Abstract
Recent observations of superconductivity in infinite-layer nickelates offer insights into high-temperature superconductivity mechanisms. However, defects and dislocations in doped films complicate the realization of superconductivity, limiting current research to superconducting nickelate heterostructures. The lack of research on superconductivity in superlattices composed of ultrathin nickelates severely impedes not only the exploration of the interface effect on superconductivity, but also the utilization of heterostructure engineering for exploring higher superconducting temperature Tc. Here, we demonstrated superconducting infinite-layer nickelate superlattices [(Nd0.8Sr0.2NiO2)8/(SrTiO3)2]10 via topotactic reduction. Our study uncovered that only above a critical thickness can high-quality superlattices be achieved, with structural formation dependent on nickelate layer thickness. The superconducting superlattice showed a Tc of 12.5 K and a 2D superconducting feature, indirectly indicate the intrinsic superconductivity of infinite-layer nickelates. Our study offers promising avenues for delving into the superconducting mechanism and for exploring multilevel interface engineering of infinite-layer nickelates, thus opening new horizons for the study of infinite-layer nickelates.
Subject terms: Superconducting properties and materials; Surfaces, interfaces and thin films
So far infinite-layer nickelates have only been observed to superconduct in single-thin-film form. Here, the authors demonstrate superconductivity in [(Nd0.8Sr0.2NiO2)8/(SrTiO3)2]10 superlattices.
Introduction
Since the discovery of high-temperature superconductivity (HTSC) in copper oxides in 19861, there has been an increasing interest in searching for similar unconventional HTSC in other systems. In 2019, Li et al. firstly realized 9–15 K superconductivity in the infinite-layer Nd0.8Sr0.2NiO2 thin films by soft-chemistry topotactic reduction2. Subsequently, some extended rare-earth nickelate superconducting films were discovered3–7. The recent discovery of superconductivity in pressure-induced La3Ni2O78 and La4Ni3O109 single crystals has positioned nickelates as a novel addition to the HTSC family. Infinite-layer nickelates and cuprates share many similarities, such as two-dimensional infinite-layer structure, superconducting dome determined by doping level3,4,10,11, d-wave systems12,13, and band dispersion14,15, etc. However, nickelates show some intrinsic differences. For parent compounds, nickelates are closer to Mott-Hubbard feature, while cuprates are in the charge-transfer regime of the Zaanen–Sawatzky–Allen scheme16,17. The nickelate parent compounds exhibit a metallic ground state2, contrasting the mott insulating state of cuperates18. Other features, including the absence of long-range antiferromagnetic orders19,20 and multiorbital physics21–23 are also different from that in cuprates. Nevertheless, employing comparable methods utilized in the study of cuprates to investigate the underlying mechanism of superconductivity in infinite layers of nickelates can still be a valuable approach.
Interface engineering serves as a potent method for modulating the nature of superconductivity. Induced or enhanced superconductivity has been observed at various interfaces, including LaAlO3/SrTiO3 2DEGs24, La1.55Sr0.45CuO4/La2CuO425, FeSe/SrTiO326 and CaCuO2/SrCuO2 superlattices27. Therefore, it promotes the puzzles whether interfaces play a role in influencing the superconductivity in infinite-layer nickelates. However, superconductivity in infinite-layer nickelates can currently only be achieved in a single film of finite thickness. Some efforts had been made in infinite-layer LaNiO2+x/LaGaO328 and NdNiO2/SrTiO329 superlattices but no superconducting feature was found. One contributing factor could be the formation of a mixed Ruddlesden–Popper (RP) secondary phase during the growth process30. Additionally, the de-embedding of oxygen31 and the presence of hydrogen32 during the reduction process also affect superconductivity. Furthermore, significant interfacial polarization and surface reconstruction occur at the interface between the polar infinite-layer nickelates and the non-polar oxides29,33,34. All of these issues make the realization of high-quality superconducting infinite-layer nickelate superlattices and the interface modulation very challenging. Nevertheless, the achievement of superlattice superconductivity and the exploration of its distinctions from single thin films remain crucial for advancing our understanding of the mechanisms underlying interfacial interactions in the superconductivity of infinite-layer nickelates thin films.
Here, we demonstrated the realization of superconductivity in (Nd0.8Sr0.2NiO2)n/(SrTiO3)m superlattices. By reducing the pristine (Nd0.8Sr0.2NiO3)n/(SrTiO3)m superlattices which were grown by pulsed laser deposition with CaH2, we found that the as-reduced films exhibit structure changes which are dependent on the thickness of nickelate layers. In (Nd0.8Sr0.2NiO2)3/(SrTiO3)2 superlattice with a thickness below a critical value of 5 u.c., mixed phases were observed, while the 8 u.c. nickelates in (Nd0.8Sr0.2NiO2)8/(SrTiO3)2 superlattice exhibit a well-defined infinite-layer structure. Only the superlattices with infinite-layer nickelates exhibit superconductivity with an onset transition temperature Tc of 12.5 K, whereas the mixed phase superlattices show insulating behaviors. Detailed experimental investigations of transport properties show that the superconducting superlattice displays a 2D superconducting feature with anisotropic cahracteristic35–39. The consistent characterization with a single film suggests that the interface has little effect on its superconducting properties, indirectly demonstrating that the superconductivity of infinite-layer nickelate is intrinsic.
Results
Thickness driven structure change in superlattices
Nickelate superlattices (SL) [(Nd0.8Sr0.2NiO3)n/(SrTiO3)2]10 (referred as Nn/S2) were initially deposited on SrTiO3 substrates by pulsed laser deposition (PLD) and subsequently fully reduced using CaH2, as detailed in the methods section. The as-reduced SLs are denoted as R-Nn/S2. The thicknesses of the pristine and as-reduced SLs thicknesses are controlled by the in-situ reflection high-energy electron diffraction and further confirmed by fitting X-ray reflectivity (XRR) (Fig. 1b and Supplementary Fig. S8c). The crystal structures were characterized by X-ray diffraction (XRD) θ–2θ symmetric scans, as depicted in Fig. 1a. The as-grown N3/S2 (black) and N8/S2 (blue) SLs exhibit distinct superlattice features with satellite peaks, indicating the absence of secondary phase. Reciprocal space mapping (RSM) corresponding to SrTiO3 (103) reflections in Fig. 1c and Supplementary Fig. S8a confirm coherent strain of the SLs on the SrTiO3 substrates. Even after an extended reduction process, the XRD patterns retain clear SL main peaks and satellite peaks. The RSMs in Fig. 1d and Supplementary Fig. S8b demonstrate that the as-reduced SLs still remain fully strained to SrTiO3 substrates.
Compared to the pristine SLs, the peaks of R-N8/S2 (green) exhibit a rightward shift towards higher values, indicating the formation of infinite-layer Nd0.8Sr0.2NiO22. The out-of-plane lattice constant c(R-N8/S2) is extracted to be 3.34 Å according to the main peak position. In contrast, the rightward shift value for R-N3/S2 (red) is much smaller than that of R-N8/S2, with a peak position of 49.75°, corresponding to c(R-N3/S2) = 3.66 Å. This observation is reminiscent of findings in SrCuO2 SLs, where discrepancies between plane-type and chain-type crystal structures were noted40,41. However, the chain-type LaNiO2 shows a larger c-axis constant than the SrTiO3 substrate42, contrary to our XRD data, suggesting that the chain-type structure does not account for the difference in lattice constants. Nevertheless, the presence of the SrTiO3 layer in the N3/S2 SL significantly hinders the reduction condition of the embedded Nd0.8Sr0.2NiO2 layers, which was not observed in N8/S2 SL. Therefore, question will arise that is there a critical thickness in Nd0.8Sr0.2NiO2 SL for the structure change, similar to that occurring in SrCuO2 SL41,43 ? Our preliminary data indeed indicates evidence of nickelate thickness-driven structural changes and the existence of a critical thickness of approximately ~5 u.c. (see Supporting Information, Supplementary Fig. S3). Below and above 5 u.c., the superlattice exhibit significantly different lattice constants. What is more, the change in lattice structure is not attributed to inadequate reduction processes (see the detailed discussion in the Supporting Information).
Figure 1e shows the temperature-dependent resistivity of pristine N3/S2, N8/S2 and as-reduced R-N3/S2, R-N8/S2. Both the as-grown N3/S2 and N8/S2 exhibit metallic behavior, consistent with previous report2. After reduction, R-N8/S2 displays metallic temperature dependence above 40 K and transitions to a superconducting state with an onset temperature of 12.5 K (Tc,onset), reaching zero resistance at 5.5 K (Tc,0). This transition temperature aligns with previous observations in 10 nm Nd0.8Sr0.2NiO2/SrTiO3 single films2. It is noteworthy that our superlattice allows a thinner thickness of the nickelate while still maintaining a high Tc value, whereas for thinner single films, the Tc value decreases monotonically with thickness (for a single Nd0.8Sr0.2NiO2/SrTiO3 film of 4.6 nm, i.e., ~14 uc, the Tc,onset is only ~6.5 K)44. Conversely, R-N3/S2 demonstrates insulating transport behavior across the entire temperature range from 30 to 300 K (refer to the inset in Fig. 1b). Below 30 K, the resistance measurement falls outside of our electronic measurement range.
The structural differences between the two superlattices were further investigated using scanning transmission electron microscopy (STEM) techniques. Figures 2a and 2e schematically shows the superlattice stacking sequence. For R-N3/S2, the first layer is of 4 u.c. Nd0.8Sr0.2NiO2, which was used to reflect the evolution for structure change. The high annular angle dark field (HAADF) STEM images for large region of R-N3/S2 and R-N8/S2 are shown in Supplementary Fig. S9, which reveal very sharp interfaces between Nd0.8Sr0.2NiO2 and SrTiO3 in the as-reduced superlattices. In both R-N3/S2 and R-N8/S2, only very few RP defects were observed, demonstrating the stabilizing effect of SrTiO3 interfacial layer on nickelate monostructures, which may be the reason that our superlattices have higher Tc values compared to single films of similar thicknesses. The atomic structures of the as-reduced films are shown by zoom in HAADF-STEM images (see Fig. 2b, f). For the bright field image, the bright intensity of spots in an atomically resolved image is positively correlated with the square of its atomic number Z (). In this way, light and heavy atoms can be clearly distinguished according to their image contrast. It can be seen that the Nd0.8Sr0.2NiO2 and SrTiO3 layers grow epitaxially with respect to each other along the [001] growth direction.
For as-reduced nickelate layers, oxygen de-embedding is a noteworthy issue. The iDPC-STEM is an effective method to avoid electron damage while preserving the lining of both light and heavy atoms, and has greater advantages over the HAADF-STEM technique in studying the distribution and defects of oxygen atoms45. Atomically resolved iDPC-STEM images corresponding to the HAADF-STEM images are shown in Fig. 2c, g. The oxygen atoms in <TiO6> octahedron of SrTiO3 and qusi−2D infinite-layer NiO2 plane of Nd0.8Sr0.2NiO2 are clearly identified. Local atomic distributions are highlighted with different colored spheres. In the case of R-N8/S2 shown in Fig. 2g, j, it is evident that the infinite-layer Nd0.8Sr0.2NiO2 alternates with the perovskite SrTiO3, and a collapse of the Nd0.8Sr0.2NiO2 unit cell along the c-axis direction is observable. This observation suggests that the nickelate layers, when capped by the SrTiO3 intermediate layer, are more prone to lose apical oxygen atoms during hydrogenation-chemical reduction, leading to the collapses into an infinite-layer structure, as previously reported2,46. Instead, only partial apical oxygen atoms are de-embedding in R-N3/N2 (Fig. 2c, i), resulting in a disordered mixture of planar-type infinite-layer, pyramidal and octahedron structures. Such disordering structure is also observed in 4 u.c. Nd0.8Sr0.2NiO2 layer which was deliberately inserted first into N3/N2 SL to obtain more thickness effect microstructure. It can be concluded that STEM also revealed same critical thickness for structure change from disordering structure to ordering infinite layer structure. This disorder increases as the nickelate layer approaches the SrTiO3 interlayer at interface, where removing apical oxygen becomes more difficult. Such disordered distribution and inhomogeneous structure may be responsible for its transformation into an insulator, which was also observed in partially reduced Pr0.8Sr0.2NiO2+x34. In addition, Nd-O layers are observed at the Nd0.8Sr0.2NiO2/SrTiO3 interface in both two SLs, serving to alleviate the strong charge field induced by the polar interface33.
Figures 2d and 2h illustrates the profiles of Nd-Nd out-of-plane distance across R-N3/S2 and R-N8/S2 superlattices derived from HAADF images. The data are averaged over seven vertical profiles. The Nd-Nd out-of-plane distance represents the lattice constant of the c-axis per unit cell. It can be seen that the c-axis lattice parameters of the SrTiO3 interlayers in both two SLs are distributed around 0.39 nm, which is consistent with the SrTiO3 substrate. The uniform in-plane dimensions indicate that the SrTiO3 interlayers maintain a cubic perovskite structure, unaffected by prolonged topotactic reduction. For the nickelates layers, the c-axis lattice constants of R-N3/S2 fluctuate around 0.36 nm, whereas it is 0.33 nm in R-N8/S2, aligning with the values obtained from XRD patterns. The characterizations of crystal structure based on the XRD and STEM reveal that the presence of residual apical oxygen directly hinders the reduction of the out-of-plane lattice constant of nickelate layers. Furthermore, local lattice extensions of 0.4 nm are observed at the SrTiO3 layer interfaces, which is attributed to the mixing stoichiometry of Ti and Ni in the interface layer33.
Large orbital polarizations are frequently observed in complex oxides with quasi−2D structures, including high-Tc cuprates like YBa2Cu3O747, SrCuO2 heterostructures40,41 and nickelates featuring infinite-layer configurations17,28,48. Such orbital polarization can be explored by measurement of linearly polarized X-ray absorption spectroscopy. In Fig. 3a, the schematic diagram illustrates the X-ray linear dichroism (XLD) measurement using two light polarizations. The intensity with electric vector E along the in-plane polarization (E//ab) corresponds to the occupation of orbitals, while the out-of-plane polarization (E//c) reflects the occupation of orbitals. As shown in Fig. 3b, we compare the polarization dependence of the XAS spectra at the Ni L2,3-edge. Firstly, XAS for R-N3/S2 and R-N8/S2 both show a main absorption peak closely resembling Ni1+ in NdNiO217. The Ni L3-edge of R-N3/S2 shows a higher peak energy (~0.2 eV) than R-N8/S2, suggesting a higher Ni valence state in R-N3/S2, corresponding to the fewer oxygen vacancies in R-N3/S2. Moreover, significant differences are observed in the XAS spectra along different polarizations in both SLs, highlighting the preferred orbital occupation in both infinite-layers. The XLD intensity (defined by IE//ab - IE//c) shown in Fig. 3c reveals strong positive dichroism in both SLs, which can be attributed to the unoccupied orbital in the low spin state picture48. It indicates that the orbitals along z-axis are more occupied in both mixed-structure R-N3/S2 and infinite-layer R-N8/S2. This phenomenon is attributed to the lower Ni-O bonding energy of apical oxygen in <NiO6> octahedra42, facilitating square-planar coordination of Ni atoms after topochemical reduction. On this basis, the crystal field splitting in the eg orbitals is greatly enhanced, leading to the lower energy levels and higher electron occupancy in the orbitals compared to the orbitals, thereby inducing significant orbital polarization. Notably, unlike the findings in SrCuO2 superlattices40,41, no significant thickness-dependent XLD intensity differences were observed in the infinite-layer nickelate superlattices. This observation further supports the absence of chain-like structures in R-N3/S2, as evidenced by STEM images and XRD patterns.
Superconducting properties of [(Nd0.8Sr0.2NiO2)8/(SrTiO3)2]10 superlattice
The superconducting properties of R-N8/S2 were further investigated by measuring the magnetoresistance. Figures 4a and 4b display the temperature-dependent magnetoresistance (MR) under the out-of-plane (μ0Hc//c) and in-plane (μ0Hc//ab) magnetic fields ranging from 0 to 9 T. Little magnetoresistance is observed in the normal state, whereas the superconducting state at lower temperature is significantly suppressed by the magnetic field. Moreover, the effect of magnetic field orientations on superconducting suppression varies considerably. The ρ(T) curves reveal a denser behavior when the magnetic field is applied transversely. Similar distinctions are observed in the magnetic-field-dependent MR shown in Supplementary Fig. S11a, S11b. These observations indicate the anisotropy in the superconductivity of the R-N8/S2 superlattice.
To further comprehend the superconductivity anisotropy in R-N8/S2, we extracted the variation of the upper critical field Hc2 with Tc,mid, where the resistivity reaches 50% of the normal state (Fig. 4d). Figure 4c illustrates the critical field in both directions versus the temperature point near Tc,0 (μ0Hc = 0). For μ0Hc//c, the upper critical field follows a linear relationship with temperature near Tc,0, while a (Tc-T)1/2-dependence is followed for μ0Hc//ab. This is commonly observed in two-dimensional superconductors49–51. It implies the possible 2D superconducting characteristic of R-N8/S2, which is consistent with the square-planar NiO2 plane geometry. This behavior can be well described by the 2D Ginzburg-Landau (G-L) formula52 as
1 |
2 |
where is the flux quantum, ξab(0) is the in-plane G-L coherence length in the zero-temperature and is the superconducting thickness. Using the above equations to fit the data, we derived = 48.7 Å, and = 223.5 Å, which are consistent with previous reports on Nd0.775Sr0.225NiO2 films35. Furthermore, both the in-plane and the out-of-plane upper critical fields remain below the Bardeen–Cooper–Schrieffer Pauli limit, as shown by the fitted curves (H Pauli, μ = 1 μB = 1.86 × Tc,0).
To confirm the two-dimensional superconducting feature of R-N8/S2, we measured the angular dependence of Hc2 of R-N8/S2 at 5 K. Figure 4e shows the MR curves at different θ values, and the inset shows a configuration for measurements. By extracting the values of the critical field at different angles, the angular-dependent Hc2(θ) curves can be obtained, as shown in Fig. 4f. A clear cusp-like peak can be observed when θ approach 90° (μ0H//ab), which can be described by the 2D-Tinkham formula for 2D superconductors (red solid line)50, which is expressed as
3 |
In contrast, the peak cannot be reproduced by 3D anisotropic G-L model (blue dotted line), which is expressed as
4 |
where anisotropy ratio . These results qualitatively indicate the 2D feature of R-N8/S2. However, unlike the large of conventional 2D superconductors50, the of R-N8/S2 approaches 1 at low temperature. This behavior may represent a unique characteristic of nickelate thin-film superconductors, which was also found in Nd0.8Sr0.2NiO2/SrTiO3 single films with 2D superconducting feature38,53.
Figure 5a presents the current-voltage (I-V) characteristics at different temperatures. A zero-voltage flat state in the I-V characteristic curves is observed at low temperatures, which indicates the presence of a superconducting state. As temperature increases towards the normal state, the plateau disappears and is changed to a linear characteristic. For superconducting states, the increasing bias current would induce a superconducting transition at a critical value, i.e., the critical current value Ic. The evolution of Ic values from 2 K-5 K summarized in Fig. 5b shows robust current tolerance at the milliamp level in R-N8/S2. Furthermore, the Berezinskii-Kosterlitz-Thouless (BKT) transition was studied in R-N8/S2. This transition is a notable feature in 2D superconductors, indicating a shift from unpaired vortices and antivortices to bound vortex–antivortex pairs54. The BKT transition appears as a power-law exponential jump at the zero-current limit of the current-voltage characteristic curves, with the BKT temperature (TBKT) typically defined as the temperature where V∝I3. Figure 5c illustrates the I-V characteristic curves in the second quadrant near the critical current on a logarithmic scale. As the temperature changes from normal to the superconducting states, the I-V characteristic changes from linear (V∝I) to exponential (V∝Iα) dependence, demonstrating the typical 2D superconductor BKT transition feature of R-N8/S2. We further extracted the power index values α as a function of temperature from slopes in the log-log scale I-V characteristic curves at different temperatures (Fig. 5d), yielding an extrapolated TBKT of ~6.76 K. In addition, the R(T) curve can be reproduced by BKT transition using Halperin-Nelson equation55 (red solid curve in the inset of Fig. 5d), , where R0 and b are material parameters. The fitting results give a BKT transition temperature of TBKT = 6.80 K, which is consistent with the data extrapolated from the I-V curves.
Discussion
Overall, the R-N8/S2 superlattice exhibits superconductivity with an onset transition temperature of 12.5 K, which is consistent with the Tc range of 9–15 K in 10-nm-thick Nd0.8Sr0.2NiO2 films grown on SrTiO3 substrates2. Different magnetoresistive responses to magnetic fields along in-plane and out-of-plane directions suggest anisotropic superconducting behavior in the R-N8/S2 superlattice. Additionally, a BKT transition with a TBKT of 6.80 K was also observed. Above all imply the presence of 2D superconducting feature in the R-N8/S2, which is consistent with the infinite-layer nickelates reported previously37,38,56. However, the superconducting anisotropy in R-N8/S2 is much smaller compared to the significant anisotropy observed in La0.8Sr0.2NiO256, which had also been observed in the Nd-based infinite-layer nickelates35. The superconducting anisotropy difference between La-based and Nd-based nickelates may due to the influence from rare-earth 4f moments37,39. All the experimental results for the superconducting properties of R-N8/S2 superlattice are in great agreement with those of Nd0.8Sr0.2NiO2 films. These facts indicate that the SrTiO3 interlayers have no evident enhancement to the superconductivity of the infinite-layer nickelates in this artificial superlattice. It is different from the FeSe/SrTiO3 interface, where the interface effect significantly enhances superconductivity26. This suggests that the superconductivity in the superlattice originates from the intrinsic properties of the infinite-layer nickelates, it is also confirmed in recently free-standing infinite-layer nickelates57,58.
In summary, we had successfully achieved superconductivity with a Tc,onset of 12.5 K in the (Nd0.8Sr0.2NiO2)8/(SrTiO3)2(R-N8/S2) superlattice using the two-step method2. We observed a structural transition modulated by the thickness of the nickelate layers within the reduced nickelate superlattice. High-resolution HADDF-STEM and iDPC images show that the superlattice R-N3/S2 with the thinner nickelate-thickness exhibits partially reduced and disordered phases even after the fully reduction process, whereas the thicker R-N8/S2 displays a purely periodic infinite-layer structure. Measurements of anisotropic magnetoresistance and current-voltage characteristic curves indicate a 2D superconducting feature of R-N8/S2. Collectively, the findings on the superconducting properties of the infinite-layer nickelate superlattice closely mirror those of Nd0.8Sr0.2NiO2 single films. This suggests that the interface between SrTiO3 and the infinite-layer nickelate has no substantial impact on its superconducting behavior, serving as indirect evidence of the intrinsic superconductivity of infinite-layer nickelates. The realization of superconducting infinite-layer superlattice nickelates opens up possibilities for further interfacial modulation of superconducting films in infinite-layer nickelates and the fabrication of devices. In addition, it also provides a new perspective for understanding the effect of interfaces on the superconductivity of infinite-layer nickelates, which is promising for further studies of the mechanisms of unconventional superconductivity.
Methods
Film growth
The pristine perovskite [(Nd0.8Sr0.2NiO3)n/(SrTiO3)m]10 superlattices were grown on TiO2-terminated (001) SrTiO3 substrates of size 5 × 5 × 0.5 mm3 by pulsed laser deposition using a KrF excimer laser (λ = 248 nm). The laser fluence and repetition rate were mixed at 2.5 J/cm2 and 4 Hz. Oxygen partial pressure and substrate temperature during growth were controlled at 600 °C and 0.1 mbar, respectively. During the growth process, Reflection High-Energy Electron Diffraction (RHEED) was used to in situ monitor the thickness of Nd0.8Sr0.2NiO3 and SrTiO3 layers, respectively. A single crystal SrTiO3 and a Nd0.8Sr0.2Ni1.15O3 polycrystalline target were used during the deposition. The Nd0.8Sr0.2Ni1.15O3 polycrystalline target precursors were prepared by high temperature sintering. Nd2O3, SrCO3 and NiO powder were mixed and dispersed with ethanol and ball milling for 24 h. The dried powder was then decarburized at 1200 °C for 12 h, re-ground and mixed, then pressed into a cylindrical target, and sintered at 1350 °C for 12h2.
Reduction process
The pristine samples were cut into four pieces with a size of 2.5 × 2.5 × 0.5 mm3. Each piece was put in an alumina crucible with CaH2 powder (~1 g) without directly contact. Then the crucible was heated to 320 °C and maintained for 10−20 h in a vacuum chamber with a background air pressure of 10−5 mbar, with heating and cooling rates of 10 °C/min. In order to ensure that the superlattice samples of different thicknesses are fully reduced, the reduction time was continuously increased. After reduction, the vacuum chamber was restored to atmospheric pressure.
Sample characterization
Surface morphologies were acquired using an atomic force microscope (AFM) in contact mode, and their root-mean-square (RMS) roughness values were analyzed using Gwyddion software. The structural properties of films were characterized by the X-ray diffraction (XRD) patterns and reciprocal space mapping (RSM) using a monochromated Cu-Kα radiation (λ = 0.154 nm). Film thicknesses were measured by X-ray reflectivity (XRR) and fitted by Genx software. The transport behaviors were measured in a van-der-Pauw geometry with Al wire bonded contacts using a Quantum Design Physical Property Measurement System (PPMS-9T). Current-voltage characteristic curves were measured by four-probe method using a source meter unit (Keithley 2400). HAADF-STEM and iDPC-STEM experiments were carried out at an aberration-corrected FEI Titan Themis G2 operated at 300 kV.
X-ray absorption measurements
X-ray absorption spectroscopy (XAS) and X-ray linear dichroism (XLD) were measured at the beamlines MCD-A and MCD-B (Soochow Beamline for Energy Materials) at the Hefei National Synchrotron Radiation Laboratory (NSRL). In the Total Electron Yield (TEY) mode, the Ni L2,3-edge was measured along the direction of normal incidence and grazing incidence, the grazing incidence angle is 15°.
Supplementary information
Source data
Acknowledgements
This work was supported by the National Key R&D Program of China (Grant No. 2022YFA1403000 (Z.L.)), CAS Project for Young Scientists in Basic Research (Grant No. YSBR-100 (Z.L.)), the Fundamental Research Funds for the Central Universities (Grant No. WK2140000019 (Z.L.)), National Nature Science Foundation of China (Grant No. 52272095 (Z.L.) and 12275272 (K.C.)), and Collaborative Innovation Program of Hefei Science Center, Chinese Academy of Sciences (Grant No. 2022HSC-CIP005 (K.C.)). The authors would like to appreciate Beamlines MCD-A and MCD-B (Soochow Beamline for Energy Materials) at NSRL.
Author contributions
Z.L. conceived the research and designed the experiment. W.X. synthesized, reduced the superlattice thin films. W.X., Z.Y and S.H characterized the structures of pristine and as-reduced thin films. W.X., X.G. and J.L. explored the reduction process. Y.He and Q.Z. performed the STEM experiments. W.X., Z.D., Y.Hong, L.Wang and T.W measured electrical properties. W.X., L.Wei, Z.S., L.L. and K.C. performed measurement and analysis of synchrotron soft X-ray absorption spectroscopy. W.X. and Z.L. wrote this manuscript. Y.G., Q.Z., K.C. and Z.L. extensively discussed the results and were involved in writing of the manuscript. All authors contributed to the analysis of the results.
Peer review
Peer review information
Nature Communications thanks the anonymous reviewers for their contribution to the peer review of this work. A peer review file is available.
Data availability
The data that support the findings of this study are available from the corresponding author upon request. Source data are provided with this paper.
Code availability
The code used in this study (for XRR fit) is available from the corresponding author on request. Or it can be accessed at https://aglavic.github.io/genx/doc/tutorials/simple_reflectivity.html#getting-started.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Contributor Information
Yulin Gan, Email: ylgan@ustc.edu.cn.
Kai Chen, Email: kaichen2021@ustc.edu.cn.
Qinghua Zhang, Email: zqh@iphy.ac.cn.
Zhaoliang Liao, Email: zliao@ustc.edu.cn.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-024-54660-w.
References
- 1.Bednorz, J. G. & Müller, K. A. Possible high Tc superconductivity in the Ba−La−Cu−O system. Z. f.ür. Phys. B Condens. Matter64, 189–193 (1986). [Google Scholar]
- 2.Li, D. et al. Superconductivity in an infinite-layer nickelate. Nature572, 624–627 (2019). [DOI] [PubMed] [Google Scholar]
- 3.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 (2020). [Google Scholar]
- 4.Osada, M. et al. Nickelate Superconductivity without Rare-Earth Magnetism: (La,Sr)NiO2. Adv. Mater.33, 2104083 (2021). [DOI] [PubMed] [Google Scholar]
- 5.Zeng, S. et al. Superconductivity in infinite-layer nickelate La1-xCaxNiO2 thin films. Sci. Adv.8, eabl9927 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Wei, W., Vu, D., Zhang, Z., Walker, F. J. & Ahn, C. H. Superconducting Nd1-xEuxNiO2 thin films using in situ synthesis. Sci. Adv.9, eadh3327 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Pan, G. A. et al. Superconductivity in a quintuple-layer square-planar nickelate. Nat. Mater.21, 160–164 (2022). [DOI] [PubMed] [Google Scholar]
- 8.Sun, H. L. et al. Signatures of superconductivity near 80 K in a nickelate under high pressure. Nature621, 493–498 (2023). [DOI] [PubMed] [Google Scholar]
- 9.Zhu, Y. et al. Superconductivity in pressurized trilayer La4Ni3O10−δ single crystals. Nature631, 531–536 (2024). [DOI] [PubMed] [Google Scholar]
- 10.Zeng, S. et al. Phase Diagram and Superconducting Dome of Infinite-Layer Nd1-xSrxNiO2 Thin Films. Phys. Rev. Lett.125, 147003 (2020). [DOI] [PubMed] [Google Scholar]
- 11.Li, D. et al. Superconducting Dome in Nd1−xSrxNiO2 Infinite Layer Films. Phys. Rev. Lett.125, 027001 (2020). [DOI] [PubMed] [Google Scholar]
- 12.Gu, Q. et al. Single particle tunneling spectrum of superconducting Nd1-xSrxNiO2 thin films. Nat. Commun.11, 6027 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Cheng, B. et al. Evidence for d-wave superconductivity of infinite-layer nickelates from low-energy electrodynamics. Nat. Mater.23, 775–781 (2024). [DOI] [PubMed] [Google Scholar]
- 14.Ding, X. et al. Cuprate-like electronic structures in infinite-layer nickelates with substantial hole dopings. Natl Sci. Rev.11, nwae194 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Sun, W. et al. Electronic Structure of Superconducting Infinite-Layer Lanthanum Nickelates. Preprint at https://arxiv.org/abs/2403.07344 (2024).
- 16.Zaanen, J., Sawatzky, G. A. & Allen, J. W. Band gaps and electronic structure of transition-metal compounds. Phys. Rev. Lett.55, 418–421 (1985). [DOI] [PubMed] [Google Scholar]
- 17.Hepting, M. et al. Electronic structure of the parent compound of superconducting infinite-layer nickelates. Nat. Mater.19, 381–385 (2020). [DOI] [PubMed] [Google Scholar]
- 18.Keimer, B., Kivelson, S. A., Norman, M. R., Uchida, S. & Zaanen, J. From quantum matter to high-temperature superconductivity in copper oxides. Nature518, 179–186 (2015). [DOI] [PubMed] [Google Scholar]
- 19.Hayward, M. A., Green, M. A., Rosseinsky, M. J. & Sloan, J. Sodium Hydride as a Powerful Reducing Agent for Topotactic Oxide Deintercalation: Synthesis and Characterization of the Nickel(I) Oxide LaNiO2. J. Am. Chem. Soc.121, 8843–8854 (1999). [Google Scholar]
- 20.Lu, H. et al. Magnetic excitations in infinite-layer nickelates. Science373, 213–216 (2021). [DOI] [PubMed] [Google Scholar]
- 21.Werner, P. & Hoshino, S. Nickelate superconductors: Multiorbital nature and spin freezing. Phys. Rev. B101, 041104 (2020). [Google Scholar]
- 22.Hu, L. H. & Wu, C. J. Two-band model for magnetism and superconductivity in nickelates. Phys. Rev. Res.1, 032046 (2019). [Google Scholar]
- 23.Zhang, Y.-H. & Vishwanath, A. Type-II t−J model in superconducting nickelate Nd1−xSrxNiO2. Phys. Rev. Res.2, 023112 (2020). [Google Scholar]
- 24.Reyren, N. et al. Superconducting Interfaces Between Insulating Oxides. Science317, 1196–1199 (2007). [DOI] [PubMed] [Google Scholar]
- 25.Gozar, A. et al. High-temperature interface superconductivity between metallic and insulating copper oxides. Nature455, 782–785 (2008). [DOI] [PubMed] [Google Scholar]
- 26.Wang, Q.-Y. et al. Interface-Induced High-Temperature Superconductivity in Single Unit-Cell FeSe Films on SrTiO3. Chin. Phys. Lett.29, 037402 (2012). [Google Scholar]
- 27.Norton, D. P. et al. Superconductivity in SrCuO2-BaCuO2 Superlattices: Formation of Artificially Layered Superconducting Materials. Science265, 2074–2077 (1994). [DOI] [PubMed] [Google Scholar]
- 28.Ortiz, R. A. et al. Superlattice approach to doping infinite-layer nickelates. Phys. Rev. B104, 165137 (2021). [Google Scholar]
- 29.Yang, C. et al. Thickness-Dependent Interface Polarity in Infinite-Layer Nickelate Superlattices. Nano Lett.23, 3291–3297 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Lee, K. et al. Aspects of the synthesis of thin film superconducting infinite-layer nickelates. APL Mater.8, 041107 (2020). [Google Scholar]
- 31.Zeng, S. et al. Origin of a Topotactic Reduction Effect for Superconductivity in Infinite-Layer Nickelates. Phys. Rev. Lett.133, 066503 (2024). [DOI] [PubMed] [Google Scholar]
- 32.Ding, X. et al. Critical role of hydrogen for superconductivity in nickelates. Nature615, 50–55 (2023). [DOI] [PubMed] [Google Scholar]
- 33.Goodge, B. H. et al. Resolving the polar interface of infinite-layer nickelate thin films. Nat. Mater.22, 466–473 (2023). [DOI] [PubMed] [Google Scholar]
- 34.Yang, C. et al. Direct observation of strong surface reconstruction in partially reduced nickelate films. Nat. Commun.15, 378 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Wang, B. Y. et al. Isotropic Pauli-limited superconductivity in the infinite-layer nickelate Nd0.775Sr0.225NiO2. Nat. Phys.17, 473–477 (2021). [Google Scholar]
- 36.Xiang, Y. et al. Physical Properties Revealed by Transport Measurements for Superconducting Nd0.8Sr0.2NiO2 Thin Films. Chin. Phys. Lett.38, 047401 (2021). [Google Scholar]
- 37.Wang, B. Y. et al. Effects of rare-earth magnetism on the superconducting upper critical field in infinite-layer nickelates. Sci. Adv.9, eadf6655 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Ji, H. et al. Rotational symmetry breaking in superconducting nickelate Nd0.8Sr0.2NiO2 films. Nat. Commun.14, 7155 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Chow, L. E. et al. Dimensionality control and rotational symmetry breaking superconductivity in square-planar layered nickelates. Preprint at https://arxiv.org/abs/2301.07606 (2023).
- 40.Samal, D. et al. Experimental evidence for oxygen sublattice control in polar infinite layer SrCuO2. Phys. Rev. Lett.111, 096102 (2013). [DOI] [PubMed] [Google Scholar]
- 41.Liao, Z. L. et al. Large orbital polarization in nickelate-cuprate heterostructures by dimensional control of oxygen coordination. Nat. Commun.10, 589 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Kawai, M. et al. Orientation Change of an Infinite-Layer Structure LaNiO2 Epitaxial Thin Film by Annealing with CaH2. Cryst. Growth Des.10, 2044–2046 (2010). [Google Scholar]
- 43.Zhong, Z., Koster, G. & Kelly, P. J. Prediction of thickness limits of ideal polar ultrathin films. Phys. Rev. B85, 121411 (2012). [Google Scholar]
- 44.Zeng, S. W. et al. Observation of perfect diamagnetism and interfacial effect on the electronic structures in infinite layer Nd0.8Sr0.2NiO2 superconductors. Nat. Commun.13, 743 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Song, D. et al. Visualization of Dopant Oxygen Atoms in a Bi2Sr2CaCu2O8+δ Superconductor. Adv. Funct. Mater.29, 1903843 (2019). [Google Scholar]
- 46.Ikeda, A., Krockenberger, Y., Irie, H., Naito, M. & Yamamoto, H. Direct observation of infinite NiO2 planes in LaNiO2 films. Appl. Phys. Express9, 061101 (2016). [Google Scholar]
- 47.Hawthorn, D. G. et al. Resonant elastic soft x-ray scattering in oxygen-ordered YBa2Cu3O6+δ. Phys. Rev. B84, 075125 (2011). [Google Scholar]
- 48.Zhang, J. et al. Large orbital polarization in a metallic square-planar nickelate. Nat. Phys.13, 864–869 (2017). [Google Scholar]
- 49.Hua, X. et al. Superconducting stripes induced by ferromagnetic proximity in an oxide heterostructure. Nat. Phys.20, 957–963 (2024). [Google Scholar]
- 50.Saito, Y., Kasahara, Y., Ye, J., Iwasa, Y. & Nojima, T. Metallic ground state in an ion-gated two-dimensional superconductor. Science350, 409–413 (2015). [DOI] [PubMed] [Google Scholar]
- 51.Xing, Y. et al. Ising Superconductivity and Quantum Phase Transition in Macro-Size Monolayer NbSe2. Nano Lett.17, 6802–6807 (2017). [DOI] [PubMed] [Google Scholar]
- 52.Harper, F. E. & Tinkham, M. The Mixed State in Superconducting Thin Films. Phys. Rev.172, 441–450 (1968). [Google Scholar]
- 53.Zhao, Q. et al. Isotropic Quantum Griffiths Singularity in Nd0.8Sr0.2NiO2 Infinite-Layer Superconducting Thin Films. Phys. Rev. Lett.133, 36003 (2024). [DOI] [PubMed] [Google Scholar]
- 54.Saito, Y., Nojima, T. & Iwasa, Y. Highly crystalline 2D superconductors. Nat. Rev. Mater.2, 16094 (2016). [Google Scholar]
- 55.Nelson, D. R. & Halperin, B. I. Dislocation-mediated melting in two dimensions. Phys. Rev. B19, 2457–2484 (1979). [Google Scholar]
- 56.Sun, W. et al. Evidence for Anisotropic Superconductivity Beyond Pauli Limit in Infinite-Layer Lanthanum Nickelates. Adv. Mater.35, 2303400 (2023). [DOI] [PubMed] [Google Scholar]
- 57.Yan, S. et al. Superconductivity in Freestanding Infinite‐Layer Nickelate Membranes. Adv. Mater.36, 2402916 (2024). [DOI] [PubMed] [Google Scholar]
- 58.Lee, Y. et al. Millimeter-scale freestanding superconducting infinite-layer nickelate membranes. Preprint at https://arxiv.org/abs/2402.05104 (2024).
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data that support the findings of this study are available from the corresponding author upon request. Source data are provided with this paper.
The code used in this study (for XRR fit) is available from the corresponding author on request. Or it can be accessed at https://aglavic.github.io/genx/doc/tutorials/simple_reflectivity.html#getting-started.