Abstract
The search for superconductivity in infinite-layer nickelates was motivated by analogy to the cuprates, and this perspective has framed much of the initial consideration of this material. However, a growing number of studies have highlighted the involvement of rare-earth orbitals; in that context, the consequences of varying the rare-earth element in the superconducting nickelates have been much debated. Here, we show notable differences in the magnitude and anisotropy of the superconducting upper critical field across the La-, Pr-, and Nd-nickelates. These distinctions originate from the 4f electron characteristics of the rare-earth ions in the lattice: They are absent for La3+, nonmagnetic for the Pr3+ singlet ground state, and magnetic for the Nd3+ Kramer’s doublet. The unique polar and azimuthal angle-dependent magnetoresistance found in the Nd-nickelates can be understood to arise from the magnetic contribution of the Nd3+ 4f moments. Such robust and tunable superconductivity suggests potential in future high-field applications.
Rare-earth moments control the magnitude and anisotropy of superconducting upper critical fields in layered nickelates.
INTRODUCTION
Following the initial discovery of superconductivity in Nd0.8Sr0.2NiO2 (1), the subsequent development of rare-earth (R) variants with Pr and La in place of Nd (2–5) has created an emerging family of superconducting infinite-layer nickelates (Fig. 1A) (6–8). Unlike the cuprates, with which the nickelates are often compared, R-site variations are expected to have appreciable impact beyond ionic size effects. Specifically, rare-earth effects in cuprates are usually overshadowed by the much larger energy scale of superconductivity (9, 10). One exception is in PrBa2Cu3O7−δ, where coincidental hybridization between the Pr 4f and oxygen 2p states is believed to compete with the emergence of superconductivity (11). In contrast, in the nickelate superconductors, there is consistent transport, spectroscopic, and theoretical evidence for hybridization between a broad range of R 5d orbitals and the Ni 3d orbitals (2–8, 12–24). An open question currently under debate is the potential role of the R 4f electrons (25–28). Here, we report that the R-site directly controls the magnitude and anisotropy of the superconducting upper critical field Hc2.
RESULTS
Figure 1 shows a central result motivating this study—that the magnetic field suppression of superconductivity differs widely among R-variants, which does not scale with Tc0, the superconducting transition temperature in zero external magnetic field. Magnetoresistance measurements up to 35 T of representative (R,Sr)NiO2 (R = La, Pr, Nd) samples at optimal doping are shown in Fig. 1 (B to D), for both out-of-plane (H⊥) and in-plane (H||) field orientations. The corresponding resistively determined upper critical fields (Hc2⊥ and Hc2||; defined as 50% of the normal-state resistivity) are shown (Fig. 1E), where the drastically different scales for Hc2 are evident.
For a superconductor in the dirty limit, which describes the nickelates (19, 29), the effective Ginzburg-Landau coherence length ξ is the geometric mean of the Pippard coherence length ξ0 and the mean free path (30). Then, as , Hc2⊥ ∝ 1/ξ0 ∝ Δ0 ∝ Tc0. Here, ϕ0 is the magnetic flux quantum and Δ0 the superconducting gap. As for Hc2||, the observed (1 − T/Tc0)1/2 temperature dependence indicates dominant paramagnetic depairing near Tc0 (19, 29). Hc2|| is also proportional to Tc0 at temperatures near Tc0, described by (31), where kB is the Boltzmann constant, μ0 is the vacuum magnetic permeability, and η(x) is the Riemann zeta function. Therefore, we normalize both Hc2 and T by Tc0 (Fig. 1F) for all samples. We observe a systematic variation in Hc2/Tc0, which is largest for La-nickelates, somewhat smaller for Pr, and substantially reduced for Nd. Furthermore, this variation between R-sites is persistent across the respective superconducting domes via Sr doping (Fig. 1F; see figs. S1 and S2 for complete datasets).
To probe the origin of this large variation in Hc2, we measure in detail the magnetoresistance as a function of the polar (θ) and azimuthal (ϕ) angle of the applied magnetic field direction, as illustrated in Fig. 2A. In the normal state, negligible angular dependence is observed for all samples (Fig. 2, C and D). Therefore, we focus on the superconducting transition region with finite sample resistance below the normal-state value. We first present the characteristic features through representative measurements (Fig. 2E; see figs. S3 to S5 for complete datasets). The La- and Pr-nickelates both show conventional angular dependence; their “∞”-shaped θ-dependence shows that superconductivity is more robust to in-plane magnetic fields than perpendicular fields, as commonly seen in layered superconductors (32). Their “figure-of-eight”–shaped ϕ-dependence, arising from vortex motion, also follows previous studies of other type II superconductors (32, 33). In sharp contrast, the Nd-nickelates show qualitatively distinct magnetoresistance. The orientation of the θ-dependence is opposite to that of La-/Pr-nickelates and other layered superconductors, i.e., superconductivity is more suppressed by an in-plane magnetic field, indicating a complete reversal of the superconducting anisotropy. Furthermore, a pronounced clover-leaf ϕ-dependence is observed for Nd. As shown in fig. S6, we can rule out a vortex-related origin of these features.
How can we understand this clear difference between the Nd-nickelates versus the La-/Pr-nickelates? In all cases, the in-plane lattice parameters are clamped to that of the SrTiO3 substrate, while for a specific Sr doping, the c-axis variations reflect the average R-site ionic radii (1–3): La3+ is the largest, ~4.5% larger than the closely similar Pr3+ and Nd3+ (within 1.5% of each other) (34). In addition, the range of the c-axis lattice constant variation of the three R-variants across the superconducting dome substantially overlap. Thus, cation size effects do not seem to account for the unique Hc2 features in Nd. Rather, the clover-leaf ϕ-dependent magnetoresistance in Nd is more suggestive of a potential role of local magnetic moments, leading to a consideration of the 4f electrons and their crystal electric field (CEF) structure.
To examine this, we use theoretical approaches including ab initio density functional theory (DFT) calculations (35, 36), Bader charge analysis (37), and symmetry arguments to determine the 4f levels split by the tetragonal CEF (see the Supplementary Materials for details). In Fig. 3A, we show the electron density isosurface of the Ni 3d, 4s, O 2p, and R 5d, 6s orbital electrons within a unit cell where the contours correspond to 2.5% of the maximum electron density. Despite the hybridization, especially between the O 2p and Ni 3d orbitals, most of the electron density remains within the vicinity of the parent ion. This motivates the use of a point charge approximation in determining the CEF-split 4f energy levels, as shown for Pr3+ and Nd3+ as a function of c-axis lattice constant (Fig. 3, B and C), where the range of experimental values is highlighted by the vertical blue bands and the bulk values are marked by dashed lines (38). While La3+ has no 4f electrons, we find a singlet crystal field ground state for Pr3+ that to first order is also nonmagnetic. By contrast, Nd3+ has a doublet ground state as expected for a Kramer’s ion, making this the lone magnetic R-site ion among the three. Furthermore, by adding an external field term to the CEF perturbation Hamiltonian, we find that the Nd 4f moment shows an easy-plane magnetic anisotropy (fig. S7). All of these conclusions are robust to a very large variation of the c-axis lattice parameter, far exceeding the experimentally observed range (Fig. 3).
Following these CEF considerations, we can construct a simple model that quantitatively describes all of the angular magnetoresistance data, based on the assumption that the Nd 4f moment contributes purely to the background magnetization and magnetic permeability of the lattice. For the θ-dependent magnetoresistance, all the nickelates have in common anisotropic orbital depairing (AOD; Fig. 4A) as a natural consequence of the shorter c-axis coherence length (30), and hence an Hc2|| larger than Hc2⊥. This can be seen to dominate at all temperatures for La-/Pr-nickelates (fig. S3). For Nd-nickelates, there is an additional competing term due to the enhanced magnetic permeability (EMP; Fig. 4B) arising from the 4f moments, which has crucially the opposite anisotropy given the easy-plane magnetic anisotropy.
The temperature-dependent balance of this competition (Fig. 4F) underlies the remarkable reversal of the θ-dependent magnetoresistance anisotropy we observe for highly doped Nd-nickelates (Fig. 4G)—i.e., the temperature-dependent crossing of Hc2⊥ and Hc2||. Our minimal model is unexpectedly quantitative; the measured magnetoresistance of all samples (La, Pr, and Nd), doping levels, temperatures, and magnetic fields can be described by one generic form for the resistance RAOD(θ) arising from AOD and another generic form for the resistance REMP(θ) from EMP for Nd, such that the total θ-dependent resistance RPL(θ) can be expressed as
(1) |
Here, C is a constant, and αAOD and αEMP are the temperature- and magnetic field–dependent coefficients of the two functional forms (fig. S8). The dashed lines in Fig. 4G and fig. S4 show the accuracy of this two-component description. In particular, as the polar anisotropy reverses with temperature, the competing AOD and EMP terms produce complex θ-dependences with up to 6 local resistance maxima (fig. S4), which can be well described by Eq. 1.
As for the ϕ-dependent magnetoresistance, we see a competition between a twofold symmetric (C2) ϕ-dependence and a fourfold symmetric (C4) ϕ-dependence (Fig. 4D, E, and H). The C2 response has been seen previously (32, 33) and is a generic result of the ϕ-dependence of the Lorentz force exerted by the measurement current on vortices, because only the perpendicular component contributes. Thus, this ϕ-dependence is seen in all R-variants of the nickelates. However, uniquely in the Nd-nickelates, we observe an additional sin(4ϕ) component (clover-leaf pattern, Fig. 4I), such that the total ϕ-dependent resistance RAZ(ϕ) can be then expressed as
(2) |
Here, C, ϕc2, and ϕc4 are constants, and αc2 and αc4 are the temperature- and magnetic field–dependent coefficients (figs. S5 and S8). We attribute the C4 symmetric dependence again to the 4f moments, because the occupied CEF levels are necessarily invariant to C4 symmetry operations. This implies an in-plane magnetic easy axis along the [110] and equivalent directions. This interpretation is bolstered by the experimental observation that ϕc2 follows the applied current direction, while ϕc4 is locked to the lattice (fig. S9).
The extracted coefficients αAOD, αEMP, αc2, and αc4 exhibit peaks in magnetic field for a given constant temperature (fig. S8), which is a natural result of the boundary conditions: In either the fully superconducting or normal state, no angle-dependent magnetoresistance is seen. With decreasing temperature, both the peaks in αEMP and αAOD shift to higher magnetic fields, tracing the superconducting phase boundary in the temperature-field plane (Fig. 2B). However, the magnitude of αAOD drops substantially with decreasing temperature, contrasting with the increase in αEMP.
DISCUSSION
The pair-breaking framework of superconductivity is useful to discuss these observations. To leading order, the external magnetic field suppresses superconductivity through either orbital or paramagnetic depairing (30). While orbital depairing has an intrinsic θ-dependence rooted in the superconducting anisotropy, paramagnetic depairing is largely isotropic as it suppresses superconductivity by the Zeeman energy cost. Therefore, although both effects are enhanced by the local magnetic field arising from the Nd 4f moments, an underlying θ-dependence of the local magnetic field intensity is more readily visible when paramagnetic depairing is dominant. With this in mind, we can associate αAODRAOD(θ) with orbital depairing, and αEMPREMP(θ) with paramagnetic depairing, which dominates at low temperatures.
In contrast, neither depairing effect has intrinsic ϕ-dependence, and both would faithfully exhibit the structure of the local magnetic field. Therefore, the clover-leaf pattern is a direct manifestation of the ϕ-dependence of the local magnetic field intensity, particularly because αc4 is consistently an order of magnitude larger than αc2 where it is observed (figs. S8 and S10; see the Supplementary Materials for more details on an additional C2 symmetric response). With decreasing temperature, the onset and marked increase of αc4 signal the growing contribution of the 4f moments, which in turn enhances αEMP. Note that while the magnetic response of the 4f moments likely exists over a broader range of temperature and magnetic field, it only appears in the magnetoresistance in the vicinity of Hc2 through pair-breaking effects and, in this process, becomes intertwined with the superconducting response. Therefore, the exact temperature or field dependence of the 4f magnetic moment magnetization cannot be simply extracted from these data. However, the impact of Nd on Hc2 is suggestive of low-temperature Nd 4f antiferromagnetic exchange, as is typically observed for complex materials hosting magnetic R-sites (fig. S11) (39).
With these developments, we lastly return to the differences in Hc2 across R-variants shown in Fig. 1. With important 4f moment enhancements of the field, Nd has the lowest Hc2/Tc0; in their absence, Hc2/Tc0 is highest for La. Pr is intermediate, consistent with the contribution of a Pr3+ 4f moment induced by magnetic field. Thus, we see that the underlying intrinsic behavior of Hc2 appears to be similar across the infinite-layer nickelates, once the additional suppression due to the R-site magnetism has been taken into account. This leads to the conclusion that the conventional weak-coupling Pauli limit Hc2 = 1.86 Tc0 (40) is broadly exceeded, at least with respect to Tc0. Although the (Nd,Sr)NiO2 samples obey this limit (Fig. 1F) (29), we now understand this to occur because of Nd 4f moment contributions. We observe that the Pauli limit is surpassed by up to a factor of 3.5 in the La series (Fig. 1F). Note that we use a 50% resistance criterion for Hc2, to minimize contributions from superconducting fluctuations and vortex creep (29); adopting higher resistance thresholds would make this discrepancy even larger.
While the magnitude of Hc2 presents an important question for future research, we discuss potentially relevant scenarios. We observe that the angular magnetoresistance in the Nd-nickelates can be fully captured by considering only the 4f moment contribution to the background magnetization and magnetic permeability. However, it is important to note that this does not preclude other contributions from the hybridization between the R 5d and the Ni 3d orbitals—in the presence of finite spin-orbit coupling, multiorbital mixing could enhance Hc2. Perhaps the most straightforward explanation is that the experimentally observed Tc0 via transport occurs in the presence of substantial phase fluctuations/disorder and does not reflect the pairing scale Δ0. Tunneling spectroscopy indicates a large Δ0-to-Tc0 ratio, although surface effects and variations remain to be clarified (41). Further evidence for this scenario may be found in a recent study of La0.8Sr0.2NiO2 with higher Tc0 but comparable extrapolations of Hc2 to what we observe (42). The fact that experimentally, nickelate superconductivity is in the dirty limit appears to us to disfavor more exotic mean-field enhancements. Independent of origin, the unusually large Hc2/Tc0 ratios in the nickelates, combined with sensitive control through the R-site, give properties that would be useful for high-field applications. Provided that substantial further advances can be made in materials synthesis, nickelates could be used for superconducting wires and components, particularly in chemically reducing environments (43).
MATERIALS AND METHODS
Materials
Thin-film nickelate samples (7 nm thick) were grown on SrTiO3 substrates using pulsed-laser deposition under growth and reduction conditions previously reported (1, 2, 3, 44).
Methods
For angular magnetoresistance measurements, the samples were contacted using wire-bonded aluminum wires in a 6-point Hall bar geometry. The 50% resistive criterion is applied to both temperature-sweep and field-sweep data to determine Tc0 and Hc2. For high-field measurements, transport measurements were performed using the standard four-probe lock-in technique, with an alternating current excitation between 1 and 10 μA. Continuous magnetic fields up to 35 T were generated using a Bitter magnet, coupled with a He-3 cryostat, at the High Field Magnet Laboratory at Nijmegen, the Netherlands. The magnetic field angle was varied in situ using a customized rotator stage.
Note
During the preparation and review of this manuscript, we became aware of two reports on Hc2 in La-nickelates (42, 45) and three reports on Nd-nickelates (46–48). All the reported behavior is qualitatively consistent with our results, independent of variations in the magnitude of the resistivity (see also fig. S12), and they further demonstrate that the presence or absence of a capping layer has minimal impact on the observed angular magnetoresistance. We also note that our crystal field assignment for the doped Nd system is very consistent with recent bulk neutron diffraction measurements (49).
Acknowledgments
We thank S. A. Kivelson, C. Murthy, S. Raghu, and Y. Yu for fruitful discussions and Y. Yu and S. P. Harvey for critical reading of the manuscript. We acknowledge the support of the HFML-RU/NWO, a member of the European Magnetic Field Laboratory (EMFL). IRB and/or IACUC guidelines were followed with human or animal subjects.
Funding: This work was supported by the US Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering grant DE-AC02-76SF00515 (to B.Y.W., T.C.W., M.O., K.L., C.J., D.L., T.P.D., I.R.F., and H.Y.H.); Gordon and Betty Moore Foundation’s Emergent Phenomena in Quantum Systems Initiative grant GBMF9072 (to J.F. and H.Y.H.); the Swiss National Science Foundation through Postdoc.Mobility P400P2199297 (to J.F.); and the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement no. 835279-Catch-22) (to Y.-T.H., C.D., and N.E.H.).
Author contributions: Conceptualization: B.Y.W. and H.Y.H. Sample synthesis: M.O., K.L., D.L., and J.F. Transport characterization: B.Y.W., T.C.W., Y.-T.H., and C.D. Theoretical calculation: B.Y.W. and C.J. Funding acquisition: H.Y.H., T.P.D., and N.E.H. Supervision: M.R.B., T.P.D., N.E.H., I.R.F., and H.Y.H. Writing: B.Y.W., T.C.W., and H.Y.H.
Competing interests: The authors declare that they have no competing interests.
Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.
Supplementary Materials
This PDF file includes:
REFERENCES AND NOTES
- 1.Li D., Lee K., Wang B. Y., Osada M., Crossley S., Lee H. R., Cui Y., Hikita Y., Hwang H. Y., Superconductivity in an infinite-layer nickelate. Nature 572, 624–627 (2019). [DOI] [PubMed] [Google Scholar]
- 2.Osada M., Wang B. Y., Lee K., Li D., Hwang H. Y., Phase diagram of infinite layer praseodymium nickelate Pr1-xSrxNiO2 thin films. Phys. Rev. Mater. 4, 121801(R) (2020). [Google Scholar]
- 3.Osada M., Wang B. Y., Goodge B. H., Harvey S. P., Lee K., Li D., Kourkoutis L. F., Hwang H. Y., Nickelate superconductivity without rare-earth magnetism: (La,Sr)NiO2. Adv. Mater. 33, 2104083 (2021). [DOI] [PubMed] [Google Scholar]
- 4.Zeng S. W., Li C. J., Chow L. E., Cao Y., Zhang Z. T., Tang C. S., Yin X. M., Lim Z. S., Hu J. X., Yang P., Ariando A., Superconductivity in infinite-layer nickelate La1-xCaxNiO2 thin films. Sci. Adv. 8, eabl9927 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Zhao D., Zhou Y. B., Fu Y., Wang L., Zhou X. F., Cheng H., Li J., Song D. W., Li S. J., Kang B. L., Zheng L. X., Nie L. P., Wu Z. M., Shan M., Yu F. H., Ying J. J., Wang S. M., Mei J. W., Wu T., Chen X. H., Intrinsic spin susceptibility and pseudogaplike behavior in infinite-layer LaNiO2. Phys. Rev. Lett. 126, 197001 (2021). [DOI] [PubMed] [Google Scholar]
- 6.Zhang J., Pajerowski D. M., Botana A. S., Zheng H., Harriger L., Rodriguez-Rivera J., Ruff J. P., Schreiber N. J., Wang B., Chen Y. S., Chen W. C., Norman M. R., Rosenkranz S., Mitchell J. F., Phelan D., Spin stripe order in a square planar trilayer nickelate. Phys. Rev. Lett. 122, 247201 (2019). [DOI] [PubMed] [Google Scholar]
- 7.Lin J. Q., Arribi P. V., Fabbris G., Botana A. S., Meyers D., Miao H., Shen Y., Mazzone D. G., Feng J., Chiuzbăian S. G., Nag A., Walters A. C., García-Fernández M., Ke-Jin Zhou J., Pelliciari I., Jarrige J. W., Freeland J. Z., Mitchell J. F., Bisogni V., Liu X., Norman M. R., Dean M. P. M., Strong superexchange in a d9−δ Nickelate revealed by resonant inelastic X-ray scattering. Phys. Rev. Lett. 126, 087001 (2021). [DOI] [PubMed] [Google Scholar]
- 8.Pan G. A., Ferenc Segedin D., LaBollita H., Song Q., Nica E. M., Goodge B. H., Pierce A. T., Doyle S., Novakov S., Córdova Carrizales D., N’Diaye A. T., Shafer P., Paik H., Heron J. T., Mason J. A., Yacoby A., Kourkoutis L. F., Erten O., Brooks C. M., Botana A. S., Mundy J. A., Superconductivity in a quintuple-layer square-planar nickelate. Nat. Mater. 21, 160–164 (2022). [DOI] [PubMed] [Google Scholar]
- 9.Hundley M. F., Thompson J. D., Cheong S.-W., Fisk Z., Oseroff S. B., Specific heat and anisotropic magnetic susceptibility of Pr2CuO4, Nd2CuO4 and Sm2CuO4 crystals. Phys. C Supercond. 158, 102–108 (1989). [Google Scholar]
- 10.Lynn J. W., Sumarlin I. W., Skanthakumar S., Li W.-H., Shelton R. N., Peng J. L., Fisk Z., Cheong S.-W., Magnetic ordering of Nd in (Nd,Ce)2CuO4. Phys. Rev. B 41, 2569–2572 (1990). [DOI] [PubMed] [Google Scholar]
- 11.Fehrenbacher R., Rice T. M., Unusual electronic structure of PrBa2Cu3O7. Phys. Rev. Lett. 70, 3471–3474 (1993). [DOI] [PubMed] [Google Scholar]
- 12.Norman M. R., Entering the nickel age of superconductivity. Phys. Ther. 13, 85 (2020). [Google Scholar]
- 13.Pickett W. E., The dawn of the nickel age of superconductivity. Nat. Rev. Phys. 3, 7–8 (2021). [Google Scholar]
- 14.Botana A. S., Bernardini F., Cano A., Nickelate superconductors: An ongoing dialog between theory and experiments. J. Exp. Theor. Phys. 132, 618–627 (2021). [Google Scholar]
- 15.Nomura Y., Arita R., Superconductivity in infinite-layer nickelates. Rep. Prog. Phys. 85, 052501 (2022). [DOI] [PubMed] [Google Scholar]
- 16.Lechermann F., Late transition metal oxides with infinite-layer structure: Nickelates versus cuprates. Phys. Rev. B 101, 081110 (2020). [Google Scholar]
- 17.Li D., Wang B. Y., Lee K., Harvey S. P., Osada M., Goodge B. H., Kourkoutis L. F., Hwang H. Y., Superconducting dome in Nd1-xSrxNiO2 infinite layer films. Phys. Rev. Lett. 125, 027001 (2020). [DOI] [PubMed] [Google Scholar]
- 18.Zeng S., Tang C. S., Yin X., Li C., Li M., Huang Z., Hu J., Liu W., Omar G. J., Jani H., Lim Z. S., Han K., Wan D., Yang P., Pennycook S. J., Wee A. T. S., Ariando A., Phase diagram and superconducting dome of infinite-layer Nd1-xSrxNiO2 thin films. Phys. Rev. Lett. 125, 147003 (2020). [DOI] [PubMed] [Google Scholar]
- 19.Xiang Y., Li Q., Li Y., Yang H., Feng N., Wen H. -H., Physical properties revealed by transport measurements for superconducting Nd0.8Sr0.2NiO2 thin films. Chin. Phys. Lett. 38, 047401 (2021). [Google Scholar]
- 20.Hsu Y. -T., Wang B. Y., Berben M., Li D., Lee K., Duffy C., Ottenbros T., Kim W. J., Osada M., Wiedmann S., Hwang H. Y., Hussey N. E., Insulator-to-metal crossover near the edge of the superconducting dome in Nd1-xSrxNiO2. Phys. Rev. Res. 3, L042015 (2021). [Google Scholar]
- 21.Wang N. N., Yang M. W., Yang Z., Chen K. Y., Zhang H., Zhang Q. H., Zhu Z. H., Uwatoko Y., Gu L., Dong X. L., Sun J. P., Jin K. J., Cheng J.-G., Pressure-induced monotonic enhancement of Tc to over 30 K in superconducting Pr0.82Sr0.18NiO2 thin films. Nat. Commun. 13, 4367 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Hepting M., Dean M. P. M., Lee W. S., Soft X-ray spectroscopy of low-valence nickelates. Front. Phys. 9, 808683 (2021). [Google Scholar]
- 23.Goodge B. H., Li D., Lee K., Osada M., Wang B. Y., Sawatzky G. A., Hwang H. Y., Kourkoutis L. F., Doping evolution of the Mott–Hubbard landscape in infinite-layer nickelates. Proc. Natl. Acad. Sci. U.S.A. 118, e2007683118 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Rossi M., Lu H., Nag A., Li D., Osada M., Lee K., Wang B. Y., Agrestini S., Garcia-Fernandez M., Kas J. J., Chuang Y.-D., Shen Z. X., Hwang H. Y., Moritz B., Zhou K.-J., Devereaux T. P., Lee W. S., Orbital and spin character of doped carriers in infinite-layer nickelates. Phys. Rev. B 104, L220505 (2021). [Google Scholar]
- 25.Jiang P., Si L., Liao Z., Zhong Z., Electronic structure of rare-earth infinite-layer RNiO2 (R = La, Nd). Phys. Rev. B 100, 201106 (2019). [Google Scholar]
- 26.Bandyopadhyay S., Adhikary P., Das T., Dasgupta I., Saha-Dasgupta T., Superconductivity in infinite-layer nickelates: Role of f orbitals. Phys. Rev. B 102, 220502 (2020). [Google Scholar]
- 27.Choi M. -Y., Lee K. -W., Pickett W. E., Role of 4 f states in infinite-layer NdNiO2. Phys. Rev. B 101, 020503 (2020). [Google Scholar]
- 28.Zhang R., Lane C., Singh B., Nokelainen J., Barbiellini B., Markiewicz R. S., Bansil A., Sun J., Magnetic and f-electron effects in LaNiO2 and NdNiO2 nickelates with cuprate-like 3dx2-y2 band. Commun. Phys. 4, 118 (2021). [Google Scholar]
- 29.Wang B. Y., Li D., Goodge B. H., Lee K., Osada M., Harvey S. P., Kourkoutis L. F., Beasley M. R., Hwang H. Y., Isotropic Pauli-limited superconductivity in the infinite-layer nickelate Nd0.775Sr0.225NiO2. Nat. Phys. 17, 473–477 (2021). [Google Scholar]
- 30.M. Tinkham, Introduction to Superconductivity (McGraw-Hill Inc., ed. 2, 1996). [Google Scholar]
- 31.Maki K., Tsuneto T., Pauli paramagnetism and superconducting state. Prog. Theor. Phys. 31, 945–956 (1964). [Google Scholar]
- 32.Iye Y., Fukushima A., Tamegai T., Terashima T., Bando Y., Anisotropy and dissipation in the mixed state of high temperature superconductors. Phys. C Supercond. 185–189, 297–302 (1991). [Google Scholar]
- 33.Iye Y., Watanabe A., Nakamura S., Tamegai T., Terashima T., Yamamoto K., Bando Y., Dissipation in the mixed state of conventional and high temperature superconductors. Phys. C Supercond. 167, 278–286 (1990). [Google Scholar]
- 34.Baloch A. A. B., Alqahtani S. M., Mumtaz F., Muqaibel A. H., Rashkeev S. N., Alharbi F. H., Extending Shannon’s ionic radii database using machine learning. Phys. Rev. Mater. 5, 043804 (2021). [Google Scholar]
- 35.Perdew J. P., Burke K., Ernzerhof M., Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). [DOI] [PubMed] [Google Scholar]
- 36.Kresse G., Hafner J., Ab initio molecular dynamics for liquid metals. Phys. Rev. B 47, 558–561 (1993). [DOI] [PubMed] [Google Scholar]
- 37.Tang W., Sanville E., Henkelman G., A grid-based Bader analysis algorithm without lattice bias. J. Phys. Condens. Matter 21, 084204 (2009). [DOI] [PubMed] [Google Scholar]
- 38.Lin H., Gawryluk D. J., Klein Y. M., Huangfu S., Pomjakushina E., von Rohr F., Schilling A., Universal spin-glass behaviour in bulk LaNiO2, PrNiO2 and NdNiO2. New J. Phys. 24, 013022 (2022). [Google Scholar]
- 39.Buzdin A. I., Bulaevskiĭ L. N., Antiferromagnetic superconductors. Sov. Phys. Uspekhi 29, 412–425 (1986). [Google Scholar]
- 40.Clogston A. M., Upper limit for the critical field in hard superconductors. Phys. Rev. Lett. 9, 266–267 (1962). [Google Scholar]
- 41.Gu Q., Li Y., Wan S., Li H., Guo W., Yang H., Li Q., Zhu X., Pan X., Nie Y., Wen H. -H., Single particle tunneling spectrum of superconducting Nd1-xSrxNiO2 thin films. Nat. Commun. 11, 6027 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.W. Sun, Y. Li, R. Liu, J. Yang, J. Li, S. Yan, H. Sun, W. Guo, Z. Gu, Y. Deng, X. Wang, Y. Nie, Evidence for quasi-two-dimensional superconductivity in infinite-layer nickelates. arXiv:2204.13264 [cond-mat.supr-con] (28 April 2022).
- 43.Xu X., A review and prospects for Nb3Sn superconductor development. Supercond. Sci. Tech. 30, 093001 (2017). [Google Scholar]
- 44.Lee K., Goodge B. H., Li D., Osada M., Wang B. Y., Cui Y., Kourkoutis L. F., Hwang H. Y., Aspects of the synthesis of thin film superconducting infinite-layer nickelates. APL Mater. 8, 041107 (2020). [Google Scholar]
- 45.L. E. Chow, K. Y. Yip, M. Pierre, S. W. Zeng, Z. T. Zhang, T. Heil, J. Deuschle, P. Nandi, S. K. Sudheesh, Z. S. Lim, Z. Y. Luo, M. Nardone, A. Zitouni, P. A. van Aken, M. Goiran, S. K. Goh, W. Escoffier, A. Ariando, Pauli-limit violation in lanthanide infinite-layer nickelate superconductors. arXiv:2204.12606 [cond-mat.supr-con] (26 April 2022).
- 46.Krieger G., Raji A., Schlur L., Versini G., Bouillet C., Lenertz M., Robert J., Gloter A., Viart N., Preziosi D., Synthesis of infinite-layer nickelates and influence of the capping-layer on magnetotransport. J. Phys. D App. Phys. 56, 024003 (2022). [Google Scholar]
- 47.H. Ji, Y. Li, Y. Liu, X. Ding, Z. Xie, S. Qi, L. Qiao, Y. Yang, G.-M. Zhang, J. Wang, Rotational symmetry breaking in superconducting nickelate Nd0.8Sr0.2NiO2 films. arXiv:2210.17023. [cond-mat.supr-con] (31 October 2022).
- 48.L. E. Chow, K. Rubi, K. Y. Yip, M. Pierre, M. Leroux, X. Liu, Z. Luo, S. Zeng, C. Li, M. Goiran, N. Harrison, W. Escoffier, S. K. Goh, A. Ariando, Dimensionality control and rotational symmetry breaking superconductivity in square-planar layered nickelates. arXiv:2301.07606 [cond-mat.supr-con] (18 January 2023).
- 49.S. Rosenkranz, D. Phelan, B. Wang, H. Zheng, D. L. Abernathy, A. S. Botana, J. F. Mitchell, Inelastic neutron scattering study of the infinite layer nickelate Nd0.8Sr0.2NiO2. Bull. Am. Phys. Soc., (2023). https://meetings.aps.org/Meeting/MAR23/Session/S19.10
- 50.Smith J. A., Cima M. J., Sonnenberg N., High critical current density thick MOD-derived YBCO films. IEEE Trans. Appl. Supercond. 9, 1531–1534 (1999). [Google Scholar]
- 51.M. T. Hutchings, Point-charge calculations of energy levels of magnetic ions in crystalline electric field, in Solid State Physics, F. Seitz, D. Turnbull, Eds. (Elsevier Amsterdam, 1964), vol. 16, pp. 227–273. [Google Scholar]
- 52.Yu M., Trinkle D. R., Accurate and efficient algorithm for Bader charge integration. J. Chem. Phys. 134, 064111 (2011). [DOI] [PubMed] [Google Scholar]
- 53.Gelessus A., Thiel W., Weber W., Multipoles and symmetry. J. Chem. Educ. 72, 505–508 (1995). [Google Scholar]
- 54.K. Lee, B. Y. Wang, M. Osada, B. H. Goodge, T. C. Wang, Y. Lee, S. Harvey, W. J. Kim, Y. Yu, C. Murthy, S. Raghu, L. F. Kourkoutis, H. Y. Hwang, Character of the "normal state" of the nickelate superconductors. arXiv:2203.02580v1 [cond-mat.supr-con] (4 March 2022).
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.