Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Investigation of spin excitations and charge order in bulk crystals of the infinite-layer nickelate LaNiO2

S. Hayashida s.hayashida@fkf.mpg.de Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany    V. Sundaramurthy Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany    P. Puphal Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany    M. Garcia-Fernandez Diamond Light Source, Harwell Campus, Didcot, Oxfordshire OX11 0DE, United Kingdom    Ke-Jin Zhou Diamond Light Source, Harwell Campus, Didcot, Oxfordshire OX11 0DE, United Kingdom    B. Fenk Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany    M. Isobe Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany    M. Minola Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany    Y.-M. Wu Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany    Y. E. Suyolcu Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany    P. A. van Aken Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany    B. Keimer b.keimer@fkf.mpg.de Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany    M. Hepting hepting@fkf.mpg.de Max-Planck-Institute for Solid State Research, Heisenbergstraβ𝛽\betaitalic_βe 1, 70569 Stuttgart, Germany
(June 4, 2024)
Abstract

Recent x-ray spectroscopic studies have revealed spin excitations and charge density waves in thin films of infinite-layer (IL) nickelates. However, clarifying whether the origin of these phenomena is intrinsic to the material class or attributable to impurity phases in the films has presented a major challenge. Here we utilize topotactic methods to synthesize bulk crystals of the IL nickelate LaNiO2 with crystallographically oriented surfaces. We examine these crystals using resonant inelastic x-ray scattering (RIXS) at the Ni L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge to elucidate the spin and charge correlations in the bulk of the material. While we detect the presence of prominent spin excitations in the crystals, fingerprints of charge order are absent at the ordering vectors identified in previous thin-film studies. These results contribute to the understanding of the bulk properties of LaNiO2 and establish topotactically synthesized crystals as viable complementary specimens for spectroscopic investigations.

I Introduction

Unconventional superconductors, including cuprates, iron pnictides, and heavy fermion systems, are at the forefront of condensed-matter research. While the microscopic mechanisms behind their superconductivity remain a subject of debate, spin and charge instabilities are recurrent features in their phase diagrams [1, 2, 3, 4, 5]. The magnetic phase is typically most pronounced in the parent materials, while charge carrier doping and/or external pressure suppress the long-range magnetic order and facilitate the emergence of superconductivity [6]. This suppression of long-range magnetic order transforms well-defined spin-wave excitations (magnons) of the parent materials into heavily damped excitations (paramagnons)  [7, 8, 9], which can persist over wide regions of the phase diagrams [10, 11, 12, 13]. In cuprates, the spin excitations arise from the two-dimensional CuO2 planes, exhibiting a notable bandwidth of several hundred meV [14, 15, 16, 17] due to strong exchange couplings (J𝐽Jitalic_J), and are a prime candidate for the superconducting pairing glue [1].

In addition, the ubiquitous charge density wave (CDW) ordering phenomenon might have a profound impact on the emergence of high-temperature conductivity in cuprates. These density waves are characterized by periodic electron density modulations within the CuO2 planes, which compete with the superconducting state, at least in the underdoped regime [18, 19, 20, 21]. A more complex interplay between superconductivity and various types of short-range and dynamic charge fluctuations has been observed at higher doping levels, and is a current focus of research on the cuprates [22, 23, 24, 25].

In this context, recent reports of charge order in the undoped to lightly doped regime  [26, 27, 28, 29] of the phase diagram of infinite-layer (IL) nickelates superconductors with composition (La,Pr,Nd)1-xSrxNiO2 [30, 31, 32, 33, 34, 35] were not entirely unexpected, given their numerous shared basic characteristics with cuprate superconductors [36]. In particular, their structural motif of square-lattice NiO2 planes along with a nominal 3d9superscript𝑑9d^{9}italic_d start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT electronic configuration draws close parallels to cuprates [37]. However, this class of materials also features significant distinctions [38], such as a lack of the cuprate-typical 3d3𝑑3d3 italic_d-2p2𝑝2p2 italic_p orbital hybridization between the transition metal ion and oxygen. Instead, rare-earth 5d5𝑑5d5 italic_d states contribute to the low-energy electronic structure of IL nickelates [39, 40, 41]. Moreover, long-range antiferromagnetic order is absent in the parent compounds of IL nickelates [42, 43, 44, 45, 46, 47, 48, 49], although they display dispersive magnetic excitations [50, 26, 27, 36, 51] that are reminiscent of paramagnons in doped cuprates [10, 11]. As a consequence, fundamental questions about the nature of spin excitations in nickelate superconductors have remained open.

Yet, the CDW phenomenon in IL nickelates is also debated controversially [52, 53, 54, 55]. One set of studies reported charge order in NdNiO2 thin films without a capping layer [26, 27], while the corresponding signal at approximately 𝐪=(±1/3,0)𝐪plus-or-minus130\mathbf{q}=(\pm 1/3,0)bold_q = ( ± 1 / 3 , 0 ) was absent in capped films. By contrast, spin-wave excitations materialized in films with capping layer  [27, 50], but were elusive in uncapped films [26, 27]. On the other hand, a different study reported the simultaneous presence of both charge order and spin excitations in LaNiO2 films with capping [28], while charge order signal was not detected in PrNiO2 and NdNiO2 films [29]. Along these lines, recent experiments using scanning transmission electron microscopy (STEM) and x-ray scattering have cast doubt on the intrinsic origin of the CDW [54, 55]. One proposed explanation for the observed charge ordering signal involves an impurity phase, in which excess oxygen ions fill certain vacant apical positions in the IL structure, aligning in a pattern that is commensurate with a 𝐪=(±1/3,0)𝐪plus-or-minus130\mathbf{q}=(\pm 1/3,0)bold_q = ( ± 1 / 3 , 0 ) ordering vector.

To examine if previously reported spin and charge instabilities in thin-film samples originated from external factors, including interfacial effects with capping layers, oxygen off-stoichiometries, and/or damages from the invasive topochemical synthesis, we focus here on bulk crystals of the IL nickelate LaNiO2. These bulk specimens with mechanically polished surfaces offer a contrasting approach to thin films, as the material exposed after polishing was neither affected by the invasive topotactic reduction atmosphere nor by contact with a capping layer. Using resonant inelastic x-ray scattering (RIXS) on a polished LaNiO2 surface, we detect low-energy spectral features that are attributable to spin excitations. We demonstrate that the observed spin excitation spectrum can be qualitatively reproduced with linear spin-wave theory and the magnetic exchange parameters known from thin films and that the lack of apparent dispersion is a consequence of the presence of three twin domains in our crystals. Hence, we conclude that the underlying spin excitation phenomenology in LaNiO2 crystals aligns with that of thin films. Conversely, a CDW signal at the ordering vectors reported in previous thin-film studies is not observed in our RIXS experiment on LaNiO2 crystals.

II Materials and methods

Single crystals of perovskite LaNiO3 were grown by the high-pressure optical floating zone (OFZ) method, as described in Refs. [56, 57]. The as-grown centimeter-sized crystals were oriented with x-ray Laue diffraction and cut into smaller cube-shaped crystals with a wire saw, such that each surface of a cube corresponds to a pseudocubic (100) plane. The lateral dimensions of the surfaces of the cut crystals were approximately 0.8×0.80.80.80.8\times 0.80.8 × 0.8 mm2.

For the topotactic oxygen deintercalation, several cube-shaped LaNiO3 crystals with a total mass of about 100 mg were placed in direct contact with roughly 250 mg of powder of the reducing agent CaH2 in a DURAN glass tube. The glass tube was evacuated to similar-to\sim10-7 mbar and sealed. The reduction process was initiated by heating to 300C in a low-temperature furnace. The temperature was maintained for a duration of approximately 13 days, which is comparable to the conditions used for previous reductions of nickelate single crystals  [58, 56]. The surfaces of the obtained LaNiO2 crystals were polished using a diamond suspension, removing the uppermost layers (similar-to\sim30 μμ\upmuroman_μm) to expose a pristine surface free of residues and secondary phases that formed during the invasive topotactic reduction.

The reduced crystals were examined by STEM, which is a sensitive method for detecting impurity phases, stacking faults in the crystal lattice, and residual oxygen in nickelates [59, 58, 60, 61]. Electron-transparent specimens were prepared on a Thermo Fisher Scios I focused ion beam (FIB) using the standard liftout method. The lateral dimensions of the specimens were 20 μμ\upmuroman_μm by 1.5 μμ\upmuroman_μm with thicknesses between 50 and 100 nm. The STEM studies were performed with a JEOL JEM-ARM200F scanning transmission electron microscope equipped with a cold-field emission electron source, and a probe abberation-corrected Cs corrector (DCOR, CEOS GmbH) at 200 kV. STEM imaging was performed with probe semiconvergence angles of 20 mrad, resulting in probe sizes of 0.8 Å. The collection angles for high-angle annular dark-field (HAADF) and annular bright field (ABF) STEM images were 75 to 310 and 11 to 23 mrad, respectively. To improve the signal-to-noise ratio of the data while minimizing sample damage, a high-speed time series (2 μμ\upmuroman_μs per pixel) was acquired and then aligned and summed.

To investigate the microstructural properties of the LaNiO2 crystals, such as the twin domain distribution, we utilized a Zeiss Merlin electron microscope equipped with electron backscatter diffraction (EBSD) capabilities from Oxford Instruments (Symmetry EBSD detector). The AZtec and AZtecCrystal software were used for data acquisition and analysis, respectively. For the EBSD measurements, the crystal was tilted to a high angle (70superscript7070^{\circ}70 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT). The detector was positioned at a 5 angle relative to the untilted crystal surface. Probe currents ranged 1-5 nA, and the accelerating voltage was 20 kV for most measurements. Further details about the EBSD measurements are given in the Supplemental Material (SM) [62].

X-ray absorption spectroscopy (XAS) and RIXS measurements were conducted at the I21-RIXS beamline of the Diamond Light Source, UK [63]. All RIXS spectra were collected at temperatures between 22 and 28 K with an effective energy resolution of ΔE=37Δ𝐸37\Delta E=37roman_Δ italic_E = 37 meV and a scattering angle of 154superscript154154^{\circ}154 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. Incident photons were linearly polarized in either the vertical (σ𝜎\sigmaitalic_σ) or horizontal (π𝜋\piitalic_π) direction with respect to the scattering plane. The x-ray beam footprint is approximately 2.5 μμ\upmuroman_μm ×\times× 40 μμ\upmuroman_μm.

Refer to caption
Figure 1: (a) Electronic transport of a LaNiO2 crystal. (Inset) A side-view SEM-SE image of the polished surface of the LaNiO2 crystal used in the RIXS experiment. (b) XRD pattern acquired from the polished surface of a LaNiO2 crystal. The Bragg peaks are indexed and the inset shows the tetragonal P4/mmm𝑃4𝑚𝑚𝑚P4/mmmitalic_P 4 / italic_m italic_m italic_m unit cell of LaNiO2. The asterisk symbol indicates an artifact originating from a carbon-based paste used to mount the LaNiO2 crystal on the sample holder.
Refer to caption
Figure 2: (a) Top-view SEM-SE image of the polished surface of a LaNiO2 crystal. (b) Cross-sectional view of the crystal in a low-magnification STEM-HAADF image. The yellow arrow marks the same dark line as the arrow in panel (a). [(c),(d)] Magnified view of the area indicated by the green (orange) box in panel (b), which consists of noncrystalline (crystalline) LaNiO2. (Inset) The fast Fourier transformation. (e) High-magnification STEM HAADF and (f) STEM ABF images acquired within a single domain of a LaNiO2 crystal. The inset in (f) zooms into the STEM ABF image, revealing the ideal infinite-layer structure without visible traces of apical oxygen (above and below the Ni atoms) remaining from the perovskite LaNiO3 structure.

III Results

III.1 Properties of LaNiO2 bulk crystals

Refer to caption
Figure 3: (a) SEM-SE image (top panel) and EBSD map (bottom panel) of the same surface region. The color code (green, purple, and brown) indicates the three twin domains (A𝐴Aitalic_A, B𝐵Bitalic_B, and C𝐶Citalic_C) of LaNiO2 with the tetragonal P4/mmm𝑃4𝑚𝑚𝑚P4/mmmitalic_P 4 / italic_m italic_m italic_m unit cell. White areas correspond to pixels where the EBSD pattern could not be indexed (see the SM [62]). (b) Schematic of the LaNiO2 crystal structure with three twin domains, along with the scattering geometry of the RIXS experiment. The a𝑎aitalic_a, b𝑏bitalic_b, and c𝑐citalic_c axes of each domain are indicated, and the square-lattice coordination of Ni (gray) and O (red) is shown for two NiO2 planes in each domain, while La ions between the planes are omitted for clarity. The incident photons (𝐤isubscript𝐤i\mathbf{k}_{\rm i}bold_k start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT) are linearly polarized either parallel (π𝜋\piitalic_π-pol.) or perpendicular (σ𝜎\sigmaitalic_σ-pol.) to the scattering plane (yellow). A polarization analysis of the the outgoing photons (𝐤fsubscript𝐤f\mathbf{k}_{\rm f}bold_k start_POSTSUBSCRIPT roman_f end_POSTSUBSCRIPT) was not performed. The incident angle θ𝜃\thetaitalic_θ is defined as the angle between 𝐤isubscript𝐤i\mathbf{k}_{\rm i}bold_k start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT and the sample surface.

In a previous study, millimeter-sized LaNiO2 crystals with randomly oriented surfaces were obtained via the topotactic reduction of LaNiO3 single crystals, utilizing a direct-contact method with CaH2 as the reducing agent [56]. Here we employ a similar topotactic method to obtain LaNiO2 bulk crystals with crystallographically oriented surfaces (for details see Sec. II). In addition, we polish the crystals to expose fresh surfaces with the LaNiO2 phase, which were not affected by the topotactic reduction atmosphere. Figure 1(a) shows the electronic transport of a LaNiO2 crystal measured in a four-point probe geometry on a polished surface [see inset in Fig. 1(a)]. The transport behavior is metallic with a subtle upturn below 50 K, which is consistent with that reported for high-quality LaNiO2 thin films [64]. In contrast, the onset of the upturn in LaNiO2 crystals with randomly oriented and unpolished surfaces was approximately 100 K [56].

In the x-ray diffraction (XRD) pattern from an oriented and polished surface we observe the characteristic Bragg peaks of LaNiO2 in the P4/mmm𝑃4𝑚𝑚𝑚P4/mmmitalic_P 4 / italic_m italic_m italic_m space group, while no additional peaks due to impurities or insufficiently reduced phases are detected [Fig. 1(b)]. However, the simultaneous presence of (H,0,0)𝐻00(H,0,0)( italic_H , 0 , 0 ) and (0,0,L)00𝐿(0,0,L)( 0 , 0 , italic_L )-type Bragg peaks suggests that the LaNiO2 crystals contain three twin domains of the tetragonal P4/mmm𝑃4𝑚𝑚𝑚P4/mmmitalic_P 4 / italic_m italic_m italic_m unit cell. The emergence of these three domains is likely a consequence of the pseudocubic symmetry of the perovskite LaNiO3 phase, which lacks a unique preferential direction for oxygen deintercalation during the topotactic reduction process.

A closer inspection of the LaNiO2 crystals reveals a distinctive texture on the polished surfaces, characterized by parallel or perpendicular dark lines [Fig. 2(a)] that mostly align with the directions of the crystal edges. This texture is both visible in secondary electron (SE) images acquired with a scanning electron microscope (SEM) [Fig. 2(a)] and optical microscopy images (see the SM [62]). A low-magnification STEM-HAADF image showing a cross-sectional view across a dark line and its vicinity is presented in Fig. 2(b). A zoom into the dark area [Fig. 2(c)] indicates a porous structure in this region, and the fast Fourier transformation (FFT) of the image suggests that crystalline order is absent [inset in Fig. 2(c)]. In contrast, a magnified view of a representative area surrounding the dark lines [Fig. 2(d)] reveals a well-ordered crystal lattice. The corresponding FFT image exhibits a pattern that is consistent with the IL structure [58]. Further magnification of a crystalline region is provided by high-resolution STEM-HAADF [Fig. 2(e)] and ABF imaging [Fig. 2(f)], confirming high crystalline quality on a local scale, without any stacking faults or secondary phase inclusions in the lattice. Moreover, the inset in Fig. 2(f) illustrates the absence of residual apical oxygen from the LaNiO3 phase, indicating that the topotactic transformation to the IL phase is complete.

Notably, the texture featuring dark lines is not present in our unreduced LaNiO3 crystals. This suggests that the disordered regions develop during the topotactic reduction, potentially acting as channels that facilitate the oxygen transport from the crystal’s interior to its surface. Similar disordered areas separating highly crystalline regions were also identified in a previous STEM study on topotactically reduced Pr0.92Ca0.08NiO2.75 crystals [60]. Elucidating the nature of these disordered regions and their role in the topotactic transformation process presents a compelling subject for future STEM investigations.

To further characterize our LaNiO2 crystals, we map the distribution of twin domains across the surface using EBSD. The top panel of Fig. 3(a) displays an SEM-SE image of a local area on the crystal surface after mechanical and low-angle Ar ion polishing, which is necessary to prepare the surface for EBSD measurements. The bottom panel of Fig. 3(a) shows the acquired EBSD map from the same surface area, where we identify three twin domains of the tetragonal space group P4/mmm𝑃4𝑚𝑚𝑚P4/mmmitalic_P 4 / italic_m italic_m italic_m of LaNiO2, color coded in green, purple, and brown. We denote these domains as A𝐴Aitalic_A, B𝐵Bitalic_B, and C𝐶Citalic_C in the following. Specifically, the EBSD map reveals extended regions of single-domain character, with domain C𝐶Citalic_C predominating. Around the center and the top left corner of the map, significant areas of domain A𝐴Aitalic_A are observed, interspersed with smaller islands of domain B𝐵Bitalic_B. A mapping of other areas on the crystal surface (see the SM [62]), however, reveals varying domain distributions, including regions where domain A𝐴Aitalic_A or B𝐵Bitalic_B are dominant. This variation in domain population across areas and length scales of tens of micrometers has important implications for the RIXS experiment, considering that the x-ray beam’s footprint on the sample surface is approximately 2.5 μμ\upmuroman_μm ×\times× 40 μμ\upmuroman_μm, which is of comparable size. This aspect will be elucidated later in the text.

Interestingly, we find that the nonindexed pixels (white color) in the EBSD maps mostly coincide with dark lines in the associated SE-SEM images, which according to the analysis in Fig. 2 correspond to disordered regions. The orientation of the twin domains is often identical on both sides of streaks of nonindexed pixels, hence suggesting that these regions are not corresponding to twin domain boundaries. Instead, the boundaries between domains [see different colors in Fig. 3(a)], exhibit a more complex structure varying on finer length scales, which also presents a subject for future STEM investigations.

III.2 RIXS experiment

The scattering geometry of the RIXS experiment together with the orientation of the three twin domains is depicted in Fig. 3(b). We denote the momentum transfer (H,K,L)𝐻𝐾𝐿(H,K,L)( italic_H , italic_K , italic_L ) in reciprocal lattice units (2π/a,2π/b,2π/c)2𝜋𝑎2𝜋𝑏2𝜋𝑐(2\pi/a,2\pi/b,2\pi/c)( 2 italic_π / italic_a , 2 italic_π / italic_b , 2 italic_π / italic_c ), with the lattice parameters a,b=3.9642(3)𝑎𝑏3.96423a,b=3.9642(3)italic_a , italic_b = 3.9642 ( 3 ) Å, and c=3.3561(3)𝑐3.35613c=3.3561(3)italic_c = 3.3561 ( 3 ) Å determined from the powder XRD refinement of a crushed LaNiO2 crystal (see the SM [62]). For simplicity, we mostly focus on domain A𝐴Aitalic_A in the following, where the scattering plane is parallel to the crystallographic a𝑎aitalic_a-c𝑐citalic_c plane [see Fig. 3(b)].

III.3 Orbital excitations

Figure 4(a) shows the RIXS spectra of the polished LaNiO2 crystal, with the incident photon energy varied across the Ni L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge. The major spectral weight is centered around 1.2 eV energy loss for incident energies between Ei=852.4subscript𝐸𝑖852.4E_{i}=852.4italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 852.4 and 853.4 eV, which roughly coincides with the energy range of the main Ni L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge peak in the XAS [see solid white line in Fig. 4(a)]. In addition, a sequence of excitations at higher energy losses can be observed in the RIXS intensity map, along with a fluorescence emission feature that streaks across the map up to the highest energies. Overall, the RIXS features in Fig. 4(a) are closely reminiscent of those reported for LaNiO2 thin films [39], attributable to dd𝑑𝑑dditalic_d italic_d excitations that emerge roughly between 1 and 3 eV energy loss. Subtle differences between the relative intensities of individual dd𝑑𝑑dditalic_d italic_d excitation peaks are likely associated with the presence of three tetragonal twin domains in our LaNiO2 crystals, whereas the RIXS map in Ref. [39] was obtained from a single-domain LaNiO2 film.

Importantly, a close inspection of the region around 0.6 eV energy loss in Fig. 4(a) reveals the presence of spectral intensity for incident energies around Ei=852.3subscript𝐸𝑖852.3E_{i}=852.3italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 852.3 eV. In prior thin film studies, this feature has been identified as a hallmark of the infinite-layer nickelates, originating from hybridized La-5d5𝑑5d5 italic_d and Ni-3d3𝑑3d3 italic_d orbital states [39, 41]. The 0.6 eV feature becomes better discernible when the RIXS spectrum for Ei=852.3subscript𝐸𝑖852.3E_{i}=852.3italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 852.3 eV is plotted separately [see dark blue arrow in Fig. 4(b)]. In addition, the RIXS spectra for Ei=853.3subscript𝐸𝑖853.3E_{i}=853.3italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 853.3 and 853.9 eV are displayed in Fig. 4(b). While spectral weight in the latter spectrum shifts towards energy losses above 1.5 eV, we observe a relatively sharp accumulation of spectral intensity below 0.3 eV in the former spectrum (see light-blue arrow). The maximum of this spectral weight is situated around similar-to\sim0.15 eV. This energy range is above that of phonons, but it is compatible with the bandwidth of spin excitations reported in previous thin-film studies of infinite-layer nickelates [50, 28, 27, 29].

Refer to caption
Figure 4: (a) RIXS intensity map of the LaNiO2 crystal for incident photon energies varied across the Ni L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT edge, taken at θ=51.2𝜃superscript51.2\theta=51.2^{\circ}italic_θ = 51.2 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. The corresponding XAS measured in total electron yield mode is superimposed as the solid white line. The increase of the XAS signal towards low incident energies is due to the tail of the strong La M4subscript𝑀4M_{4}italic_M start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT-line centered below 851 eV. (b) RIXS spectra for three selected incident photon energies. Arrows at similar-to\sim0.6 and similar-to\sim0.15 eV indicate characteristic low energy excitations, as described in the text.

III.4 Spin excitations

To elucidate the characteristics of the possible spin excitations, we investigate the momentum-dependence of this spectral weight (Fig. 5). To account for photon absorption effects in the crystal for different incident angles, we perform a self-absorption correction of the RIXS data [16, 12]. Details about the employed procedure can be found in the SM [62]. Notably, the RIXS spectra along the (H,0)𝐻0(H,0)( italic_H , 0 ) [Fig. 5(a)] and (H,H)𝐻𝐻(H,H)( italic_H , italic_H ) [Fig. 5(b)] directions reveal that the spectral weight is broadly distributed and extends up to similar-to\sim300 meV for all measured momenta. This is in contrast to the highly dispersive spin wave signal in the RIXS spectra of the nickelate thin films [50, 28, 27], although the maximum bandwidth across the Brillouin zone is comparable. A subtle increase in spectral intensity is discernible towards higher momenta along both directions, mostly within the energy range below 100 meV, which is close to the energy regime of phonons. In particular, a prominent phonon peak centered around 60 meV has been reported in Ni L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge RIXS experiments on LaNiO2 thin films [28]. Nonetheless, when comparing the RIXS spectra for representative momenta along the (H,0)𝐻0(H,0)( italic_H , 0 ) direction [Fig. 5(a)] it appears that the most significant change occurs between 80 and 100 meV. Hence, it is plausible that the broad distribution of low-energy spectral weight in Figs. 5(a) and 5(b) contains contributions from elastic scattering mostly below 40 meV, a phonon around 60 meV, and spin excitations or other contributions at higher energies up to 300 meV.

Refer to caption
Figure 5: (a) RIXS spectra of the LaNiO2 crystal for varied momenta along the (H,0)𝐻0(H,0)( italic_H , 0 ) direction, acquired with π𝜋\piitalic_π-polarized photons. The dashed curves are Gaussian fits of the elastic line centered at zero-energy loss, using an energy resolution of 37 meV. Substantial spectral weight extends up to similar-to\sim300 meV, but shows essentially no dispersion as a function of momentum transfer. (b) RIXS spectra along the (H,H)𝐻𝐻(H,H)( italic_H , italic_H ) direction.

A difference to single-domain thin films  [50, 28, 27] is that the RIXS signal in our experiment contains the superposition of the three twin domains. To test whether this superposition is responsible for the observed “washed-out” character of the spin excitations in Figs. 5(a) and 5(b), we simulate spin-wave spectra based on the linear spin-wave theory for the simultaneous presence of three orthogonal crystallographic orientations. Specifically, we employ a J1subscript𝐽1J_{1}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT-J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT model for a square-lattice Heisenberg antiferromagnet, which was also used to model the dispersion measured from an IL nickelate film [50]. As an approximation, we assume that the RIXS intensity is proportional to the dynamical structure factor of the spin-wave excitations [65, 66]. The magnetic Hamiltonian is formulated as follows:

=J1i,j𝐒^i𝐒^j+J2i,k𝐒^i𝐒^k,subscript𝐽1subscript𝑖𝑗subscript^𝐒𝑖subscript^𝐒𝑗subscript𝐽2subscript𝑖𝑘subscript^𝐒𝑖subscript^𝐒𝑘{\cal H}=J_{1}\sum_{i,j}\hat{\mathbf{S}}_{i}\cdot\hat{\mathbf{S}}_{j}+J_{2}% \sum_{i,k}\hat{\mathbf{S}}_{i}\cdot\hat{\mathbf{S}}_{k},caligraphic_H = italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_i , italic_k end_POSTSUBSCRIPT over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , (1)

where 𝐒^isubscript^𝐒𝑖\hat{\mathbf{S}}_{i}over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is a spin operator with spin S=1/2𝑆12S=1/2italic_S = 1 / 2 at a site i𝑖iitalic_i. J1subscript𝐽1J_{1}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are the nearest and next-nearest-neighbor exchange interactions in the ab𝑎𝑏abitalic_a italic_b plane, respectively. We adopt the reported values J1=63.6subscript𝐽163.6J_{1}=63.6italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 63.6 and J2=10.3subscript𝐽210.3J_{2}=-10.3italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - 10.3 meV from a NdNiO2 thin film study [50], where the spin waves observed in the RIXS spectra exhibited close similarities to those in square-lattice cuprates. Accordingly, we define a collinear magnetic structure with an antiferromagnetic ordering vector 𝐐=(1/2,1/2)𝐐1212\mathbf{Q}=(1/2,1/2)bold_Q = ( 1 / 2 , 1 / 2 ) [15, 67]. Using the linear spin-wave theory, the dispersion relation is given by

ω𝐪=2ZcA𝐪2B𝐪2,Planck-constant-over-2-pisubscript𝜔𝐪2subscript𝑍𝑐superscriptsubscript𝐴𝐪2superscriptsubscript𝐵𝐪2\hbar\omega_{\mathbf{q}}=2Z_{c}\sqrt{A_{\mathbf{q}}^{2}-B_{\mathbf{q}}^{2}},roman_ℏ italic_ω start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT = 2 italic_Z start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT square-root start_ARG italic_A start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_B start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (2)

where

A𝐪subscript𝐴𝐪\displaystyle A_{\mathbf{q}}italic_A start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT =\displaystyle== J1J2{1cos(2πH)cos(2πK)}subscript𝐽1subscript𝐽212𝜋𝐻2𝜋𝐾\displaystyle J_{1}-J_{2}\left\{1-\cos{(2\pi H)}\cos{(2\pi K)}\right\}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT { 1 - roman_cos ( 2 italic_π italic_H ) roman_cos ( 2 italic_π italic_K ) } (3)
B𝐪subscript𝐵𝐪\displaystyle B_{\mathbf{q}}italic_B start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT =\displaystyle== J1{cos(2πH)+cos(2πK)}/2.subscript𝐽12𝜋𝐻2𝜋𝐾2\displaystyle J_{1}\left\{\cos{(2\pi H)}+\cos{(2\pi K)}\right\}/2.italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT { roman_cos ( 2 italic_π italic_H ) + roman_cos ( 2 italic_π italic_K ) } / 2 . (4)

The renormalization factor Zcsubscript𝑍𝑐Z_{c}italic_Z start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is fixed at 1.18 [68], which is expected for a two-dimensional S=1/2𝑆12S=1/2italic_S = 1 / 2 square-lattice antiferromagnet. Further, the inelastic part of the dynamical structure factor is given by:

S(𝐪,ω)A𝐪B𝐪A𝐪+B𝐪δ(ωω𝐪).proportional-to𝑆𝐪𝜔subscript𝐴𝐪subscript𝐵𝐪subscript𝐴𝐪subscript𝐵𝐪𝛿𝜔subscript𝜔𝐪S(\mathbf{q},\omega)\propto\sqrt{\frac{A_{\mathbf{q}}-B_{\mathbf{q}}}{A_{% \mathbf{q}}+B_{\mathbf{q}}}}\delta\left(\omega-\omega_{\mathbf{q}}\right).italic_S ( bold_q , italic_ω ) ∝ square-root start_ARG divide start_ARG italic_A start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT - italic_B start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT end_ARG start_ARG italic_A start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT + italic_B start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT end_ARG end_ARG italic_δ ( italic_ω - italic_ω start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT ) . (5)

In analogy to paramagnons in cuprates, which emerge in the absence of long-range order, also the spin excitations in IL nickelates feature a large linewidth [50]. Hence, we convolute our calculated spin spectra with a damped harmonic oscillator function. We employ a damping factor of 74 meV, in accord with the average of the momentum-dependent damping factors in Ref. [50]. Furthermore, our damped spectra are convoluted with Gaussian functions along the energy and momentum directions according to the RIXS experimental energy resolution of 37 meV and the q𝑞qitalic_q resolution of 0.09 Å-1, respectively The q𝑞qitalic_q resolution is estimated from the mosaicity of the LaNiO2 crystal according to the full width at half maximum of the (001)001(001)( 001 ) Bragg peak.

Refer to caption
Figure 6: Spin wave simulation for a system with three twin domains and the scattering geometry used in the RIXS experiment, with varying incident angle θ𝜃\thetaitalic_θ, fixed scattering angle at 154, and Ei=853subscript𝐸𝑖853E_{i}=853italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 853 eV. (a)–(c) False-color plots (top panels) of the simulated spin-wave spectra of the twin domains A𝐴Aitalic_A, B𝐵Bitalic_B, and C𝐶Citalic_C, respectively. The intensity scale corresponds to the dynamical structure factor S(𝐪,ω)𝑆𝐪𝜔S(\mathbf{q},\omega)italic_S ( bold_q , italic_ω ) and the solid black lines are the dispersion relations. The solid red lines (bottom panels) indicate the corresponding momentum transfer for each domain within the HL𝐻𝐿H-Litalic_H - italic_L, LH𝐿𝐻L-Hitalic_L - italic_H, and KH𝐾𝐻K-Hitalic_K - italic_H planes, respectively. (d)–(g) Superposition of S(𝐪,ω)𝑆𝐪𝜔S(\mathbf{q},\omega)italic_S ( bold_q , italic_ω ) from the three domains assuming different populations of the domains A𝐴Aitalic_A, B𝐵Bitalic_B, and C𝐶Citalic_C. For better comparability to the RIXS data in Fig. 5(a), the momenta in panel (d) are indexed according to domain A𝐴Aitalic_A with variation along the (H,0)𝐻0(H,0)( italic_H , 0 ) direction, although the calculated S(𝐪,ω)𝑆𝐪𝜔S(\mathbf{q},\omega)italic_S ( bold_q , italic_ω ) contains the contributions from all three domains. The domain population ratios in panels (d)–(f) are 1:1:1:11:11:1:11 : 1 : 1, 1:1:0.3:11:0.31:1:0.31 : 1 : 0.3, and 0.4:1:0.1:0.41:0.10.4:1:0.10.4 : 1 : 0.1, respectively. In panel (f), an almost non-dispersive behavior of the spin wave is reproduced, which can be further enhanced by taking into account a 25% volume fraction with randomly oriented grains or disordered crystal structure (powder).

Figures 6(a)–6(c) present the simulated spin wave spectra for three orthogonal crystal orientations, each contributing a distinct spin-wave signal to the RIXS spectrum. The three orientations represent the twin domains illustrated in Fig. 3(b). In our RIXS experiment, the momentum transfer is varied by changing the incident angle θ𝜃\thetaitalic_θ [see Fig. 3(b)], while keeping the scattering angle fixed at 154superscript154154^{\circ}154 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. As a consequence, a change of the in-plane momentum transfer also involves a change of the out-of-plane momentum. For a single-domain sample, such as an IL nickelate thin film, the change in the out-of-plane momentum is usually neglected since the dispersion of the spin waves along the L𝐿Litalic_L direction is flat for a quasi-two-dimensional square lattice antiferromagnet. However, for the simultaneous presence of three twin domains the situation is more involved and hence requires the separate consideration of each domain [Figs.6(a)–6(c)]. The corresponding trajectories in reciprocal space for the three domains when the incident angle is varied from θ=20𝜃superscript20\theta=20^{\circ}italic_θ = 20 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT to 134 are shown in the bottom panels in Figs. 6(a)–6(c).

Assuming an equal population (1:1:1):11:1(1:1:1)( 1 : 1 : 1 ) of the A𝐴Aitalic_A, B𝐵Bitalic_B, and C𝐶Citalic_C twin domains, we obtain a composite spectrum shown in Fig. 6(d). The simulated peak positions disperse significantly more weakly as a function of H𝐻Hitalic_H than the spin waves of single-domain thin films  [50]. However, a strong increase in intensity for the highest momenta H𝐻Hitalic_H in Fig. 6(d) is not present in our RIXS data in Fig. 5(a). Considering that the typical lateral dimensions of the domains in the LaNiO2 crystal are between a few and several tens of micrometers (see Fig. 3(a) and the SM [62]) while the footprint of the x-ray beam on the sample surface is 2.5 μμ\upmuroman_μm in the vertical direction and more than 40 μμ\upmuroman_μm in the horizontal direction (depending on the incident angle θ𝜃\thetaitalic_θ), it is plausible to assume that the population of the three domains in the probed sample volume was not equal. Accordingly, we next simulate unequal domain populations of 1:1:0.3:11:0.31:1:0.31 : 1 : 0.3 and 0.4:1:0.1:0.41:0.10.4:1:0.10.4 : 1 : 0.1 in Figs. 6(e) and 6(f), respectively. A notable trend in these simulations is a further reduction of the dispersive character and intensity enhancement, aligning more closely with the RIXS data presented in Fig 5(a).

As a final step, we take into account the presence of the textures with dark lines in our LaNiO2 crystals [Figs. 2(a)–2(c)], indicative of regions with a disordered crystal structure. From Fig. 2 and optical microscope images (see the SM [62]) we estimate that these regions contribute at most 25% of the the sample’s volume fraction, and we incorporate this sample characteristic as powder-averaged spin-wave spectra in our simulation. Although the simulation result [Fig. 6(g)] does still not exactly match our RIXS data, the trend is notably consistent. Considering that we used the damping and magnetic exchange parameters from previous thin film studies as sole input for our simulation, it suggests that the magnetic correlations within a single domain of our crystals closely resemble those in single-domain films. The remaining discrepancy observed in Fig. 6(g) might be due to increased damping of the spin waves in the crystals or other factors, highlighting the need for further investigation.

III.5 Charge order

Next, we examine whether the charge order phenomenology reported for IL nickelate thin films similarly occurs in crystals. Specifically, a number of previous RIXS studies on thin films performed under comparable experimental conditions, reported the occurrence of a CDW signal in the form of enhanced quasielastic scattering around H=±1/3𝐻plus-or-minus13H=\pm 1/3italic_H = ± 1 / 3 [28, 27, 26].

Figure 7(a) shows a RIXS intensity map of the region of quasielastic scattering of the LaNiO2 crystal for different momenta along the (H,0)𝐻0(H,0)( italic_H , 0 ) direction, utilizing σ𝜎\sigmaitalic_σ-polarized incident photons to enhance the charge signal. Note that these RIXS spectra were obtained from a different surface spot on the LaNiO2 crystal compared to those in Fig. 5. Hence, the underlying domain population is likely distinct, and for simplicity the momentum transfer in Fig. 7 is indexed with respect to domain A𝐴Aitalic_A. The self-absorption correction [16, 12] was applied to all spectra in Fig. 7(a).

Remarkably, we find that the data from our LaNiO2 crystal lack clear charge ordering signatures around H=±1/3𝐻plus-or-minus13H=\pm 1/3italic_H = ± 1 / 3 [Fig. 7(a)]. This absence of an enhanced quasielastic signal is corroborated when examining the integrated RIXS intensity between 5050-50- 50 meV E50absent𝐸50\leq E\leq 50≤ italic_E ≤ 50 meV energy loss [red data points in Fig. 7(b)]. For a direct contrast, we also plot LaNiO2 film data from Ref. [28] in Fig. 7(b). Note that we have normalized both our RIXS data and those of Ref. [28] by the integrated intensity of the respective dd𝑑𝑑dditalic_d italic_d excitations to ensure direct comparability.

Refer to caption
Figure 7: (a) RIXS intensity map of the LaNiO2 crystal for momenta varied along the (±H,0)plus-or-minus𝐻0(\pm H,0)( ± italic_H , 0 ) direction, acquired with σ𝜎\sigmaitalic_σ-polarized incident photons. (b) Integrated intensity of the quasielastic scattering along the (±H,0)plus-or-minus𝐻0(\pm H,0)( ± italic_H , 0 ) direction. The integration range is 5050-50- 50 meV E50absent𝐸50\leq E\leq 50≤ italic_E ≤ 50 meV energy loss. Error bars are determined by the standard deviation of six RIXS spectra measured at each momentum. In addition, the quasielastic intensity from a RIXS experiment on a LaNiO2 thin film in Ref. [28] is superimposed as gray symbols, exhibiting a distinct CDW peak around H=0.344𝐻0.344H=0.344italic_H = 0.344. The RIXS intensities of the crystal and the film are both normalized to the integrated intensity of the dd𝑑𝑑dditalic_d italic_d excitations (see the SM [62]). In addition, a vertical offset is added to the film intensity to match the baseline of the crystal data.

While the pronounced H=±1/3𝐻plus-or-minus13H=\pm 1/3italic_H = ± 1 / 3 peaks of Ref. [28] are clearly absent in our data, we observe a broad peak-like feature around H0.27similar-to𝐻0.27H\sim 0.27italic_H ∼ 0.27, which is subtle but exceeds the range of our experimental error bars. Yet, for momentum transfer in the opposite direction, a corresponding peak feature at H0.27similar-to𝐻0.27H\sim-0.27italic_H ∼ - 0.27 is not visible. Instead, the quasielastic scattering intensity in this direction increases continuously from H0.15similar-to𝐻0.15H\sim-0.15italic_H ∼ - 0.15 to 0.30.3-0.3- 0.3. To verify that these behaviors are not associated with CDW signals, we carried out additional RIXS measurements on other locations on the surface of the same crystal as well as on two additional crystals (see the SM [62]), which did not reproduce the peak at H±0.27plus-or-minus𝐻0.27H\pm 0.27italic_H ± 0.27. In consequence, we attribute such deviations from the baseline intensity of quasielastic scattering to artifacts resulting from the partially uneven surface topography of our LaNiO2 crystals (see inset in Fig. 1(a) and the SM [62]).

Nonetheless, due to the presence of three twin domains in our LaNiO2 crystals, we cannot fully rule out that a CDW signal has remained undetected in our experiment. For instance, one scenario could involve an unfavorable domain distribution within the surface area probed by RIXS in Fig. 7 where the domains A𝐴Aitalic_A and B𝐵Bitalic_B would be essentially unpopulated, while for the dominating domain C𝐶Citalic_C our applied momentum transfer range would not cover the nominal 𝐪=(±1/3,0)𝐪plus-or-minus130\mathbf{q}=(\pm 1/3,0)bold_q = ( ± 1 / 3 , 0 ) ordering vector. However, this scenario seems implausible, as our EBSD analysis of various crystal surface areas (see the SM [62]) revealed that typically at least two twin domains exhibit significant population within an area corresponding to the x-ray beam footprint of approximately 2.5 μμ\upmuroman_μm ×\times× 40 μμ\upmuroman_μm. Moreover, our search for a CDW signal encompassed different spots on multiple LaNiO2 crystals (see the SM [62]), making it even less likely that we consistently encountered unfavorable domain distributions.

The absence of the 𝐪=(±1/3,0)𝐪plus-or-minus130\mathbf{q}=(\pm 1/3,0)bold_q = ( ± 1 / 3 , 0 ) CDW peaks in our LaNiO2 sample is particularly noteworthy when considered in conjunction with the STEM characterization of our crystal [Fig. 2(f)]. Specifically, the absence of any traces of residual apical oxygen indicates that the ideal IL structure is realized in our crystals. This concomitant absence of apical oxygen and a CDW signal is compatible with the notion that excess oxygen arranged in patterns characterized by a 𝐪=(±1/3,0)𝐪plus-or-minus130\mathbf{q}=(\pm 1/3,0)bold_q = ( ± 1 / 3 , 0 ) ordering vector is a prerequisite for the emergence of charge order in nickelates [54, 55].

IV Conclusions

In summary, our results have several important implications for the ongoing debate about spin and charge instabilities in the material class of IL nickelates. In particular, our observation of spin excitations with a bandwidth of more than 200 meV akin to those reported for thin films  [50, 27, 28] suggests that the associated spin dynamics are an ubiquitous feature of the pure IL phase. Nevertheless, the absence of a discernible spin-wave dispersion in our data, presumably due to a distribution of twin domains within the probed region on the crystal surface, calls for further investigations. In particular, the synthesis of single-domain LaNiO2 crystals becomes imperative for future experimental work, potentially achievable by applying strain or pressure during the topotactic reduction, akin to the epitaxial strain conditions that yield single-domain IL thin films [69, 70, 71].

Furthermore, the lack of a 𝐪=(±1/3,0)𝐪plus-or-minus130\mathbf{q}=(\pm 1/3,0)bold_q = ( ± 1 / 3 , 0 ) CDW signal in our data a suggests that charge order is not intrinsic to the bulk phase of IL nickelates. This notion is compatible with the scenario that a secondary phase of residual apical oxygen ions hosts the CDW [54, 55], or that epitaxial strain from a substrate plays a role in the emergence of charge order in IL nickelates [29]. In both cases, further investigations are warranted to clarify if analogies between the 𝐪=(±1/3,0)𝐪plus-or-minus130\mathbf{q}=(\pm 1/3,0)bold_q = ( ± 1 / 3 , 0 ) nickelate charge order phenomenology and that in cuprates [18, 19, 20, 21, 22, 23, 24, 25] exist.

Given our focus on the undoped parent compound LaNiO2 and axial directions up to wave vectors of 𝐪=(±0.4,0)𝐪plus-or-minus0.40\mathbf{q}=(\pm 0.4,0)bold_q = ( ± 0.4 , 0 ), our results do not rule out the existence of alternative charge ordering patterns in IL nickelates with different vectors and/or an emergence upon hole doping. This highlights the need for future studies on IL nickelate bulk crystals and thin films with assured stoichiometric oxygen content, with a focus either on exploring the phase space suggested by the CDW in cuprates, or undertaken as a comprehensive survey of the entire experimentally accessible reciprocal space and doping parameters.

Acknowledgements.
We thank W.-S. Lee, J. Weis, and A. Krajewska for helpful discussions, and C. Busch, U. Waizmann, and M. Hagel for technical support. We thank J. Deuschle for FIB lamella preparation. We acknowledge Diamond Light Source for providing the beamtime under the Proposal No. MM33241-1, and are grateful to the technical staff for assistance during the experiment.

References

  • Scalapino [2012] D. J. Scalapino, A common thread: The pairing interaction for unconventional superconductors, Rev. Mod. Phys. 84, 1383 (2012).
  • Keimer et al. [2015] B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida, and J. Zaanen, From quantum matter to high-temperature superconductivity in copper oxides, Nature 518, 179 (2015).
  • Fernandes et al. [2022] R. M. Fernandes, A. I. Coldea, H. Ding, I. R. Fisher, P. J. Hirschfeld, and G. Kotliar, Iron pnictides and chalcogenides: a new paradigm for superconductivity, Nature 601, 35–44 (2022).
  • Gegenwart et al. [2008] P. Gegenwart, Q. Si, and F. Steglich, Quantum criticality in heavy-fermion metals, Nat. Phys. 4, 186 (2008).
  • Si and Steglich [2010] Q. Si and F. Steglich, Heavy fermions and quantum phase transitions, Science 329, 1161 (2010).
  • Lee et al. [2006] P. A. Lee, N. Nagaosa, and X.-G. Wen, Doping a mott insulator: Physics of high-temperature superconductivity, Rev. Mod. Phys. 78, 17 (2006).
  • J. Birgeneau et al. [2006] R. J. Birgeneau, C. Stock, J. M. Tranquada, and K. Yamada, Magnetic neutron scattering in hole-doped cuprate superconductors, J. Phys. Soc. Jpn. 75, 111003 (2006).
  • Dai [2015] P. Dai, Antiferromagnetic order and spin dynamics in iron-based superconductors, Rev. Mod. Phys. 87, 855 (2015).
  • Smidman et al. [2023] M. Smidman, O. Stockert, E. M. Nica, Y. Liu, H. Yuan, Q. Si, and F. Steglich, Colloquium: Unconventional fully gapped superconductivity in the heavy-fermion metal CeCu2Si2subscriptCeCu2subscriptSi2{\mathrm{CeCu}}_{2}{\mathrm{Si}}_{2}roman_CeCu start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_Si start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPTRev. Mod. Phys. 95, 031002 (2023).
  • Le Tacon et al. [2011] M. Le Tacon, G. Ghiringhelli, J. Chaloupka, M. M. Sala, V. Hinkov, M. W. Haverkort, M. Minola, M. Bakr, K. J. Zhou, S. Blanco-Canosa, C. Monney, Y. T. Song, G. L. Sun, C. T. Lin, G. M. De Luca, M. Salluzzo, G. Khaliullin, T. Schmitt, L. Braicovich, and B. Keimer, Intense paramagnon excitations in a large family of high-temperature superconductors, Nat. Phys. 7, 725 (2011).
  • Dean et al. [2013] M. P. M. Dean, G. Dellea, R. S. Springell, F. Yakhou-Harris, K. Kummer, N. B. Brookes, X. Liu, Y.-J. Sun, J. Strle, T. Schmitt, L. Braicovich, G. Ghiringhelli, I. Božović, and J. P. Hill, Persistence of magnetic excitations in La2-xSrxCuO4 from the undoped insulator to the heavily overdoped non-superconducting metal, Nat. Mater. 12, 1019 (2013).
  • Minola et al. [2015] M. Minola, G. Dellea, H. Gretarsson, Y. Y. Peng, Y. Lu, J. Porras, T. Loew, F. Yakhou, N. B. Brookes, Y. B. Huang, J. Pelliciari, T. Schmitt, G. Ghiringhelli, B. Keimer, L. Braicovich, and M. Le Tacon, Collective Nature of Spin Excitations in Superconducting Cuprates Probed by Resonant Inelastic X-Ray Scattering, Phys. Rev. Lett. 114, 217003 (2015).
  • Peng et al. [2018] Y. Y. Peng, E. W. Huang, R. Fumagalli, M. Minola, Y. Wang, X. Sun, Y. Ding, K. Kummer, X. J. Zhou, N. B. Brookes, B. Moritz, L. Braicovich, T. P. Devereaux, and G. Ghiringhelli, Dispersion, damping, and intensity of spin excitations in the monolayer (Bi,Pb)2(Sr,La)2CuO6+δsubscriptBi,Pb2subscriptSr,La2subscriptCuO6𝛿{(\text{Bi,Pb})}_{2}{(\text{Sr,La})}_{2}{\mathrm{CuO}}_{6+\delta}( Bi,Pb ) start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( Sr,La ) start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CuO start_POSTSUBSCRIPT 6 + italic_δ end_POSTSUBSCRIPT cuprate superconductor family, Phys. Rev. B 98, 144507 (2018).
  • Peng et al. [2017] Y. Y. Peng, G. Dellea, M. Minola, M. Conni, A. Amorese, D. Di Castro, G. M. De Luca, K. Kummer, M. Salluzzo, X. Sun, X. J. Zhou, G. Balestrino, M. Le Tacon, B. Keimer, L. Braicovich, N. B. Brookes, and G. Ghiringhelli, Influence of apical oxygen on the extent of in-plane exchange interaction in cuprate superconductors, Nat. Phys. 13, 1201 (2017).
  • Coldea et al. [2001] R. Coldea, S. M. Hayden, G. Aeppli, T. G. Perring, C. D. Frost, T. E. Mason, S.-W. Cheong, and Z. Fisk, Spin Waves and Electronic Interactions in La2CuO4subscriptLa2subscriptCuO4{\mathrm{La}}_{2}{\mathrm{CuO}}_{4}roman_La start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CuO start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPTPhys. Rev. Lett. 86, 5377 (2001).
  • Robarts et al. [2021] H. C. Robarts, M. García-Fernández, J. Li, A. Nag, A. C. Walters, N. E. Headings, S. M. Hayden, and K.-J. Zhou, Dynamical spin susceptibility in La2CuO4subscriptLa2subscriptCuO4{\mathrm{La}}_{2}{\mathrm{CuO}}_{4}roman_La start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_CuO start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT studied by resonant inelastic x-ray scattering, Phys. Rev. B 103, 224427 (2021).
  • Lee et al. [2014] W. S. Lee, J. J. Lee, E. A. Nowadnick, S. Gerber, W. Tabis, S. W. Huang, V. N. Strocov, E. M. Motoyama, G. Yu, B. Moritz, H. Y. Huang, R. P. Wang, Y. B. Huang, W. B. Wu, C. T. Chen, D. J. Huang, M. Greven, T. Schmitt, Z. X. Shen, and T. P. Devereaux, Asymmetry of collective excitations in electron- and hole-doped cuprate superconductors, Nat. Phys. 10, 883 (2014).
  • Ghiringhelli et al. [2012] G. Ghiringhelli, M. L. Tacon, M. Minola, S. Blanco-Canosa, C. Mazzoli, N. B. Brookes, G. M. D. Luca, A. Frano, D. G. Hawthorn, F. He, T. Loew, M. M. Sala, D. C. Peets, M. Salluzzo, E. Schierle, R. Sutarto, G. A. Sawatzky, E. Weschke, B. Keimer, and L. Braicovich, Long-Range Incommensurate Charge Fluctuations in (Y,Nd)Ba2Cu3O6+xScience 337, 821 (2012).
  • Chang et al. [2012] J. Chang, E. Blackburn, A. T. Holmes, N. B. Christensen, J. Larsen, J. Mesot, R. Liang, D. A. Bonn, W. N. Hardy, A. Watenphul, M. v. Zimmermann, E. M. Forgan, and S. M. Hayden, Direct observation of competition between superconductivity and charge density wave order in YBa2Cu3O6.67Nat. Phys. 8, 871 (2012).
  • Blanco-Canosa et al. [2014] S. Blanco-Canosa, A. Frano, E. Schierle, J. Porras, T. Loew, M. Minola, M. Bluschke, E. Weschke, B. Keimer, and M. Le Tacon, Resonant x-ray scattering study of charge-density wave correlations in YBa2Cu3O6+xsubscriptYBa2subscriptCu3subscriptO6𝑥{\mathrm{YBa}}_{2}{\mathrm{Cu}}_{3}{\mathrm{O}}_{6+x}roman_YBa start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_Cu start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT roman_O start_POSTSUBSCRIPT 6 + italic_x end_POSTSUBSCRIPTPhys. Rev. B 90, 054513 (2014).
  • Li et al. [2020a] J. Li, A. Nag, J. Pelliciari, H. Robarts, A. Walters, M. Garcia-Fernandez, H. Eisaki, D. Song, H. Ding, S. Johnston, R. Comin, and K.-J. Zhou, Multiorbital charge-density wave excitations and concomitant phonon anomalies in Bi2Sr2LaCuO6+δProc. Natl. Acad. Sci. U.S.A. 117, 16219 (2020a).
  • Frano et al. [2020] A. Frano, S. Blanco-Canosa, B. Keimer, and R. J. Birgeneau, Charge ordering in superconducting copper oxides, J. Phys.: Condens. Matter 32, 374005 (2020).
  • Arpaia and Ghiringhelli [2021] R. Arpaia and G. Ghiringhelli, Charge order at high temperature in cuprate superconductors, J. Phys. Soc. Jpn. 90, 111005 (2021).
  • Lee et al. [2021] W. S. Lee, K.-J. Zhou, M. Hepting, J. Li, A. Nag, A. C. Walters, M. Garcia-Fernandez, H. C. Robarts, M. Hashimoto, H. Lu, B. Nosarzewski, D. Song, H. Eisaki, Z. X. Shen, B. Moritz, J. Zaanen, and T. P. Devereaux, Spectroscopic fingerprint of charge order melting driven by quantum fluctuations in a cuprate, Nat. Phys. 17, 53 (2021).
  • Lee [2021] W.-S. Lee, X-ray Studies of the CDW Ground State and Excitations in High-TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT Cuprates, J. Phys. Soc. Jpn. 90, 111004 (2021).
  • Tam et al. [2022] C. C. Tam, J. Choi, X. Ding, S. Agrestini, A. Nag, M. Wu, B. Huang, H. Luo, P. Gao, M. García-Fernández, L. Qiao, and K.-J. Zhou, Charge density waves in infinite-layer NdNiO2 nickelates, Nat. Mater. 21, 1116 (2022).
  • Krieger et al. [2022] G. Krieger, L. Martinelli, S. Zeng, L. E. Chow, K. Kummer, R. Arpaia, M. Moretti Sala, N. B. Brookes, A. Ariando, N. Viart, M. Salluzzo, G. Ghiringhelli, and D. Preziosi, Charge and Spin Order Dichotomy in NdNiO2subscriptNdNiO2{\mathrm{NdNiO}}_{2}roman_NdNiO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Driven by the Capping Layer, Phys. Rev. Lett. 129, 027002 (2022).
  • Rossi et al. [2022] M. Rossi, M. Osada, J. Choi, S. Agrestini, D. Jost, Y. Lee, H. Lu, B. Y. Wang, K. Lee, A. Nag, Y.-D. Chuang, C.-T. Kuo, S.-J. Lee, B. Moritz, T. P. Devereaux, Z.-X. Shen, J.-S. Lee, K.-J. Zhou, H. Y. Hwang, and W.-S. Lee, A broken translational symmetry state in an infinite-layer nickelate, Nat. Phys. 18, 869 (2022).
  • Rossi et al. [2024] M. Rossi, H. Lu, K. Lee, B. H. Goodge, J. Choi, M. Osada, Y. Lee, D. Li, B. Y. Wang, D. Jost, S. Agrestini, M. Garcia-Fernandez, Z. X. Shen, K.-J. Zhou, E. Been, B. Moritz, L. F. Kourkoutis, T. P. Devereaux, H. Y. Hwang, and W. S. Lee, Universal orbital and magnetic structures in infinite-layer nickelates, Phys. Rev. B 109, 024512 (2024).
  • Li et al. [2019] D. Li, K. Lee, B. Y. Wang, M. Osada, S. Crossley, H. R. Lee, Y. Cui, Y. Hikita, and H. Y. Hwang, Superconductivity in an infinite-layer nickelate, Nature 572, 624 (2019).
  • Zeng et al. [2020] S. Zeng, C. S. Tang, X. Yin, C. Li, M. Li, Z. Huang, J. Hu, W. Liu, G. J. Omar, H. Jani, Z. S. Lim, K. Han, D. Wan, P. Yang, S. J. Pennycook, A. T. S. Wee, and A. Ariando, Phase Diagram and Superconducting Dome of Infinite-Layer Nd1xSrxNiO2subscriptNd1𝑥subscriptSr𝑥subscriptNiO2{\mathrm{Nd}}_{1-x}{\mathrm{Sr}}_{x}{\mathrm{NiO}}_{2}roman_Nd start_POSTSUBSCRIPT 1 - italic_x end_POSTSUBSCRIPT roman_Sr start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_NiO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Thin Films, Phys. Rev. Lett. 125, 147003 (2020).
  • Li et al. [2020b] D. Li, B. Y. Wang, K. Lee, S. P. Harvey, M. Osada, B. H. Goodge, L. F. Kourkoutis, and H. Y. Hwang, Superconducting Dome in Nd1xSrxNiO2subscriptNd1𝑥subscriptSr𝑥subscriptNiO2{\mathrm{Nd}}_{1-x}{\mathrm{Sr}}_{x}{\mathrm{NiO}}_{2}roman_Nd start_POSTSUBSCRIPT 1 - italic_x end_POSTSUBSCRIPT roman_Sr start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_NiO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Infinite Layer Films, Phys. Rev. Lett. 125, 027001 (2020b).
  • Osada et al. [2020] M. Osada, B. Y. Wang, K. Lee, D. Li, and H. Y. Hwang, Phase diagram of infinite layer praseodymium nickelate Pr1xSrxNiO2subscriptPr1𝑥subscriptSr𝑥subscriptNiO2{\mathrm{Pr}}_{1-x}{\mathrm{Sr}}_{x}{\mathrm{NiO}}_{2}roman_Pr start_POSTSUBSCRIPT 1 - italic_x end_POSTSUBSCRIPT roman_Sr start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_NiO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT thin films, Phys. Rev. Mater. 4, 121801(R) (2020).
  • Gao et al. [2021] Q. Gao, Y. Zhao, X.-J. Zhou, and Z. Zhu, Preparation of Superconducting Thin Films of Infinite-Layer Nickelate Nd0.8Sr0.2NiO2Chinese Physics Letters 38, 077401 (2021).
  • Zeng et al. [2022] S. Zeng, C. Li, L. E. Chow, Y. Cao, Z. Zhang, C. S. Tang, X. Yin, Z. S. Lim, J. Hu, P. Yang, and A. Ariando, Superconductivity in infinite-layer nickelate La1-xCaxNiO2 thin films, Sci. Adv. 8, eabl9927 (2022).
  • Benckiser et al. [2022] E. Benckiser, M. Hepting, and B. Keimer, Neighbours in charge, Nat. Mater. 21, 1102 (2022).
  • Anisimov et al. [1999] V. I. Anisimov, D. Bukhvalov, and T. M. Rice, Electronic structure of possible nickelate analogs to the cuprates, Phys. Rev. B 59, 7901 (1999).
  • Hepting et al. [2021] M. Hepting, M. P. M. Dean, and W.-S. Lee, Soft X-Ray Spectroscopy of Low-Valence Nickelates, Front. Phys. 9:808683 (2021).
  • Hepting et al. [2020] M. Hepting, D. Li, C. J. Jia, H. Lu, E. Paris, Y. Tseng, X. Feng, M. Osada, E. Been, Y. Hikita, Y.-D. Chuang, Z. Hussain, K. J. Zhou, A. Nag, M. Garcia-Fernandez, M. Rossi, H. Y. Huang, D. J. Huang, Z. X. Shen, T. Schmitt, H. Y. Hwang, B. Moritz, J. Zaanen, T. P. Devereaux, and W. S. Lee, Electronic structure of the parent compound of superconducting infinite-layer nickelates, Nat. Mater. 19, 381 (2020).
  • Goodge et al. [2021] B. H. Goodge, D. Li, K. Lee, M. Osada, B. Y. Wang, G. A. Sawatzky, H. Y. Hwang, and L. F. Kourkoutis, Doping evolution of the mott–hubbard landscape in infinite-layer nickelates, Proc. Natl. Acad. Sci. U.S.A. 118, e2007683118 (2021).
  • Rossi et al. [2021] M. Rossi, H. Lu, A. Nag, D. Li, M. Osada, K. Lee, B. Y. Wang, S. Agrestini, M. Garcia-Fernandez, J. J. Kas, Y.-D. Chuang, Z. X. Shen, H. Y. Hwang, B. Moritz, K.-J. Zhou, T. P. Devereaux, and W. S. Lee, Orbital and spin character of doped carriers in infinite-layer nickelates, Phys. Rev. B 104, L220505 (2021).
  • Hayward et al. [1999] M. A. Hayward, M. A. Green, M. J. Rosseinsky, and J. Sloan, Sodium Hydride as a Powerful Reducing Agent for Topotactic Oxide Deintercalation Synthesis and Characterization of the Nickel(I) Oxide LaNiO2J. Am. Chem. Soc. 121, 8843 (1999).
  • Hayward and Rosseinsky [2003] M. A. Hayward and M. J. Rosseinsky, Synthesis of the infinite layer Ni(I) phase NdNiO2+x by low temperature reduction of NdNiO3 with sodium hydride, Solid State Sci. 5, 839 (2003).
  • Crespin et al. [2005] M. Crespin, O. Isnard, F. Dubois, J. Choisnet, and P. Odier, LaNiO2: Synthesis and structural characterization, J. Solid State Chem. 178, 1326 (2005).
  • Wang et al. [2020] B.-X. Wang, H. Zheng, E. Krivyakina, O. Chmaissem, P. P. Lopes, J. W. Lynn, L. C. Gallington, Y. Ren, S. Rosenkranz, J. F. Mitchell, and D. Phelan, Synthesis and characterization of bulk Nd1xSrxNiO2subscriptNd1𝑥subscriptSr𝑥subscriptNiO2{\mathrm{Nd}}_{1-x}{\mathrm{Sr}}_{x}\mathrm{Ni}{\mathrm{O}}_{2}roman_Nd start_POSTSUBSCRIPT 1 - italic_x end_POSTSUBSCRIPT roman_Sr start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_NiO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and Nd1xSrxNiO3subscriptNd1𝑥subscriptSr𝑥subscriptNiO3{\mathrm{Nd}}_{1-x}{\mathrm{Sr}}_{x}\mathrm{Ni}{\mathrm{O}}_{3}roman_Nd start_POSTSUBSCRIPT 1 - italic_x end_POSTSUBSCRIPT roman_Sr start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_NiO start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPTPhys. Rev. Mater. 4, 084409 (2020).
  • Ortiz et al. [2022] R. A. Ortiz, P. Puphal, M. Klett, F. Hotz, R. K. Kremer, H. Trepka, M. Hemmida, H.-A. Krug von Nidda, M. Isobe, R. Khasanov, H. Luetkens, P. Hansmann, B. Keimer, T. Schäfer, and M. Hepting, Magnetic correlations in infinite-layer nickelates: An experimental and theoretical multimethod study, Phys. Rev. Research 4, 023093 (2022).
  • Klett et al. [2022] M. Klett, P. Hansmann, and T. Schäfer, Magnetic Properties and Pseudogap Formation in Infinite-Layer Nickelates: Insights From the Single-Band Hubbard Model, Front. Phys. 10:834682 (2022).
  • Fowlie et al. [2022] J. Fowlie, M. Hadjimichael, M. M. Martins, D. Li, M. Osada, B. Y. Wang, K. Lee, Y. Lee, Z. Salman, T. Prokscha, J.-M. Triscone, H. Y. Hwang, and A. Suter, Intrinsic magnetism in superconducting infinite-layer nickelates, Nat. Phys. 18, 1043 (2022).
  • Puphal et al. [2022] P. Puphal, V. Pomjakushin, R. A. Ortiz, S. Hammoud, M. Isobe, B. Keimer, and M. Hepting, Investigation of hydrogen incorporations in bulk infinite-layer nickelates, Front. Phys. 10:842578 (2022).
  • Lu et al. [2021] H. Lu, M. Rossi, A. Nag, M. Osada, D. F. Li, K. Lee, B. Y. Wang, M. Garcia-Fernandez, S. Agrestini, Z. X. Shen, E. M. Been, B. Moritz, T. P. Devereaux, J. Zaanen, H. Y. Hwang, K.-J. Zhou, and W. S. Lee, Magnetic excitations in infinite-layer nickelates, Science 373, 213 (2021).
  • Worm et al. [2023] P. Worm, Q. Wang, M. Kitatani, I. Biało, Q. Gao, X. Ren, J. Choi, D. Csontosová, K.-J. Zhou, X. Zhou, Z. Zhu, L. Si, J. Chang, J. M. Tomczak, and K. Held, Spin fluctuations sufficient to mediate superconductivity in nickelates, arXiv preprint arXiv:2312.08260  (2023).
  • Pelliciari et al. [2023] J. Pelliciari, N. Khan, P. Wasik, A. Barbour, Y. Li, Y. Nie, J. M. Tranquada, V. Bisogni, and C. Mazzoli, Comment on newly found charge density waves in infinite layer nickelates, arXiv preprint arXiv:2306.15086  (2023).
  • Tam et al. [2023] C. C. Tam, J. Choi, X. Ding, S. Agrestini, A. Nag, M. Wu, B. Huang, H. Luo, P. Gao, M. Garcia-Fernandez, L. Qiao, and K.-J. Zhou, Reply to ”Comment on newly found Charge Density Waves in infinite layer Nickelates”, arXiv preprint arXiv:2307.13569  (2023).
  • Raji et al. [2023] A. Raji, G. Krieger, N. Viart, D. Preziosi, J.-P. Rueff, and A. Gloter, Charge distribution across capped and uncapped infinite-layer neodymium nickelate thin films, Small 19, 2304872 (2023).
  • Parzyck et al. [2024] C. T. Parzyck, N. K. Gupta, Y. Wu, V. Anil, L. Bhatt, M. Bouliane, R. Gong, B. Z. Gregory, A. Luo, R. Sutarto, F. He, Y.-D. Chuang, T. Zhou, G. Herranz, L. F. Kourkoutis, A. Singer, D. G. Schlom, D. G. Hawthorn, and K. M. Shen, Absence of 3a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT charge density wave order in the infinite-layer nickelate NdNiO2Nat. Mater.  (2024).
  • Puphal et al. [2023a] P. Puphal, B. Wehinger, J. Nuss, K. Küster, U. Starke, G. Garbarino, B. Keimer, M. Isobe, and M. Hepting, Synthesis and physical properties of LaNiO2subscriptLaNiO2{\mathrm{LaNiO}}_{2}roman_LaNiO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT crystals, Phys. Rev. Mater. 7, 014804 (2023a).
  • Puphal et al. [2023b] P. Puphal, V. Sundaramurthy, V. Zimmermann, K. Küster, U. Starke, M. Isobe, B. Keimer, and M. Hepting, Phase formation in hole- and electron-doped rare-earth nickelate single crystals, APL Materials 11, 081107 (2023b).
  • Puphal et al. [2021] P. Puphal, Y.-M. Wu, K. Fürsich, H. Lee, M. Pakdaman, J. A. N. Bruin, J. Nuss, Y. E. Suyolcu, P. A. van Aken, B. Keimer, M. Isobe, and M. Hepting, Topotactic transformation of single crystals: From perovskite to infinite-layer nickelates, Sci. Adv. 7, eabl8091 (2021).
  • Suyolcu et al. [2021] Y. E. Suyolcu, K. Fürsich, M. Hepting, Z. Zhong, Y. Lu, Y. Wang, G. Christiani, G. Logvenov, P. Hansmann, M. Minola, B. Keimer, P. A. van Aken, and E. Benckiser, Control of the metal-insulator transition in NdNiO3subscriptNdNiO3{\mathrm{NdNiO}}_{3}roman_NdNiO start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT thin films through the interplay between structural and electronic properties, Phys. Rev. Mater. 5, 045001 (2021).
  • Wu et al. [2023] Y.-M. Wu, P. Puphal, H. Lee, J. Nuss, M. Isobe, B. Keimer, M. Hepting, Y. E. Suyolcu, and P. A. van Aken, Topotactically induced oxygen vacancy order in nickelate single crystals, Phys. Rev. Mater. 7, 053609 (2023).
  • Hepting et al. [2018] M. Hepting, R. J. Green, Z. Zhong, M. Bluschke, Y. E. Suyolcu, S. Macke, A. Frano, S. Catalano, M. Gibert, R. Sutarto, F. He, G. Cristiani, G. Logvenov, Y. Wang, P. A. van Aken, P. Hansmann, M. Le Tacon, J.-M. Triscone, G. A. Sawatzky, B. Keimer, and E. Benckiser, Complex magnetic order in nickelate slabs, Nat. Phys. 14, 1097 (2018).
  • [62] See Supplemental Material for the further characterization of LaNiO2 crystals; details of the analysis of twin domains by electron backscatter diffraction (EBSD); the self-absorption correction of RIXS spectra; and complementary RIXS spectra further discussing the absence of charge order signal. The Supplemental Material also contains Refs. [56, 57, 44, 16, 12, 28, 26].
  • Zhou et al. [2022] K.-J. Zhou, A. Walters, M. Garcia-Fernandez, T. Rice, M. Hand, A. Nag, J. Li, S. Agrestini, P. Garland, H. Wang, S. Alcock, I. Nistea, B. Nutter, N. Rubies, G. Knap, M. Gaughran, F. Yuan, P. Chang, J. Emmins, and G. Howell, I21: an advanced high-resolution resonant inelastic X-ray scattering beamline at Diamond Light Source, J. Synchrotron Rad. 29, 563 (2022).
  • Hsu et al. [2022] Y.-T. Hsu, M. Osada, B. Y. Wang, M. Berben, C. Duffy, S. P. Harvey, K. Lee, D. Li, S. Wiedmann, H. Y. Hwang, and N. E. Hussey, Correlated Insulating Behavior in Infinite-Layer Nickelates, Front. Phys. 10:846639 (2022).
  • Ament et al. [2011] L. J. P. Ament, M. van Veenendaal, T. P. Devereaux, J. P. Hill, and J. van den Brink, Resonant inelastic x-ray scattering studies of elementary excitations, Rev. Mod. Phys. 83, 705 (2011).
  • Jia et al. [2014] C. J. Jia, E. A. Nowadnick, K. Wohlfeld, Y. F. Kung, C.-C. Chen, S. Johnston, T. Tohyama, B. Moritz, and T. P. Devereaux, Persistent spin excitations in doped antiferromagnets revealed by resonant inelastic light scattering, Nat. Commun. 5, 3314 (2014).
  • Christensen et al. [2007] N. B. Christensen, H. M. Rønnow, D. F. McMorrow, A. Harrison, T. G. Perring, M. Enderle, R. Coldea, L. P. Regnault, and G. Aeppli, Quantum dynamics and entanglement of spins on a square lattice, Proc. Natl. Acad. Sci. U.S.A. 104, 15264 (2007).
  • Singh [1989] R. R. P. Singh, Thermodynamic parameters of the T=0𝑇0T=0italic_T = 0, spin-1/2 square-lattice Heisenberg antiferromagnet, Phys. Rev. B 39, 9760 (1989).
  • Lee et al. [2023] K. Lee, B. Y. Wang, M. Osada, B. H. Goodge, T. C. Wang, Y. Lee, S. Harvey, W. J. Kim, Y. Yu, C. Murthy, S. Raghu, L. F. Kourkoutis, and H. Y. Hwang, Linear-in-temperature resistivity for optimally superconducting (Nd,Sr)NiO2Nature 619, 288 (2023).
  • Sun et al. [2023] W. Sun, Y. Li, R. Liu, J. Yang, J. Li, W. Wei, G. Jin, S. Yan, H. Sun, W. Guo, Z. Gu, Z. Zhu, Y. Sun, Z. Shi, Y. Deng, X. Wang, and Y. Nie, Evidence for Anisotropic Superconductivity Beyond Pauli Limit in Infinite-Layer Lanthanum Nickelates, Adv. Mater. 35, 2303400 (2023).
  • Chow and Ariando [2022] L. E. Chow and A. Ariando, Infinite-Layer Nickelate Superconductors: A Current Experimental Perspective of the Crystal and Electronic Structures, Front. Phys. 10:834658 (2022).