This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Quantifying the stability of the anion ordering in SrVO2H

Masayuki Ochi    Kazuhiko Kuroki Department of Physics, Osaka University, Machikaneyama-cho, Toyonaka, Osaka 560-0043, Japan
Abstract

We investigate a text-book mixed-anion compound SrVO2H using first-principles calculation, to theoretically pin down the factors that stabilize its anion ordering. We find that the trans preference by the characteristic crystal field in the VO4H2 octahedron in addition to a coherent shrinkage along the V-H-V direction, taking place when such direction is consistent among neighboring hydrogens, stabilize the anion ordering observed in experiment. Our study gives an important clue for controlling the anion ordering in transition metal oxyhydrides.

I Introduction

Transition metal oxides are of particular importance in condensed matter physics SC due to their wide variety of physical properties. Transition metal cations therein play a dominant role in determining their physical properties; anions, on the other hand, have been regarded as subsidiary constituents. However, mixed-anion strategy – to say, employing multiple anions for controlling materials properties – have recently attracted increasing attention mixed_review . In particular, transition metal oxyhydride is a special class of mixed-anion compounds, a unique electronic structure of which originates from several differences between O2- and H- ions. Their different valence numbers enable the carrier control and lead us to the unexplored phases of iron-based superconductors FeAs1 ; FeAs2 . A unique crystal field for cation dd orbitals is realized by the inequivalency between the two anions SVOH_Angew ; SVOH_epitaxial ; SVOH_JACS ; SVOH_Yamamoto ; SVOH_theory ; SVOH_theory_wan ; cRPAochi . Their different compressibility results in the strongly anisotropic deformation by external pressure for SrVO2SVOH_Yamamoto . Also, their different valence orbitals, pp for oxygen and ss for hydrogen, drastically changes the dimensionality of the electronic structure in early transition metal oxyhydrides because the t2gt_{2g} orbitals of cation cannot form a chemical bond with the H-ss orbital unlike with the O-pp orbitals, due to their different spatial symmetry SVOH_Angew ; SVOH_epitaxial ; SVOH_JACS ; SVOH_Yamamoto ; SVOH_theory ; SVOH_theory_wan ; cRPAochi . Some recent theoretical studies focus on a unique electronic structure realized in transition metal oxyhydrides in the context of nickelate superconductivity Held ; Kitamine .

Multiple anions, however, also bring an inevitable but crucial difficulty in treating a huge variety of anion configurations. For molecules, a stable configuration of different kinds of ligands and its origin were investigated in many systems (e.g., Refs. mol1, ; mol2, ). For mixed-anion solids, however, the number of possible anion configurations exponentially increases by increasing the size of the simulation cell, which makes its theoretical analysis very difficult. In usual cases, the anion configuration in mixed-anion solids observed in experiment is interpreted on the basis of the knowledge of molecular systems. For example, in perovskite oxynitrides AMAMO2N (AA: alkaline earth metal, MM: transition metal), an energy profit by strong covalent bonding of MM dπd_{\pi}-N pπp_{\pi} is considered to be maximized in a cis-MMO4N2 octahedron rather than trans one, on the basis of the knowledge of complex chemistry mol1 ; mol2 , and thus is regarded to be the origin of the cis-preference in this system oxynit_order . As another example, the stability of the anion ordering in Srn+1VnO2n+1Hn (n=1,2,n=1,2,\inftySVOH_Angew ; SVOH_JACS , (Fig. 1(a) for the n=n=\infty case, SrVO2H) is usually explained by the trans preference of the d2d^{2} filling in the isolated VO4H2 octahedron (e.g., Ref. SVOH_Angew, ), as shown in Fig. 1(b). These explanations are tested by some first-principles calculations Fang ; Wolff ; SVTiOH , but the number of configurations investigated in their studies is very limited, owing to the theoretical difficulty mentioned above. Thus, it is indispensable to perform first-principles calculation for more extensive set of anion configurations. In fact, such a calculation has recently become possible BaVOH ; Kaneko ; BaScHO . Firm physical insight on the anion ordering is of great help to control anion ordering in experiments, which is possible only for very few cases strain_control ; kageyama_svoh_strain in the present day.

In this paper, we investigate the relation between the total energy and the anion configurations in SrVO2H, a text-book mixed-anion compound, using the first-principles calculation, in order to theoretically pin down the factors that stabilize the anion ordering in this compound. We find that the trans preference by the characteristic crystal field in the VO4H2 octahedron is indeed important for realizing the anion ordering, but is not sufficient to identify the stable structure. In fact, we identify another important factor: whether the structure allows a coherent shrinkage along the V-H-V direction, which can be induced by the different size of the O2- and H- ions and yields a sizable stabilization of the system. This study gives an important clue for controlling the anion ordering in transition metal oxyhydrides.

This paper is organized as follows. Section II presents our computational model and calculation conditions. In Secs. III A–C, we present our calculation results performed under several conditions and discuss what stabilizes the anion ordering in SrVO2H. Additional discussions that reinforce our view including a rough estimate of some minor effects not included in our simulation are shown in Secs. III D–F. Section IV summarizes this study.

Refer to caption
Figure 1: (a) Crystal structure of SrVO2H, (b) trans and cis configurations of hydrogens in VO4H2 octahedra shown together with their crystal fields for the V-t2gt_{2g} orbitals with d2d^{2} filling, (c) definition of a pair of two hydrogens, the number of which are counted as nneighborn_{neighbor}, shown with a solid circle, together with a pair of hydrogens with their distance being a/2a/\sqrt{2} shown with a dotted circle (see details in the main text). The green, blue, red, and white spheres represent Sr, V, O, and H atoms, respectively. Although t2gt_{2g} is an inappropriate name for oxyhydrides with a lowered crystal-field symmetry in the strict sense of the term, we call the dxy,yz,xzd_{xy,yz,xz} orbitals the ‘t2gt_{2g} orbitals’ for simplicity. The crystal structures shown in this paper were depicted using the VESTA software VESTA .

II Computational model and methods

II.1 Computational model

As a computational model, we considered possible anion configurations in Sr8V8O16H8, by replacing one third of oxygens with hydrogens in the 2×2×22\times 2\times 2 supercell of SrVO3, and optimized their structures in the way described in Sec. II.2. Because of a formidable number of possible anion configurations even for our relatively small-sized simulation cell, we restricted our analysis to the configurations satisfying that each vanadium atom has two nearest-neighboring hydrogen atoms. This constraint is energetically natural and often applied for analysis of anion (dis)ordering in materials with MMOX24{}_{4}X_{2} octahedra (XX: anion other than oxygen) corr_disorder . By using this assumption, the number of possible (symmetrically-inequivalent) crystal structures becomes around 250.

We focused on the following two quantities to characterize anion configurations. One is the number of trans configurations of hydrogens around vanadiums, ntransn_{trans}, which is up to eight since our supercell includes eight vanadium atoms. The other one is the number of a pair of two hydrogens with specific relative positions as shown in Fig. 1(c) with a solid circle, nneighborn_{neighbor}. Note that the atoms used for defining nneighborn_{neighbor} are two vanadiums and two hydrogens surrounded with the solid circle in Fig. 1(c), and the other atoms shown in this figure are irrelevant to the definition. In other words, nneighborn_{neighbor} is the number of the square-shaped relative position of these four atoms: V-V-H-H. Here, we avoid double-counting pairs of the equivalent hydrogens due to the periodic boundary condition, and then, e,g., nneighborn_{neighbor} is eight for the 2×2×22\times 2\times 2 supercell constructed from the experimental structure of SrVO2H shown in Fig. 1(a).

These two quantities were chosen by the following reason. First, we can represent the hydrogen configuration in Sr8V8O16H8 by a vector 𝒙={x1,x2,,x24}{\bm{x}}=\{x_{1},x_{2},\dots,x_{24}\} where xix_{i} denotes whether the ii-th oxygen site in Sr8V8O24 (i.e., a 2×2×22\times 2\times 2 supercell of SrVO3) is replaced with hydrogen (xi=1x_{i}=1) or not (xi=0x_{i}=0). One can also uniquely specify a hydrogen configuration by a set of the values of all the kk-th monomial xi1xi2xikx_{i_{1}}x_{i_{2}}\dots x_{i_{k}} up to 8th order, where kk-th monomial corresponds to the kk-body correlation among hydrogen (e.g., x5x7x9=1x_{5}x_{7}x_{9}=1 means that hydrogens occupy the 5th, 7th, and 9th anion sites). Thus, we expect the relation,

E(𝒙)=E0+icixi+i<jcijxixj+,E({\bm{x}})=E_{0}+\sum_{i}c_{i}x_{i}+\sum_{i<j}c_{ij}x_{i}x_{j}+\dots, (1)

where E(𝒙)E({\bm{x}}) is the total energy with the hydrogen configuration 𝒙{\bm{x}} and E0,ci,cij,E_{0},c_{i},c_{ij},\dots are unknown parameters (coefficients). This is the way used in the popular cluster-expansion method cluster , where atomic species in multicomponent systems such as alloy are denoted by a discrete variable and the energy of the system is represented with polynomials of those variables. There is no need to consider terms including xinx_{i}^{n} (n2n\geq 2) since xin=xix_{i}^{n}=x_{i}. Because a small supercell was used in this study under the periodic boundary condition, we should concentrate on short-range correlations represented with a low-degree monomial. There is no first order monomial that can be used to characterize anion configurations. This is because the equivalency among all the oxygen sites in cubic SrVO3 requires that not xix_{i} but i=124xi\sum_{i=1}^{24}x_{i} should be used to represent the total energy of this system (to say, all the coefficients cic_{i} in E(𝒙)E({\bm{x}}) should be the same by symmetry), but i=124xi\sum_{i=1}^{24}x_{i} is constant (8) in our simulation. By a similar discussion for the second order monomials, the number of equivalent hydrogen pairs ijxixj\sum_{\langle ij\rangle}x_{i}x_{j} can be used to represent the energy of this sytem, where the ii-th and jj-th hydrogens having specific relative positions that are equivalent by the crystal symmetry, are summed over. Then, only the two quantities, ntransn_{trans} and nneighborn_{neighbor}, are allowed if one considers monomials representing a pair occupation of hydrogens with their distance less than or equal to aa, a lattice constant of cubic primitive cell. Note that the inter-hydrogen length for the initial cubic structure is used in explanation just for simplicity. We also note that the number of hydrogen pairs with their distance being a/2a/\sqrt{2} as shown with a dotted circle in Fig. 1(c), is nothing but the number of cis configurations, and thus is uniquely determined by ntransn_{trans} because we restrict the anion configurations to those where all the vanadiums have two adjacent hydrogens. For simplicity, we omit the higher-order monomials in the following analysis. It is an interesting future task to perform more elaborated statistical analysis such as the multiple regression analysis using many factors omitted in this study, while it might require a larger supercell without the restriction introduced in our simulation. In the latter analysis, we shall see that, even by our simplification, the total energies well correlate with the two quantities focused in this study with a clear physical interpretation.

II.2 Conditions of first-principles calculations

We performed the structural optimization using the density functional theory with the Perdew-Burke-Ernzerhof parametrization of the generalized gradient approximation (PBE-GGA) PBE and the projector augmented wave (PAW) method paw as implemented in the Vienna ab initio simulation package (VASP) vasp1 ; vasp2 ; vasp3 ; vasp4 . Both cell deformation and optimization of the atomic coordinates were allowed unless noted. The following orbitals for each element are treated as core electrons in the PAW potentials: [Ar]3d103d^{10} for Sr, [Ne] for V, and [He] for O. We started our calculation from the G-type antiferromagnetic configurations of the V-dd orbitals. Here, we need to include the spin polarization to describe the d2d^{2} occupation into the crystal field shown in Fig. 1(b), which was pointed out to be a key to realize the trans preference of this system in the previous studies as mentioned in the introduction. We included the +U+U correction DFTU1 ; DFTU2 with UJ=3U-J=3 eV for the V-dd orbitals, which is a typical value for the 3d3d orbitals and is also consistent with our first-principles estimation by constrained random-phase approximation cRPAochi . Here, we used the simplified rotationally invariant approach to the DFT+U+U method introduced by Dudarev et alDFTU2 , where the correction term only depends on UJU-J rather than the values of UU and JJ themselves. Note that SrVO2H becomes gapless in simple PBE-GGA calculation without the +U+U correction even when the spin polarization is taken into account, owing to the well-known underestimation of the band gap in PBE-GGA, while the system is insulating in experiments SVOH_Angew . Further discussion on our choice of the UJU-J parameter is presented in Sec. III.1. While our simple DFT+U+U calculation cannot describe the electron correlation effects completely, we consider that our simulation captures important factors determining the anion ordering as we shall see: a large structural change by introduction of hydrogen and the different crystal fields between trans and cis configurations. This is to some extent guaranteed by the spin-polarized insulating state observed in experimental studies at low temperature, where such a static approximation of the electron-electron interaction is somewhat validated. Electron correlation effects beyond our treatment, e.g., renormalization of the size of the crystal-field splitting, is an interesting issue but requires high computational cost and then out of scope of this study. The initial crystal structure was generated in the way described before with a small random displacement, around 0.010.01 in the crystal coordinate of the supercell for each direction, added to break the symmetry before the structural optimization. The plane-wave cutoff energy of 550 eV and a 6×6×66\times 6\times 6 𝒌\bm{k}-mesh were used. We applied Gaussian smearing with a smearing width of 0.1 eV. Structural optimization was continued until the Hellmann-Feynman force on each atom becomes less than 0.01 eVÅ-1 for each direction. The spin-orbit coupling was not taken into account throughout this paper.

Refer to caption
Figure 2: (a) Total energy (eV) of all the calculated structures plotted against ntransn_{trans}, where color of each point represents nneighborn_{neighbor}. The total energy for the 2×2×22\times 2\times 2 supercell relative to that for structure A are shown throughout this paper. Δ\Delta defined in the main text, and the two all-trans structures called structures A and A in this paper, are also shown. (b) Crystal structure B, which has the lowest energy except structures A and A, and so used for calculating Δ\Delta, in panel (a). (c) Plot similar to panel (a) but shown against nneighborn_{neighbor} instead of ntransn_{trans}.

III Results and discussions

III.1 Total energy and characteristic quantities

Figure 2(a) presents the total energies of computed structures plotted against ntransn_{trans} colored using a value of nneighborn_{neighbor}. Here, the two crystal structures with all-trans configurations called structures A and A in this paper are also shown. We define Δ\Delta as the total energy of the most stable structure except A and A, relative to that for structure A. The corresponding crystal structure with ntrans=4n_{trans}=4, called structure B in this paper, is shown in Fig. 2(b). Note that structure B is not used for calculating Δ\Delta under different conditions where another structure with ntrans8n_{trans}\neq 8 is more stable than structure B, as shown in the following sections. Because of the periodic boundary condition applied to our small supercell, ntransn_{trans} always takes an even value in our simulation. We find that structure A has the lowest energy in our simulation, which is consistent with experiments. The total energies of the calculated configurations tend to be lowered by increasing ntransn_{trans}. However, the all-trans structure A has a sizably higher energy than that for structure A. This means that the explanation of the anion ordering in the previous studies based on the stable d2d^{2} crystal field for the trans configurations, is partially correct from the viewpoint of the negative correlation between the total energy and ntransn_{trans} as mentioned above but insufficient to describe the stability of anion ordering in this system. We shall come back to detailed analysis on the ntransn_{trans} dependence later in this paper, but first we focus on the difference between structures A and A, which should be a key to understand what is missing in the description by ntransn_{trans}.

One noticeable difference between structures A and A is whether hydrogens are placed onto the same (abab) plane or not, which can be captured by nneighborn_{neighbor}. The colors of data points in Fig. 2(a) show that not only these two structures but also the other data points exhibit a tendency that a large nneighborn_{neighbor} stabilizes the energy. In Fig. 2(c), the total energy is plotted against nneighborn_{neighbor} instead of ntransn_{trans}. For each set of data points with the same ntransn_{trans}, we can verify the negative correlation between the total energy and nneighborn_{neighbor}. A possible interpretation is as follows. It is known that the size of the H- and O2- ions are different and thus the lattice constants of SrVO2H are quite anisotropic: a=3.9290(6)a=3.9290(6)Å and c=3.6569(5)c=3.6569(5)Å  in experiment SVOH_Angew . In other words, introduction of hydrogen inclines the crystal to shrink along the V-H-V direction. In structure A, the hydrogen occupation along the same direction allows the crystal to coherently shrink along the same direction (cc direction), while such coherent shrinkage is not allowed in structure A. The coherent shrinkage by concomitant occupation of the neighboring hydrogen sites takes place when nneighborn_{neighbor} is large. We note that an existence of a hydrogen pair captured by nneighborn_{neighbor} in the present small supercell means that hydrogens occupy anion sites in a row with an infinite length. Thus, here we see that, not only a two-dimensional occupation of hydrogen in structure A, i.e., all oxygens in some abab planes are fully replaced with hydrogens, but also a one-dimensional occupation of hydrogen as captured by nneighborn_{neighbor} effectively lowers the total energy. While the small supercell used in this study is only capable of representing limited patterns of hydrogen configurations, we can partially see how the ordering of hydrogen occupation stabilizes the crystal structure.

Before moving on to further analysis, to check the robustness of our conclusion against the UJU-J value, we calculated the total energy difference between structures A, A, and B, using UJ=5U-J=5 eV. For this purpose, we did not optimize the crystal structure but just used the optimized structure using UJ=3U-J=3 eV. As a result, we find that the energy difference between structures A and B is 1.101.10 eV for UJ=5U-J=5 eV while it is 1.131.13 eV for UJ=3U-J=3 eV, and that between structures A and A is 0.970.97 eV for UJ=5U-J=5 eV while it is 0.990.99 eV for UJ=3U-J=3 eV. Therefore, we believe that our conclusion is less affected by the choice of the UJU-J value in simulation.

Refer to caption
Figure 3: (a) Total energy (eV) of all the calculated structures using the fixed lattice constant a=b=c=a=b=c= 3.864 Å for (a) and 3.999 Å for (b), plotted in the same way as Fig. 2(a).

III.2 Effect of coherent shrinkage of V-H-V: insight from cubic-cell calculation

To verify our view, we also calculated the total energies using a fixed cubic lattice, with a=b=c=a=b=c= 3.864 Å and 3.999 Å for Figs. 3(a) and (b), respectively, while atomic coordinates were relaxed. Although the overall correlation between the total energies and ntransn_{trans}, nneighborn_{neighbor} are similar to Fig. 1(a), we find that the energy difference between structures A and A are drastically decreased: from 0.990.99 eV in Fig. 2(a) to 0.290.29 and 0.160.16 eV in Figs. 3(a)–(b), respectively. This is because the tetragonal deformation, i.e., a=bca=b\neq c, is preferred for structure A but is not allowed in the cubic lattice. It is interesting that first-principles calculation of unsynthesized KTiO2H, although with restricted number of sampled structures, also pointed out that the different size between O2- and H{-} ions and a resulting shrinkage of the lattice along Ti-H-Ti are important in stabilizing structure A Tsuneyuki , where the crystal-field effect cannot take place for Ti-d0d^{0} occupation therein.

Note that, structure A with the cubic lattice constants exhibits an octahedral rotation with the V-O-V angle of 168.5 and 170.1 degrees for Figs. 3(a) and (b), respectively, unlike the that for the relaxed tetragonal lattice shown in Fig. 2(a). It is interesting that the lowest energy among the all-cis configurations with ntrans=0n_{trans}=0 relative to the total energy of structure A becomes much lower in Fig. 3(a) (0.200.20 eV) than in Fig. 3(b) (0.360.36 eV). A definitive conclusion cannot be reached within our limited calculation data, but it seems that a smaller ionic radius of H- alleviates a cramped environment for atoms, and then stabilizes the crystal structure with the cis configurations, while it is not so active for structure A with VO2 planes without hydrogen. In the experimental study on SrCrO2H, it was pointed out that the introduction of hydrogen into SrCrO3 makes its tolerance factor closer to unity and then raises the Néel temperature CrOH . It is an interesting future issue to investigate the effect of the octahedral rotation and tilt onto the stability of anion configuration, requiring some quantities to characterize such kinds of local distortion. Here we just point out that this issue has some relevance to the coherent shrinkage captured by nneighborn_{neighbor} dependence of the total energy and by the energy difference between structures A and A: both of them originate from the different anion size and the resultant structural change.

Refer to caption
Figure 4: (a) Total energy (eV) of all the calculated structures using the number of electrons (a) decreased or (b) increased by two, from the nominal Sr8V8O16H8, plotted in the same way as Fig. 2(a). (c) Schematic picture representing the d1d^{1} filling into the trans and cis crystal fields, where removed electrons from Fig. 1(b) are shown by red broken arrows. (d) That for the d3d^{3} filling where additional electrons from Fig. 1(b) are shown by red bold arrows.

III.3 Effect of the crystal-field splitting: calculations with different numbers of electrons

To get insight into the ntransn_{trans} dependency, we performed the same simulation as Fig. 2(a) but with different numbers of electrons. This is because the energy stabilization by the trans crystal field should be affected by a change of the number of electrons. Figures 4(a) and (b) present the total energies calculated using the number of electrons increased or decreased by two in the supercell, respectively. In other words, the electron occupation in the V-dd bands is d1.75d^{1.75} and d2.25d^{2.25} in average, for Figs. 4(a) and (b), respectively. For these calculation, the same amount of the background positive charge was introduced. We find that Δ\Delta is decreased from 1.131.13 eV in Fig. 2(a) to 0.610.61 and 0.810.81 eV in Figs. 4(a)–(b), respectively. The enhanced stability of the trans configuration in the d2d^{2} occupation pointed out and tested for some limited anion configurations in previous studies, is verified by our simulation for more intensive set of anion configurations. In fact, experimental studies showed that hydrogens are randomly distributed in AATiO3-xHx (A=A= Ca, Sr, Ba, Eu) TiOH1 ; TiOH2 ; TiOH3 ; TiOH4 , SrV1-xTixO1.5H1.5 SVTiOH , and SrCrO2CrOH , where the number of dd electrons are increased or decreased from SrVO2H. Note that Δ\Delta can be decreased if one uses a larger supercell that can represent anion configurations not representable in our simulation. In this sense, Δ\Delta estimated in our study is its upper bound, and so the stability of the trans configuration against the change of the electron number can be lost more easily than our simulation.

Here we point out that a fewer electron occupation than d2d^{2} can induce a Jahn-Teller instability in the trans crystal field that makes dxzd_{xz} and dyzd_{yz} inequivalent, and in fact it takes place in our simulation: V-O length becomes different by up to 0.09 Å along the aa and bb directions in the optimized crystal structure. It is interesting that a small reduction of electrons can provide an interesting platform of Jahn-Teller physics, while reducing a large number of electron will destabilize the crystal structures with the trans configuration. We note that a Jahn-Teller instability can take place also for the dnd^{n} (1<n<31<n<3) occupation in the cis configuration, but the crystal structure with cis configurations has a low symmetry in general even without the Jahn-Teller instability, and so it is difficult to quantify such an instability for the cis configurations.

III.4 Alternative estimation of the crystal-field effect

To verify our view on the crystal-field effect, we here make an alternative estimation of the energy stabilization by the crystal-field splitting. In our previous study cRPAochi , we evaluated the size of the crystal-field splitting between the dxyd_{xy} and dxz/yzd_{xz/yz} orbitals, Δcf=0.45\Delta_{cf}=0.45 eV, by constructing the Wannier functions Wannier1 ; Wannier2 . Suppose for simplicity that the energy levels of the t2gt_{2g} orbitals in the VO4H2 octahedron are simply determined by the number of hydrogens on the plane where the orbitals are extended (e.g., the xyxy plane for the dxyd_{xy} orbital), the energy diagrams for the trans and cis configurations are obtained as shown in Fig. 4(c). Therefore, the energy difference between the trans and cis configurations with d2d^{2} occupation is around Δcf/2=0.23\Delta_{cf}/2=0.23 eV per vanadium. This estimate corresponds to 0.23-0.23 ntransn_{trans} (eV) dependence of the total energy in the supercell, which amounts to the energy stabilization of 1.91.9 eV for ntrans=8n_{trans}=8 compared with ntrans=0n_{trans}=0. This is roughly consistent with the energy separation among the data set of ntrans=0n_{trans}=0 and ntrans=8n_{trans}=8 at nneighbor=4n_{neighbor}=4 and nneighbor=8n_{neighbor}=8 in Fig. 2(c). On the other hand, at nneighbor=4n_{neighbor}=4 in Fig. 2(c), the energy separation between the data set of ntrans=2n_{trans}=2 and ntrans=4n_{trans}=4, and that between ntrans=4n_{trans}=4 and ntrans=8n_{trans}=8, are around 0.2 eV, which is a few times smaller than our rough estimate. One possible reason is that, as we have seen, the introduction of hydrogen can induce a large structural change, which might alter the size of the crystal-field splitting from our rough estimate. In addition, we can do a similar estimate of how Δ\Delta changes by adding or removing two electrons from or into the supercell. By considering the cis and trans crystal fields shown in Figs. 4(c)–(d), the resultant change of Δ\Delta is Δcf=0.45\Delta_{cf}=0.45 eV for the both cases (i.e., making d1d^{1} or d3d^{3} filling for two vanadiums as we verified in the crystal structures used for calculating Δ\Delta). This rough estimate is again consistent with our simulation for the change of Δ\Delta, 0.520.52 eV for Fig. 4(a) and 0.320.32 eV for Fig. 4(b). These observations, to some extent, validate our interpretation regarding the crystal field effect. Because the introduction of hydrogen can induce a large structural change, further investigation using more extensive data sets and characteristic quantities in a larger supercell would be an important issue to distinguish the crystal field effect and others more clearly, while the trans-preference is clearly verified in our simulation.

III.5 Stabilization of random anion configuration by entropy effects

Although both our simulation and experiment show that the anion-ordered structure A is the most stable, random anion distribution can be to some extent stabilized by entropy. Here, we provide a rough estimate on the effects of possible two kinds of entropy as follows.

First, the configuration entropy for fully random anion configuration is given by Sconf=kBixilnxiS_{\mathrm{conf}}=-k_{B}\sum_{i}x_{i}\ln x_{i} per single anion site, where xix_{i} is the fraction of the anion ii, i.e., xi=2/3x_{i}=2/3 for oxygen and 1/31/3 for hydrogen in our case. Thus, an energetic gain TSconf-TS_{\mathrm{conf}} for fully random distribution compared with structure A with perfect anion ordering is 390-390 meV for the 2×2×22\times 2\times 2 supercell at T=300T=300 K, which is much smaller than Δ\Delta in Fig. 2(a) (1.13 eV). Note that the structures with random anion configurations have, in average, a larger energy by around 2 eV than structure A as shown in Fig. 2(a).

Second, the d2d^{2} occupation in the cis configuration realized for random anion configuration can yield the electronic entropy. To evaluate the upper bound of the electronic entropy effect, we assume that all the vanadiums have the cis configuration of hydrogens without a sufficiently large Jahn-Teller instability that breaks the degeneracy of the energy levels. We also assume that the G-type antiferromagnetic ordering with a high-spin state S=1S=1 is kept, where d2d^{2} occupation is two-fold degenerate in the t2gt_{2g} manifold of the cis configuration (see Fig. 4(c)). Then, the electronic entropy is given by Sel=kBln2S_{el}=k_{B}\ln 2 per vanadium, resulting in 143-143 meV stabilization for the 2×2×22\times 2\times 2 supercell at T=300T=300 K. This is in the same order of magnitude but smaller than the stabilization by the configuration entropy described above. While we assume the high-spin state (S=1S=1) here, note that the low-spin state (S=0S=0) rather stabilizes the all-trans configurations in terms of the electronic entropy effect. Additional entropy released for the paramagnetic state is not considered here because this is also acquired for the all-trans configurations above the transition temperature.

By considering the two types of entropies here, we can conclude that the entropy effects can be regarded as a relatively small correction to the total energy in SrVO2H, although this effect can be important when Δ\Delta is small, e.g., in the case of different transition metal species and/or at high temperatures.

III.6 Zero-point vibrational energy of hydrogen

It is often the case that a relatively large zero-point vibrational energy due to a light mass of hydrogen can affect the stability of the crystal structure including hydrogen. To check how the zero-point vibrational energy of hydrogens affects our conclusion, we performed phonon calculation for structures A and B because these two structures determine the value of Δ\Delta for Figs. 2(a) and Figs. 4(a)–(b), while the lowest-energy structure with ntrans8n_{trans}\neq 8 is different from structure B in Fig. 3. For simplicity, we calculated the Γ\Gamma-point phonon frequencies to give a rough estimate of the zero-point vibrational energy: iωi/2\sum_{i}\hbar\omega_{i}/2, where ii and ωi\omega_{i} denote the index of the phonon mode at the Γ\Gamma point and its frequency, respectively. For this purpose, we employed the finite displacement method as implemented in the Phonopy phonopy software in combination with VASP. Other computational conditions are the same as those for the structural optimization presented in Sec. II.2.

Figure 5 presents the calculated phonon frequencies at the Γ\Gamma point for structures A and B. While the phonon frequencies are similar as a whole between the two structures, we find that the eight highest frequencies exhibit a sizable change of around 20 meV. As a result, the zero-point vibrational energy, iωi/2\sum_{i}\hbar\omega_{i}/2, for these eight phonon modes in structure B is smaller by 79.6 meV than that for structure A. Because these high-frequency phonon modes represent the hydrogen moving toward the V-H-V direction, such an increase of the phonon frequency should take place when the V-H-V length becomes shorter. Therefore, the energy stabilization by the shrinkage of the V-H-V length that can be captured by nneighborn_{neighbor}, is partially canceled by the increase of the zero-point vibrational energy. However, the effect of the zero-point vibrational energy seems to be smaller than the energy profits discussed so far, hence does not affect our conclusion.

Refer to caption
Figure 5: Phonon frequencies at the Γ\Gamma point for structures A and B.

IV Conclusion

We have investigated how the anion ordering is realized in SrVO2H using the first-principles evaluation of the total energies for the structures with different anion configurations. We have found that there are two important factors that stabilize the experimental crystal structures: one is the trans preference by the characteristic crystal field splitting in the VO4H2 octahedron and the other one is the coherent shrinkage along the V-H-V direction. The importance of the latter is demonstrated by the large energy difference between the all-trans structures A and A. In other words, structure A observed in experiment is stabilized by the trans preference at each VO4H2 octahedron together with the effective inter-octahedron interaction that favors the parallel alignment of the V-H directions. Our study offers an important knowledge to control the anion ordering, which is indispensable for extracting desirable functionalities from the mixed-anion materials, and thus expands the possibility of materials design.

V Acknowledgments

Part of the numerical calculations were performed using the large-scale computer systems provided by the following institutions: the supercomputer center of the Institute for Solid State Physics, the University of Tokyo, the Information Technology Center, the University of Tokyo, and the Cybermedia Center, Osaka University. Computational resource from the Cybermedia Center was provided through the HPCI System Research Project (Project ID hp190022 and hp200007). This study was supported by JSPS KAKENHI (Grants No. JP19H04697 and No. JP19H05058) and JST CREST (Grant No. JPMJCR20Q4).

References

  • (1) M. Imada, A. Fujimori, and Y. Tokura, Rev. Mod. Phys. 70, 1039 (1998).
  • (2) H. Kageyama, K. Hayashi, K. Maeda, J. P. Attfield, Z. Hiroi, J. M. Rondinelli, and K. R. Poeppelmeier, Nat. Commun. 9, 772 (2018).
  • (3) S. Iimura, S. Matsuishi, H. Sato, T. Hanna, Y. Muraba, S. W. Kim, J. E. Kim, M. Takata, and H. Hosono, Nat. Commun. 3, 943 (2012).
  • (4) M. Hiraishi, S. Iimura, K. M. Kojima, J. Yamaura, H. Hiraka, K. Ikeda, P. Miao, Y. Ishikawa, S. Torii, M. Miyazaki, I. Yamauchi, A. Koda, K. Ishii, M. Yoshida, J. Mizuki, R. Kadono, R. Kumai, T. Kamiyama, T. Otomo, Y. Murakami, S. Matsuishi, and H. Hosono, Nat. Phys. 10, 300 (2014).
  • (5) F. D. Romero, A. Leach, J. S. Möller, F. Foronda, S. J. Blundell, and M. A. Hayward, Angew. Chem. Int. Ed. 53, 7556 (2014).
  • (6) J. Bang, S. Matsuishi, H. Hiraka, F. Fujisaki, T. Otomo, S. Maki, J. Yamaura, R. Kumai, Y. Murakami, and H. Hosono, J. Am. Chem. Soc. 136, 7221 (2014).
  • (7) T. Yamamoto, D. Zeng, T. Kawakami, V. Arcisauskaite, K. Yata, M. A. Patino, N. Izumo, J. E. McGrady, H. Kageyama, and M. A. Hayward, Nat. Commun. 8, 1217 (2017).
  • (8) T. Katayama, A. Chikamatsu, K. Yamada, K. Shigematsu, T. Onozuka, M. Minohara, H. Kumigashira, E. Ikenaga, and T. Hasegawa, J. Appl. Phys. 120, 085305 (2016).
  • (9) Y. Wei, H. Gui, X. Li, Z. Zhao, Y.-H. Zhao, and W. Xie, J. Phys.: Condens. Matter 27, 206001 (2015).
  • (10) K. Liu, Y. Hou, X. Gong, and H. Xiang, Sci. Rep. 6, 19653 (2016).
  • (11) M. Ochi and K. Kuroki, Phys. Rev. B 99, 155143 (2019).
  • (12) L. Si, W. Xiao, J. Kaufmann, J. M. Tomczak, Y. Lu, Z. Zhong, and K. Held, Phys. Rev. Lett. 124, 166402 (2020).
  • (13) N. Kitamine, M. Ochi, and K. Kuroki, arXiv:2007.01553 (2020).
  • (14) K. Tatsumi and R. Hoffmann, Inorg. Chem. 19, 2656 (1980).
  • (15) P. Barrie, T. A. Coffey, G. D. Forster, and G. Hogarth, J. Chem. Soc., Dalton Trans., 4519 (1999).
  • (16) M. Yang, J. Oró-Solé, J. A. Rodgers, A. B. Jorge, A. Fuertes, and J. P. Attfield, Nat. Chem. 3, 47 (2011).
  • (17) C. M. Fang, G. A. de Wijs, E. Orhan, G. de With, R. A. de Groot, H. T. Hintzen, and R. Marchand, J. Phys. Chem. Solids 64, 281 (2003).
  • (18) H. Wolff and R. Dronskowski, J. Comput. Chem. 29, 2260 (2008).
  • (19) M. A. Patino, D. Zeng, S. J. Blundell, J. E. McGrady, and M. A. Hayward, Inorg. Chem. 57, 2890 (2018).
  • (20) T. Yamamoto, K. Shitara, S. Kitagawa, A. Kuwabara, M. Kuroe, K. Ishida, M. Ochi, K. Kuroki, K. Fujii, M. Yashima, C. M. Brown, H. Takatsu, C. Tassel, and H. Kageyama, Chem. Mater. 30, 1566 (2018).
  • (21) M. Kaneko, M. Fujii, T. Hisatomi, K. Yamashita, K. Domen, J. Energy Chem. 36, 7 (2019).
  • (22) F. Takeiri, A. Watanabe, A. Kuwabara, H. Nawaz, N. I. P. Ayu, M. Yonemura, R. Kanno, and G. Kobayashi, Inorg. Chem. 58, 4431 (2019).
  • (23) D. Oka, Y. Hirose, F. Matsui, H. Kamisaka, T. Oguchi, N. Maejima, H. Nishikawa, T. Muro, K. Hayashi, and T. Hasegawa, ACS Nano 11, 3860 (2017).
  • (24) M. Namba, H. Takatsu, W. Yoshimune, A. Daniel, S. Itoh, T. Terashima, and H. Kageyama, Inorganics 8, 26 (2020).
  • (25) K. Momma and F. Izumi, J. Appl. Crystallogr. 44, 1272 (2011).
  • (26) P. J. Camp, A. Fuertes, and J. P. Attfield, J. Am. Chem. Soc. 134, 6762 (2012).
  • (27) J. M. Sanchez, F. Ducastelle, and D. Gratias, Physica A 128, 334 (1984).
  • (28) J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
  • (29) G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).
  • (30) G. Kresse and J. Hafner, Phys. Rev. B 47, 558(R) (1993).
  • (31) G. Kresse and J. Hafner, Phys. Rev. B 49, 14251 (1994).
  • (32) G. Kresse and J. Furthmüller, Comput. Mater. Sci. 6, 15 (1996).
  • (33) G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 (1996).
  • (34) A. I. Liechtenstein, V. I. Anisimov, and J. Zaanen, Phys. Rev. B 52, R5467(R) (1995).
  • (35) S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J. Humphreys, and A. P. Sutton, Phys. Rev. B 57, 1505 (1998)
  • (36) Y. Kobayashi, O. J. Hernandez, T. Sakaguchi, T. Yajima, T. Roisnel, Y. Tsujimoto, M. Morita, Y. Noda, Y. Mogami, A. Kitada, M. Ohkura, S. Hosokawa, Z. Li, K. Hayashi, Y. Kusano, J. E. Kim, N. Tsuji, A. Fujiwara, Y. Matsushita, K. Yoshimura, K. Takegoshi, M. Inoue, M. Takano, and H. Kageyama, Nat. Mater. 11, 507 (2012).
  • (37) T. Sakaguchi, Y. Kobayashi, T. Yajima, M. Ohkura, C. Tassel, F. Takeiri, S. Mitsuoka, H. Ohkubo, T. Yamamoto, J. E. Kim, N. Tsuji, A. Fujihara, Y. Matsushita, J. Hester, M. Avdeev, K. Ohoyama, and H. Kageyama, Inorg. Chem. 51, 11371 (2012).
  • (38) T. Yamamoto, R. Yoshii, G. Bouilly, Y. Kobayashi, K. Fujita, Y. Kususe, Y. Matsushita, K. Tanaka, and H. Kageyama, Inorg. Chem. 54, 1501 (2015).
  • (39) N. Masuda, Y. Kobayashi, O. Hernandez, T. Bataille, S. Paofai, H. Suzuki, C. Ritter, N. Ichijo, Y. Noda, K. Takegoshi, C. Tassel, T. Yamamoto, and H. Kageyama, J. Am. Chem. Soc. 137, 15315 (2015).
  • (40) C. Tassel, Y. Goto, Y. Kuno, J. Hester, M. Green, Y. Kobayashi, and H. Kageyama, Angew. Chem. Int. Ed. 53, 10377 (2014).
  • (41) N. Marzari and D. Vanderbilt, Phys. Rev. B 56, 12847 (1997).
  • (42) I. Souza, N. Marzari, and D. Vanderbilt, Phys. Rev. B 65, 035109 (2001).
  • (43) N. Sato and S. Tsuneyuki, Appl. Phys. Lett. 109, 172903 (2016).
  • (44) A. Togo and I. Tanaka, Scr. Mater. 108, 1 (2015).