Skip to main content

REVIEW article

Front. Phys., 21 January 2022
Sec. Condensed Matter Physics
This article is part of the Research Topic Advances in Superconducting Infinite-Layer and Related Nickelates View all 16 articles

Phase Diagram of Nickelate Superconductors Calculated by Dynamical Vertex Approximation

  • 1Institute for Solid State Physics, Vienna University of Technology, Vienna, Austria
  • 2School of Physics, Northwest University, Xi’an, China
  • 3Institute for Theoretical Solid State Physics, Leibniz IFW Dresden, Dresden, Germany
  • 4RIKEN Center for Emergent Matter Sciences (CEMS), Wako, Japan
  • 5Department of Applied Physics, the University of Tokyo, Bunkyo, Japan
  • 6Key Laboratory of Magnetic Materials and Devices and Zhejiang Province Key Laboratory of Magnetic Materials and Application Technology, Ningbo Institute of Materials Technology and Engineering (NIMTE), Chinese Academy of Sciences, Ningbo, China

We review the electronic structure of nickelate superconductors with and without effects of electronic correlations. As a minimal model, we identify the one-band Hubbard model for the Ni 3dx2y2 orbital plus a pocket around the A-momentum. The latter, however, merely acts as a decoupled electron reservoir. This reservoir makes a careful translation from nominal Sr-doping to the doping of the one-band Hubbard model mandatory. Our dynamical mean-field theory calculations, in part already supported by the experiment, indicate that the Γ pocket, Nd 4f orbitals, oxygen 2p, and the other Ni 3d orbitals are not relevant in the superconducting doping regime. The physics is completely different if topotactic hydrogen is present or the oxygen reduction is incomplete. Then, a two-band physics hosted by the Ni 3dx2y2 and 3d3z2r2orbitals emerges. Based on our minimal modeling, we calculated the superconducting Tc vs. Sr-doping x phase diagram prior to the experiment using the dynamical vertex approximation. For such a notoriously difficult to determine quantity as Tc, the agreement with the experiment is astonishingly good. The prediction that Tc is enhanced with pressure or compressive strain has been confirmed experimentally as well. This supports that the one-band Hubbard model plus an electron reservoir is the appropriate minimal model.

1 Introduction

Twenty years ago, Anisimov, Bukhvalov, and Rice [1] suggested high-temperature (Tc) superconductivity in nickelates based on material calculations that showed apparent similarities to cuprates. Subsequent calculations [24] demonstrated the potential to further engineer the nickelate Fermi surface by heterostructuring. Two years ago, Li, Hwang et al. [5] discovered superconductivity in Sr-doped NdNiO2 films grown on a SrTiO3 substrate and protected by a SrTiO3 capping layer. These novel SrxNd1–xNiO2 films are isostructural and formally isoelectric to the arguably simplest, but certainly not the best superconducting cuprate: infinite layer CaCuO2 [69].

However, the devil is in the details, and here cuprates and nickelates differ. For revealing such material-specific differences, band-structure calculations based on density functional theory (DFT) are the method of choice. They serve as a starting point for understanding the electronic structure and, subsequently, the phase diagram of nickelate superconductors. Following the experimental discovery of nickelate superconductivity, and even before that, numerous DFT studies have been published [1020]. Based on these DFT calculations, various models for the low-energy electronic structure for nickelates and the observed superconductivity have been proposed. Besides the cuprate-like Ni 3dx2y2 band, DFT shows an A and a Γ pocket which originate from Nd 5dxy and 5d3z2r2 bands, but with major Ni 3d admixture in the region of the pocket. The importance of the Ni 3d3z2r2 orbital has been suggested in some studies [2124], and that of the Nd-4f orbitals in others [20, 25]. Furthermore, there is the question regarding the relevance of the oxygen 2p orbitals. For cuprates, these are, besides the Cu 3dx2y2 orbitals, the most relevant. Indeed, cuprates are generally believed to be charge-transfer insulators [26]. This leads to the three-orbital Emery model [27] visualized in Figures 1A,C as the minimal model for cuprates. The much more frequently investigated Hubbard model [2830] may, in the case of cuprates, only be considered as an effective Hamiltonian mimicking the physics of the Zhang–Rice singlet [31].

FIGURE 1
www.frontiersin.org

FIGURE 1. Crystal lattice and most important orbitals for (A) cuprates and (B) nickelates. (C) For cuprates, the arguably simplest model is the Emery model with a half-filled copper 3dx2y2 band and holes in the oxygen 2p orbitals that can hop to other oxygen (tpp) and copper sites (tdp) where double occupations are suppressed by the interaction U. (D) For nickelates, we have a Ni-3dx2y2-band Hubbard model which, however, only accommodates part of the holes induced by Sr-doping. The others go to the A pocket stemming from the Nd 5dxy band and acting as a decoupled reservoir.

At first glance, nickelates appear to be much more complicated with more relevant orbitals than in the case of the cuprates. In this article, we review the electronic structure of nickelates in comparison to that of cuprates and the arguments for a simpler description of nickelate superconductors, namely, a Hubbard model for the Ni 3dx2y2 band plus a largely decoupled reservoir corresponding to the A pocket. This A pocket is part of the Nd 5dxy band which has, however, a major admixture of Ni 3dxz/yz and O 2pz states around the momentum A. This leaves us with Figures 1B,D as the arguably simplest model for nickelates [32, 33]. This (our) perspective is still controversially discussed in the literature. However, as we will point out below, a number of experimental observations already support this perspective against some of the early suggestions that other orbitals are relevant. Certainly, other perspectives will be taken in other articles of this series on “Advances in Superconducting Infinite-Layer and Related Nickelates.” The simple picture of a one-band Hubbard model, whose doping needs to be carefully calculated since part of the holes in Sr-doped SrxNd1−xNdO2 go to the A pocket, allowed us [33] to calculate Tc, see Figure 5, at a time when only the Tc for a single doping x = 20% was experimentally available. To this end, state-of-the-art dynamical vertex approximation (DΓA) [3638], a Feynman diagrammatic extension of dynamical mean-field theory (DMFT) [3942] has been used. For such a notoriously difficult to calculate physical quantity as Tc, the agreement of the single-orbital Hubbard model calculation with subsequent experiments [34, 35] is astonishingly good. This further supports the modeling by a single-orbital Hubbard model which, thus, should serve at the very least as a good approximation or a starting point.

The outline of this article is as follows: In Section 2,we first compare the electronic structure of nickelates to that of cuprates, starting from DFT but also discussing effects of electronic correlations as described, for example, by DMFT. Subsequently, we argue in Section 3, orbital-by-orbital, that the other orbitals besides the Ni 3dx2y2 and the A pocket are, from our perspective, not relevant. This leaves us with the one-3dx2y2-band Hubbard model plus an electron reservoir representing the A pocket of Figures 1B,D, which is discussed in Section 4, including the translation of Sr-doping to the filling in the Hubbard model and the reservoir. In Section 5, we discuss the effect of non-local correlations as described in DΓA and the calculated superconducting phase diagram. Section 6 shows that topotactic hydrogen, which is difficult to detect in the experiment, completely overhauls the electronic structure and the prevalence of superconductivity. Finally, Section 7 summarizes the article.

2 Electronic Structure: Nickelates vs. Cuprates

Let us start by looking into the electronic structure in more detail and start with the DFT results. On a technical note, the calculations presented have been performed using the wien2k [43, 44], VASP [45], and FPLO [46] program packages, with the PBE [47] version of the generalized gradient approximation (GGA). For further details, see the original work [33]. Figure 2 compares the bandstructure of the two simple materials: CaCuO2 and NdNiO2. Here, we restrict ourselves to only the Brillouin zone path along the most relevant momenta for these compounds: Γ (0, 0, 0), X (π, 0, 0), and A (π, π, π). In DFT, both the cuprate and nickelate parent compounds are metals with a prominent Cu or Ni 3dx2y2 band crossing the Fermi energy. In other aspects, both materials differ (for a review cf [48]): In the case of cuprates, the oxygen bands are much closer to the Fermi energy. Hence, if electronic correlations split the DFT bands into two Hubbard bands as indicated in Figure 2 by the arrows and the spectral function in the left side panel, we get a charge-transfer insulator [26]. For this charge-transfer insulator, the oxygen 2p orbitals are the first orbitals below the Fermi level (EF) and receive the holes that are induced by doping. The Cu 3dx2y2 lower Hubbard band is below these oxygen orbitals, and the Cu 3dx2y2 upper Hubbard band is above the Fermi level. Let us note that we here refer to oxygen 2p orbitals and Cu 3dx2y2 orbitals even though the hybridization between both is very strong. Indeed, the two sets of orbitals strongly mix in the resulting effective DFT bands of Figure 2.

FIGURE 2
www.frontiersin.org

FIGURE 2. Electronic structures of CaCuO2 (A) and NdNiO2 (B) exemplifying superconducting cuprates and nickelates. Top: The bars indicate the center of energy for the most important DFT bands. The dashed arrows indicate the correlation-induced splitting of the Cu or Ni 3dx2y2 band into Hubbard bands, leading to the (schematic) spectral function at the side. Cuprates are charge-transfer insulators with a gap between the O 2p and the upper Cu 3dx2y2 Hubbard band as the lower Cu 3dx2y2 Hubbard band is below the O 2p band. For nickelates, bands with a large La (Nd) 5d contribution cross the Fermi energy at Γ and A, self-doping the Ni 3dx2y2 band away from half-filling. (C,D): Corresponding DFT Fermi surface. For cuprates, DFT shows a single Cu 3dx2y2 hole–like Fermi surface; for nickelates, there is a similar, slightly more warped Ni 3dx2y2 Fermi surface and additional small pockets around the Γ and A momentum.

Because cuprates are charge-transfer insulators, the one-band Hubbard model can only be considered an effective Hamiltonian mimicking the Zhang–Rice singlet [31]. As already pointed out in the Introduction, more appropriate is the Emery model of Figure 1. The correlation-induced splitting into the Hubbard bands [40], as well as the Zhang–Rice singlet [49], can be described already by DMFT [39, 40, 42]. Two-dimensional spin-fluctuations and superconductivity, however, cannot be described. For describing such physics, non-local correlations beyond DMFT are needed.

For the nickelates, the oxygen bands are at a much lower energy. Hence, as indicated in the right side panel of Figure 2, the lower Ni 3dx2y2 Hubbard band can be expected to be closer to the Fermi energy than the oxygen p orbitals [32, 33]. Consequently, undoped nickelates would be Mott–Hubbard insulators if it was not for two additional bands that cross EF around the Γ- and A-momentum. These form electron pockets, as visualized in Figure 2 (bottom right), and self-dope the Ni 3dx2y2 band away from half-filling. As the 3dx2y2 is doped, it develops, even when the Coulomb interaction is large, a quasiparticle peak at the Fermi energy as displayed in the right side panel of Figure 2.

3 Irrelevance of Various Orbitals

Next, we turn to various orbitals that may appear relevant at first glance but turn out to be irrelevant for the low-energy physics when taking electronic correlations properly into account. To account for the latter, we use DFT + DMFT [5054], which is the state of the art for calculating correlated materials.

Oxygen Orbitals

For nickelates, the oxygen 2p orbitals are approximately 3 eV lower in energy than in cuprates within DFT. Hence, some DFT + DMFT calculations did not include these from the beginning [32, 33], and those that did [22] also found the oxygen 2p orbitals at a lower energy than the lower Ni 3dx2y2 Hubbard band. Hence, while there is still some hybridization and mixing between the O 2p states and the Ni 3dx2y2 states, a projection onto a low-energy set of orbitals without oxygen appears possible.

Ni 3d3z2r2 and t2g Orbitals

Instead of the oxygen 2p orbitals, the DFT calculation in Figure 2 and elsewhere [10, 11, 17, 48] show other Ni 3d orbitals closely below the Ni 3dx2y2 band. In fact, these other 3d orbitals are somewhat closer to the Fermi level than in the case of cuprates. Electronic correlations can strongly modify the DFT band structure. In particular, the Hund’s exchange J tends to drive the system toward a more equal occupation of different orbitals, especially if there is more than one hole (more than one unpaired electron) in the Ni 3d orbitals. This is not only because a larger local spin is made possible but also because the inter-orbital Coulomb interaction U′ between two electrons in two different orbitals is smaller than the intra-orbital Coulomb interaction U = U′ + 2J for two electrons in the same orbital. This tendency is countered by the crystal field splitting (local DFT potentials) which puts the 3dx2y2 orbital above the Ni 3d3z2r2 orbital and the other (t2g) Ni 3d orbitals because of the absence of apical O atoms in NiO4 squares.

Figure 3 shows the DFT + DMFT spectral function for SrxLa1−xNiO2 from 0 to 25% Sr-doping. In these calculations [33], all Ni 3d and all La 5d orbitals have been taken into account in a wien2wannier [56] projection supplemented by interactions calculated within the constrained random phase approximation (cRPA) [55] to be U′ = 3.10 eV (2.00 eV) and Hund’s exchange J = 0.65 eV (0.25 eV) for Ni (La). On a technical note, the DMFT self-consistency equations [39] have been solved here at room temperature (300 K) by continuous-time quantum Monte Carlo simulations in the hybridization expansions [57] using the w2dynamics implementation [58, 59] and the maximum entropy code of ana_cont [60] for analytic continuation.

FIGURE 3
www.frontiersin.org

FIGURE 3. DMFT k-integrated (A–C) and k-resolved (D,E) spectral functions A(ω) and A (k, ω) of undoped LaNiO2 (A), 20% Sr-doped LaNiO2 (Sr0.2La0.8NiO2) (B), and 25% Sr-doped LaNiO2 (Sr0.25La0.75NiO2) (C). The k-resolved spectral function A (k, ω) of La0.8Sr0.2NiO2 is shown in (D); (E) is a zoom-in of (D). Data partially obtained from [33, 55].

Clearly, Figure 3 indicates that for up to 20% Sr-doping, the other Ni 3d orbitals besides the 3dx2y2 orbital are not relevant for the low-energy physics of SrxLa1−xNiO2; they are fully occupied below the Fermi energy. With doping, these other Ni 3d orbitals, however, shift more and more upward in energy. At around 25% Sr-doping, they touch the Fermi energy and hence a multi-orbital Ni description becomes necessary at larger dopings. That is, between 20 and 30% Sr-doping, the physics of SrxLa1−xNiO2 turns from single to multi-orbital. In the case of SrxNd1−xNiO2, this turning point is at slightly larger doping [33]. Later, in Section 6, we will see that for the Ni 3d8 configuration, which in Section 6 is induced by topotactic hydrogen and here would be obtained for 100% Sr doping; the two holes in the Ni 3d orbitals form a spin-1 and occupy two orbitals: 3dx2y2 and 3d3z2r2. In Figure 3,we see at 30% doping, the first steps into this direction. Importantly, within the superconducting doping regime which, noteworthy, is below 24% Sr-doping for SrxNd1−xNiO2 [34, 35] and 21% for SrxLa1−xNiO2 [61], a single 3dx2y2 Ni-orbital is sufficient for the low-energy modeling. A one-band Hubbard model description based on DMFT calculations was also concluded in [32] for the undoped parent compound.

DFT + DMFT calculations by Lechermann [22, 23] stress, on the other hand, the relevance of the 3d3z2r2 orbital. Let us note that also in [22] the number of holes in the 3d3z2r2 orbital is considerably less than in that in the 3dx2y2. However, for low Sr-doping, also a small quasiparticle peak develops for the 3d3z2r2 band [22, 23]. An important difference to [33, 55] is that the 5d Coulomb interaction has been taken into account in [33, 55] and that the Coulomb interaction of [22] is substantially larger. The 5d Coulomb interaction pushes the Γ pocket above the Fermi energy (see next paragraph). As much of the holes in the 3d3z2r2 orbital stem from the admixture of this orbital to the Γ pocket, this difference is very crucial for the occupation of the 3d3z2r2 orbital in some calculations [21, 23]. On the other hand, in GW + extended DMFT calculations by Petocchi et al. [24], the 3d3z2r2 orbital is pushed to the Fermi energy for large kz instead, that is, around the R, Z, and A point. Except for this large kz deviation, the Fermi surface, the effective mass of the Ni 3dx2y2 orbital etc. of [24] are similar to our calculation [33, 62].

First experimental hints on the (ir) relevance of the 3d3z2r2 can be obtained from resonant inelastic X-ray scattering (RIXS) experiments [63, 64]. Higashi et al. [65] analyzed these RIXS data by comparison with DFT + DMFT and obtained good agreement with the experiment. They conclude that NdNiO2 is slightly doped away from 3d9 because of a small self-doping from the Nd 5d band and that only the 3dx2y2 Ni orbital (not the 3d3z2r2 orbital) is partially filled and that the Ni–O hybridization plays a less important role than for the cuprates.

Γ Pocket

A feature clearly present in DFT calculations for the nickelate parent compounds LaNiO2 and NdNiO2 is the Γ pocket, see Figure 2D. However, when the Coulomb interaction on the La or Nd sites is included, it is shifted upward in energy. Furthermore, Sr-doping depopulates the Ni 3dx2y2 orbital and the A and Γ pocket, and, thus, also helps pushing the Γ pocket above the Fermi energy. Clearly in the DFT + DMFT k-resolved spectrum of Figure 3, the Γ pocket is above the Fermi energy. To some extent, the presence or absence of the Γ pocket also depends on the rare-earth cation. For NdNiO2, we obtain a Γ pocket for the undoped compound [33], which only shifts above the Fermi energy with Sr-doping in the superconducting region, whereas for LaNiO2, it is already above the Fermi level without Sr-doping. We can hence conclude that while there might be a Γ pocket without Sr-doping, DFT + DMFT results suggest that it is absent in the superconducting doping regime.

Briefly after the discovery of superconductivity in nickelates, it has also been suggested that the Nd 5d orbitals of the pockets couple to the Ni 3dx2y2 spin giving rise to a Kondo effect [20, 66]. However, Table 1 shows that the hybridization between the relevant Ni 3dx2y2 and the most important La or Nd 5dxy and 5d3z2r2 vanishes by symmetry. Also, the full 5 Ni and 5 Nd band DMFT calculations in Figure 3 do not show a hybridization (gap) between A pocket and Ni bands. This suggests that the Γ and A pocket are decoupled from the 3dx2y2 orbitals. There is no hybridization and hence no Kondo effect.

TABLE 1
www.frontiersin.org

TABLE 1. Hybridization (hopping amplitude in eV) between the partially occupied Ni 3dx2y2 and the La/Nd 5d orbitals [33]. Here, the results are obtained from Wannier projections onto 17-bands (La/Nd-4f + La/Nd-5d + Ni-3d) and 10-bands (La/Nd-5d + Ni-3d).

Nd 4f Orbitals

Finally, the importance of the Nd 4f orbitals has been suggested in the literature. Treating these 4f orbitals in DFT is not trivial because DFT puts them in the vicinity of the Fermi level. This neglects that electronic correlations split the Nd 4f into upper and lower Hubbard bands as they form a local spin. This effect is beyond DFT. One way to circumvent this difficulty is to put the Nd 4f orbitals in the core instead of having them as valence states close to the Fermi energy. This is denoted as “GGA open core” instead of standard “GGA” in Table 1. The localized Nd 4f spins might, in principle, be screened through a Kondo effect. However, the hybridization of the Nd 4f with the Ni 3dx2y2 orbital at the Fermi energy is extremely small, see Table 2 and [67]. Hence, the Kondo temperature is zero for all practical purposes. In spin-polarized DFT + U, there is, instead, a local exchange interaction between the Nd 4f and the predominately Nd 5d Γ pocket [68]. However, as pointed out in the previous paragraph, the Γ pocket is shifted above the Fermi level in the superconducting Sr-doping regime. Hence in [33], we ruled out that the Nd 4f is relevant for superconductivity. This has been spectacularly confirmed experimentally by the discovery of superconductivity in nickelates without f electrons: CaxLa1−xNiO2 [69] and SrxLa1−xNiO2 [61] have a similar Tc.

TABLE 2
www.frontiersin.org

TABLE 2. Hybridization (hopping amplitude in eV) between the Ni 3dx2y2 and the Nd (La) 4f orbitals, as obtained from Wannier projections onto 17-bands (La/Nd-4f + La/Nd-5d + Ni-3d), including the 4f as valence states in DFT (GGA) [33].

4 One-Band Hubbard model Plus Reservoir

Altogether, this leaves us with Figures 1B,D as the arguably simplest model for nickelate superconductors, consisting of a strongly correlated Ni 3dx2y2 band and an A pocket. This A pocket is derived from the Nd 5dxy band which, however, crosses the Ni 3d orbitals and hybridizes strongly with the Ni t2g orbitals so that at the bottom of the A pocket, that is, at the momentum A, it is made up primarily from Ni t2g, whereas the Nd 5dxy contribution is here at a lower energy. This makes the A pocket much more resistive to shifting up in energy than the Γ pocket.

On the other hand, the A pocket does not interact with the Ni 3dx2y2 band; that is, does not hybridize in Table 1. Hence, we can consider the A pocket as a mere hole reservoir which accommodates part of the holes induced by Sr-doping, whereas the other part goes into the correlated Ni 3dx2y2 band which is responsible for superconductivity. Figure 4 shows the thus obtained Ni 3dx2y2 occupation as a function of Sr-doping in the DFT + DMFT calculation with 5 Ni and 5 Nd(La) orbitals.

FIGURE 4
www.frontiersin.org

FIGURE 4. Occupation of the Ni 3dx2y2 orbital (blue; right y-axis) and its effective mass enhancement m*/m = 1/Z (black; left y-axis in panels) vs. Sr-doping for LaNiO2 (A) and NdNiO2 (B) as calculated in DFT + DMFT. From [33] (Supplemental material).

It is noted that NdNiO2 shows for Sr-doping below about 10% more holes in the Ni 3dx2y2 orbital and a weaker dependence on the Sr-doping since here the Γ pocket is still active, not only taking away electrons from Ni but also first absorbing some of the holes from the Sr-doping until it is completely depopulated (shifted above the Fermi energy) before superconductivity sets in.

In the subsequent one-band calculation, presented in the next paragraph, we employ the occupation from the Ni 3dx2y2 orbital as calculated in this full DMFT calculation with 5 Ni and 5 Nd orbitals. This accounts not only for the electron pocket in the DMFT calculation but also for minor hybridization effects between the Ni 3dx2y2 and 3d3z2r2 orbital, for example, along the Γ-X direction. In principle, this hybridization effect, which inter-mixes the orbital contribution to the bands, should not be taken into account in the one-band Hubbard model. This aims at modeling the effective 3dx2y2 band which is crossing the Fermi level and which is predominantly Ni 3dx2y2 (but also has admixtures from the other orbitals because of the hybridization). In the case of the Ni 3dx2y2 orbital, this hybridization is very weak [33] (Supplemental Material) and can be neglected as a first approximation [62] (Supplemental Material). For other Ni 3d orbitals, this hybridization has a sizeable effect on their respective occupation. For example, the Ni 3d3z2r2 orbital which is strongly hybridizing with the Nd 5d3z2r2 orbital only has an occupation of 1.85 electrons per site in our multi-orbital calculation [62] (Supplemental material), whereas the effective Ni 3d3z2r2 orbital, including contributions from the hybridization, is fully occupied with two electrons per site as it is completely below the Fermi level.

The hopping parameters for the Ni-3dx2y2 model from a one-band Wannier projection are shown in Table 3 and compared to that of the same orbital in a 10-band and 17-band Wannier projection. Here, tRx,Ry,Rz denotes the hopping by Ri unit cells in the i direction. That is, t000 is the on-site potential, t = − t100; t' = − t110, and t = − t200 are the nearest, next-nearest, and next-next-nearest neighbor hopping, respectively; and tz = − t001 is the hopping in the z-direction perpendicular to the NiO2 planes. The hopping parameters are strikingly similar for LaNiO2 and NdNiO2 and the different Wannier projections.

TABLE 3
www.frontiersin.org

TABLE 3. Major hopping elements (in units of eV) of the Ni-3dx2y2 orbital from 1-band (Ni-3dx2y2), 10-bands (La/Nd-d + Ni-d), and 17-bands (La/Nd-f + La/Nd-d + Ni-d) Wannier projections. The DFT-relaxed lattice parameters are as follows: LaNiO2 (a = b = 3.88 Å, c = 3.35 Å), NdNiO2 (a = b = 3.86 Å, c = 3.24 Å) [33].

Besides the doping from Figure 4 and the hopping for the one-band Wannier projection from Table 3, we only need the interaction parameter for doing realistic one-band Hubbard model calculations for nickelates. In cRPA for a single 3dx2y2 orbital, one obtains U = 2.6 eV [15, 17] at zero frequency. But, the cRPA interaction has strong frequency dependence because of the screening of all the other Ni and Nd (La) orbitals close by. To mimic this frequency dependence, the static U parameter needs to be slightly increased. Expertise from many DFT + DMFT calculations for transition metal oxides shows that it typically needs to be about 0.5 eV larger, so that U = 3.2 eV = 8t is reasonable. Altogether, this defines a one-band Hubbard model for nickelates at various dopings. For the conductivity and other transport properties, the A pocket may be relevant as well, but superconductivity should arise from the correlated 3dx2y2 band that is hardly coupled to the A pocket.

5 Non-Local Correlations and Superconducting Phase Diagram

DFT provides a first picture of the relevant orbitals, and DMFT adds to this effect of strong local correlations, such as the splitting into Hubbard bands, the formation of a quasiparticle peak, and correlation-induced orbital shifts, such as the upshift of the Γ pocket. However, at low temperatures, non-local correlations give rise to additional effects. Relevant factors are as follows: the emergence of strong spin fluctuations and their impact on the spectral function and superconductivity.

For including such non-local correlations, diagrammatic extensions of DMFT, such as the dynamical vertex approximation (DΓA) [3638, 70], have been proven extremely powerful. Such calculations are possible down to the temperatures of the superconducting phase transition in the correlated regime and for very large lattices so that the long-range correlations close to a phase transition can be properly described. Even (quantum) critical exponents can be calculated [7174]. DΓA has proven reliable compared with the numerically exact calculations where these are possible [75] and, in particular, provide for a more accurate determination of Tc [76] since the full local frequency dependence of the two-particle vertex is included. Such local frequency dependence can affect even the non-local pairing through spin fluctuations. In, for example, RPA, this frequency dependence and the suppression of the pairing vertex for small frequencies can only be improperly mimicked by (quite arbitrarily) adjusting the static U.

This simple one-band Hubbard model in DΓA has been the basis for calculating the phase diagram Tc vs. Sr-doping in Figure 5 [33]. At the time of the calculation, only a single experimental Tc at 20% Sr-doping was available [5]. The physical origin of the superconductivity in these calculations is strong spin fluctuations which form the pairing glue for high-temperature superconductivity. Charge fluctuations are much weaker; the electron–phonon coupling has not been considered and is also too weak for transition metal oxides to yield high-temperature superconductivity. The theoretical Tc in Figure 5 at 20% doping was from the very beginning slightly larger than in the experiment. Most likely, this is because in the ladder DΓA [36, 37] calculation of Tc, the spin fluctuations are first calculated and then enter the superconducting particle–particle channel [76]. This neglects the feedback effect of these particle–particle fluctuations on the self-energy and the spin fluctuations, which may, in turn, suppress the tendency toward superconductivity somewhat. Such effects would be only included in a more complete parquet DΓA calculation [7779]. Also, the ignored weak three-dimensional dispersion will suppress Tc. Let us note that antiferromagnetic spin fluctuations have recently been observed experimentally [64, 80].

FIGURE 5
www.frontiersin.org

FIGURE 5. Superconducting phase diagram Tc vs. x for SrxNd1−xNiO2 as predicted by DΓA [33] and experimentally confirmed a posteriori in [34, 35]. A priori, that is, at the time of the calculation, only one experimental data point [5] was available. For such a difficult to determine quantity as the superconducting Tc and without adjusting parameters, the accuracy is astonishing. The bottom x-axis shows the Sr-doping and the top x-axis the calculated hole doping of the 3dx2y2 band according to Figure 4. Adjusted from [33].

Given the slight overestimation of Tc from the very beginning, the agreement with the subsequently obtained experimental Tc vs. Sr-doping x phase diagram [34, 35] in Figure 5 is astonishingly good. We further see that the superconducting doping regime also concurs with the doping regime where a one-band Hubbard model description is possible for SrxNd1−xNiO2, as concluded from a full DFT + DMFT calculation for 5 Ni plus 5 Nd bands. This regime is marked dark blue in Figure 5 and, as already noted, extends to somewhat larger dopings [33] than for SrxLa1−xNiO2 shown in Figure 3. Concomitant with this is the fact that the experimental superconducting doping range for SrxLa1−xNiO2 extends to a larger x than for SrxNd1−xNiO2. For dopings larger than the dark blue regime in Figure 5, two Ni 3d bands need to be included. As we will show in the next section, this completely changes the physics and is not favorable for superconductivity. For dopings smaller than the dark blue regime in Figure 5, on the other hand, the Γ pocket may become relevant for SrxNd1−xNiO2, as well as its exchange coupling to the 4f moments.

Our theoretical calculations also reveal ways to enhance Tc. Particularly promising is to enhance the hopping parameter t. This enhances Tc because (i) t sets the energy scale of the problem, and a larger t means a larger Tc if U/t, t′/t, and t/t and doping are kept fixed. Furthermore, the ratio U/t = 8 for nickelates is not yet optimal. Indeed, (ii) a somewhat smaller ratio U/t would imply a larger Tc at fixed t [33]. Since the interaction U is local, it typically varies much more slowly when, for example, applying compressive strain or pressure and can be assumed to be constant as a first approximation (for secondary effects, see [81, 82]). Thus, compressive strain or pressure enhance (i) t and reduce (ii) U/t. Both effects enhance Tc. This prediction made in [33] has been confirmed experimentally: applying a pressure of 12 GPa increases Tc from 18K to 31 K in Sr0.18Pr0.82NiO2 [83]. This is so far the record Tc for nickelates, and there are yet no signs for saturation or maximum, indicating that even higher Tc’s are possible at higher pressures.

Alternatives to enhance t are 1) to substitute the SrTiO3 substrate by a substrate with smaller in-plane lattice constants since the nickelate film in-plane axis parameters will be locked to that of the substrate. Furthermore, one can 2) replace 3d Ni by 4d Pd, that is, try to synthesize Nd (La)PdO2 [11]. Since the Pd 4d orbitals are more extended than the 3d Ni orbitals, this should enhance t as well.

Next, we turn to the DΓA spectra, more precisely Fermi surfaces, in Figure 6. Here, beyond quasiparticle renormalizations of DMFT, non-local spin fluctuations can further impact the spectrum. Only the spectral function of the Hubbard model is shown, describing the 3dx2y2 band. Please keep in mind that on top of the Fermi surface in Figure 6, there is also a weakly correlated A pocket. As one can see in Figure 6, antiferromagnetic spin-fluctuations lead to a pseudogap at the antinodal momenta (± π, 0) (0, ± π) if the filling of the 3dx2y2 band is close to half-filling. Indeed n3dx2y2=0.95 is the filling for the undoped parent compound NdNiO2 where the A- and Γ pocket have taken 5% of the electrons away from the Ni 3dx2y2 band. An Sr-doping of 20% is in between n3dx2y2=0.85 and n3dx2y2=0.8, see Figure 4. Comparing these theoretical predictions with the experimental Fermi surface, even the k-integrated spectrum is very much sought after. However, here, we face the difficulty that the superconducting samples require a SrTiO3 capping layer or otherwise may oxidize out of vacuum. This hinders photoemission spectroscopy (PES) experiments as these are extremely surface-sensitive. Hitherto, PES is only available without the capping layer for SrxPr1−xNiO2 [84]. These show a surprisingly low spectral density at the Fermi energy despite the metallic behavior of the doped system, raising the question of how similar these films are to the superconducting films.

FIGURE 6
www.frontiersin.org

FIGURE 6. DΓA k-resolved spectrum at the Fermi energy for T = 0.02t = 92 K (A) and T = 0.01t = 46 K (B) and four different dopings ndx2y2 of the Ni-3dx2y2 band (left to right). From [33].

6 Topotactic Hydrogen: Turning the Electronic Structure Upside Down

The fact that it took 20 years from the theoretical prediction of superconductivity in rare-earth nickelates to the experimental realization already suggests that the synthesis is far from trivial. This is because nickel has to be in the unusually low oxidation state Ni+1. The recipe of success for nickelate superconductors is a two-step process [85]: First, doped perovskite films Sr (Ca)xNd (La,Pr)1−xNiO3 films are deposited on a SrTiO3 substrate by pulsed laser deposition. Already, this first step is far from trivial, not least because the doped material has to be deposited with homogenous Sr (Ca) concentration. Second, Sr(Ca)xNd(La,Pr)1−xNiO3 needs to be reduced to Sr(Ca)xNd(La,Pr)1−xNiO2. To this end, the reducing agent CaH2 is used. Here, the problem is that this reduction might be incomplete with excess oxygen remaining or that hydrogen from CaH2 is topotactically intercalated in the Sr (Ca)xNd (La,Pr)1−xNiO2 structure. A particular difficulty is that the light hydrogen is experimentally hard to detect, for example, it evades conventional X-ray structural detection.

In [55], we studied the possibility to intercalate hydrogen, that is, to synthesize unintendedly SrxNd(La)1−xNiO2H instead of SrxNd(La)1−xNiO2. For the reduction of, for example, SrVO3 with CaH2, it is well-established that SrVO2H may be obtained as the end product [86]. Both possible end products are visualized in Figure 7. The extra H takes away one more electron from the Ni sites. Hence, we have two holes on the Ni sites which in a local picture are distributed to two orbitals and form a spin-1, due to Hund’s exchange.

FIGURE 7
www.frontiersin.org

FIGURE 7. Reduction of SrxNd(La)1−xNiO3 with CaH2 may result not only in the pursued end product SrxNd(La)1−xNiO2 but also in SrxNd(La)1−xNiO2H, where H atoms occupy the vacant O sites between the layers. This has dramatic consequences for the electronic structure. In a first, purely local picture, visualized on the right side, there is instead of Ni 3d9 with one-hole in the 3dx2y2 orbital, two holes in the 3dx2y2 and 3d3z2r2 orbital forming a local spin-1.

The first question is how susceptible the material is to bind topotactic H. To answer this question, one can calculate the binding energy E (ABNiO2) + 1/2 E (H2)- E (ABNiO2H) in DFT [55, 87]. The result is shown in Figure 8, which clearly shows that early transition metal oxides are prone to intercalate hydrogen, whereas for cuprates, the infinite layer compound without H is more stable. Nickelates are in-between. For the undoped compounds NdNiO2, and even a bit more for LaNiO2, it is favorable to intercalate H. However, for the Sr-doped nickelates, the energy balance is inverted. Here, it is unfavorable to bind hydrogen.

FIGURE 8
www.frontiersin.org

FIGURE 8. Binding energy for topotactical H as calculated by non–spin-polarized DFT and spin-polarized DFT + U for various transition metals B (x-axis). Positive binding energies indicate that ABNiO2H is energetically favored; for negative binding energies, ABNiO2 is more stable. The data for non–spin-polarized DFT calculations are from [55].

Let us emphasize that this is only the enthalpy balance. In the actual synthesis also the reaction kinetics matter, and the entropy which is large for the H2 gas. Nonetheless, this shows that undoped nickelates are very susceptible to topotactic H. This possibly means that, experimentally, not a complete H-coverage as in ABNiO2H of Figure 8 is realized, but some hydrogen may remain in the nickelates because of an incomplete reduction with CaH2. Indeed, hydrogen remainders have later been detected experimentally by nuclear magnetic resonance (NMR) spectroscopy, and they have even been used to analyze the antiferromagnetic spin fluctuations [80].

Now that we have established that remainders of hydrogen can be expected for nickelates at low doping, the question is how this affects the electronic structure. The local picture of Figure 7 already suggested a very different electronic configuration. This is further corroborated by DFT + DMFT calculations for LaNiO2H presented in Figure 9. Here, the DFT band structure shows a metallic behavior with two orbitals, Ni 3dx2y2 and 3d3z2r2, crossing the Fermi level. There are no rare-earth electron pockets any longer. Thus, we have an undoped Ni 3d8 configuration without Sr-doping. If electronic correlations are included in DMFT, the DFT bands split into two sets of Hubbard bands. Above the Fermi level, one can identify the upper 3dx2y2 and 3d3z2r2 Hubbard band below the flat f bands in Figure 9, with some broadening because of the electronic correlations. The lower Hubbard bands intertwine with the Ni t2g orbitals below the Fermi energy.

FIGURE 9
www.frontiersin.org

FIGURE 9. DFT (white lines) and DMFT (color bar) k-resolved spectral function for LaNiO2H. Here, a model with La-d + La-f + Ni-d was used, which is beyond the Ni-d only and La-d + Ni-d models shown in [55].

Even if we dope LaNiO2H, this electronic structure is not particularly promising for superconductivity. First, it is not two-dimensional because of the 3d3z2r2 orbitals, which makes the system more three-dimensional. More specifically, there is a considerable hopping process from Ni 3d3z2r2 via H to the Ni 3d3z2r2 on the vertically adjacent layer, as evidenced in Figure 9 by the DFT dispersion of this band in the Γ-Z direction; the other 3dx2y2 band is (as expected) flat in this direction. Second, the tendency to form local magnetic moments of spin-1 counteracts the formation of Cooper pairs from two spin-1/2s. Hence, altogether, we expect topotactic H to prevent high-temperature superconductivity.

7 Conclusion

In this article, we have discussed the physics of nickelate superconductors from the perspective of a one-band Hubbard model for the Ni 3dx2y2 band plus an A pocket. Because of symmetry, this A pocket does not hybridize with the 3dx2y2 band and merely acts as a decoupled electron reservoir. Hence, once the filling of the 3dx2y2 band is calculated as a function of Sr- or Ca-doping in Sr(Ca)xNd(La,Pr)1−xNiO2, we can, for many aspects, concentrate on the physics of the thus doped Hubbard model. This includes antiferromagnetic spin fluctuations and the onset of superconductivity. Other physical properties, such as transport and the Hall conductivity, depend, as a matter of course, also on the A pocket. This is in stark contrast to the cuprates, where the oxygen p orbitals are much closer to the Fermi level so that we have a charge-transfer insulator that needs to be modeled by the more complex Emery model.

The one-band Hubbard model picture for nickelates was put forward early on for nickelates [11, 19, 32, 33], and its proper doping, including correlation effects, has been calculated in [33]. This picture has been confirmed by many experimental observations so far. The Nd 4f states are, from the theoretical perspective, irrelevant because they form a local spin and barely hybridize with the 3dx2y2 band. This has been confirmed experimentally by the observation of superconductivity in Sr (Ca)xLa1−xNiO2. The minor importance of the other Ni 3d orbitals, in particular the 3d3z2r2 orbital, is indicated through the careful analysis [65] of RIXS data [63, 64]. Not confirmed experimentally is hitherto the prediction that the Γ pocket is shifted above the Fermi level in the superconducting doping regime.

Strong evidence for the one-band Hubbard model picture is the prediction of the superconducting phase diagram [33], confirmed experimentally in [34, 35]. A further prediction was the increase of Tc with pressure or compressive strain [33], which was subsequently found in the experiment with a record Tc = 31 K for nickelates under pressure [83]. The strength of antiferromagnetic spin-fluctuations as obtained in RIXS [64] also roughly agrees with that of the calculation [33]. Altogether, this gives us quite some confidence in the one-band Hubbard model scenario, which even allowed for a rough calculation of Tc. Notwithstanding, further theoretical calculations, in particular including non-local correlations also in a realistic multi-orbital setting [37, 70, 88], are eligible. On the experimental side more detailed, for example, k-resolved information is desirable as are further close comparisons between the experiment and theory.

A good analysis of the quality of the samples is also mandatory, especially against the background that superconducting nickelates have been extremely difficult to synthesize. Incomplete oxygen reduction and topotactic hydrogen [55, 87] are theoretically expected to be present because this is energetically favored, at least for low Sr-doping. This leads to two holes in two orbitals forming a high-spin state and a three-dimensional electronic structure, thus obstructing the intrinsic physics of superconducting nickelates.

Author Contributions

All authors listed have made a substantial, direct, and intellectual contribution to the study and approved it for publication.

Funding

We acknowledge funding through the Austrian Science Funds (FWF), Project numbers P 32044 and P 30213, and Grant-in-Aids for Scientific Research (JSPS KAKENHI), Grant numbers 19H05825, JP20K22342, and JP21K13887. OJ was supported by the Leibniz Association through the Leibniz Competition. Calculations were partially performed on the Vienna Scientific Cluster (VSC).

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Acknowledgments

We thank Atsushi Hariki, Motoaki Hirayama, Josef Kaufmann, Yusuke Nomura, and Terumasa Tadano for valuable discussions.

References

1. Anisimov VI, Bukhvalov D, Rice TM. Electronic Structure of Possible Nickelate Analogs to the Cuprates. Phys Rev B (1999) 59:7901–6. doi:10.1103/PhysRevB.59.7901

CrossRef Full Text | Google Scholar

2. Chaloupka J, Khaliullin G. Orbital Order and Possible Superconductivity in LaNiO3/LaMO3 Superlattices. Phys Rev Lett (2008) 100:016404. doi:10.1103/PhysRevLett.100.016404

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Hansmann P, Toschi A, Yang X, Andersen O, Held K. Electronic Structure of Nickelates: From Two-Dimensional Heterostructures to Three-Dimensional Bulk Materials. Phys Rev B (2010) 82:235123. doi:10.1103/PhysRevB.82.235123

CrossRef Full Text | Google Scholar

4. Hansmann P, Yang X, Toschi A, Khaliullin G, Andersen OK, Held K. Turning a Nickelate Fermi Surface into a Cupratelike One through Heterostructuring. Phys Rev Lett (2009) 103:016401. doi:10.1103/PhysRevLett.103.016401

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Li D, Lee K, Wang BY, Osada M, Crossley S, Lee HR, et al. Superconductivity in an Infinite-Layer Nickelate. Nature (2019) 572:624–7. doi:10.1038/s41586-019-1496-5

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Balestrino G, Medaglia PG, Orgiani P, Tebano A, Aruta C, Lavanga S, et al. Very Large Purely Intralayer Critical Current Density in Ultrathin Cuprate Artificial Structures. Phys Rev Lett (2002) 89:156402. doi:10.1103/physrevlett.89.156402

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Di Castro D, Cantoni C, Ridolfi F, Aruta C, Tebano A, Yang N, et al. High-Tc Superconductivity at the Interface between the CaCuO2 and SrTiO3 Insulating Oxides. Phys Rev Lett (2015) 115:147001. doi:10.1103/PhysRevLett.115.147001

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Orgiani P, Aruta C, Balestrino G, Born D, Maritato L, Medaglia PG, et al. Direct Measurement of Sheet Resistance R in Cuprate Systems: Evidence of a Fermionic Scenario in a Metal-Insulator Transition. Phys Rev Lett (2007) 98:036401. doi:10.1103/PhysRevLett.98.036401

PubMed Abstract | CrossRef Full Text | Google Scholar

9. Siegrist T, Zahurak S, Murphy D, Roth R. The Parent Structure of the Layered High-Temperature Superconductors. Nature (1988) 334:231–2. doi:10.1038/334231a0

CrossRef Full Text | Google Scholar

10. Botana AS, Norman MR. Similarities and Differences between LaNiO2 and CaCuO2 and Implications for Superconductivity. Phys Rev X (2020) 10:011024. doi:10.1103/PhysRevX.10.011024

CrossRef Full Text | Google Scholar

11. Hirayama M, Tadano T, Nomura Y, Arita R. Materials Design of Dynamically Stable D9 Layered Nickelates. Phys Rev B (2020) 101:075107. doi:10.1103/PhysRevB.101.075107

CrossRef Full Text | Google Scholar

12. Hu L-H, Wu C. Two-band Model for Magnetism and Superconductivity in Nickelates. Phys Rev Res (2019) 1:032046. doi:10.1103/PhysRevResearch.1.032046

CrossRef Full Text | Google Scholar

13. Jiang M, Berciu M, Sawatzky GA. Critical Nature of the Ni Spin State in Doped NdNiO2. Phys Rev Lett (2020) 124:207004. doi:10.1103/PhysRevLett.124.207004

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Lee K-W, Pickett W. Infinite-layer LaNiO2: Ni1+ Is Not Cu2+. Phys Rev B (2004) 70:165109. doi:10.1103/PhysRevB.70.165109

CrossRef Full Text | Google Scholar

15. Nomura Y, Hirayama M, Tadano T, Yoshimoto Y, Nakamura K, Arita R. Formation of a Two-Dimensional Single-Component Correlated Electron System and Band Engineering in the Nickelate Superconductor NdNiO2. Phys Rev B (2019) 100:205138. doi:10.1103/PhysRevB.100.205138

CrossRef Full Text | Google Scholar

16. Pavarini E, Dasgupta I, Saha-Dasgupta T, Jepsen O, Andersen OK. Band-structure Trend in Hole-Doped Cuprates and Correlation with T (C max). Phys Rev Lett (2001) 87:047003. doi:10.1103/PhysRevLett.87.047003

PubMed Abstract | CrossRef Full Text | Google Scholar

17. Sakakibara H, Usui H, Suzuki K, Kotani T, Aoki H, Kuroki K. Model Construction and a Possibility of Cupratelike Pairing in a New D9 Nickelate Superconductor (Nd, Sr)NiO2. Phys Rev Lett (2020) 125:077003. doi:10.1103/PhysRevLett.125.077003

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Werner P, Hoshino S. Nickelate Superconductors: Multiorbital Nature and Spin Freezing. Phys Rev B (2020) 101:041104. doi:10.1103/PhysRevB.101.041104

CrossRef Full Text | Google Scholar

19. Wu X, Di Sante D, Schwemmer T, Hanke W, Hwang HY, Raghu S, et al. Robust Dx2-y2 -wave Superconductivity of Infinite-Layer Nickelates. Phys Rev B (2020) 101:060504. doi:10.1103/PhysRevB.101.060504

CrossRef Full Text | Google Scholar

20. Zhang G-M, Yang Y-F, Zhang F-C. Self-Doped Mott Insulator for Parent Compounds of Nickelate Superconductors. Phys Rev B (2020) 101:020501. doi:10.1103/PhysRevB.101.020501

CrossRef Full Text | Google Scholar

21. Adhikary P, Bandyopadhyay S, Das T, Dasgupta I, Saha-Dasgupta T. Orbital-selective Superconductivity in a Two-Band Model of Infinite-Layer Nickelates. Phys Rev B (2020) 102:100501. doi:10.1103/PhysRevB.102.100501

CrossRef Full Text | Google Scholar

22. Lechermann F. Late Transition Metal Oxides with Infinite-Layer Structure: Nickelates versus Cuprates. Phys Rev B (2020) 101:081110. doi:10.1103/PhysRevB.101.081110

CrossRef Full Text | Google Scholar

23. Lechermann F. Multiorbital Processes Rule the Nd1−xSrxNiO2 normal State. Phys Rev X (2020) 10:041002. doi:10.1103/PhysRevX.10.041002

CrossRef Full Text | Google Scholar

24. Petocchi F, Christiansson V, Nilsson F, Aryasetiawan F, Werner P. Normal State of Nd1−xSrxNiO2 from Self-Consistent GW + EDMFT. Phys Rev X (2020) 10:041047. doi:10.1103/PhysRevX.10.041047

CrossRef Full Text | Google Scholar

25. Bandyopadhyay S, Adhikary P, Das T, Dasgupta I, Saha-Dasgupta T. Superconductivity in Infinite-Layer Nickelates: Role of F Orbitals. Phys Rev B (2020) 102:220502. doi:10.1103/PhysRevB.102.220502

CrossRef Full Text | Google Scholar

26. Zaanen J, Sawatzky GA, Allen JW. Band Gaps and Electronic Structure of Transition-Metal Compounds. Phys Rev Lett (1985) 55:418–21. doi:10.1103/PhysRevLett.55.418

PubMed Abstract | CrossRef Full Text | Google Scholar

27. Emery VJ. Theory of High–Tc Superconductivity in Oxides. Phys Rev Lett (1987) 58:2794–7. doi:10.1103/PhysRevLett.58.2794

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Gutzwiller MC. Effect of Correlation on the Ferromagnetism of Transition Metals. Phys Rev Lett (1963) 10:159–62. doi:10.1103/PhysRevLett.10.159

CrossRef Full Text | Google Scholar

29. Hubbard J. Electron Correlations in Narrow Energy Bands. Proc R Soc Lond Ser A, Math Phys Sci (1963) 276:238–57. doi:10.1098/rspa.1963.0204

CrossRef Full Text | Google Scholar

30. Kanamori J. Electron Correlation and Ferromagnetism of Transition Metals. Prog Theor Phys (1963) 30:275–89. doi:10.1143/ptp.30.275

CrossRef Full Text | Google Scholar

31. Zhang FC, Rice TM. Effective Hamiltonian for the Superconducting Cu Oxides. Phys Rev B (1988) 37:3759–61. doi:10.1103/PhysRevB.37.3759

PubMed Abstract | CrossRef Full Text | Google Scholar

32. Karp J, Botana AS, Norman MR, Park H, Zingl M, Millis A. Many-Body Electronic Structure of NdNiO2 and CaCuO2. Phys Rev X (2020) 10:021061. doi:10.1103/PhysRevX.10.021061

CrossRef Full Text | Google Scholar

33. Kitatani M, Si L, Janson O, Arita R, Zhong Z, Held K. Nickelate Superconductors – a Renaissance of the One-Band Hubbard Model. npj Quan Mater (2020) 5:59. arXiv:2002.12230. doi:10.1038/s41535-020-00260-y

CrossRef Full Text | Google Scholar

34. Li D, Wang BY, Lee K, Harvey SP, Osada M, Goodge BH, et al. Superconducting Dome in Nd1−xSrxNiO2 Infinite Layer Films. Phys Rev Lett (2020) 125:027001. arxiv:2003.08506. doi:10.1103/PhysRevLett.125.027001

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Zeng S, Tang CS, Yin X, Li C, Li M, Huang Z, et al. Phase Diagram and Superconducting Dome of Infinite-Layer Nd1−xSrxNiO2 Thin Films, arxiv:2004.11281. Phys Rev Lett (2020) 125:147003. doi:10.1103/physrevlett.125.147003

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Katanin AA, Toschi A, Held K. Comparing Pertinent Effects of Antiferromagnetic Fluctuations in the Two- and Three-Dimensional Hubbard Model. Phys Rev B (2009) 80:075104. doi:10.1103/PhysRevB.80.075104

CrossRef Full Text | Google Scholar

37. Rohringer G, Hafermann H, Toschi A, Katanin AA, Antipov AE, Katsnelson MI, et al. Diagrammatic Routes to Nonlocal Correlations beyond Dynamical Mean Field Theory. Rev Mod Phys (2018) 90:025003. doi:10.1103/revmodphys.90.025003

CrossRef Full Text | Google Scholar

38. Toschi A, Katanin AA, Held K. Dynamical Vertex Approximation; a Step Beyond Dynamical Mean-Field Theory. Phys Rev B (2007) 75:045118. doi:10.1103/PhysRevB.75.045118

CrossRef Full Text | Google Scholar

39. Georges A, Kotliar G, Krauth W, Rozenberg MJ. Dynamical Mean-Field Theory of Strongly Correlated Fermion Systems and the Limit of Infinite Dimensions. Rev Mod Phys (1996) 68:13. doi:10.1103/RevModPhys.68.13

CrossRef Full Text | Google Scholar

40. Georges A, Krauth W. Numerical Solution of the D = Hubbard Model: Evidence for a Mott Transition. Phys Rev Lett (1992) 69:1240–3. doi:10.1103/PhysRevLett.69.1240

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Jarrell M. Hubbard Model in Infinite Dimensions: A Quantum Monte Carlo Study. Phys Rev Lett (1992) 69:168–71. doi:10.1103/PhysRevLett.69.168

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Metzner W, Vollhardt D. Correlated Lattice Fermions in D = Dimensions. Phys Rev Lett (1989) 62:324–7. doi:10.1103/PhysRevLett.62.324

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Blaha P, Schwarz K, Madsen G, Kvasnicka D, Luitz J. wien2k: An Augmented Plane Wave+ Local Orbitals Program for Calculating crystal Properties (2001). Wien: TU Wien.

Google Scholar

44. Schwarz K, Blaha P, Madsen GKH (2002). Electronic Structure Calculations of Solids Using the Wien2k Package for Material Sciences. Comp Phys Comm 147:71–6. doi:10.1016/S0010-4655(02)00206-0

CrossRef Full Text | Google Scholar

45. Kresse G, Hafner J. Ab Initio molecular Dynamics for Open-Shell Transition Metals. Phys Rev B (1993) 48:13115–8. doi:10.1103/PhysRevB.48.13115

CrossRef Full Text | Google Scholar

46. Koepernik K, Eschrig H. Full-potential Nonorthogonal Local-Orbital Minimum-Basis Band-Structure Scheme. Phys Rev B (1999) 59:1743–57. doi:10.1103/PhysRevB.59.1743

CrossRef Full Text | Google Scholar

47. Perdew JP, Burke K, Ernzerhof M. Generalized Gradient Approximation Made Simple. Phys Rev Lett (1996) 77:3865–8. doi:10.1103/PhysRevLett.77.3865

PubMed Abstract | CrossRef Full Text | Google Scholar

48. Nomura Y, Arita R. Superconductivity in Infinite-Layer Nickelates (2021). arXiv:2107.12923.

Google Scholar

49. Hansmann P, Parragh N, Toschi A, Sangiovanni G, Held K. Importance of D–P Coulomb Interaction for High Tc Cuprates and Other Oxides. New J Phys (2014) 16:033009. doi:10.1088/1367-2630/16/3/033009

CrossRef Full Text | Google Scholar

50. Anisimov VI, Poteryaev AI, Korotin MA, Anokhin AO, Kotliar G. First-principles Calculations of the Electronic Structure and Spectra of Strongly Correlated Systems: Dynamical Mean-Field Theory. J Phys Condens Matter (1997) 9:7359–67. doi:10.1088/0953-8984/9/35/010

CrossRef Full Text | Google Scholar

51. Held K. Electronic Structure Calculations Using Dynamical Mean Field Theory. Adv Phys (2007) 56:829–926. doi:10.1080/00018730701619647

CrossRef Full Text | Google Scholar

52. Held K, Nekrasov IA, Keller G, Eyert V, Blümer N, McMahan AK, et al. Realistic Investigations of Correlated Electron Systems with LDA+DMFT. physica status solidi (b) (2006) 243:2599–631. Previously appeared as Psi-k Newsletter No. 56 (April 2003). doi:10.1002/pssb.200642053

CrossRef Full Text | Google Scholar

53. Kotliar G, Savrasov SY, Haule K, Oudovenko VS, Parcollet O, Marianetti CA. Electronic Structure Calculations with Dynamical Mean-Field Theory. Rev Mod Phys (2006) 78:865. doi:10.1103/RevModPhys.78.865

CrossRef Full Text | Google Scholar

54. Lichtenstein AI, Katsnelson MI. Ab Initio calculations of Quasiparticle Band Structure in Correlated Systems: LDA++ Approach. Phys Rev B (1998) 57:6884–95. doi:10.1103/PhysRevB.57.6884

CrossRef Full Text | Google Scholar

55. Si L, Xiao W, Kaufmann J, Tomczak JM, Lu Y, Zhong Z, et al. Topotactic Hydrogen in Nickelate Superconductors and Akin Infinite-Layer Oxides ABO2. Phys Rev Lett (2020) 124:166402. doi:10.1103/PhysRevLett.124.166402

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Kuneš J, Arita R, Wissgott P, Toschi A, Ikeda H, Held K. Wien2wannier: From Linearized Augmented Plane Waves to Maximally Localized Wannier Functions. Comput Phys Commun (2010) 181:1888–95. doi:10.1016/j.cpc.2010.08.005

CrossRef Full Text | Google Scholar

57. Gull E, Millis AJ, Lichtenstein AI, Rubtsov AN, Troyer M, Werner P. Continuous-time Monte Carlo Methods for Quantum Impurity Models. Rev Mod Phys (2011) 83:349–404. doi:10.1103/RevModPhys.83.349

CrossRef Full Text | Google Scholar

58. Parragh N, Toschi A, Held K, Sangiovanni G. Conserved Quantities of SU(2)-Invariant Interactions for Correlated Fermions and the Advantages for Quantum Monte Carlo Simulations. Phys Rev B (2012) 86:155158. doi:10.1103/PhysRevB.86.155158

CrossRef Full Text | Google Scholar

59. Wallerberger M, Hausoel A, Gunacker P, Kowalski A, Parragh N, Goth F, et al. w2dynamics: Local One- and Two-Particle Quantities from Dynamical Mean Field Theory. Comp Phys Comm (2019) 235:388–99. doi:10.1016/j.cpc.2018.09.007

CrossRef Full Text | Google Scholar

60. Kaufmann J, Held K. ana_cont: Python Package for Analytic Continuation (2021). arXiv:2105.11211.

Google Scholar

61. Osada M, Wang BY, Goodge BH, Harvey SP, Lee K, Li D, et al. Nickelate Superconductivity without Rare-Earth Magnetism: (La,Sr)NiO2. Adv Mater (2021) 33:2104083. doi:10.1002/adma.202104083

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Worm P, Si L, Kitatani M, Arita R, Tomczak JM, Held K. Correlations Turn Electronic Structure of Finite-Layer Nickelates Upside Down (2021). arXiv:2111.12697.

Google Scholar

63. Hepting M, Li D, Jia CJ, Lu H, Paris E, Tseng Y, et al. Electronic Structure of the Parent Compound of Superconducting Infinite-Layer Nickelates. Nat Mater (2020) 19:381. doi:10.1038/s41563-019-0585-z

PubMed Abstract | CrossRef Full Text | Google Scholar

64. Lu H, Rossi M, Nag A, Osada M, Li DF, Lee K, et al. Magnetic Excitations in Infinite-Layer Nickelates. Science (2021) 373:213–6. doi:10.1126/science.abd7726

PubMed Abstract | CrossRef Full Text | Google Scholar

65. Higashi K, Winder M, Kuneš J, Hariki A. Core-level X-ray Spectroscopy of Infinite-Layer Nickelate: LDA + DMFT Study. Phys Rev X (2021) 11:041009. doi:10.1103/PhysRevX.11.041009

CrossRef Full Text | Google Scholar

66. Gu Y, Zhu S, Wang X, Hu J, Chen H. A Substantial Hybridization between Correlated Ni-D Orbital and Itinerant Electrons in Infinite-Layer Nickelates. Commun Phy (2020) 3:84. doi:10.1038/s42005-020-0347-x

CrossRef Full Text | Google Scholar

67. Jiang P, Si L, Liao Z, Zhong Z. Electronic Structure of Rare-Earth Infinite-Layer RNiO2 (R = La, Nd). Phys Rev B (2019) 100:201106. doi:10.1103/PhysRevB.100.201106

CrossRef Full Text | Google Scholar

68. Choi M-Y, Lee K-W, Pickett WE. Role of 4f States in Infinite-Layer NdNiO2. Phys Rev B (2020) 101:020503. doi:10.1103/PhysRevB.101.020503

CrossRef Full Text | Google Scholar

69. Zeng SW, Li CJ, Chow LE, Cao Y, Zhang ZT, Tang CS, et al. Superconductivity in Infinite-Layer Lanthanide Nickelates (2021). arXiv:2105.13492.

Google Scholar

70. Galler A, Thunström P, Gunacker P, Tomczak JM, Held K. Ab Initio dynamical Vertex Approximation. Phys Rev B (2017) 95:115107. doi:10.1103/PhysRevB.95.115107

CrossRef Full Text | Google Scholar

71. Antipov AE, Gull E, Kirchner S. Critical Exponents of Strongly Correlated Fermion Systems from Diagrammatic Multiscale Methods. Phys Rev Lett (2014) 112:226401. doi:10.1103/PhysRevLett.112.226401

PubMed Abstract | CrossRef Full Text | Google Scholar

72. Rohringer G, Toschi A, Katanin A, Held K. Critical Properties of the Half-Filled Hubbard Model in Three Dimensions. Phys Rev Lett (2011) 107:256402. doi:10.1103/PhysRevLett.107.256402

PubMed Abstract | CrossRef Full Text | Google Scholar

73. Schäfer T, Katanin AA, Held K, Toschi A. Interplay of Correlations and Kohn Anomalies in Three Dimensions: Quantum Criticality with a Twist. Phys Rev Lett (2017) 119:046402. doi:10.1103/PhysRevLett.119.046402

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Schäfer T, Katanin AA, Kitatani M, Toschi A, Held K. Quantum Criticality in the Two-Dimensional Periodic anderson Model. Phys Rev Lett (2019) 122:227201. doi:10.1103/PhysRevLett.122.227201

PubMed Abstract | CrossRef Full Text | Google Scholar

75. Schäfer T, Wentzell N, Šimkovic F, He Y-Y, Hille C, Klett M, et al. Tracking the Footprints of Spin Fluctuations: A Multimethod, Multimessenger Study of the Two-Dimensional Hubbard Model. Phys Rev X (2021) 11:011058. doi:10.1103/PhysRevX.11.011058

CrossRef Full Text | Google Scholar

76. Kitatani M, Schäfer T, Aoki H, Held K. Why the Critical Temperature of High-Tc Cuprate Superconductors Is So Low: The Importance of the Dynamical Vertex Structure. Phys Rev B (2019) 99:041115. doi:10.1103/PhysRevB.99.041115

CrossRef Full Text | Google Scholar

77. Li G, Kauch A, Pudleiner P, Held K. The Victory Project v1.0: An Efficient Parquet Equations Solver. Comp Phys Comm (2019) 241:146–54. doi:10.1016/j.cpc.2019.03.008

CrossRef Full Text | Google Scholar

78. Li G, Wentzell N, Pudleiner P, Thunström P, Held K. Efficient Implementation of the Parquet Equations: Role of the Reducible Vertex Function and its Kernel Approximation. Phys Rev B (2016) 93:165103. doi:10.1103/physrevb.93.165103

CrossRef Full Text | Google Scholar

79. Valli A, Schäfer T, Thunström P, Rohringer G, Andergassen S, Sangiovanni G, et al. Dynamical Vertex Approximation in its Parquet Implementation: Application to Hubbard Nanorings. Phys Rev B (2015) 91:115115. doi:10.1103/PhysRevB.91.115115

CrossRef Full Text | Google Scholar

80. Cui Y, Li C, Li Q, Zhu X, Hu Z, feng Yang Y, et al. NMR Evidence of Antiferromagnetic Spin Fluctuations in Nd0.85Sr0.15NiO2. Chin Phys Lett (2021) 38:067401. doi:10.1088/0256-307x/38/6/067401

CrossRef Full Text | Google Scholar

81. Ivashko O, Horio W, Wan M, Christensen N, McNally D, Paris E, et al. Strain-engineering Mott-insulating La2CuO4. Nat Comm (2019) 10:786. doi:10.1038/s41467-019-08664-6

PubMed Abstract | CrossRef Full Text | Google Scholar

82. Tomczak JM, Miyake T, Sakuma R, Aryasetiawan F. Effective Coulomb Interactions in Solids under Pressure. Phys Rev B (2009) 79:235133. doi:10.1103/PhysRevB.79.235133

CrossRef Full Text | Google Scholar

83. Wang NN, Yang MW, Chen KY, Yang Z, Zhang H, Zhu ZH, et al. Pressure-induced Monotonic Enhancement of Tc to over 30 K in the Superconducting Pr0.82Sr0.18NiO2 Thin Films (2021). arXiv:2109.12811.

Google Scholar

84. Chen Z, Osada M, Li D, Been EM, Chen S-D, Hashimoto M, et al. Electronic Structure of Superconducting Nickelates Probed by Resonant Photoemission Spectroscopy (2021). arXiv:2106.03963.

Google Scholar

85. Lee K, Goodge BH, Li D, Osada M, Wang BY, Cui Y, et al. Aspects of the Synthesis of Thin Film Superconducting Infinite-Layer Nickelates. APL Mater (2020) 8:041107. doi:10.1063/5.0005103

CrossRef Full Text | Google Scholar

86. Katayama T, Chikamatsu A, Yamada K, Shigematsu K, Onozuka T, Minohara M, et al. Epitaxial Growth and Electronic Structure of Oxyhydride SrVO2H Thin Films. J Appl Phys (2016) 120:085305. doi:10.1063/1.4961446

CrossRef Full Text | Google Scholar

87. Malyi OI, Varignon J, Zunger A. Bulk NdNiO2 Is Thermodynamically Unstable with Respect to Decomposition while Hydrogenation Reduces the Instability and Transforms it from Metal to Insulator (2021). arXiv:2107.01790.

Google Scholar

88. Tomczak JM, Liu P, Toschi A, Kresse G, Held K. Merging GW with DMFT and Non-local Correlations beyond. Eur Phys J Spec Top (2017) 226:2565–90. doi:10.1140/epjst/e2017-70053-1

CrossRef Full Text | Google Scholar

Keywords: electronic structure calculations, dynamical mean field theory, electronic correlation, high-temperature superconductivity, solid state theory

Citation: Held K, Si L, Worm P, Janson O, Arita R, Zhong Z, Tomczak JM and Kitatani M (2022) Phase Diagram of Nickelate Superconductors Calculated by Dynamical Vertex Approximation. Front. Phys. 9:810394. doi: 10.3389/fphy.2021.810394

Received: 06 November 2021; Accepted: 10 December 2021;
Published: 21 January 2022.

Edited by:

Matthias Hepting, Max Planck Institute for Solid State Research, Germany

Reviewed by:

Fabio Bernardini, University of Cagliari, Italy
Liviu Chioncel, University of Augsburg, Germany

Copyright © 2022 Held, Si, Worm, Janson, Arita, Zhong, Tomczak and Kitatani. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Karsten Held, held@ifp.tuwien.ac.at; Liang Si, liang.si@ifp.tuwien.ac.at; Motoharu Kitatani, motoharu.kitatani@riken.jp

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.