(
1954
).
154.
H. P.
Hjalmarson
,
P.
Vogl
,
D. J.
Wolford
, and
J. D.
Dow
, “
Theory of substitutional deep traps in covalent semiconductors
,”
Phys. Rev. Lett.
44
,
810
(
1980
).
155.
Q.
Liu
,
G. M.
Dalpian
, and
A.
Zunger
, “
Antidoping in insulators and semiconductors having intermediate bands with trapped carriers
,”
Phys. Rev. Lett.
122
,
106403
(
2019
).
156.
M.
Kotiuga
and
K. M.
Rabe
, “
High-density electron doping of SmNiO3 from first principles
,”
Phys. Rev. Mater.
3
,
115002
(
2019
).
157.
Z.
Zhang
,
D.
Schwanz
,
B.
Narayanan
,
M.
Kotiuga
,
J. A.
Dura
,
M.
Cherukara
,
H.
Zhou
,
J. W.
Freeland
,
J.
Li
,
R.
Sutarto
,
F.
He
,
C.
Wu
,
J.
Zhu
,
Y.
Sun
,
K.
Ramadoss
,
S. S.
Nonnenmann
,
N.
Yu
,
R.
Comin
,
K. M.
Rabe
,
S. K. R. S.
Sankaranarayanan
, and
S.
Ramanathan
, “
Perovskite nickelates as electric-field sensors in salt water
,”
Nature
553
,
68
(
2018
).
158.
Y.
Sun
,
M.
Kotiuga
,
D.
Lim
,
B.
Narayanan
,
M.
Cherukara
,
Z.
Zhang
,
Y.
Dong
,
R.
Kou
,
C.-J.
Sun
,
Q.
Lu
,
I.
Waluyo
,
A.
Hunt
,
H.
Tanaka
,
A. N.
Hattori
,
S.
Gamage
,
Y.
Abate
,
V. G.
Pol
,
H.
Zhou
,
S. K. R. S.
Sankaranarayanan
,
B.
Yildiz
,
K. M.
Rabe
, and
S.
Ramanathan
, “
Strongly correlated perovskite lithium ion shuttles
,”
Proc. Natl. Acad. Sci. U.S.A.
115
,
9672
(
2018
).
159.
N.
Lu
,
P.
Zhang
,
Q.
Zhang
,
R.
Qiao
,
Q.
He
,
H.-B.
Li
,
Y.
Wang
,
J.
Guo
,
D.
Zhang
,
Z.
Duan
,
Z.
Li
,
M.
Wang
,
S.
Yang
,
M.
Yan
,
E.
Arenholz
,
S.
Zhou
,
W.
Yang
,
L.
Gu
,
C.-W.
Nan
,
J.
Wu
,
Y.
Tokura
, and
P.
Yu
, “
Electric-field control of tri-state phase transformation with a selective dual-ion switch
,”
Nature
546
,
124
(
2017
).
160.
M.
Brahlek
,
L.
Zhang
,
C.
Eaton
,
H.-T.
Zhang
, and
R.
Engel-Herbert
, “
Accessing a growth window for SrVO3 thin films
,”
Appl. Phys. Lett.
107
,
143108
(
2015
).
161.
R. G.
Gordon
, “
Criteria for choosing transparent conductors
,”
MRS Bull.
25
,
52
(
2000
).
162.
T.
Minami
, “
Present status of transparent conducting oxide thin-film development for indium-tin-oxide (ITO) substitutes
,”
Thin Solid Films
516
,
5822
(
2008
).
163.
G.
Brunin
,
F.
Ricci
,
V.-A.
Ha
,
G.-M.
Rignanese
, and
G.
Hautier
, “
Transparent conducting materials discovery using high-throughput computing
,”
npj Comput. Mater.
5
,
63
(
2019
).
164.
N.
Suzuki
,
S.
Ohira
,
M.
Tanaka
,
T.
Sugawara
,
K.
Nakajima
, and
T.
Shishido
, “
Fabrication and characterization of transparent conductive Sn-doped β-Ga2O3 single crystal
,”
Phys. Status Solidi C
4
,
2310
(
2007
).
165.
A.
Prakash
,
P.
Xu
,
A.
Faghaninia
,
S.
Shukla
,
J. W.
Ager
,
C. S.
Lo
, and
B.
Jalan
, “
Wide band gap BaSnO3 films with room temperature conductivity exceeding 104 S cm−1
,”
Nat. Commun.
8
,
15167
(
2017
).