[1]\fnmHoshang \surSahib [1]\fnmDaniele \surPreziosi

1]Universitè de Strasbourg, CNRS, IPCMS UMR 7504, F-67034 Strasbourg, France 3]Laboratoire de Physique des Solides, CNRS, Université Paris-Saclay, 91405 Orsay, France 4]Synchrotron SOLEIL, L’Orme des Merisiers, BP 48 St Aubin, 91192 Gif sur Yvette, France 2]Dipartimento di Fisica, Politecnico di Milano, Piazza Leonardo da Vinci 32, I-20133 Milano, Italy 5]European XFEL, Holzkoppel 4, Schenefeld, D-22869, Germany 6]CNR-SPIN Complesso di Monte S. Angelo, via Cinthia - I-80126 Napoli, Italy

Superconductivity in PrNiO2 infinite-layer nickelates

hoshang.sahib@ipcms.unistra.fr    \fnmFrancesco \surRosa    \fnmAravind \surRaji    \fnmGiacomo \surMerzoni    \fnmGiacomo \surGhiringhelli    \fnmMarco \surSalluzzo    \fnmAlexandre \surGloter    \fnmNathalie \surViart    daniele.preziosi@ipcms.unistra.fr [ [ [ [ [ [
Abstract

Several reports about infinite-layer nickelate thin films suggest that the superconducting critical temperature versus chemical doping phase diagram has a dome-like shape, similar to cuprates. Here, we demonstrate a highly reproducible superconducting state in undoped PrNiO2 thin films grown onto SrTiO3. Scanning transmission electron microscopy measurements demonstrate coherent and defect-free infinite-layer phase, a high structural quality with no unintentional chemical doping and a total absence of interstitial oxygen. X-ray absorption measurements show very sharp features at the Ni L3,2-edges with a large linear dichroism, indicating the preferential hole-occupation of Ni1+-3dx2y2superscript𝑥2superscript𝑦2{}_{x^{2}-y^{2}}start_FLOATSUBSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_FLOATSUBSCRIPT orbitals in a square planar geometry. Resonant inelastic X-ray scattering measurements reveal sharp magnon excitations of 200 meV energy at magnetic Brillouin zone boundary, highly resonant at the Ni1+ absorption peak. The results indicate that, when properly stabilized, infinite-layer nickelate thin films are superconducting without chemical doping.

keywords:
superconductivity, infinite-layer nickelates, STEM, RIXS

1 Introduction

Superconductivity in Infinite-Layer (IL) nickelate thin films is under scrutiny since its discovery [1], and several crucial questions remain still unanswered. Contrarily to cuprates, IL nickelates do not show a long range antiferromagnetic (AFM) order nor bulk superconductivity [2], which suggests that a template is strictly necessary to properly stabilize the IL phase. The latter is obtained by following a delicate synthesis process, which poses severe problems of reproducibility [3]. Because of the difficulties in the optimization of the growth and reduction parameters, the electronic and magnetic properties of these new compounds are not yet full established. From the synthesis point of view, the presence of a capping layer emerged as a relevant step to reduce the defect density, although superconductivity was demonstrated in both capped [1] and uncapped [4] chemically doped IL films. In particular, the absence of a capping layer largely limits the quality of the samples [5, 6]. Uncapped NdNiO2 IL nickelate are only partially reduced, and show a (303)-stripe-organized interstitial oxygen pattern, which can be linked to the origin of the early observed charge ordering [7]. Similarly, the bad-metal low temperature resistance upturn observed in undoped IL nickelates, attributed to strong electron correlation effects [8], might not be an intrinsic property of the IL phase. Indeed, NdNiO2 thin films realized by in-situ oxygen de-intercalation by atomic hydrogen did not show a pronounced low temperature resistance upturn [6, 9].
So far, superconductivity in IL nickelates was reported by several groups only in chemically doped samples, regardless of the approaches used to stabilize the IL phase, including the solid state Al-reduction method used in the case of Nd1-xEuxNiO3 thin films [10], or the atomic-hydrogen method in Nd1-xSrxNiO2 [6] and La1-xSrxNiO2 thin films [11]. In this context, it is worth mentioning that superconductivity was first demonstrated in La0.8Sr0.2NiO2 only after a significant optimization of the CaH2-based process by Osada et𝑒𝑡etitalic_e italic_t al.𝑎𝑙al.italic_a italic_l .[12]. In the same report, hints of a superconducting transition, at lower temperature, were provided also for undoped LaNiO2 thin films. So far, this result is not yet reproduced in other laboratories. Only very recently, an incomplete superconducting transition has been reported in the case of the undoped NdNiO2 films as well [13]. The observations of superconductivity in undoped IL nickelates, while needing independent verification, suggest that the reported phase diagram of nickelates, also based on the notion that undoped compounds are weakly-insulating bad-metals, [14, 15, 16, 17], might need a revision.

Here, we show that highly crystalline undoped PrNiO2 thin films are superconducting, with onset transition temperature (TC) in the 7-11 K range, and a zero resistance temperature up to 4 K. The highly reproducible zero resistance state is obtained via consecutive CaH2-based topotactic reduction steps for samples prepared with a STO-capping-layer larger than 6 unit-cells (uc), while all PrNiO2 films, irrespective of the presence/absence of a capping layer and/or the number of reduction cycles, show the onset of a superconducting transition. Scanning Transmission Electron Microscopy (STEM) shows very limited amount of defects and/or spurious phases in the entire observed volume, while divergence of the Center of Mass (dCOM) four-dimensional (4D)-STEM measurements reveals a complete absence of apical oxygens and NiO2 planes much more ordered than those of NdNiO2 samples [18]. O K-edge electron energy loss spectroscopy (EELS) further support the absence of any unintentional chemical doping. X-ray Absorption Spectroscopy (XAS) at the Ni L3,2-edges show very sharp peaks with a relatively large dichroism, demonstrating a complete reduction of our thin films [19]. Additionally, Resonant Inelastic X-ray Scattering (RIXS) shows that PrNiO2 are characterized by well defined magnons with a bandwidth of ca.𝑐𝑎ca.italic_c italic_a . 200 meV. Our results suggest that the superconductivity in undoped IL nickelates might be mostly hampered by subtle details accompanying the topotactic reduction process, instead of being controlled by a threshold value of chemical doping. The data show that PrNiO2 nickelates are self-hole-doped, as suggested by early theoretical reports [20]. Finally, we show that the robust stabilization of the IL-phase is a consequence of an high-quality growth of the perovskite-phase, which is an indispensable requisite for superconductivity in undoped infinite-layer nickelates.

2 Results

Precursor perovskite PrNiO3 (PNO3) films were deposited onto SrTiO3 (STO) single crystal as substrate and capped by an epitaxial STO film grown in-situ (See Methods and Supporting Information). A correct cation stoichiometry is mandatory to reduce extended defects and to achieve bulk-like transport properties, while granting a Ni3+ valence state [21, 22]. It is well known that a large epitaxial mismatch can lower the energetic barrier for defects formation, that in the case of nickelates are usually identified as Ruddlesden–Popper (RP)-like stacking faults [23]. These type of defects largely hampered reproducible superconductivity in Nd-based IL nickelate thin films, due to the lower crystallinity of the tensile strained perovskite precursor [24]. According to the perovskite nickelates phase diagram, bulk PNO3 exhibits a less distorted unit cell akin to a larger ionic radius (rPr3+𝑃superscript𝑟limit-from3{}_{Pr^{3+}}start_FLOATSUBSCRIPT italic_P italic_r start_POSTSUPERSCRIPT 3 + end_POSTSUPERSCRIPT end_FLOATSUBSCRIPT = 0.111 nm) compared to Nd (rNd3+𝑁superscript𝑑limit-from3{}_{Nd^{3+}}start_FLOATSUBSCRIPT italic_N italic_d start_POSTSUPERSCRIPT 3 + end_POSTSUPERSCRIPT end_FLOATSUBSCRIPT = 0.110 nm) [25]. The increased Ni3d3𝑑3d3 italic_d-O2p2𝑝2p2 italic_p orbital overlap leads to a metal-to-insulator transition (MIT) at a relatively lower temperature (TMITsubscript𝑇𝑀𝐼𝑇T_{MIT}italic_T start_POSTSUBSCRIPT italic_M italic_I italic_T end_POSTSUBSCRIPT = 130 K), making the material more metallic compared to NdNiO3 (TMITsubscript𝑇𝑀𝐼𝑇T_{MIT}italic_T start_POSTSUBSCRIPT italic_M italic_I italic_T end_POSTSUBSCRIPT = 200 K). The room temperature orthorhombic lattice parameters of bulk PNO3 are a𝑎aitalic_a = 0.541 nm, b𝑏bitalic_b = 0.538 nm, and c𝑐citalic_c = 0.762 nm, corresponding to a pseudo-cubic lattice with apcsubscript𝑎𝑝𝑐a_{pc}italic_a start_POSTSUBSCRIPT italic_p italic_c end_POSTSUBSCRIPT = 0.382 nm approximately [26]. On the other hand, the PrNiO2 (PNO2) larger volume of the tetragonal (P4/mmm) unit cell (a𝑎aitalic_a = 0.394 nm and c𝑐citalic_c = 0.328 nm) [27] results in a higher compressive strain of PNO2 thin films onto STO (-0.9%) compared to NdNiO2 (-0.5%). This should not be seen as a limiting factor for the stabilization of the IL phase, as there are reports of superconductivity in Nd0.8Sr0.2NiO2 [28] and Pr0.8Sr0.2NiO2 [29] thin films grown onto (LaAlO3)0.3(Sr2TaAlO6)0.7 (LSAT), with hints of an enhancement of the Tc by strain. Furthermore, calculations based on a dynamical vertex approximation technique predict that, under hydrostatic pressure, the PNO2 parent compound could become a high-temperature superconductor by experiencing enhanced self-doping of the Ni orbitals [30]. We have used several experimental techniques to study the structural, electronic and magnetic properties of our fully reduced PNO2 samples. Figure 1a displays the temperature dependence of the resistivity for one of our first superconducting PNO2 film, composed of 16 unit-cells (uc)-thick PNO2 film capped with 6 uc of STO (hereafter referred as STO6uc-PNO2). Despite a relatively large resistivity at room temperature, compared to other reports in literature [31, 32], the metallic T-linear behavior is followed by a slight upturn below ca.𝑐𝑎ca.italic_c italic_a . 50 K and a complete superconducting transition around 4 K. The measured onset critical temperature TC (maximum curvature) is slightly below 11 K. This represents the main result of this work. In the top-left inset of Figure 1a we show the temperature-dependence of the resistivity as a function of the perpendicularly applied magnetic field, up to 9 Tesla. The normal state resistivity is insensitive to the magnetic field, while superconductivity is progressively suppressed. In the bottom-right inset of Figure 1a we show the temperature dependence of Hall coefficient, which is negative in the whole temperature range. The Hall coefficient shows a tendency towards a change of sign down to 50 K, and then an opposite trend. Nearby TC-onset, its negative value further increases, which is remarkably different from change of sign, from negative to positive, observed in superconducting chemically doped samples [1, 16, 33]. On the other hand, the Hall coefficient temperature dependence is very similar to that one reported on undoped (non-superconducting) samples.

Refer to caption
Figure 1: (a) Temperature-dependent resistivity of one of the STO(6uc)-capped, 16-unit-cell thick PNO2 film. Upper left inset shows the temperature dependence of the resistivity under applied magnetic fields; lower right inset displays the temperature-dependent Hall coefficient. (b) Temperature-dependent resistivity of several superconducting films grown in nominal similar conditions. The inset shows the same data around the superconducting transition. To note that samples with a lower residual resistance have a much broader superconducting transition.

In Figure 1b we show the transport properties of a series of STO6uc-PNO2 samples prepared in nominal similar conditions. All the samples display a zero resistance state below 4 K. The TC-onset varies from sample-to-sample in a relatively small range, with a minimum value of 7 K and maximum value of 11 K, while the normal state residual resistivity and its value at room temperature varies in a much larger range, likely reflecting subtle details in the topotactic reduction process. The superconductivity in these undoped PNO2 films is very robust. In particular, an incomplete superconducting transition, with similar TC-onset, is observed also in uncapped PNO2 samples, although a zero resistance state is obtained only on PNO2 capped with at least 6uc of STO (See Supplemental Information for additional data). This confirms the key role played by the capping layer in stabilizing a clean and robust IL phase.

Refer to caption
Figure 2: (a) HAADF-STEM image of a 6-unit-cell STO capped PNO2 film (16uc thick). (b) HAADF image from a zoom-in region. (c) EELS Sr, Pr, Ni, and Ti elemental maps, and (d) elemental profiles integrated over the zoom-in region showing a Pr/Ni ratio close to 1 over the whole thin film. Both top and bottom interfaces are not sharp. (e) A magnified HAADF image from the bottom interface region, (f) the corresponding 4D-STEM dCOM analysis showing clean infinite-layer phase without the presence of any apical oxygen. (h) Real space evolution of the O-K edge EELS fine structure from regions as labelled in the HAADF image in (g).

We resorted to STEM measurements with a High-Angle Annular Dark-Field (HAADF) imaging technique and Electron-Energy-Loss Spectroscopy (EELS) to study the precise atomic stack, including both bottom and top interfaces. The atomic-resolved HAADF-STEM cross-sectional image shown in Figure 2a displays a clear infinite-layer phase with a very high structural quality, thus demonstrating the overall absence of RP-like defects in the precursor phase, as confirmed by HAADF-STEM cross-sectional images for STO6uc-PNO3 samples. The EELS data acquired over the entire STO/PNO2//STO stack of a selected area of the HAADF image (Fig. 2b), show a clear intermixing at the bottom film-substrate and at the top film-STO interfaces (Figure 2c). In particular, the interface with the TiO2-terminated substrate is characterized by a one unit cell (Ni,Ti) intermixing layer, confirming other reports [34, 35]. Additionally, the total element map intensity profile shown in Figure 2d clearly indicates that the top interface exhibits a complex Ni-Pr-Ti-Sr atomic stack, which differs from other recent reports of more abrupt interfaces in Sr-doped PNO3 and related IL thin films [35]. It is worth to recall here that, early theoretical calculations proposed the formation of a quasi-2D electron gas, with possible superconducting properties, in the case of a clean interface between the infinite-layer nickelate and the STO substrate [36]. Our HAADF-STEM results, in agreement with Goodge et𝑒𝑡etitalic_e italic_t al𝑎𝑙alitalic_a italic_l., fully rules out the formation of a quasi-2D electron gas since the intermixed (Ni,Ti)O3 interface-layer fully compensates the polar discontinuity [34].

Unintentional chemical doping can arise from Pr/Ni off stoichiometry, Sr-diffusion and interstitial/apical oxygen incorporated into the nickelate layer. Our elemental EELS mapping shows a near unitary Pr/Ni ratio and no unintentional Sr-doping (within the experimental resolution of the order of 10%) in the PNO2 sample. In order to identify any traces of these interstitial/apical oxygen, we employed 4D-STEM dCOM imaging. Some of the authors have already shown that this powerful technique allows high-resolution real space atomic mapping with good oxygen contrast [5]. The HAADF image in Figure 2e, is obtained at the bottom interface, and the corresponding 4D-STEM dCOM analysis in Figure 2f unveils the occupied oxygen sites not visible in the HAADF image. Excluding the intermixed interfacial unit-cell, characterized by a (Ni,Ti)Ox composition in an octahedral coordination, we can easily distinguish occupied oxygen-sites only within the NiO2 planes as, indeed, expected for a properly stabilized IL-phase. This directly excludes traces of apical oxygen within the PNO2 samples and, moreover, demonstrates a rather clean Ni-sites in square planar arrangement. To further evaluate the absence of any chemical source of doping with higher sensitivity, we performed atomically resolved EELS fine structure analysis at the Oxygen K-edge. This is shown in Figures 2g,h. Figure 2h shows spatially resolved STEM-EELS of O K-edge fine structures in different regions of the sample, including the two interfacial regions. As expected for fully reduced IL [37, 5], the O K-edge profiles do not show any pre-peak feature, with the only exception of the area in proximity of the substrate, where a certain degree of Ni3d3𝑑3d3 italic_d-O2p2𝑝2p2 italic_p hybridization is still observed because of the (Ni,Ti) intermixing layer [38], and nearby the top, intermixed, interface due to the STO-capping-layer. The absence of any pre-edge is a clear indication that no unintentional chemical doping is present in our samples. In particular, a pre-peak shoulder is always observed, also locally, in Sr-doped IL-nickelates, as shown in previous works [39, 37, 35]. From this analysis, we can confidently conclude that the residual chemical doping is below the sensitivity of the multiple techniques used, and thus much lower than the minimum threshold needed so-far to trigger superconductivity in chemically doped IL-nickelates.

In order to get information about the electronic structure of our superconducting PNO2 thin films, we performed XAS and RIXS measurements at the Ni L3,2-edges and L3-edge, respectively (see Methods for details). Overall, the Ni L3 XAS spectra show features similar to properly optimized NdNiO2 thin films [18, 39, 19] with a dominant peak due to the 2p63d9 \rightarrow 2p53d10 transition. Figure 3a shows XAS spectra acquired with linearly-polarized light in the parallel and nearly perpendicular direction with respect to the NiO2 planes, i.e.formulae-sequence𝑖𝑒i.e.italic_i . italic_e . σ𝜎\sigmaitalic_σ-pol and π𝜋\piitalic_π-pol, respectively. The more than 50% dichroism is a direct consequence of the very robust infinite-layer phase, where the majority of the holes resides in the Ni1+-3dx2y2superscript𝑥2superscript𝑦2{}_{x^{2}-y^{2}}start_FLOATSUBSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_FLOATSUBSCRIPT orbitals. These results confirm a full oxygen reduction after the topotactic process. Moreover, the Ni L3 peak is very sharp with a small shoulder at higher photon energy, speaking against adventitious sources of doping due to incomplete reduction (Ni2+) and/or excess oxygen [39, 18]. For comparison, the inset in Figure 3a reports the XAS at the Ni L3-edge peak acquired for a STO-capped NdNiO2 with, indeed, very similar features. In Figures 3b,c we show the mid-infrared region of RIXS spectra acquired in σ𝜎\sigmaitalic_σ-pol and π𝜋\piitalic_π-pol, respectively. We can easily recognize both the phonon and magnon excitations, which we fit with a Gaussian and a Damped Harmonic Oscillator susceptibility, respectively (please refer to the Supporting information for further details about the fitting procedure). A linear background from the tails of higher-energy excitation is also included. The phonon is apparently enhanced by σ𝜎\sigmaitalic_σ incident polarization and, therefore, mostly associated to in-plane Ni-O stretching modes, similarly to cuprates [40]. The magnetic peak occupies an energy range around ca.𝑐𝑎ca.italic_c italic_a . 200 meV, consistently with previous results on IL nickelates [41, 42, 43, 44]. In the insets we show the same RIXS spectra in the full energy loss range, enough to capture the Ni 3d orbital excitations betweeen 1 and 3 eV [39]. The peak at 0.6 eV, which is more pronounced in π𝜋\piitalic_π-pol, is attributed to out-of-plane Pr-Ni hybridization, in analogy with other undoped nickelates [45, 39, 46, 42]. Finally, the shape of the magnetic excitations, among the sharpest ever measured by RIXS in undoped IL nickelates, though not resolution-limited, indicates that superconductivity is reached in the presence of long-range spin-spin correlations. This means that either a self doping mechanism is present, which does not perturb the antiferromagnetic background, or that superconductivity is obtained at extremely low levels of hole doping.

Refer to caption
Figure 3: (a) Linearly polarized XAS spectra at the Ni L3,2 edges for a STO6uc-PNO2 sample, with electric field parallel (σ𝜎\sigmaitalic_σ-pol) and perpendicular (π𝜋\piitalic_π-pol) to the NiO2 planes. The inset shows the comparison with a XAS acquired in σ𝜎\sigmaitalic_σ-pol at the Ni L3-edge for a NdNiO2 sample. (b) Mid-infrared region of the σ𝜎\sigmaitalic_σ-pol RIXS spectrum along the transferred momentum 𝐐𝐐\bf{Q}bold_Q (-0.36,0). The spectrum was obtained as a sum of three single spectra, with incident energies 852.3 eV, 852.5 eV and 853.7 eV, to reduce the noise. The full-range spectrum is reported in the inset. (c) Mid-infrared region of the π𝜋\piitalic_π-pol RIXS spectrum acquired at 852.5 eV photon energy. The spectrum was obtained as a sum of four single spectra, with exchange momentum H = 0.4, 0.425, 0.45, 0.475 r.l.u. along the [H,0] direction of the Brillouin zone, to reduce the noise. The shaded colorful areas in panels b-c are the results of a fitting procedure as presented in Methods. All the measurements were performed at T = 20 K ca.𝑐𝑎ca.italic_c italic_a .

3 Discussion

The main finding of this study is the observation of a superconducting ground state in undoped PrNiO2 thin films: below, we briefly discuss our finding and its implications. The synthesis of infinite-layer nickelates remains a real challenge, and one has to make sure that the reported superconductivity in nickelates parent compounds PNO2 thin films is not due to fortuitous sources of doping. There are at least three sources of possible unintentional chemical doping: adventitious Sr-doping; Pr/Ni off-stoichiometry; and excess oxygen due to incomplete de-intercalation during the CaH2-based topotactic reduction. The combination of STEM-EELS, layer resolved O K-edge EELS, and 4D-STEM analysis on PNO2 samples in Figure 2, largely discussed in the previous section, clearly demonstrate that, locally, no relevant chemical-doping could be detected. Moreover, XAS spectra (cf.𝑐𝑓cf.italic_c italic_f . to Fig. 3a) show very sharp Ni1+ peaks, at odds with possible doping by chemical substitution, which will give signatures of Ni2+ features [18].

Here now we discuss the relevance of the structural and chemical perfection of the perovskite precursor phase, which allows to provide further support of the absence of a relevant unintentional chemical doping related to cation off stoichiometry. From the STEM-EELS map on PNO2 in Figure 2, we found that the Pr/Ni ratio is close to one. However, to further exclude Pr/Ni off-stoichiometry, in Figures 4a-d we report atomic-resolved HAADF-STEM (Figs. 4a,b) and EELS maps (Figs. 4c,d) of our STO6uc-PNO3 precursor samples. We find that the Pr/Ni ratio is equal to one within the experimental uncertainty, thus further excluding off-stoichiometry issues. Additionally, in Figure 4e we report the transport properties of STO6uc-PNO3 precursor samples deposited onto STO, and on well-matched NdGaO3 (110) single crystal grown in the same deposition conditions, thus characterized by the same Pr/Ni composition [47]. In these nickelates, the sharpness of the metal-to-insulator transition (MIT) is a valid proxy to gain major information about the Pr/Ni ratio [21]. The data show a MIT with a reduced jump in the case of the PNO3//STO samples due to strain but still a very high MIT temperature, in agreement with other studies [16], while a very large resistance jump for the PNO3//NGO film, comparable to PrNiO3 single crystals [48]. It is worth noting that the observation of a MIT in our precursor phase is a further indication of the absence of any Pr/Ni off-stoichiometry and even Sr-doping. Indeed, it is well known that the MIT is suppressed by a small amount of Pr/Ni off-stoichiometry as well as by few percent of Sr-doping. All these results imply a nearly unitary Pr/Ni ratio in our films, ruling out non-unitary Pr/Ni ratio as a source of unintentional doping, and also Sr-doping.

Refer to caption
Figure 4: (a) HAADF-STEM image of the STO6uc-PNO3 film. (b) A defect-free zoom-in region. (c) Elemental STEM-EELS maps of Sr, Pr, Ni, and Ti. (d) Elemental profiles integrated over the zoom-in region showing a Pr/Ni ratio close to one throughout the investigated sample volume. Importantly, no Sr-interdiffusion could be resolved. (e) Temperature-dependent resistivity of 16 unit-cells thick PNO3 film grown onto STO and NdGaO3 substrates, and both capped with 6 unit-cells of STO.

Beyond material imperfections, superconductivity in undoped nickelates, on the other hand, might be related to self-doping of holes due to the R5d5𝑑5d5 italic_d-Ni3d3𝑑3d3 italic_d hybridization. According to theoretical calculations, the R5d5𝑑5d5 italic_d states gives rise to partially occupied electron pockets at the ΓΓ\Gammaroman_Γ-point, and the hybridization with Ni-3dz2superscript𝑧2z^{2}italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT states at the Fermi level provides self-doped holes in the the quasi-2D 3dx2y2superscript𝑥2superscript𝑦2x^{2}-y^{2}italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT band [20]. This specific electronic configuration, while setting one of the major differences with the cuprates’ fermiology, may be responsible of a superconducting state already at zero chemical doping [49].

To summarize, we report clear experimental evidences that undoped PrNiO2 IL-nickelates are superconducting, at odds with the observation of superconductivity only in IL doped by Sr,Ca cations, which provide extra-holes in the NiO2 planes [1, 3, 33, 16]. Our results, together with the report of the onset of superconductivity in undoped LaNiO2 thin films [12], and more recently in undoped NdNiO2 thin films [13], represent an important breakthrough for a full understanding of infinite-layer nickelates physics. According to these studies, due to the multi-orbital nature of the electronic properties of nickelates, self-doped holes in the Ni-3dx2y2superscript𝑥2superscript𝑦2x^{2}-y^{2}italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT orbital-derived bands, are enough to set a superconducting state. This requires a reconsideration of the IL nickelate phase-diagram.

Methods

Thin Film Growth:

The epitaxial growth of perovskite PrNiO3 films on 5×5mm255𝑚superscript𝑚25\times 5mm^{2}5 × 5 italic_m italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT SrTiO3 substrates was performed by pulsed laser deposition using a 248 nm KrF excimer laser, with ceramic targets from Toshima Manufacturing Co. Ltd. The SrTiO3 substrates (Shinkosha Co. Ltd) were prepared by etching in an NH4F-buffered HF solution, followed by annealing for 2 hours at 950C in air to achieve a well-defined TiO2-terminated step-terraced surface. Prior to growth, the substrate was pre-annealed for 1 hour at 800C under a pressure of 0.3 mbar in oxygen flow to ensure a very clean and sharp step-and-terrace surface. The layer-by-layer growth of the PrNiO3 was monitored via Reflection High Energy Electron Diffraction (RHEED) technique. The 15-20 unit cells thick PrNiO3 films were grown at a substrate temperature of 675C with an oxygen partial pressure PO2subscript𝑂2{}_{O_{2}}start_FLOATSUBSCRIPT italic_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_FLOATSUBSCRIPT of 0.3 mbar, using a laser fluence of 4 J/cm2 and a 1×1.411.41\times 1.41 × 1.4 mm2 laser spot size on the target.For capped samples, the SrTiO3 top layers (from 1 to 12 unit cell thickness), were grown at a substrate temperature of 575C and PO2subscript𝑂2{}_{O_{2}}start_FLOATSUBSCRIPT italic_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_FLOATSUBSCRIPT = 0.3 mbar, with a laser fluence of 1.3 J/cm2 and a 1×1.411.41\times 1.41 × 1.4 mm2 laser spot size. After growth, the samples were cooled to room temperature at a rate of 5C/min in the same oxidizing growth conditions.

Topochemical Reduction (infinite-layer formation):

After growth, each sample was cut into two pieces of size 5×2.552.55\times 2.55 × 2.5 mm2 using a precision diamond wire saw (Well 3242). This method is used to cut the samples to avoid stress during the cutting process, which can introduce defects. The pieces of each sample to be reduced were placed in an evacuated silica tube sealed with a membrane valve, with 0.5 g of CaH2 powder in direct contact, as used in prior studies [50]. The tube was heated to 260C at a rate of 5C/min, and held at this temperature for varying durations (2-8 hours) depending on the PrNiO3 and SrTiO3 capping thicknesses; then it was cooled to room temperature at a rate of 5C/min. The process was optimized through a systematic series of steps, with θ𝜃\thetaitalic_θ-2θ2𝜃2\theta2 italic_θ scans and electrical transport measurements performed ex situ after each step to assess the degree of reduction. In particular, the highly reproducible zero resistance state is obtained via consecutive CaH2-based topotactic reduction steps for samples prepared with a STO capping layer (from 3 up to 12 unit-cells), while a superconducting transition is always encountered irrespective of the presence/absence of the capping layer and/or number of reduction cycles.

Characterization:

The surface morphology of the samples was examined using a Park XE7 (Park System) atomic force microscope (AFM) in true non-contact mode. XRD measurements were performed using a Rigaku Smartlab diffractometer with a Cu-Kα𝛼\alphaitalic_α radiation source (0.154056 nm). For transport measurements, the 2.5×\times×5 mm2 samples were wire-bonded directly with Al, without the use of top-electrodes. The wire bonds were placed at the four edges of the samples, and the measurements were conducted using the Van der Pauw method, with a current amplitude of 10 μ𝜇\muitalic_μA, with a cryo-free Dynacool system (Quantum Design).

High resolution STEM-EELS:

The cross-sectional focused ion beam (FIB) transmission electron microscopy (TEM) lamellae were prepared at C2N, University of Paris-Saclay, France. Before FIB lamellae preparation, around 20 nm of amorphous carbon was deposited on top for protection. For additional protection, electron beam-induced deposition of platinum and ion-beam beam-induced deposition of platinum were done. The HAADF imaging and EELS were carried out in a NION UltraSTEM 200 C3/C5-corrected STEM. The experiments were done at 200 keV with a probe current of approximately 14 pA and convergence semi-angles of 30 mrad. The EELS spectra were obtained using the full 4 × 1 configuration of a MerlinEM detector (Quantum Detectors Ltd) installed on a Gatan ENFINA spectrometer mounted on the microscope [51]. The EELS spectrometer was set into non-energy dispersive trajectories for 4D-STEM experiments. Data was collected with only one chip of the MerlinEM detector in 6-bit mode that enables faster acquisition without compromising on the signal-to-noise ratio. The high resolution EELS fine structure analysis was done in a monochromated and C3/C5 corrected NION Chromatem microscope operating at 100 keV, with a probe current of 30 pA, convergence semi-angles of 25 mrad.

XAS, RIXS:

Measurements were performed at the ID32 soft x-ray beamline of the ESRF, Grenoble, France. XAS spectra were measured at 10 grazing incidence, with linear polarization forming an angle of 0 or 80 with respect to the NiO2 planes, and labelled for brevity parallel-to\parallelNiO2 and perpendicular-to\perpNiO2, respectively. For RIXS, the combined resolution at the Ni L3-edge was of 40 meV. The scattering angle 2θ𝜃\thetaitalic_θ was fixed at 149.5. RIXS energy-resolved maps were acquired at an incident angle of θ=30𝜃superscript30\theta=30^{\circ}italic_θ = 30 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT (grazing-in geometry), corresponding to an exchanged in-plane momentum of about 0.36 relative lattice units (r.l.u.). Momentum-resolved maps were taken at the XAS resonance energy of 852.4 eV ca. in π𝜋\piitalic_π incident polarization, which is known to enhance magnetic excitations. Grazing-out geometry was adopted in this case, with incident angles θ𝜃\thetaitalic_θ between 80 and 140. All RIXS and XAS measurements were performed at 20 K, the lowest temperature safely reachable by the ID32 cooling apparatus.

Supporting Information

Growth of SrTiO3(d)/PrNiO3//SrTiO3 heterostructures

Refer to caption
Figure S1: a) X-ray diffraction θ𝜃\thetaitalic_θ-2θ𝜃\thetaitalic_θ symmetric scans of PNO3 sample series with varying thicknesses of the STO capping layer. Below we show a zoomed-in around the (002) diffraction peak showing the peculiar modulation as described in the text together the fitting curves (red dotted lines), obtained by considering the presence of an epitaxial and coherent STO layer of 6uc (b) and (c) 12 uc as thickness. The fitting curves are obtained by using the software described in Ref. [52].

In Figure S1a, we show the XRD θ𝜃\thetaitalic_θ-2θ𝜃\thetaitalic_θ symmetric scans of 16 unit cells (uc) thick PNO3//STO films capped with varying thicknesses (d) of STO, ranging from 0 to 12 uc. For all samples, intense (00\ellroman_ℓ) diffraction peaks (\ellroman_ℓ = 1, 2, 3) are observed, confirming the high quality of the perovskite nickelate phase [50]. From the (00\ellroman_ℓ) peak positions, we calculate a c-axis lattice parameter of 0.375 nm, consistent with a fully strained and stoichiometric PNO3 thin film. Accounting for a Poisson ratio of 0.3, the expected out-of-plane lattice parameter for a tensile-strained PNO3 thin film onto STO is approximately of 0.375 nm, thus confidently ruling out any possible presence of off-stoichiometry and/or oxygen vacancies [21, 22]. A zoom-in around the (002) diffraction peaks (Figures S1b,c) reveals an intensity modulation of the peak profile for the PNO3 sample prepared with an STO capping layer thicker than 6uc. The HAADF-STEM results for the STO6uc sample (Figure 4b) clearly demonstrate that this intensity modulation of the (002) XRD peak profile is not due to secondary phases or hole-vacancies, but is most likely a direct consequence of the STO capping layer itself as, indeed, we carefully demonstrated by performing a fitting procedure by using a tool already used in literature [52]. The red dotted lines superimposed to the XRD pattern of our STO6uc-PNO3 and STO12uc-PNO3 samples properly reproduce the observed PNO3(002) XRD peak modulation.

Structural and transport properties of SrTiO3(d)/PrNiO2222//SrTiO3 heterostructures

Refer to caption
Figure S2: a) X-ray diffraction θ𝜃\thetaitalic_θ-2θ𝜃\thetaitalic_θ symmetric scans of STO6uc- and STO1uc-PNO2 samples, clearly showing the formation of the infinite-layer (IL) phase. b) Reciprocal space mapping performed around the asymmetric (103) STO diffraction peak, indicating that the samples are fully strained and exhibit a 0.333 nm c-axis parameter.

To stabilize the PrNiO2 (PNO2) infinite-layer (IL) phase, we employed a CaH2-based topotactic reduction process. Figure S2a shows the XRD θ𝜃\thetaitalic_θ-2θ𝜃\thetaitalic_θ symmetric scans of fully reduced samples capped with STO of 6 unit cells (STO6uc) and 1 unit cell (STO1uc), exhibiting robust (00\ellroman_ℓ) family peaks at the expected positions, with a c𝑐citalic_c-axis of 0.333 nm, confirming the complete formation of the IL phase. Notably, features on the left of the main (00\ellroman_ℓ) diffraction peaks in the STO6uc-capped sample are likely attributed to Laue fringes, indicating the high quality of the IL phase, stabilized coherently by the appropriate thickness of the capping layer [35, 3].

Refer to caption
Figure S3: (Top) HAADF-STEM images for a PNO2 sample capped with 6 unit cells of STO, with (Bottom) GPA analysis applied, confirming a 15-16% reduction in the out-of-plane c-axis and full in-plane strain of the IL-phase. This reduction confirm the overall XRD result and shown in S2

Reciprocal Space Mapping (RSM) measurements around the asymmetric STO(-103) and PNO2(-103) diffraction peaks confirm that the PNO2 films are fully strained to the substrate and show no other features, as shown in Figure S2b. This is further corroborated by a geometrical phase analysis (GPA) map obtained via HAADF-STEM measurements, shown in Figure S3, which displays a homogeneous 15% reduction in the c𝑐citalic_c-axis (GPA-ϵxxsubscriptitalic-ϵ𝑥𝑥\epsilon_{xx}italic_ϵ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT), supporting the macroscopic observations via XRD and indicating a fully strained IL-phase to the substrate (GPA-ϵyysubscriptitalic-ϵ𝑦𝑦\epsilon_{yy}italic_ϵ start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT).

In Figure S4 we show the transport properties of the full PNO2 series prepared with different STO(d) capping layer thicknesses (left), and also a series of PNO2(d) samples with the same STO6uc capping layer and overall similar conditions for the topotactic reduction. While the superconducting transition is always present, only samples with a capping layer thicker than 6 unit cells reach the zero resistance state. On the other side the zero-resistance state in encountered only for sample 16uc thick. This largely demonstrate that the structural quality of the precursor phase (eventually worsened for thicker sample) is a key parameter to obtain a superconducting state with a clear zero resistance state. This demonstrates, beyond the HAADF-STEM data, that the PNO2 superconductivity is not linked to any possible source of Sr interdiffusion.

Refer to caption
Figure S4: Temperature-dependent resistivity of the superconducting PNO2 sample series, prepared with (a) varying unit cells of the STO-capping layer but constant PNO2 thickness and (b) varying unit cells of PNO2 but constant STO-capping layer. The zero resistance state is achieved for samples with moderate PNO2 thickness (16 uc) and STO layers thicker than 6 unit cells.

Energy dependence of the magnon

As already suggested by the colour maps in Fig. 3(b), the energy dependence of the single magnon follows a peculiar trend with respect to the other main spectral features (e.g., elastic peak, phonon). In particular, as shown in Fig. S5, the magnon intensity resonates before the absorption edge, 0.4similar-to-or-equalsabsent0.4\simeq 0.4≃ 0.4 eV de-tuned from the XAS peak, as opposed to the elastic and the phonon intensity. Moreover, the magnon resonance appears to be significantly narrower than the other two features presented in Fig. S5(b).

Refer to caption
Figure S5: (a) RIXS spectrum at the magnon resonance, taken with π𝜋\piitalic_π incident polarization. The arrows indicate the main spectral features in the low energy scale: the elastic line, the phonon peak and the magnon spectral weight. (b) Energy dependence of the different spectral features highlighted in panel (a) across the Ni L3 edge. The integration intervals are [-0.03 eV, 0.03 eV], [0.03 eV, 0.1 eV] and [0.1 eV, 0.35 eV] for the elastic, phonon and magnon respectively.

Fitting of RIXS spectra

Here we introduce the fitting procedure for the results for the RIXS spectra showed in Figure 3c of the main text. A Damped Harmonic Oscillator susceptibility is employed for the magnetic peak, while two resolution-wide Gaussians fit the elastic and phonon contribution. The tail of the broad 0.6 eV peak due to Pr5d5𝑑5d5 italic_d hybridization is taken into account with a linear background. Fittings close to the Gamma point are complicated by the softening magnon, which tends to merge with other low-energy features. This is testified by the large error bars on the damping at such values of Q. Nevertheless, our fitting results are in good agreement with literature [43, 53], including our very recent measurements on NdNiO2 [42].

\bmhead

Acknowledgments The authors thank enriching discussions with K.M. Shen. This work was funded by the French National Research Agency (ANR) through the ANR-JCJC FOXIES ANR-21-CE08-0021. This work was also done as part of the Interdisciplinary Thematic Institute QMat, ITI 2021 2028 program of the University of Strasbourg, CNRS and Inserm, and supported by IdEx Unistra (ANR 10 IDEX 0002), and by SFRI STRAT’US project (ANR 20 SFRI 0012) and EUR QMAT ANR-17-EURE-0024 under the framework of the French Investments for the Future Program. The authors would like to thank the PLD, XRD, MEB-CRO, TEM and MagTransCS platforms of the IPCMS. The RIXS experiments were performed at ESRF Synchrotron facility in Grenoble (France) under proposal number SC-5438. The ID32 staff is also acknowledged for technical support.

References

  • \bibcommenthead
  • Li et al. [2019] 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) https://doi.org/10.1038/s41586-019-1496-5
  • Li et al. [2020] Li, Q., He, C., Si, J., Zhu, X., Zhang, Y., Wen, H.-H.: Absence of superconductivity in bulk nd1-xsrxnio2. Communications Materials 1, 16 (2020) https://doi.org/10.1038/s43246-020-0018-1
  • Lee et al. [2020] 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(4), 41107 (2020) https://doi.org/10.1063/5.0005103
  • Zeng et al. [2020] 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. Physical Review Letters 125 (2020) https://doi.org/10.1103/PhysRevLett.125.147003
  • [5] Raji, A., Krieger, G., Viart, N., Preziosi, D., Rueff, J.-P., Gloter, A.: Charge distribution across capped and uncapped infinite-layer neodymium nickelate thin films. Small n/a, 2304872 https://doi.org/10.1002/smll.202304872
  • Parzyck et al. [2024] Parzyck, C.T., Gupta, N.K., Wu, Y., Anil, V., Bhatt, L., Bouliane, M., Gong, R., Gregory, B.Z., Luo, A., Sutarto, R., He, F., Chuang, Y.D., Zhou, T., Herranz, G., Kourkoutis, L.F., Singer, A., Schlom, D.G., Hawthorn, D.G., Shen, K.M.: Absence of 3a 0 charge density wave order in the infinite-layer nickelate ndnio2. Nature Materials (2024) https://doi.org/%****␣Draft.bbl␣Line␣175␣****10.1038/s41563-024-01797-0
  • Ghiringhelli [2024] Ghiringhelli, G.: A noticeable absence. Nature Materials 23, 443–444 (2024) https://doi.org/10.1038/s41563-024-01835-x
  • Lee et al. [2023] Lee, K., Wang, B.Y., Osada, M., Goodge, B.H., Wang, T.C., Lee, Y., Harvey, S., Kim, W.J., Yu, Y., Murthy, C., Raghu, S., Kourkoutis, L.F., Hwang, H.Y.: Linear-in-temperature resistivity for optimally superconducting (nd,sr)nio2. Nature 2023 619:7969 619, 288–292 (2023) https://doi.org/10.1038/s41586-023-06129-x
  • Parzyck et al. [2024] Parzyck, C.T., Anil, V., Wu, Y., Goodge, B.H., Roddy, M., Kourkoutis, L.F., Schlom, D.G., Shen, K.M.: Synthesis of thin film infinite-layer nickelates by atomic hydrogen reduction: Clarifying the role of the capping layer. APL Materials 12, 031132 (2024) https://doi.org/10.1063/5.0197304
  • Wei et al. [2023] Wei, W., Vu, D., Zhang, Z., Walker, F.J., Ahn, C.H.: Superconducting nd1-xeuxnio2 thin films using in situ synthesis. Science Advances 9(27), 3327 (2023) https://doi.org/%****␣Draft.bbl␣Line␣250␣****10.1126/sciadv.adh3327
  • Sun et al. [2024] Sun, W., Wang, Z., Hao, B., Yan, S., Sun, H., Gu, Z., Deng, Y., Nie, Y.: In situ preparation of superconducting infinite-layer nickelate thin films with atomically flat surface. Advanced Materials 36, 2401342 (2024) https://doi.org/10.1002/adma.202401342
  • Osada et al. [2021] 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. Advanced Materials 33, 2104083 (2021) https://doi.org/10.1002/adma.202104083
  • Parzyck et al. [2024] Parzyck, C.T., Wu, Y., Bhatt, L., Kang, M., Arthur, Z., Pedersen, T.M., Sutarto, R., Fan, S., Pelliciari, J., Bisogni, V., Herranz, G., Georgescu, A.B., Hawthorn, D.G., Kourkoutis, L.F., Muller, D.A., Schlom, D.G., Shen, K.M.: Superconductivity in the parent infinite-layer nickelate NdNiO2 (2024). https://arxiv.org/abs/2410.02007
  • Li et al. [2020] 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. Physical Review Letters 125, 27001 (2020) https://doi.org/10.1103/PhysRevLett.125.027001
  • Zeng et al. [2020] 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. Physical Review Letters 125 (2020) https://doi.org/10.1103/PhysRevLett.125.147003
  • Osada et al. [2020] Osada, M., Wang, B.Y., Lee, K., Li, D., Hwang, H.Y.: Phase diagram of infinite layer praseodymium nickelate {{pr}}{1{}x}{{sr}}{x}{{nio}}{2}\{\mathrm{\{}pr\}\}_{\{}1\{-\}x\}\{\mathrm{\{}sr\}\}_{\{}x\}\{\mathrm{\{}nio\}% \}_{\{}2\}{ { italic_p italic_r } } start_POSTSUBSCRIPT { end_POSTSUBSCRIPT 1 { - } italic_x } { { italic_s italic_r } } start_POSTSUBSCRIPT { end_POSTSUBSCRIPT italic_x } { { italic_n italic_i italic_o } } start_POSTSUBSCRIPT { end_POSTSUBSCRIPT 2 } thin films. Physical Review Materials 4, 121801 (2020) https://doi.org/10.1103/PhysRevMaterials.4.121801
  • Zeng et al. [2022] Zeng, S., Li, C., Chow, L.E., Cao, Y., Zhang, Z., Tang, C.S., Yin, X., Lim, Z.S., Hu, J., Yang, P., Ariando, A.: Superconductivity in infinite-layer nickelate la1-xcaxnio2 thin films. Science Advances 8, 9927 (2022) https://doi.org/10.1126/SCIADV.ABL9927/SUPPL_FILE/SCIADV.ABL9927_SM.PDF
  • Krieger et al. [2024] Krieger, G., Sahib, H., Rosa, F., Rath, M., Chen, Y., Raji, A., Pinho, P.V.B., Lefevre, C., Ghiringhelli, G., Gloter, A., Viart, N., Salluzzo, M., Preziosi, D.: Signatures of canted antiferromagnetism in infinite-layer nickelates studied by x-ray magnetic dichroism (2024). https://arxiv.org/abs/2403.16969
  • Zeng et al. [2024] Zeng, S., Tang, C.S., Luo, Z., Chow, L.E., Lim, Z.S., Prakash, S., Yang, P., Diao, C., Yu, X., Xing, Z., et al.: Origin of a topotactic reduction effect for superconductivity in infinite-layer nickelates. Physical Review Letters 133, 66503 (2024) https://doi.org/10.1103/PhysRevLett.133.066503
  • Karp et al. [2020] Karp, J., Botana, A.S., Norman, M.R., Park, H., Zingl, M., Millis, A.: Many-Body Electronic Structure of NdNiO2 and CaCuO2. Phys. Rev. X 10(2), 21061 (2020) https://doi.org/10.1103/PhysRevX.10.021061
  • Preziosi et al. [2017] Preziosi, D., Sander, A., Barthélémy, A., Bibes, M.: Reproducibility and off-stoichiometry issues in nickelate thin films grown by pulsed laser deposition. AIP Adv. 7(1), 015210 (2017) https://doi.org/10.1063/1.4975307
  • Breckenfeld et al. [2014] Breckenfeld, E., Chen, Z., Damodaran, A.R., Martin, L.W.: Effects of nonequilibrium growth, nonstoichiometry, and film orientation on the metal-to-insulator transition in ndnio3 thin films. ACS Applied Materials and Interfaces 6, 22436–22444 (2014) https://doi.org/10.1021/am506436s
  • Guo et al. [2020] Guo, Q., Farokhipoor, S., Magén, C., Rivadulla, F., Noheda, B.: Tunable resistivity exponents in the metallic phase of epitaxial nickelates. Nat. Commun. 11(1), 2949 (2020) https://doi.org/10.1038/s41467-020-16740-5
  • Li et al. [2021] Li, Y., Sun, W., Yang, J., Cai, X., Guo, W., Gu, Z., Zhu, Y., Nie, Y.: Impact of cation stoichiometry on the crystalline structure and superconductivity in nickelates. Frontiers in Physics 9 (2021) https://doi.org/10.3389/fphy.2021.719534
  • Catalano et al. [2018] Catalano, S., Gibert, M., Fowlie, J., Íñiguez, J., Triscone, J.-M., Kreisel, J.: Rare-earth nickelates r nio 3 : thin films and heterostructures. Reports on Progress in Physics 81, 046501 (2018) https://doi.org/10.1088/1361-6633/aaa37a
  • Gawryluk et al. [2019] Gawryluk, D.J., Klein, Y.M., Shang, T., Sheptyakov, D., Keller, L., Casati, N., Lacorre, P., Fernández-Díaz, M.T., Rodríguez-Carvajal, J., Medarde, M.: Distortion mode anomalies in bulk prnio3subscriptprnio3{\mathrm{prnio}}_{3}roman_prnio start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT: Illustrating the potential of symmetry-adapted distortion mode analysis for the study of phase transitions. Phys. Rev. B 100, 205137 (2019) https://doi.org/10.1103/PhysRevB.100.205137
  • Lin et al. [2022] Lin, H., Gawryluk, D.J., Klein, Y.M., Huangfu, S., Pomjakushina, E., Rohr, F., Schilling, A.: Universal spin-glass behaviour in bulk lanio2, prnio2 and ndnio2. New Journal of Physics 24(1), 013022 (2022) https://doi.org/10.1088/1367-2630/ac465e
  • Lee et al. [2023] Lee, K., Wang, B.Y., Osada, M., Goodge, B.H., Wang, T.C., Lee, Y., Harvey, S., Kim, W.J., Yu, Y., Murthy, C., Raghu, S., Kourkoutis, L.F., Hwang, H.Y.: Linear-in-temperature resistivity for optimally superconducting (nd,sr)nio2. Nature 2023 619:7969 619, 288–292 (2023) https://doi.org/10.1038/s41586-023-06129-x
  • Ren et al. [2023] Ren, X., Li, J., Chen, W.-C., Gao, Q., Sanchez, J.J., Hales, J., Luo, H., Rodolakis, F., McChesney, J.L., Xiang, T., Hu, J., Comin, R., Wang, Y., Zhou, X., Zhu, Z.: Possible strain-induced enhancement of the superconducting onset transition temperature in infinite-layer nickelates. Communications Physics 6, 341 (2023) https://doi.org/10.1038/s42005-023-01464-x
  • Cataldo et al. [2024] Cataldo, S.D., Worm, P., Tomczak, J.M., Si, L., Held, K.: Unconventional superconductivity without doping in infinite-layer nickelates under pressure. Nature Communications 15, 3952 (2024) https://doi.org/10.1038/s41467-024-48169-5
  • Osada et al. [2020] Osada, M., Wang, B.Y., Goodge, B.H., Lee, K., Yoon, H., Sakuma, K., Li, D., Miura, M., Kourkoutis, L.F., Hwang, H.Y.: A superconducting praseodymium nickelate with infinite layer structure. Nano Letters 20, 5735–5740 (2020) https://doi.org/10.1021/acs.nanolett.0c01392
  • Gutiérrez-Llorente et al. [2024] Gutiérrez-Llorente, A., Raji, A., Zhang, D., Divay, L., Gloter, A., Gallego, F., Galindo, C., Bibes, M., Iglesias, L.: Toward reliable synthesis of superconducting infinite layer nickelate thin films by topochemical reduction. Advanced Science n/a, 2309092 (2024) https://doi.org/10.1002/advs.202309092
  • Shengwei et al. [2022] Shengwei, Z., Changjian, L., Er, C.L., Yu, C., Zhaoting, Z., Sin, T.C., Xinmao, Y., Shiuh, L.Z., Junxiong, H., Ping, Y., Ariando, A.: Superconductivity in infinite-layer nickelate La1-xCaxNiO2 thin films. Sci. Adv. 8(7), 9927 (2022) https://doi.org/10.1126/sciadv.abl9927
  • Goodge et al. [2023] Goodge, B.H., Geisler, B., Lee, K., Osada, M., Wang, B.Y., Li, D., Hwang, H.Y., Pentcheva, R., Kourkoutis, L.F.: Resolving the polar interface of infinite-layer nickelate thin films. Nature Materials 22, 466–473 (2023) https://doi.org/10.1038/s41563-023-01510-7
  • Raji et al. [2024] Raji, A., Gutiérrez-Llorente, A., Zhang, D., Li, X., Bibes, M., Iglesias, L., Rueff, J.-P., Gloter, A.: Unraveling p-type and n-type interfaces in superconducting infinite-layer nickelate thin films. Advanced Functional Materials n/a, 2409930 (2024) https://doi.org/10.1002/adfm.202409930
  • Geisler and Pentcheva [2020] Geisler, B., Pentcheva, R.: Fundamental difference in the electronic reconstruction of infinite-layer versus perovskite neodymium nickelate films on srtio3subscriptsrtio3{\mathrm{srtio}}_{3}roman_srtio start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT(001). Phys. Rev. B 102, 020502 (2020) https://doi.org/10.1103/PhysRevB.102.020502
  • Goodge et al. [2021] 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. 118(2), 2007683118 (2021) https://doi.org/10.1073/pnas.2007683118
  • Bisogni et al. [2016] Bisogni, V., Catalano, S., Green, R.J., Gibert, M., Scherwitzl, R., Huang, Y., Strocov, V.N., Zubko, P., Balandeh, S., Triscone, J.M., Sawatzky, G., Schmitt, T.: Ground-state oxygen holes and the metal-insulator transition in the negative charge-transfer rare-earth nickelates. Nat. Commun. 7, 17–19 (2016) https://doi.org/10.1038/ncomms13017
  • Rossi et al. [2021] Rossi, M., Lu, H., Nag, A., Li, D., Osada, M., Lee, K., Wang, B.Y., Agrestini, S., Garcia-Fernandez, M., 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(22), 220505 (2021) https://doi.org/10.1103/PhysRevB.104.L220505
  • Braicovich et al. [2020] Braicovich, L., Rossi, M., Fumagalli, R., Peng, Y., Wang, Y., Arpaia, R., Betto, D., De Luca, G.M., Di Castro, D., Kummer, K., et al.: Determining the electron-phonon coupling in superconducting cuprates by resonant inelastic x-ray scattering: Methods and results on nd 1+ x ba 2- x cu 3 o 7- δ𝛿\deltaitalic_δ. Physical Review Research 2(2), 023231 (2020) https://doi.org/10.1103/PhysRevResearch.2.023231
  • Gao et al. [2024] Gao, Q., Fan, S., Wang, Q., Li, J., Ren, X., Biało, I., Drewanowski, A., Rothenbühler, P., Choi, J., Sutarto, R., Wang, Y., Xiang, T., Hu, J., Zhou, K.-J., Bisogni, V., Comin, R., Chang, J., Pelliciari, J., Zhou, X.J., Zhu, Z.: Magnetic excitations in strained infinite-layer nickelate prnio2 films. Nature Communications 15, 5576 (2024) https://doi.org/10.1038/s41467-024-49940-4
  • Rosa et al. [2024] Rosa, F., Martinelli, L., Krieger, G., Braicovich, L., Brookes, N.B., Merzoni, G., Sala, M.M., Yakhou-Harris, F., Arpaia, R., Preziosi, D., et al.: Spin excitations in nd1-xsrxnio2 and yba2cu3o7-delta: the influence of hubbard u. arXiv preprint arXiv:2406.09271 (2024) https://doi.org/10.48550/arXiv.2406.09271
  • Lu et al. [2021] Lu, H., Rossi, M., Nag, A., Osada, M., Li, D.F., Lee, K., Wang, B.Y., Garcia-Fernandez, M., Agrestini, S., Shen, Z.X., Been, E.M., Moritz, B., Devereaux, T.P., Zaanen, J., Hwang, H.Y., Zhou, K.-J., Lee, W.S.: Magnetic excitations in infinite-layer nickelates. Science 373(6551), 213–216 (2021) https://doi.org/10.1126/science.abd7726
  • Fan et al. [2024] Fan, S., LaBollita, H., Gao, Q., Khan, N., Gu, Y., Kim, T., Li, J., Bhartiya, V., Li, Y., Sun, W., et al.: Capping effects on spin and charge excitations in parent and superconducting nd1-xsrxnio2. arXiv preprint arXiv:2409.18258 (2024) https://doi.org/10.48550/arXiv.2409.18258
  • Hepting et al. [2020] Hepting, M., Li, D., Jia, C.J., Lu, H., Paris, E., Tseng, Y., Feng, X., Osada, M., Been, E., Hikita, Y., Chuang, Y.-D., Hussain, Z., Zhou, K.J., Nag, A., Garcia-Fernandez, M., Rossi, M., Huang, H.Y., Huang, D.J., Shen, Z.X., Schmitt, T., Hwang, H.Y., Moritz, B., Zaanen, J., Devereaux, T.P., Lee, W.S.: Electronic structure of the parent compound of superconducting infinite-layer nickelates. Nat. Mater. 19, 381–385 (2020) https://doi.org/10.1038/s41563-019-0585-z
  • Krieger et al. [2022] Krieger, G., Martinelli, L., Zeng, S., Chow, L.E., Kummer, K., Arpaia, R., Sala, M.M., Brookes, N.B., Ariando, A., Viart, N., Salluzzo, M., Ghiringhelli, G., Preziosi, D.: Charge and spin order dichotomy in ndnio2 driven by the capping layer. Phys. Rev. Lett. 129, 27002 (2022) https://doi.org/10.1103/PhysRevLett.129.027002
  • Suyolcu et al. [2021] Suyolcu, Y.E., Fürsich, K., Hepting, M., Zhong, Z., Lu, Y., Wang, Y., Christiani, G., Logvenov, G., Hansmann, P., Minola, M., Keimer, B., Aken, P.A., Benckiser, E.: Control of the metal-insulator transition in ndnio3 thin films through the interplay between structural and electronic properties. Phys. Rev. Mater. 5, 045001 (2021) https://doi.org/10.1103/PhysRevMaterials.5.045001
  • Saito et al. [2003] Saito, T., Azuma, M., Nishibori, E., Takata, M., Sakata, M., Nakayama, N., Arima, T., Kimura, T., Urano, C., Takano, M.: Monoclinic distortion in the insulating phase of prnio3. Physica B: Condensed Matter 329-333, 866–867 (2003) https://doi.org/10.1016/S0921-4526(02)02584-X
  • Pickett [2020] Pickett, W.E.: The dawn of the nickel age of superconductivity. Nat. Rev. Phys. (2020) https://doi.org/10.1038/s42254-020-00257-3
  • Krieger et al. [2023] 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. Journal of Physics D: Applied Physics 56, 024003 (2023) https://doi.org/10.1088/1361-6463/aca54a
  • Blank et al. [2014] Blank, D.H.A., Dekkers, M., Rijnders, G.: Pulsed laser deposition in twente: from research tool towards industrial deposition. Journal of Physics D: Applied Physics 47, 034006 (2014) https://doi.org/10.1088/0022-3727/47/3/034006
  • Lichtensteiger [2018] Lichtensteiger, C.: Interactivexrdfit: a new tool to simulate and fit x-ray diffractograms of oxide thin films and heterostructures. Journal of applied crystallography 51(6), 1745–1751 (2018) https://doi.org/10.1107/S1600576718012840
  • Rossi et al. [2024] Rossi, M., Lu, H., Lee, K., Goodge, B., Choi, J., Osada, M., Lee, Y., Li, D., Wang, B., Jost, D., et al.: Universal orbital and magnetic structures in infinite-layer nickelates. Physical Review B 109(2), 024512 (2024) https://doi.org/10.1103/PhysRevB.109.024512