Abstract
After the significant discovery of the hole-doped nickelate compound Nd0.8Sr0.2NiO2, analyses of the electronic structure, orbital components, Fermi surfaces and band topology could be helpful to understand the mechanism of its superconductivity. Based on first-principle calculations, we find that Ni states contribute the largest Fermi surface. The states form an electron pocket at Γ, while 5dxy states form a relatively bigger electron pocket at A. These Fermi surfaces and symmetry characteristics can be reproduced by our two-band model, which consists of two elementary band representations: B1g@1a ⊕ A1g@1b. We find that there is a band inversion near A, giving rise to a pair of Dirac points along M-A below the Fermi level upon including spin-orbit coupling. Furthermore, we perform density functional theory based Gutzwiller (DFT+Gutzwiller) calculations to treat the strong correlation effect of Ni 3d orbitals. In particular, the bandwidth of has been renormalized largely. After the renormalization of the correlated bands, the Ni 3dxy states and the Dirac points become very close to the Fermi level. Thus, a hole pocket at A could be introduced by hole doping, which may be related to the observed sign change of the Hall coefficient. By introducing an additional Ni 3dxy orbital, the hole-pocket band and the band inversion can be captured in our modified model. Besides, the nontrivial band topology in the ferromagnetic two-layer compound La3Ni2O6 is discussed and the band inversion is associated with Ni and La 5dxy orbitals.
Keywords: nickelate superconductors, band representations, topological Dirac points, DFT+Gutzwiller calculations
INTRODUCTION
After the discovery of high-Tc superconductivity in the cuprates [1,2], mixed-valent nicklates with similar crystals and electronic configurations as cuprates have attracted a lot of attention [3]. In particular, the configuration of Ni+ in infinite-layer nickelates LnNiO2 (Ln = La, Nd, Pr) is almost identical to that of Cu++ in the parent compounds of cuprates. Although much effort has been devoted along this direction in the past two decades, the possible superconductivity in mixed-valent nicklates remains elusive. Until very recently, the superconductivity with Tc = 9 ∼ 15 K was discovered in hole doped Nd0.8Sr0.2NiO2 for the first time [4]. For the parent compound NdNiO2, previous studies have established several experimental facts that are distinct from the parent compound of cuprates. First, no long-range magnetic order is observed experimentally [5,6], while an antiferromagnetically ordered state is formed in the cuprates [7]. Second, NdNiO2 exhibits a metallic behavior above 50 K [4], while the parent cuprates are Mott insulators [8]. Third, the superconductivity (so far) is only found in the hole doped Nd0.8Sr0.2NiO2, while it is found in the electron-doped cuprate Sr1−xLaxCuO2 in the same structure [9]. These experimental facts indicate that the ground state of the parent nickelates could have significant difference from the cuprates. Analyses of their electronic band structures, orbital components, Fermi surfaces and the band topology are needed. In addition, a minimal-band effective model is very helpful to further understand the mechanism of superconductivity.
In this work, we perform detailed first-principle calculations within the framework of density functional theory (DFT). The obtained band structure of the parent (undoped) compound NdNiO2 is similar to that of LaNiO2 reported previously [10] and PrNiO2 (Nd-4f and Pr-4f orbitals are treated as core states). In this article, LaNiO2 is taken as a representative and the band structures of NdNiO2 and PrNiO2 are presented in the online supplementary material. In the undoped case, there are three bands intersecting the Fermi level (EF), mainly from Ni , La 5dxy and La orbitals. They form a large cylinderlike electron pocket (EP) surrounding the Γ-Z line, a spherelike EP at A and a comparatively smaller spherelike EP at Γ, respectively. These Fermi surfaces and symmetry characteristics can be reproduced by our two-band model, which consists of two elementary band representations (EBRs): B1g@1a ⊕ A1g@1b. The EBR of B1g@1a refers to the orbital at Wyckoff site 1a[0,0,0], while the EBR of A1g@1b refers to the A1g orbital at Wyckoff site 1b[0,0,0.5], where no atoms sit.
We find that a band inversion near A happens between the Ni 3dxy states and Ln 5dxy states (the effect of Coulomb interaction U is discussed in the online supplementary material). With small U and spin-orbital coupling (SOC), this gives rise to a pair of Dirac points along M-A. After considering the renormalization of Ni 3d bands, the Dirac point becomes very close to the charge neutrality level and accessible by hole doping. As a result, a hole pocket (HP) could emerge at A, which may be responsible for the sign change of the Hall coefficient in the experiment [4]. By introducing an additional Ni 3dxy orbital, the hole-pocket band and the band inversion can be captured in the modified model. Besides, the nontrivial band topology in the ferromagnetic two-layer compound La3Ni2O6 (n=2) is discussed and a band inversion happens between the Ni and La 5dxy orbitals.
CRYSTAL STRUCTURE AND METHODOLOGY
The parent compound LaNiO2 can be obtained from the perovskites LaNiO3 by removing the apical oxygens (i.e. b site), as shown in Fig. 1(a). Consequently, it has a tetragonal lattice, and has the same planes as the cuprate superconductors with Ni+ instead of Cu++ ions. Similarly, the two-layer nickelate La3Ni2O6 (see Fig. 1(b)) can be produced [11] from the two-layer perovskite La3Ni2O7. We perform first-principle calculations using the VASP package [12,13] based on density functional theory with the projector augmented wave method [14,15]. The generalized gradient approximation (GGA) and the exchange-correlation functional of Perdew, Burke and Ernzerhof [16] are employed. The kinetic energy cutoff is set to 500 eV for the plane wave basis. A 10 × 10 × 10 k mesh for self-consistent Brillouin zone (BZ) sampling is adopted. The experimental lattice parameters of LaNiO2 and La3Ni2O6 are employed [4,11].
RESULTS AND DISCUSSIONS
Band structure and density of states
We first perform first-principle calculations on LaNiO2 without SOC. The band structure is presented in Fig. 2(a) and the total density of states is given in Fig. 2(c). The blue dashed horizontal line corresponds to the charge neutrality level of the undoped case. The red dashed line in Fig. 2(b) is the theoretically estimated chemical potential for the 20% hole-doped superconductivity Nd0.8Sr0.2NiO2. Moreover, the partial DOSs are also computed for O 2p, Ni 3d and La 5d orbitals, respectively. Since the main quantum numbers of different atom’s orbitals are distinct, we call them 2p, 3d and 5d orbitals (states) for short in the following discussion. From the plotted partial DOSs in Fig. 2(c), we note that 2p states are mainly located from −10 to −3.5 eV below EF, while 3d states are around EF, from −3.5 to 1.5 eV. The situation is much different from the situation in copper-based superconductors, where O 2p states are slightly below EF and hybridize strongly with Cu 3d states [10]. In addition, from the orbital projections in Fig. 2(a), we note that there are states at Γ and 5dxy states at A around EF, suggesting that the 5d states are more extended, compared to the Ca 3d states in CaCuO2 [17].
Evolution of Fermi surfaces
At the charge neutrality level, we find that there are three bands crossing EF, which are mainly from , and 5dxy orbitals, respectively. The weights of these orbitals are depicted by the size of the different symbols in Fig. 2(a). Therefore, as shown in Fig. 2(d), three EPs are formed: (i) the orbital forms the largest electron pocket around the Γ and Z points, which has a strong two-dimensional (2D) feature; (ii) the second larger EP is nearly a sphere around A, formed by the 5dxy orbital; (iii) the smallest EP is a sphere around Γ, formed by the orbital (which is also hybridized with the Ni orbital in the DFT calculations, yet we still call it for simplicity).
In the hole-doped superconductor Nd0.8Sr0.2NiO2, the estimated chemical potential of the 20% Sr-doped level is denoted by a red dashed line, which corresponds to 32.8 electrons per unit cell (the charge neutrality level corresponds to 33 electrons). Needless to say that all the electron pockets become smaller with hole doping. In particular, the -orbital-formed Γ-centered EP is about to be removed. On the other hand, the states from the 3dxy orbital become closer to the chemical potential, especially in the vicinity of point A.
Band inversion and Dirac points
The band crossings along R-A, A-Z and M-A are protected by m001, m110 and C4, respectively. After considering SOC, the band crossings open small gaps (blue circles) along the mirror protected R-A and A-Z lines, but remain gapless along the C4-invariant line M-A, as shown in Fig. 2(e). The gapless Dirac points along M-A [highlighted in the red circle in Fig. 2(e)] are protected by C4 symmetry. Namely, the two doubly degenerate bands belong to different 2D irreducible representations (IRs) of the C4 double group. In our GGA+U calculations, we find that the band inversion is sensitive to the value of the Coulomb interaction U, as we show in the online supplementary material.
Analysis of EBRs and orbitals
By doing an analysis of EBRs in the theory of topological quantum chemistry [18], we can also obtain orbital information in real space. An EBR of ρ@q is labeled by the Wyckoff position q and the IR ρ of its site symmetry group. The IRs of the six low-energy bands at maximal high symmetry k points [19,20] are computed without SOC. The results are listed in Table 1. In the crystal of LaNiO2, the Ni atom sits at Wyckoff position 1a[0, 0, 0], while the La atom sits at Wyckoff position 1d[0.5, 0.5, 0.5]. Both Wyckoff positions have the site-symmetry group of 4/mmm. Note that five d orbitals only support the IRs of , , B2g (dxy) and Eg (dxz, yz) under the single group of 4/mmm.
Table 1.
A | Γ | M | Z | R | X | |
---|---|---|---|---|---|---|
DFT bands | A3 −(1) | GM1+(1) | M1+(1) | Z4+(1) | R4+(1) | X1+(1) |
A1+(1) | GM4+(1) | M4+(1) | Z5+(2) | R1+(1) | X4+(1) | |
A5+(2) | GM5+(2) | M5+(2) | R2+(1) | X2+(1) | ||
Z1+(1) | R3+(1) | X3+(1) | ||||
A4 +(1) | GM2+(1) | Z2+(1) | R1+(1) | X1+(1) | ||
A2+(1) | GM1+(1) | M5 −(2) | Z1+(1) | R2 −(1) | X4 −(1) | |
EBRs | ||||||
A1g@1a | A1+ | GM1+ | M1+ | Z1+ | R1+ | X1+ |
B1g@1a | A2+ | GM2+ | M2+ | Z2+ | R1+ | X1+ |
B2g@1a | A4+ | GM4+ | M4+ | Z4+ | R2+ | X2+ |
Eg@1a | A5+ | GM5+ | M5+ | Z5+ | R3+ | X3+ |
R4+ | X4+ | |||||
A1g@1d | A2 − | GM1+ | M4+ | Z3 − | R3+ | X4 − |
B2g@1d | A3 − | GM4+ | M1+ | Z2 − | R4+ | X3 − |
A1g@1b | A3 − | GM1+ | M1+ | Z3 − | R2 − | X1+ |
The results of the EBR analysis on atomic orbitals are as follows. First, by switching the IRs of A3− and A4+ at A, we find that the four occupied bands can be represented as the sum of three EBRs: A1g@1a ⊕ B2g@1a ⊕ Eg@1a. Among them, the 3dxy-based EBR B2g@1a has highest states at A, which may intersect with the chemical potential in the hole-doping case. Second, the band of the -induced EBR B1g@1a is clearly shown by the weights in Fig. 2(a). There is no doubt that the orbital contributes the largest Fermi surface. Third, the IR GM1+ at Γ is from the -induced EBR A1g@1d. Last, the inverted IR A3− is from the 5dxy-induced EBR B2g@1d. So far, all of the orbital compounds are consistent with our DFT calculations in Fig. 2(a).
We now aim to construct a minimal effective model to reproduce the bands and symmetry characteristics near EF. Besides the -induced EBR of B1g@1a, the two IRs of A3− and GM1+ can be generated in the EBR of A1g@1b[0,0,0.5]. Therefore, we derive a two-band model, consisting of two EBRs: B1g@1a ⊕ A1g@1b, which reproduces the exact IRs of the bands near EF from the DFT calculations. Even through there is no actual atomic orbital at Wyckoff position 1b (the apical oxygens are removed), it could be formed by the hybridization of the atomic orbitals on other sites. The two-band model will be constructed for the undoped case in the following section.
Two-band effective model
Under the basis of the B1g orbital at the 1a Wyckoff position and the A1g orbital at the 1b Wyckoff position, the tight binding model is constructed as follows. The diagonal terms in the Hamiltonian are
(1) |
where α = 1, 2 represent the B1g orbital and the A1g orbital, respectively. Here stands for the hopping parameter from orbital β of the original cell to α of the (l, m, n) cell:
(2) |
In the off-diagonal term, the nearest and next-nearest hoppings between different orbitals are given as
(3) |
Thus, our two-band model Hamiltonian is written as
(4) |
where the dagger symbol means the Hermitian conjugate of the lower left part of the Hamiltonian matrix (same below). The fitting results are shown as blue dashed lines in Fig. 2(f) and the parameters can be found in Box 1. This two-band model can reproduce all the Fermi surfaces and symmetry characteristics obtained from the DFT calculations. It can be used for further study of superconductivity in the systems.
Box 1.
Parameters for the two-band model | |||||||
---|---|---|---|---|---|---|---|
0.1470 | 0.8318 | 0.0098 | |||||
−0.4125 | 0.0913 | ||||||
−0.0538 | 0.0650 | ||||||
0.0894 | −0.0606 | ||||||
0.0000 | 0.1988 | ||||||
0.0134 | 0.0281 | ||||||
Parameters for the three-band model | |||||||
0.0100 | 0.8280 | −0.7960 | 0.0098 | ||||
0.0042 | −0.0510 | −0.1655 | −0.0132 | ||||
0.0378 | −0.0079 | −0.0360 | 0.0001 | ||||
0.0043 | 0.0105 | −0.0497 | |||||
−0.1338 | 0.0000 | 0.0105 | |||||
0.0038 | 0.0018 | −0.0113 |
DFT+Gutzwiller method
To treat the correlation effect of five Ni 3d orbitals, we employ the DFT+Gutzwiller method [24,25]. The corresponding Gutzwiller trial wave function has been constructed as with
(5) |
where |0〉 is the noninteracting wave function and is the Gutzwiller local projector with |i, Γ〉 the atomic eigenvectors on site i. The are the so-called Gutzwiller variational parameters that adjust the weights of different local atomic configurations. Ground states are obtained by minimizing the total energy E = 〈G|(Htb + Hdc + Hint)|G〉 with some Gutzwiller constraints. The noninteracting Hamiltonian (Htb) is extracted using the Wannier90 package [26] from the DFT calculations without SOC, which contains two La orbitals and five Ni 3d orbitals. The double-counting term Hdc is given self-consistently. The on-site interacting term takes the Slater–Kanamori rotationally invariant atomic interaction [27]
(6) |
where creates (annihilates) an electron in the state of the orbital a and the spin σ, and .
For simplicity, we adopt diagonal variational parameters, which mean that here. Five Ni 3d orbitals are treated as correlated orbitals in the calculations (which correspond to 10 bands upon considering the spin degree of freedom). We take a Coulomb interaction U of 5 eV, Hund’s coupling J of 0.18U, where U′ = U − 2J and J′ = J. Besides, the occupancy of Ni 3d orbitals has been forced to be 8.462 obtained from the DFT calculations. The results show that the quasiparticle weight of the orbital is very small, 0.12, while the weights of the other four orbitals are about 0.85. After considering the renormalization of the correlated 3d orbitals, the modified band structure is obtained and shown in Fig. 3(a). Two significant features are found after the Gutzwiller correction: (i) the bandwidth of has been largely renormalized, leading to a DOS peak around EF, which may contribute to the large peak around zero energy in the resonant inelastic X-ray scattering spectrum (RIXS) [28]; (ii) the 3dxy states near the A point become very close to EF. As a result, a HP could be induced by hole doping, which may be related to the observed sign change of the Hall coefficient.
In addition, the band inversion between the 3dxy states and 5dxy states near the A point could be important in the hole-doped nickelate compound Nd0.8Sr0.2NiO2, since they are very close to EF. Therefore, we modify our model by simply adding an additional 3dxy orbital to capture the potential band inversion in this hole-doped compound. The modified model is written as
(7) |
with
The parameters are obtained by fitting with the renormalized bands, as shown in Box 1. The results of the modified model H3(k) are shown as blue dashed lines in Fig. 3(c), which fit very well with the DFT+Gutzwiller bands, which can be compared with the angle-resolved photoemission spectroscopy (ARPES) experimental data. In any case, a four-band model constructed totally from the real atomic orbitals is presented in the online supplementary material.
Ferromagnetic state in La3Ni2O6
To better understand the electronic structures of T′-type Lan+1NinO2n+2 compounds (n ≥ 2), we calculate the other end member La3Ni2O6 (i.e. n = 2 corresponds to the most reduced case) [29,30]. The ferromagnetic band structure of La3Ni2O6 with SOC is shown in Fig. 4(a). The -oriented magnetism [31] reduces the symmetry from type-II magnetic SG 139 to type-I magnetic SG 87. Near the M/A point, the four downward parabolic bands (i.e. 89–92 bands) are mainly from the states of Ni atoms in two planar Ni–O layers, while the two upward parabolic bands (i.e. 93–94 bands) are mainly from the 5dxy states of the La atoms sandwiched by two Ni–O planes (Fig. 1(b)). Looking closely at the crossings between the 90th and 91st bands in Fig. 4(b), we find that there is a gap along X-M (R-A), while there are two crossing points along M-Γ (A-Z). These crossing points are parts of the tiny M-protected nodal rings at the k = 0, 2π/c planes, which can be easily removed by very small perturbations (without changing the ordering of energy bands at high symmetry k points). According to symmetry indicators defined in magnetic systems [32], insulators in type-I MSG 87 are characterized by the indicators [33,34], which can be explained by two mirror Chern numbers in the k = 0 plane. By using the C4 eigenvalues [35], we find that the Chern numbers are 0 and 2 for the mirror eigenvalue +i and −i sectors, respectively, which indicate that nontrivial edge states can emerge on the M-preserving surfaces.
CONCLUSION
Based on our first-principle calculations, we find that in the undoped case there are three EPs: the largest EP coming from the Ni orbital, a small EP at Γ from the La orbital and a relatively larger EP at A from La 5dxy. These Fermi surfaces and symmetry characteristics can be reproduced by our two-band model, which consists of two elementary band representations: B1g@1a ⊕ A1g@1b. This two-band model could be used for further study of superconductivity in these systems. In the obtained band structure, a band inversion occurred at A, which gives rise to a pair of Dirac points along M-A upon including SOC. The correlation effect of Ni 3d orbitals has been estimated in our DFT+Gutzwiller calculation, and the renormalized band structure obtained, which shows that Ni 3dxy states become very close to EF around A. A hole pocket is likely induced by hole doping, which may be related to the observed sign change of the Hall coefficient. The potential hole pocket and the band inversion can be captured in the modified model by simply including another Ni 3dxy orbital. In addition, we show that the nontrivial band topology in the ferromagnetic two-layer compound La3Ni2O6 is associated with Ni and La 5dxy orbitals.
Note added. While finalizing the present paper, similar works on nickelates were published [36–40].
Supplementary Material
Acknowledgements
We thank Jinguang Cheng for drawing our attention to this subject and valuable discussions, as well as critical reading of this work.
Contributor Information
Jiacheng Gao, Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China; School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China.
Shiyu Peng, Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China; School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China.
Zhijun Wang, Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China; School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China.
Chen Fang, Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China; Songshan Lake Materials Laboratory, Dongguan 523808, China; Kavli Institute for Theoretical Sciences, Chinese Academy of Sciences, Beijing 100190, China.
Hongming Weng, Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China; School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China; Songshan Lake Materials Laboratory, Dongguan 523808, China; Physical Science Laboratory, Huairou National Comprehensive Science Center, Beijing 101407, China.
FUNDING
This work was supported by the National Natural Science Foundation of China (11974395, 11504117, 11774399, 11622435, U1832202, 11921004, 11925408 and 11674370), the Beijing Natural Science Foundation (Z180008), the Ministry of Science and Technology of China (2016YFA0300600, 2016YFA0401000, 2018YFA0305700 and 2016YFA0302400), the Strategic Priority Research Program of Chinese Academy of Sciences (XXH13506-202 and XDB33000000), the Beijing Municipal Science and Technology Commission (Z181100004218001, Z171100002017018) and the Center for Materials Genome. H.W. acknowledges support from the K. C. Wong Education Foundation (GJTD-2018-01). Z.W. acknowledges support from the CAS Pioneer Hundred Talents Program and the National Thousand-Young-Talents Program.
AUTHOR CONTRIBUTIONS
H.W., C.F. and Z.W. proposed and supervised the project. J.G. carried out the DFT calculations and built the effective models. J.G. and Z.W. did the symmetry analysis. S.P. performed the DFT+Gutzwiller calculations. All authors contributed to analyzing the results and writing the manuscript.
Conflict of interest statement. None declared.
REFERENCES
- 1.Bednorz JG, Müller KA.. Possible high Tc superconductivity in the Ba-La-Cu-O system. Eur Phys J B 1986; 64: 189–93. 10.1007/BF01303701 [DOI] [Google Scholar]
- 2.Wu MK, Ashburn JR, Torng CJet al. . Superconductivity at 93 K in a new mixed-phase Y-Ba-Cu-O compound system at ambient pressure. Phys Rev Lett 1987; 58: 908–10. 10.1103/PhysRevLett.58.908 [DOI] [PubMed] [Google Scholar]
- 3.Anisimov VI, Bukhvalov D, Rice TM.. Electronic structure of possible nickelate analogs to the cuprates. Phys Rev B 1999; 59: 7901–6. 10.1103/PhysRevB.59.7901 [DOI] [Google Scholar]
- 4.Li D, Lee K, Wang BYet al. . Superconductivity in an infinite-layer nickelate. Nature 2019; 572: 624–7. 10.1038/s41586-019-1496-5 [DOI] [PubMed] [Google Scholar]
- 5.Hayward M, Green M, Rosseinsky Met al. . Sodium hydride as a powerful reducing agent for topotactic oxide deintercalation: synthesis and characterization of the nickel (I) oxide LaNiO2. J Am Chem Soc 1999; 121: 8843–54. 10.1021/ja991573i [DOI] [Google Scholar]
- 6.Hayward M, Rosseinsky M.. Synthesis of the infinite layer Ni (I) phase NdNiO2+x by low temperature reduction of NdNiO3 with sodium hydride. Solid State Sci 2003; 5: 839–50. 10.1021/ja991573i [DOI] [Google Scholar]
- 7.Motoyama E, Yu G, Vishik Iet al. . Spin correlations in the electron-doped high-transition-temperature superconductor Nd2−xCexCuO4 ± δ. Nature 2007; 445: 186–9. 10.1038/nature05437 [DOI] [PubMed] [Google Scholar]
- 8.Lee PA, Nagaosa N, Wen XG.. Doping a Mott insulator: physics of high-temperature superconductivity. Rev Mod Phys 2006; 78: 17–85. 10.1103/RevModPhys.78.17 [DOI] [Google Scholar]
- 9.Smith M, Manthiram A, Zhou Jet al. . Electron-doped superconductivity at 40 K in the infinite-layer compound Sr1−yNdyCuO2. Nature 1991; 351: 549–51. 10.1038/351549a0 [DOI] [Google Scholar]
- 10.Lee KW, Pickett W.. Infinite-layer LaNiO2: Ni1+ is not Cu2+. Phys Rev B 2004; 70: 165109. 10.1103/PhysRevB.70.165109 [DOI] [Google Scholar]
- 11.Poltavets VV, Greenblatt M, Fecher GHet al. . Electronic properties, band structure, and Fermi surface instabilities of Ni1+/Ni2+ nickelate La3Ni2O6, isoelectronic with superconducting cuprates. Phys Rev Lett 2009; 102: 046405. 10.1103/PhysRevLett.102.046405 [DOI] [PubMed] [Google Scholar]
- 12.Kresse G, Furthmüller J.. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput Mater Sci 1996; 6: 15–50. 10.1103/PhysRevLett.102.046405 [DOI] [PubMed] [Google Scholar]
- 13.Kresse G, Furthmüller J.. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys Rev B 1996; 54: 11169–86. 10.1103/PhysRevB.54.11169 [DOI] [PubMed] [Google Scholar]
- 14.Blöchl PE.Projector augmented-wave method. Phys Rev B 1994; 50: 17953–79. 10.1103/PhysRevB.50.17953 [DOI] [PubMed] [Google Scholar]
- 15.Kresse G, Joubert D.. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys Rev B 1999; 59: 1758–75. 10.1103/PhysRevB.59.1758 [DOI] [Google Scholar]
- 16.Perdew JP, Burke K, Ernzerhof M.. Generalized gradient approximation made simple. Phys Rev Lett 1996; 77: 3865–8. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- 17.Wu H, Zheng Q-Q, Gong X-Get al. . The electronic structure of CaCuO2 and SrCuO2. J Phys Condens Matter 1999; 11: 4637–46. [Google Scholar]
- 18.Bradlyn B, Elcoro L, Cano Jet al. . Topological quantum chemistry. Nature 2017; 547: 298–305. 10.1038/nature23268 [DOI] [PubMed] [Google Scholar]
- 19.Gao J, Wu Q, Persson Cet al. . Irvsp: to obtain irreducible representations of electronic states in the VASP. arXiv:2002.04032. [Google Scholar]
- 20.Vergniory M, Elcoro L, Felser Cet al. . A complete catalogue of high-quality topological materials. Nature 2019; 566: 480–5. 10.1038/s41586-019-0954-4 [DOI] [PubMed] [Google Scholar]
- 21.Aroyo MI, Perez-Mato JM, Orobengoa Det al. . Crystallography online: Bilbao crystallographic server. Bulg Chem Commun 2011; 43: 183–97. [Google Scholar]
- 22.Aroyo MI, Perez-Mato JM, Capillas Cet al. . Bilbao crystallographic server: I. Databases and crystallographic computing programs. Zeitschrift für Kristallographie 2006; 221: 15–27. [Google Scholar]
- 23.Aroyo MI, Kirov A, Capillas Cet al. . Bilbao crystallographic server. II. Representations of crystallographic point groups and space groups. Acta Crystallogr A 2006; 62: 115–28. [DOI] [PubMed] [Google Scholar]
- 24.Deng X, Wang L, Dai Xet al. . Local density approximation combined with Gutzwiller method for correlated electron systems: formalism and applications. Phys Rev B 2009; 79: 075114. 10.1103/PhysRevB.79.075114 [DOI] [Google Scholar]
- 25.Du L, Huang L, Dai X.. Metal-insulator transition in three-band Hubbard model with strong spin-orbit interaction. Eur Phys J B 2013; 86: 94. 10.1140/epjb/e2013-31024-6 [DOI] [Google Scholar]
- 26.Marzari N, Mostofi AA, Yates JRet al. . Maximally localized Wannier functions: theory and applications. Rev Mod Phys 2012; 84: 1419–75. 10.1103/RevModPhys.84.1419 [DOI] [Google Scholar]
- 27.Lanatà N, Strand HUR, Dai Xet al. . Efficient implementation of the Gutzwiller variational method. Phys Rev B 2012; 85: 035133. 10.1103/PhysRevB.85.035133 [DOI] [Google Scholar]
- 28.Hepting M, Li D, Jia Cet al. . Electronic structure of the parent compound of superconducting infinite-layer nickelates. Nat Mater 2020; 19: 381–5. 10.1038/s41563-019-0585-z [DOI] [PubMed] [Google Scholar]
- 29.Manthiram A, Tang J, Manivannan V.. Factors influencing the stabilization of Ni+ in perovskite-related oxides. J Solid State Chem 1999; 148: 499–507. 10.1006/jssc.1999.8487 [DOI] [Google Scholar]
- 30.Poltavets VV, Lokshin KA, Dikmen Set al. . La3Ni2O6: a new double T’-type nickelate with infinite Ni1+/2+O2 layers. J Am Chem Soc 2006; 128: 9050–1. 10.1021/ja063031o [DOI] [PubMed] [Google Scholar]
- 31.Sarkar S, Dasgupta I, Greenblatt Met al. . Electronic and magnetic structures of bilayer La3Ni2O6 and trilayer La4Ni3O8 nickelates from first principles. Phys Rev B 2011; 84: 180411. 10.1103/PhysRevB.84.180411 [DOI] [Google Scholar]
- 32.Turner AM, Zhang Y, Mong RSKet al. . Quantized response and topology of magnetic insulators with inversion symmetry. Phys Rev B 2012; 85: 165120. 10.1103/PhysRevB.85.165120 [DOI] [Google Scholar]
- 33.Po HC, Vishwanath A, Watanabe H.. Symmetry-based indicators of band topology in the 230 space groups. Nat Commun 2017; 8: 50. 10.1038/s41467-017-00133-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Ono S, Watanabe H.. Unified understanding of symmetry indicators for all internal symmetry classes. Phys Rev B 2018; 98: 115150. 10.1103/PhysRevB.98.115150 [DOI] [Google Scholar]
- 35.Fang C, Gilbert MJ, Bernevig BA.. Bulk topological invariants in noninteracting point group symmetric insulators. Phys Rev B 2012; 86: 115112. 10.1103/PhysRevB.86.115112 [DOI] [Google Scholar]
- 36.Hirsch J, Marsiglio F.. Hole superconductivity in infinite-layer nickelates. Physica C Supercond 2019; 566: 1353534. 10.1016/j.physc.2019.1353534 [DOI] [Google Scholar]
- 37.Sakakibara H, Usui H, Suzuki Ket al. . Model construction and a possibility of cupratelike pairing in a new d 9 nickelate superconductor (Nd, Sr) NiO2. Phys Rev Lett 2020; 125: 077003. [DOI] [PubMed] [Google Scholar]
- 38.Botana AS, Norman MR.. Similarities and differences between LaNiO2 and CaCuO2 and implications for superconductivity. Phys Rev X 2020; 10: 011024. 10.1103/PhysRevX.10.011024 [DOI] [Google Scholar]
- 39.Wu X, Di Sante D, Schwemmer Tet al. . Robust dx2−y2 wave superconductivity of infinite-layer nickelates. Phys Rev B 2020; 101: 060504. 10.1103/PhysRevB.101.060504 [DOI] [Google Scholar]
- 40.Nomura Y, Hirayama M, Tadano Tet al. . Formation of a two-dimensional single-component correlated electron system and band engineering in the nickelate superconductor NdNiO2. Phys Rev B 2019; 100: 205138. 10.1103/PhysRevB.100.205138 [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.