Skip to main content
Nature Communications logoLink to Nature Communications
. 2024 Nov 6;15:9597. doi: 10.1038/s41467-024-53863-5

Electronic and magnetic excitations in La3Ni2O7

Xiaoyang Chen 1,#, Jaewon Choi 2,#, Zhicheng Jiang 3, Jiong Mei 4,5, Kun Jiang 4,5, Jie Li 6, Stefano Agrestini 2, Mirian Garcia-Fernandez 2, Hualei Sun 7, Xing Huang 8, Dawei Shen 3, Meng Wang 8, Jiangping Hu 4,9, Yi Lu 6,10,, Ke-Jin Zhou 2,, Donglai Feng 3,10,11,
PMCID: PMC11541582  PMID: 39505866

Abstract

High-temperature superconductivity was discovered in the pressurized nickelate La3Ni2O7 which has a unique bilayer structure and mixed valence state of nickel. The properties at ambient pressure contain crucial information of the fundamental interactions and bosons mediating superconducting pairing. Here, using X-ray absorption spectroscopy and resonant inelastic X-ray scattering, we identified that Ni 3dx2y2, Ni 3dz2, and ligand oxygen 2p orbitals dominate the low-energy physics with a small charge-transfer energy. Well-defined optical-like magnetic excitations soften into quasi-static spin-density-wave ordering, evidencing the strong electronic correlation and rich magnetic properties. Based on an effective Heisenberg spin model, we extract a much stronger inter-layer effective magnetic superexchange than the intra-layer ones and propose two viable magnetic structures. Our findings emphasize that the Ni 3dz2 orbital bonding within the bilayer induces novel electronic and magnetic excitations, setting the stage for further exploration of La3Ni2O7 superconductor.

Subject terms: Electronic properties and materials, Magnetic properties and materials


It was recently found that a certain nickelate compound, La3Ni2O7, at moderately high pressures has a superconducting phase that persists to above liquid nitrogen temperatures. Here, by studying the parent phase at ambient pressure, Chen et al uncover rich magnetic properties and show the vital role of the strong bonding of the inter-layer Ni orbitals in the magnetic and electronic excitations.

Introduction

The striking discovery of high-temperature superconductivity (HTSC) at 80 K in a bilayer nickelate La3Ni2O7 under a pressure of about 14 GPa ignited a new wave of studying HTSC in nickelates17. Unlike cuprate superconductors with a Cu2+ 3d9 electron configuration, La3Ni2O7 hosts Ni ions with mixed 2 + (3d8) and 3 + (3d7) valences with unpaired electrons in both 3dx2y2 and 3dz2 orbitals from a Ni-O bilayer structure1,2,813. In particular, the molecular bonding between the two inter-layer Ni 3dz2 orbitals through the apical O pz orbital, together with Ni 3dx2y2 orbitals, are proposed by theory as a critical ingredient for the low-energy electronic structure of La3Ni2O72,822. The orbital character governing the electronic properties of the unconventional superconductors is essential for understanding the underlying pairing mechanism. In cuprates, the small charge-transfer energy and strong hybridisation between Cu 3dx2y2 and O 2p orbitals lead to the formation of the strongly correlated Zhang-Rice singlet band, which serves as the foundation for describing the electronic properties including the superconducting pairing interaction with dx2y2 symmetry23. On the other hand, the iron-based superconductors feature relatively weaker correlation and multiple 3d bands near the Fermi surface24. The orbital-dependent correlation and the strong anisotropy in the electronic hopping result in a distinct s pairing symmetry. At first sight, La3Ni2O7 appears to be a sibling of iron-based superconductors owing to the multi-orbital nature and the bad metallicity in the undoped parental phase. However, perovskite nickelates are also known to exhibit strong electronic correlation and small charge-transfer energy25,26, resembling cuprates. Theories to date vary in their opinions on which orbitals are most relevant for the electronic properties, especially the superconductivity, in La3Ni2O78,9,1522.

The antiferromagnetic (AFM) superexchange interaction is accepted as another important ingredient of unconventional superconductors. Upon the doping of charge carriers, the long-range AFM-ordered parental phase evolves into one with short-range AFM spin fluctuations, which may mediate the superconducting pairing. In a sizable part of the phase diagram, the interplay among spin, charge, and lattice degrees of freedom often leads to exotic ordering phases such as the periodic density modulation of charge or spin. In cuprates and iron-based superconductors, charge (CDW) and spin density waves (SDW) intertwine with superconducting phase which is regarded as being closely relevant to HTSC23,24. The bilayer structure and the multi-orbital nature of La3Ni2O7 have profound impact on its magnetism as well, which plays a pivotal role in theories on this novel superconductor1,8,27,28. Some theory suggest the importance of the interlayer antiferromagnetic coupling Jz between dz2 orbitals1,8; some others advocate that the strong interlayer coupling would cause the bilayer splitting of band structure, while in-plane magnetic exchange interactions play a dominant role in superconductivity12,15. In the as-grown La3Ni2O7 crystal at ambient pressure, resistivity measurements found a kink-like transition at around 153 K implying a possible CDW or SDW state29. NMR studies found CDW order possibly mixed with SDW order in polycrystalline La3Ni2O730, and most recently SDW order was revealed in single crystal La3Ni2O731. In addition, μSR experiments suggested that a static long-range magnetic order emerges in polycrystalline La3Ni2O7  ~ 150 K32,33. Despite the proposals of potential density waves, NMR and μSR experiments reported that the magnetic moment per Ni site is  ~ 0.08 μB and 0.3–0.7 μB, respectively31,33.

Given the currently limited knowledge on the essential electronic and magnetic properties, such as the charge-transfer energy and the magnetic exchange interactions, experimental verification is indispensable. In this work, we employ X-ray absorption spectroscopy (XAS) and resonant inelastic X-ray scattering (RIXS) at both Ni L3-edge and O K-edge of La3Ni2O7 single crystal at ambient pressure. These spectroscopic and scattering techniques are sensitive to low-energy electronic and magnetic structures together with elementary excitations, and thus are ideally suited for tackling the core issues in La3Ni2O7.

Results

Electronic structure of La3Ni2O7

As-grown La3Ni2O7 crystallises in an orthorhombic structure with the space group of Amam1. We define the reciprocal space index (HKL) based on the pseudo-tetragonal unit cell (Fig. 1a, b, “Method” section). Figure 1c shows the experimental geometry, in which the incident X-ray is linearly polarised, while the scattered X-ray is typically non-polarised but otherwise polarised if stated explicitly (see “Method” section).

Fig. 1. XAS spectra and the incident energy-dependent RIXS maps in La3Ni2O7.

Fig. 1

a Schematic top view of the NiO2 plane in La3Ni2O7. The solid black square represents the pseudo-tetragonal unit cell with a lattice constant aT  ~ 3.833 Å, while the dashed black square represents the real orthorhombic in-plane unit cell when considering the tilting of Ni-O octahedra. b In-plane Brillouin zone (BZ) for the pseudo-tetragonal unit cell. c Sketch of the RIXS experimental geometry. Details of the setup are described in Method. d, e σ polarised XAS spectra of La3Ni2O7 (red filled circles) taken at the O K-edge (d) and Ni L3-edge (e), respectively. The latter is displayed after subtracting the background of La M4-edge. The calculated Ni L3-XAS (red curve) is also shown. XAS spectra measured on NiO (Ni2+) and NdNiO3 (Ni3+) (black-filled circles) are presented as references. The XAS data of Nd1.5Sr0.5NiO4 extracted from ref. 42 are also displayed. f, g RIXS intensity maps measured as a function of incident photon energy with σ- (f) and π- (g) polarised photons, respectively. The corresponding XAS spectrum is superimposed as a solid white curve on each map. Both XAS and RIXS spectra were collected at 20 K at a grazing-in incident angle of 20°. h Integral RIXS spectra in (f) and (g) over the incident energy range [851.8 eV, 853.4 eV]. The grey solid bars display the multiplet calculations for the Ni L3-RIXS.

Figure 1d, e illustrate XAS spectra of La3Ni2O7 taken near the O K-edge and Ni L3-edge, respectively. A sizable O K- pre-edge peak at  ~ 528.5 eV originates from oxygen 1s electron excitations into the unoccupied oxygen 2p ligand hole state near the Fermi level, as observed for the Zhang-Rice singlet state in cuprate superconductors34. The Ni L3-XAS data show a sharp resonant peak around 852.4 eV, followed by a broad satellite peak at a higher energy. As the Ni valence 2.5+ of La3Ni2O7 falls in between the archetypal nickelates NiO and NdNiO3, the XAS spectra of La3Ni2O7 can be qualitatively understood in relation to these two. NiO is a canonical charge-transfer insulator in the Zaanen-Sawatzky-Allen classification, whose large charge-transfer energy Δ (≈5 eV) suppresses the charge fluctuations between the Ni 3d and ligand oxygen 2p orbitals despite their large orbital hopping integral35. Consequently, its ground state is well described by α3d8+β3d9L_ (α2 + β2 ≲ 1 and L_ denotes a ligand hole) with a dominant 3d8 character (α2 ≈ 0.8)35,36. On the other hand, the perovskite NdNiO3 with a nominal 3d7 configuration is widely acknowledged as a negative charge-transfer system, where electrons from ligand oxygen spontaneously transfer onto Ni cations, resulting in a ground state with a leading 3d8L_ contribution25,26,3739. Such a substantial ligand hole concentration is underscored by the pronounced pre-edge hole peak in the O K-edge XAS of NdNiO3, similar to that of La3Ni2O7 (Fig. 1d). This is distinct from NiO, where the pre-peak is absent, and the unoccupied ligand states are at an elevated energy across the charge-transfer gap. For the Ni L3-XAS, the prominent resonant peak of La3Ni2O7 is also observed for NiO and NdNiO3 at a similar energy (Fig. 1e), which was previously identified as the Ni 2p  → 3d8 or 3d8 + 3d8L_ transitions into the half-filled eg states, respectively36,39. A broad satellite peak at a higher energy is likewise seen for NdNiO3, originating mainly from a part of its ground state wavefunction that contains additional ligand holes3941. Similar spectral profiles are present at nominally half-doped nickelate Nd1.5Sr0.5NiO4 at both the O K- and Ni L3- edges42. The above suggests a predominant 3d8 occupancy on the Ni cation in La3Ni2O7, accompanied by a significant amount of ligand holes.

Figure 1f, g shows the incident-energy-dependent RIXS measurements of La3Ni2O7 across the Ni L3-edge. A clear low-energy excitation (~70 meV) is observed near the elastic peak which will be discussed in the next section. The sharp XAS resonance at  ~852.4 eV decays mainly to a final state of a localised excitation at ~1 eV, known as the t2g → egdd orbital excitation similar to NiO and NdNiO340,43,44. The band-like fluorescence excitation, resonating across the broad satellite XAS peak, stems from the delocalised Ni-O hybridised continuum states40,41. The intensity distribution of the fluorescence contracts under π polarisation that couples stronger to the 3dz2 orbital, indicative of a smaller out-of-plane bandwidth arising from the quasi-two-dimensional structure. In addition, distinct from NdNiO3, two extra dd excitations show up in La3Ni2O7 (at around 0.4 eV and 1.6 eV). They exhibit stronger intensities under π polarisation, suggesting a more prominent involvement of the 3dz2 orbital in them.

To gain a quantitative understanding of XAS and RIXS measurements, we built a double-cluster model capturing the bilayer structure of La3Ni2O7 and then carried out multiplet calculations for Ni L3- XAS and RIXS spectra (see details in Supplementary Note 2). Systematic optimisations of the calculated spectra suggest that the charge-transfer energy Δ falls between 0 and 2 eV, pointing out the rather small-charge-transfer nature of La3Ni2O745. This result is reasonable since Δ is  ~ 5 eV and  ~ 0 for NiO and NdNiO3, respectively41,46. With the estimated range of Δ, the ground state wavefunction of La3Ni2O7 can be deduced to approximately α3d8+β3d8L_+γ3d7 with leading α2 and β2. The calculated XAS for Δ = 0.5 eV is shown in Fig. 1e, which corresponds to a ground state with (α2β2γ2)  ≈ (0.4, 0.3, 0.2).

The two sets of RIXS excitations centred around 0.4 eV and 1.0 eV in Fig. 1h are well captured in the calculated RIXS spectra. The higher-energy excitation around 1.6 eV is less prominent in calculation partly due to the limited degrees of freedom in the model. To further understand the nature of these excitations, we characterise these excited states in the double-cluster model by evaluating their corresponding orbital occupations and wave function configurations. The excitations at 0.4 eV involve charge transfers between the orbitals of z2 and x2 − y2 symmetry and are of mixed charge and spin type. The peak centred around 1 eV involves the transition between the dz2 and dxz/yz orbitals, which characterise the crystal-field splitting between the 3deg and t2g orbitals. They involve relatively small movements of the ligand states, signifying almost pure dd-type excitations. Higher-energy excitations between 1.3 ~ 1.5 eV correspond to more complex dd-type excitations, involving transitions between all 3d orbitals. The excitation energy is somewhat lower than that observed in experiment, potentially owing to the limited in-plane size of the cluster model, which may underrepresent the bandwidth of the planar orbitals. Note that there is evident fluctuation of wave function configuration weights between the local d7 and d8L_ as well as the global d7d8 and d8d8L_ over almost the entire energy range owing to the small charge-transfer energy (Supplementary Note 2). Remarkably, we found that both the XAS line shape and the lower dd excitation ( ~ 0.4 eV) in RIXS show marked difference upon tuning the inter-layer hopping strength mediated by the 3dz2 -OAP 2pz - 3dz2 orbital overlap in the calculation (OAP stands for the apical oxygen), underlining the importance of the inter-layer coupling for the electronic structure (Supplementary Note 2). This result is consistent with previous experimental report1, and lends support to several recent theoretical works emphasising on the importance of the bilayer structure2,8,1320.

Magnetic excitations

Figure 2 summarises the detailed energy-momentum dependence of low-energy excitations in La3Ni2O7 taken at the incident energy of 852.4 eV corresponding to the sharp resonance peak of Ni L3-XAS. Figure 2a, b show strongly dispersive excitations along directions illustrated in insets. The excitations reach maximal energy of about 70 meV at (0, 0) and (0.5, 0) while softening to zero energy (within the experimental energy resolution) at (0.25, 0.25) where a quasi-elastic scattering peak is formed. The latter signals the formation of translational symmetry breaking, i. e, a superstructure along (π, π) direction with a size four times of the crystal lattice structure. Similar dispersive low-energy excitations also appear when excited by π incident X-rays polarisation (Fig. S6). Along the out-of-plane direction, this mode does not exhibit sizable dispersion as a function of L, indicating its quasi-two-dimensional nature (Fig. 2c).

Fig. 2. Energy-momentum dependent magnon in La3Ni2O7.

Fig. 2

a RIXS intensity maps along high-symmetry directions as indicated in the insets. Data were collected at 20 K using σ-polarised X-ray at the Ni L3-edge of 852.4 eV. The red filled circles depict the peak positions of magnetic excitations here and throughout all panels of this figure. b RIXS spectra at representative projected in-plane momentum transfers. The weaker excitations at  ~ 120 meV may result from the multi-magnons. c L scan of RIXS spectra at the fixed q = (0.035, 0.035). d Polarimetric RIXS data at q = (0.035, 0.035, 2.727). The spectra are decomposed into ππ, πσ, σσ and σπ components.

As both magnon and phonon could contribute to the low-energy excitations, the polarimetric RIXS was employed to analyse the outgoing X-rays linear polarisation for unravelling the origin of these excitations (see Methods). Clearly, as shown in Fig. 2d, the inelastic excitation is present under the ππ, πσ, and σπ channels, while gets much reduced under the σσ channel. Such behaviour is in agreement with the assumption of a magnetic origin of the scattering and a recent polarimetric RIXS study on magnons in cuprates47,48. Our multiplet RIXS calculation of magnetic excitations in the double-cluster model confirmed the outgoing polarisation dependence (Fig. S5). Concerning phonons, in principle, their spectra weight should be present in the σσ channel. However, the corresponding polarimetric RIXS spectrum shows negligible spectral weight hence a minute contribution to the Ni L3-RIXS (Fig. 2d). We therefore conclude that the low-energy excitations observed at the Ni L3-edge are dominated by magnons. Interestingly, in the half-doped nickelate La3/2Sr1/2NiO4, which has the same nominal Ni2.5+ valence state as La3Ni2O7, an SDW order is formed near the wavevector (0.25, 0.25) from which a low-energy magnon emerges49. The similar superstructure and the magnon softening near the order wavevector suggest an SDW order exists in La3Ni2O7. The only difference is the dispersion near Γ point: the magnon in La3/2Sr1/2NiO4 is acoustic-like, whereas in La3Ni2O7 they are dominantly optical-like (Fig. 2a).

By fitting the magnon spectra to a damped harmonic oscillator (DHO) function χ(qω), we extracted the peak energy and width of the magnon (Supplementary Note 4)50. Three possible spin configurations consistent with the spin order at Q = (0.25, 0.25) can be constructed: the diagonal spin-charge stripe order as in half-doped La3/2Sr1/2NiO4 where Ni2+ spin and nominal Ni3+ charge stripes intertwined (Stripe-1, Fig. 3a)49; the SDW order could also be realised with homogeneous valence state Ni2.5+, ie. , a double-spin stripe order (Stripe-2, Fig. 3b) that is similar to the bi-collinear spin order in FeTe51; by exchanging the Stripe-1 charge stripe positions with those of a spin stripe, a third spin configuration could be achieved as a double spin-charge stripe order (Stripe-3 in Fig. S11c). For all these SDW orders, owing to the strong bilayer bonding, spins are antiferromagnetically aligned in the top and bottom NiO2 layers. To obtain the magnetic interaction parameters, we constructed an effective J1-J2-Jz Heisenberg model: H=iJzSitSib+ijαJ1SiαSjα+ijαJ2SiαSjα, where α is the layer index for the bottom (b) or top (t) layer, J1 and J2 are the nearest-neighbour and next-nearest-neighbour exchange couplings, respectively, in a single NiO2 layer, and Jz is the inter-layer exchange coupling along the c-axis. Owing to the metallic background, the exchange couplings JS should be considered as the Weiss molecular field governing the spin dynamics for a spin density wave order. All JS values are fitted to the experimental magnon dispersion by solving the semiclassic torque equations52 (Supplementary Note 6). We found that the magnon dispersions based on both Stripe-1 and Stripe-2 spin configurations agree with our RIXS result (Fig. 3c and Supplementary Note 6). Owing to the scattering matrix effect, the simulated acoustic magnon spectra are significantly weaker than the optical magnon, consistent with the experimental findings. In general, the inter-layer effective superexchange interaction is an order of magnitude larger than that of the intra-layer. The finding of a dominant magnetic interaction along the molecular bonding direction is in good accordance with previous theoretical calculation2. Interestingly, J2S here shows comparable strength to that in the half-doped La3/2Sr1/2NiO449. For the Stripe-2, the fitted J1S is negligibly weak comparing to the dominant inter-layer exchange interaction leading to a similar spin dynamics and magnon dispersion as in Stripe-1. Based on the above results and the currently limited information, we can conjecture the true spin configuration of La3Ni2O7 is either Stripe-1 or Stripe-2 or their mixture (see details in Supplementary Note 5).

Fig. 3. The spin configuration and the magnon dispersion of La3Ni2O7.

Fig. 3

a The spin configurations for the spin-charge stripe order (Stripe-1). To simplify the sketch only nickel cations are shown. The blue, red and black circles represent spin up Ni2+, spin down Ni2+, and the spinless Ni3+ sites, respectively. The solid lines illustrate the in-plane pseudo-tetragonal unit cells and the grey cubics represent the Ni-O octahedra. The fitted values of J1S, J2S, and JzS based on this spin configuration are noted (see details in Supplementary Note 6). b The spin configuration for the double spin stripe (Stripe-2), and the fitted value of J1S, J2S, and JzS. c The experimental magnon dispersion ϵq (red filled circles) and damping factor γq (black open circles) versus projected in-plane momentum transfer q along high-symmetry directions at 20 K. See fitting details in Supplementary Note 4. Error bars of ϵq were estimated by combining the uncertainty of the elastic peak position, linear background, and the standard deviation of the fits. Error bars of γq were estimated by combining the standard deviation of the fits. The horizontal dashed line marks the total energy resolution (36 meV). The results of an effective J1-J2-Jz Heisenberg model based on Stripe-1 order are overlaid. The results from the model based on Stripe-2 are also consistent with the experimental data. All calculations are performed based on single-domain configurations, without considering the effects of twinning. The blue curves represent the dispersion of two magnon modes, where the thickness of the lines and the depth of their colour represent the mode intensity. The detailed parameters are listed in Supplementary Note 6.

Spin-density-wave order

We now took an explicit examination on the SDW order. Polarimetric RIXS was used to confirm the magnetic origin of low-energy excitations, likewise, it was applied to characterise the SDW order in La3Ni2O7. The momentum-dependent quasi-elastic SDW scattering peak shows the same trend as magnon, ie. , sizable scattering intensities under ππ, πσ, and σπ except for σσ (Fig. 4a, b), confirming the magnetic origin of such SDW order. The same polarisation dependence was found in the polarimetric resonant X-ray scattering study of the magnetic order in NdNiO353. Further insight into the nature of the SDW was gained through the energy dependence of the SDW scattering at its order wavevector across the Ni L3-edge (Fig. 4c). Unlike the XAS spectra where La M4 shows a greater absorption intensity than that of Ni L3, the SDW scattering predominantly results from the Ni 3d - O 2p hybridised states. Furthermore, the SDW scattering peak exhibits a colossal polarisation dependence, namely, its intensity probed under π polarisation is ~30 times higher than that with σ polarisation. Figure 4d gives an example taken with 852.4 eV photons, which may indicate its strong association with Ni 3dz2 orbital. For that under π polarisation, the fitted peak centre value is  ~0.25 r.l.u. The half-width at half-maximum Γ = 0.0022 ± 0.0002 r.l.u. of the scattering peak corresponds to the in-plane correlation length (ξH = 1/Γ) of  ~ 27.7 nm. This is comparable to that of the long-range CDW order in La1.875Ba0.125CuO4 (~20 nm)54. A much broader peak with HWHM of about 0.3 r.l.u. is observed as a function of L along the direction of (0.25, 0.25, L), which corresponds to a short out-of-plane correlation length ξL of  ~ 0.2 nm, establishing the quasi-two-dimensional nature of the SDW order in La3Ni2O7 (Fig. 4e).

Fig. 4. SDW order at (0.25, 0.25) of La3Ni2O7.

Fig. 4

a, b Polarimetric SDW data. The spectra are decomposed into ππ, πσ, σσ and σπ components. c SDW peaks intensities as a function of incident photon energy and polarisation. The inset shows the XAS spectra at the La M4-edge and the Ni L3-edge. d SDW integrated intensity as a function of projected momentum transfer (q) along the (H, H) direction. e SDW intensity as a function of L at the fixed q = (0.25,0.25). f SDW peaks and their Lorentzian fits along the (H, H) direction at various temperatures. gi Temperature dependence of the SDW peak area (g), the correlation length (h) and the SDW wave vector position (i). Normal and polarimetric SDW intensity was integrated over an energy window equivalent to the FWHM energy resolution of 36 meV and 55 meV, respectively. The error bars represent the standard deviation derived from the Lorentzian fitting of the SDW peak.

The temperature dependence of the SDW order illustrates a substantial reduction in both the intensity and the correlation length when the temperature is raised above  ~ 150 K (Fig. 4f-h), while the SDW wavevector does not exhibit a discernible temperature dependence (Fig. 4i). The discovery of the SDW with a characteristic temperature of around 150 K is in line with the transport, NMR and μSR studies on La3Ni2O72932.

Discussion

Our RIXS and XAS measurements revealed the dispersive magnon and SDW order below 150 K in La3Ni2O7. Detailed analysis suggests that Ni 3dx2y2, Ni 3dz2, and O 2p orbitals dominate the low-energy physics with charge-transfer energy less than 2 eV, and the inter-layer effective magnetic superexchange interaction is much larger than the intra-layer ones. These give critical information for constructing the minimal orbital model for La3Ni2O7 superconductor. It is worth noting that the predominant inter-layer superexchange in La3Ni2O7 is rather unique owing to the strong molecular bonding of partially unoccupied inter-layer Ni 3dz2 orbitals. The situation can neither be realised in multilayer cuprates nor in multilayer nickelates with reduced valence states5557. In the latter two cases, the almost fully occupied Cu or Ni 3dz2 orbitals reduce substantially the inter-layer electronic hopping or the molecular bonding strength.

Apart from the extraordinary bilayer structure and the associated predominant magnetic exchange interaction, the electronic structure of La3Ni2O7 fits in general into the family of Ruddlesden-Popper (RP) nickelates. The formation of the Zhang-Rice-like hole band, the small charge-transfer energy, and the well-defined dispersive magnon allude to its nature of the strong electronic correlations58. The above are typical characteristics of the strongly correlated cuprates where charge- and spin-density modulation can take place. Moreover, the occurrence of SDW order at (0.25, 0.25) is reminiscent of that in the half-doped single-layer La3/2Sr1/2NiO4, where a spin-charge stripe order exists, and implies the same picture in La3Ni2O7 as illustrated in the scenario of Stripe-1 (Fig. 3a)45,49,59. Indeed in layered half-doped RP nickelates, manganites, and cobaltates, the spin-charge intertwined order is prevailing5961. On the other hand, the double spin stripe order accommodating homogeneous charge density (Stripe-2, Fig. 3b) may be possible too as the 3dx2y2 orbitals are more itinerant in-plane than the 3dz2 orbitals. Verifying the magnetic structure of transition metal oxides is possible via resonant soft X-ray scattering53,62. The final choice of the ground state magnetic structure depends on multiple competing interactions which may also impact on the superconductivity. Some theoretical studies taking the viewpoint of the strong inter-layer hybridisation suggest either d-wave or (d + is)-wave pairing symmetry with a dominant d-wave component1922. While in the weaker interaction regime, studies predict that La3Ni2O7 host s±-wave pairing symmetry8,9,1518.

Finally, we would like to extrapolate our findings to superconducting La3Ni2O7, here, a moderately high pressure induces a structural phase transition accompanied by a few percent shrinkage of the lattice constants, and the Ni-O-Ni bonding angles between adjacent NiO6 octahedra straighten to 180°1. Consequently, the electronic hopping is likely to increase, potentially suppressing density waves that compete with the superconductivity63,64. Furthermore, the magnetic superexchange Jz may get significantly enlarged due to the increased hopping along Ni-OAP-Ni. Despite the presence of Zhang-Rice singlet physics and competing orders as in cuprates, the reinforced molecular orbital bonding and the dominating inter-layer AFM interaction may be novel additions to the HTSC of such a bilayer nickelate superconductor.

Methods

Sample fabrication

La3Ni2O7 sample was fabricated by the high oxygen pressure floating zone technique and the details are described in29. The sample quality was checked by X-ray diffraction (XRD) and Laue diffraction (see details in Supplementary Note 1). Samples were cleaved to get a flat, clean surface before RIXS measurements.

XAS and RIXS measurements

XAS and RIXS measurements were performed at Beamline I21 at Diamond Light Source65. In this work, we describe the structural properties of La3Ni2O7 referencing to a pseudo-tetragonal unit cell with cell parameters aT = bT  ~ 3.833 Å and c = 20.45 Å. Reciprocal lattice units (r.l.u.) are defined (where 2π/aT = 2π/bT = 2π/c = 1) with Q = HaT * + KbT * + Lc*. The crystallographic aT-c (bT-c) plane of La3Ni2O7 single crystal was aligned within the horizontal scattering plane (Fig. 1c). The polar angular offsets (θ and χ) of the crystal were aligned by the (002) diffraction peak, and the azimuthal offset (ϕ) by SDW order peak, such that the c* axis lays in the scattering plane. The spectrometer arm was at a fixed position of Ω = 154° except for L scans where variable Ω was employed.

XAS spectra were collected with a grazing incidence angle of θ0 = 20° to probe both in-plane and out-of-plane unoccupied states. All XAS measurements were done at a temperature of 20 K with the exit slit opening to 30 μm. Total electron yield XAS spectra were collected using the draincurrent and normalised to the incoming beam intensity. Both linear vertical (σ) and horizontal (π) polarisations were used.

Energy-dependent RIXS measurements were performed at the grazing incidence angle of θ0 = 20° and the temperature of 20 K. The exit slit was open to 30 μm corresponding to an average energy resolution of 40 meV (FWHM). The incident energy range went from 851 to 855 eV in steps of 0.2 eV to fully capture the resonance behaviour across the Ni-L3 absorption peaks.

Momentum-dependent RIXS measurements were performed at the resonant energy of 852.4 eV at a temperature of 20 K with the exit slit opening to 20 μm corresponding to an average energy resolution of 36 meV (FWHM). The momentum resolution is 0.002 r.l.u. near the SDW wavevector at the Ni L3-edge. RIXS spectra were collected using both σ and π polarisations. The grazing out geometry (θ > Ω/2) was applied for the acquisition of RIXS spectra shown in the main text.

Polarimetric RIXS apparatus employs a graded multilayer designed for the Ni L3-edge with a grazing incidence angle of 20° lying perpendicular to the scattering plane. Measurements were performed at Q = (0.035, 0.035, L) and around (0.25, 0.25, L) to analyse the outgoing X-rays linear polarisation of the magnon and SDW ordering, respectively. The total energy resolution of the polarimetric RIXS is  ~ 55 meV (FWHM). Since the multilayer does not work at the exact Brewster’s angle, the outgoing polarised RIXS (the indirect RIXS) from the reflection of the multilayer will be a mixture of linearly polarised spectra. The direct and indirect RIXS spectral intensities are then given by the following formula:

Idirect=Iσ+Iπ 1
Iindirect=RσIσ+RπIπ 2

where Idirect and Iindirect stands for the outgoing nonpolarised and mixed polarised RIXS spectral intensity, respectively. From the above formula, the outgoing σ and π polarised RIXS spectra can be deduced:

Iπ=IindirectRσIdirectRπRσ 3
Iσ=IindirectRπIdirectRσRπ 4

In the above, Rσ (Rπ) refers to the multilayer reflectivity of the outgoing σ (π) polarised X-ray photon. At the Ni L3-edge, Rσ and Rπ is 14.1% and 9.1%, respectively, based on the calibration of the multilayer.

Theoretical calculations

The Ni L3-edge XAS and RIXS calculations shown in Fig. 1 were performed employing a fully correlated Ni2O11 cluster model, accounting for the two corner-sharing NiO6 octahedra within the pseudo-tetragonal unit cell. The noninteracting part of the Hamiltonian integrates material-specific on-site energies and hybridisations involving Ni 3d and O 2p orbitals, along with spin-orbit coupling within the Ni core 2p and 3d shells. Full Coulomb interactions within the Ni 3d shell and between the Ni 2p and 3d shells are included, with parametrization by Slater integrals scaled at 0.8 based on atomic Hartree-Fock values66. Comprehensive details regarding model construction and relevant parameters are described in Supplementary Note 2. The model was solved using the exact diagonalization method as implemented in QUANTY67.

The DFT calculations employ the Vienna ab-initio simulation package (VASP) code68 with the projector augmented wave (PAW) method69. The Perdew-Burke-Ernzerhof (PBE) exchange-correlation functional70 is used. The energy cutoff energy for expanding the wave functions into a plane-wave basis is set to be 500 eV. The Γ-centred k-mesh is used in KPOINTS files which are generated by VASPKIT71 with the KPT-resolved value equal to 0.02 for different unit cells. The SDW orders are calculated using the simplified rotation invariant approach based on the DFT+U method introduced by Dudarev et al. 72. Then, the effective Heisenberg interactions for the SDW orders are constructed. The magnon dispersions within the linear spin wave theory are calculated using the torque equation formalism52,57. The RIXS intensity for the magnon mode in the σ-π polarisation channel is calculated following the reference73. More details can be found in Supplementary Note 5 and 6.

Supplementary information

Acknowledgements

This work was supported by National Natural Science Foundation of China (Nos.11888101, U2032208, 12274207, 12174428, and 12174454), the New Cornerstone Science Foundation, and the National Key R&D Program of China (Nos. 2023YFA1406304, 2022YFA1403000, 2023YFA1406500, and 2023YFA1406002). We acknowledge Diamond Light Source for providing beamtime at I21 Beamline under Proposal MM35805 and the science commissioning beamtime for using the polarimeter at I21. Work at SYSU was as well supported by the Guangdong Basic and Applied Basic Research Funds (No. 2021B1515120015), Guangzhou Basic and Applied Basic Research Funds (Nos. 202201011123, 2024A04J6417), and Guangdong Provincial Key Laboratory of Magnetoelectric Physics and Devices (No. 2022B1212010008). J.C. acknowledges support from the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) through the Sejong Science Fellowship (Grant No. RS-2023-00252768).

Author contributions

J.C., X.Y.C., D.W.S., S.A., M.G.-F., D.L.F. and K.-J.Z. conducted XAS and RIXS experiments at Diamond Light Source. X.Y.C., S.A., J.C. and K.-J.Z. analysed the data. J.M., K.J. and J.P.H. performed DFT and stripe states calculations. J.L. and Y.L. performed multiplet calculations. H.L.S., X.H., and M.W. fabricated samples. X.Y.C. and Z.C.J. performed XRD and Laue measurements. K.-J.Z., K.J., Y.L., D.W.S., D.L.F. and X.Y.C. wrote the manuscript, with input from all authors. D.L.F. and K.-J.Z. are responsible for project direction and planning.

Peer review

Peer review information

Nature Communications thanks the anonymous reviewers for their contribution to the peer review of this work. A peer review file is available.

Data availability

All data shown in the main text are available via Zenodo data repository (10.5281/zenodo.13955595).

Code availability

All code used to perform the XAS and RIXS calculation is available from the corresponding authors upon reasonable request.

Competing interests

The authors declare no competing interests.

Footnotes

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

These authors contributed equally: Xiaoyang Chen, Jaewon Choi.

Change history

12/3/2024

A Correction to this paper has been published: 10.1038/s41467-024-54993-6

Contributor Information

Yi Lu, Email: yilu@nju.edu.cn.

Ke-Jin Zhou, Email: kejin.zhou@diamond.ac.uk.

Donglai Feng, Email: dlfeng@ustc.edu.cn.

Supplementary information

The online version contains supplementary material available at 10.1038/s41467-024-53863-5.

References

  • 1.Sun, H. et al. Signatures of superconductivity near 80 K in a nickelate under high pressure. Nature621, 493–498 (2023). [DOI] [PubMed] [Google Scholar]
  • 2.Luo, Z., Hu, X., Wang, M., Wú, W. & Yao, D.-X. Bilayer two-orbital model of La3Ni2O7 under pressure. Phys. Rev. Lett.131, 126001 (2023). [DOI] [PubMed] [Google Scholar]
  • 3.Christiansson, V., Petocchi, F. & Werner, P. Correlated electronic structure of La3Ni2O7 under pressure. Phys. Rev. Lett.131, 206501 (2023). [DOI] [PubMed] [Google Scholar]
  • 4.Yang, S. et al. Effective Bi-layer model hamiltonian and density-matrix renormalization group study for the high-Tc superconductivity in La3Ni2O7 under high pressure. Chin. Phys. Lett.40, 127401 (2023). [Google Scholar]
  • 5.Zhang, Y. et al. High-temperature superconductivity with zero resistance and strange-metal behaviour in La3Ni2O7−δ. Nat. Phys.20, 1269 (2024). [Google Scholar]
  • 6.Yang, J. et al. Orbital-dependent electron correlation in double-layer nickelate La3Ni2O7. Nat. Commun.15, 4373 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Hou, J. et al. Emergence of high-temperature superconducting phase in pressurized La3Ni2O7 crystals. Chin. Phys. Lett.40, 117302 (2023). [Google Scholar]
  • 8.Liu, Y. B., Mei, J. W., Ye, F., Chen, W. Q. & Yang, F. s± wave pairing and the destructive role of apical-oxygen deficiencies in La3Ni2O7 under pressure. Phys. Rev. Lett.131, 236002 (2023). [DOI] [PubMed] [Google Scholar]
  • 9.Zhang, Y., Lin, L.-F., Moreo, A., Maier, T. A. & Dagotto, E. Structural phase transition, s±-wave pairing, and magnetic stripe order in bilayered superconductor La3Ni2O7 under pressure. Nat. Commun.15, 2470 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Lu, C., Pan, Z., Yang, F. & Wu, C. Interplay of two E g orbitals in superconducting La3Ni2O7 under pressure. Phys. Rev. B110, 094509 (2024). [Google Scholar]
  • 11.Kaneko, T., Sakakibara, H., Ochi, M. & Kuroki, K. Pair correlations in the two-orbital Hubbard ladder: Implications for superconductivity in the bilayer nickelate La3Ni2O7. Phys. Rev. B109, 045154 (2024). [Google Scholar]
  • 12.Chen, J., Yang, F. & Li, W. Orbital-selective superconductivity in the pressurized bilayer nickelate La3Ni2O7: An infinite projected entangled-pair state study. Phys. Rev. B110, L041111 (2024). [Google Scholar]
  • 13.Yang, Y.-f, Zhang, G.-M. & Zhang, F.-C. Interlayer valence bonds and two-component theory for high-Tc superconductivity of La3Ni2O7 under pressure. Phys. Rev. B108, L201108 (2023). [Google Scholar]
  • 14.Dong, Z. et al. Visualization of oxygen vacancies and self-doped ligand holes in La3Ni2O7−δ. Nature630, 847 (2024). [DOI] [PubMed] [Google Scholar]
  • 15.Tian, Y.-H., Chen, Y., Wang, J.-M., He, R.-Q. & Lu, Z.-Y. Correlation effects and concomitant two-orbital s±-wave superconductivity in La3Ni2O7 under high pressure. Phys. Rev. B109, 165154 (2024). [Google Scholar]
  • 16.Yang, Q.-G., Wang, D. & Wang, Q.-H. Possible s±-wave superconductivity in La3Ni2O7. Phys. Rev. B108, L140505 (2023). [Google Scholar]
  • 17.Qu, X.-Z. et al. Bilayer t-J-J model and magnetically mediated pairing in the pressurized nickelate La3Ni2O7. Phys. Rev. Lett.132, 036502 (2024). [DOI] [PubMed] [Google Scholar]
  • 18.Qin, Q. & Yang, Y.-f High-Tc superconductivity by mobilizing local spin singlets and possible route to higher Tc in pressurized La3Ni2O7. Phys. Rev. B108, L140504 (2023). [Google Scholar]
  • 19.Jiang, K., Wang, Z. & Zhang, F.-C. High-temperature superconductivity in La3Ni2O7. Chin. Phys. Lett.41, 017402 (2024). [Google Scholar]
  • 20.Lu, C., Pan, Z., Yang, F. & Wu, C. Interlayer-coupling-driven high-temperature superconductivity in La3Ni2O7 under pressure. Phys. Rev. Lett.132, 146002 (2024). [DOI] [PubMed] [Google Scholar]
  • 21.Fan, Z. et al. Superconductivity in nickelate and cuprate superconductors with strong bilayer coupling. Phys. Rev. B110, 024514 (2024). [Google Scholar]
  • 22.Lechermann, F., Gondolf, J., Bötzel, S. & Eremin, I. M. Electronic correlations and superconducting instability in La3Ni2O7 under high pressure. Phys. Rev. B108, L201121 (2023). [Google Scholar]
  • 23.Keimer, B., Kivelson, S. A., Norman, M. R., Uchida, S. & Zaanen, J. From quantum matter to high-temperature superconductivity in copper oxides. Nature518, 179–186 (2015). [DOI] [PubMed] [Google Scholar]
  • 24.Chen, X., Dai, P., Feng, D., Xiang, T. & Zhang, F.-C. Iron-based high transition temperature superconductors. Natl Sci. Rev.1, 371–395 (2014). [Google Scholar]
  • 25.Medarde, M. et al. RNiO3 perovskites (R= Pr, Nd): Nickel valence and the metal-insulator transition investigated by X-ray-absorption spectroscopy. Phys. Rev. B46, 14975 (1992). [DOI] [PubMed] [Google Scholar]
  • 26.Mizokawa, T. et al. Electronic structure of PrNiO3 studied by photoemission and X-ray-absorption spectroscopy: Band gap and orbital ordering. Phys. Rev. B52, 13865 (1995). [DOI] [PubMed] [Google Scholar]
  • 27.Sachdev, S. Colloquium: order and quantum phase transitions in the cuprate superconductors. Rev. Mod. Phys.75, 913–932 (2003). [Google Scholar]
  • 28.Zhang, Y., Lin, L.-F., Moreo, A. & Dagotto, E. Electronic structure, dimer physics, orbital-selective behavior, and magnetic tendencies in the bilayer nickelate superconductor La3Ni2O7 under pressure. Phys. Rev. B108, L180510 (2023). [Google Scholar]
  • 29.Liu, Z. et al. Evidence for charge and spin density waves in single crystals of La3Ni2O7 and La3Ni2O6. Sci. China Phys. Mech. Astron.66, 217411 (2022). [Google Scholar]
  • 30.Kakoi, M. et al. Multiband metallic ground state in multilayered nickelates La3Ni2O7 and La4Ni3O10 revealed by 139La-NMR at ambient pressure. J. Phys. Soc. Jpn93, 053702 (2024). [Google Scholar]
  • 31.Dan, Z. et al. Spin-density-wave transition in double-layer nickelate La3Ni2O7. arXiv preprint arXiv:2402.03952 (2024).
  • 32.Chen, K. et al. Evidence of spin density waves in La3Ni2O7−δ. Phys. Rev. Lett.132, 256503 (2024). [DOI] [PubMed] [Google Scholar]
  • 33.Khasanov, R. et al. Pressure-induced split of the density wave transitions in La3Ni2O7−δ. arXiv preprint arXiv:2402.10485 (2024).
  • 34.Chen, C. T. et al. Electronic states in La2−xSrxCuO4+δ probed by soft-X-ray absorption. Phys. Rev. Lett.66, 104–107 (1991). [DOI] [PubMed] [Google Scholar]
  • 35.Sawatzky, G. A. & Allen, J. W. Magnitude and origin of the band gap in NiO. Phys. Rev. Lett.53, 2339–2342 (1984). [Google Scholar]
  • 36.van der Laan, G., Zaanen, J., Sawatzky, G. A., Karnatak, R. & Esteva, J.-M. Comparison of x-ray absorption with x-ray photoemission of nickel dihalides and NiO. Phys. Rev. B33, 4253–4263 (1986). [DOI] [PubMed] [Google Scholar]
  • 37.Abbate, M. et al. Electronic structure and metal-insulator transition in LaNiO3−δ. Phys. Rev. B65, 155101 (2002). [Google Scholar]
  • 38.Horiba, K. et al. Electronic structure of LaNiO3−x: an in situ soft X-ray photoemission and absorption study. Phys. Rev. B76, 155104 (2007). [Google Scholar]
  • 39.Green, R. J., Haverkort, M. W. & Sawatzky, G. A. Bond disproportionation and dynamical charge fluctuations in the perovskite rare-earth nickelates. Phys. Rev. B94, 195127 (2016). [Google Scholar]
  • 40.Bisogni, V. et al. Ground-state oxygen holes and the metal–insulator transition in the negative charge-transfer rare-earth nickelates. Nat. Commun.7, 13017 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Lu, Y. et al. Site-selective probe of magnetic excitations in rare-earth nickelates using resonant inelastic X-ray scattering. Phys. Rev. X8, 031014 (2018). [Google Scholar]
  • 42.Hu, Z. et al. Hole Distribution between the Ni 3d and O 2p Orbitals in Nd2−xSrxNiO4−δ. Phys. Rev. B61, 3739 (2000). [Google Scholar]
  • 43.Chiuzbăian, S. G. et al. Localized electronic excitations in NiO studied with resonant inelastic X-ray scattering at the Ni M threshold: Evidence of spin flip. Phys. Rev. Lett.95, 197402 (2005). [DOI] [PubMed] [Google Scholar]
  • 44.Nag, A. et al. Many-body physics of single and double spin-flip excitations in NiO. Phys. Rev. Lett.124, 067202 (2020). [DOI] [PubMed] [Google Scholar]
  • 45.Chen, X., Jiang, P., Li, J., Zhong, Z. & Lu, Y. Critical charge and spin instabilities in superconducting La3Ni2O7. arXiv preprint arXiv:2307.07154 (2023).
  • 46.Tanaka, A. & Jo, T. Resonant 3d, 3p and 3s photoemission in transition metal oxides predicted at 2p threshold. J. Phys. Soc. Jpn63, 2788–2807 (1994). [Google Scholar]
  • 47.Hill, J. & McMorrow, D. Resonant exchange scattering: polarization dependence and correlation function. Acta Crystallogr. Sect. A Found. Crystallogr.52, 236–244 (1996). [Google Scholar]
  • 48.Fumagalli, R. et al. Polarization-resolved Cu L3-edge resonant inelastic x-ray scattering of orbital and spin excitations in NdBa2Cu3O7−δ. Phys. Rev. B99, 134517 (2019). [Google Scholar]
  • 49.Freeman, P. et al. Spin dynamics of half-doped La3/2Sr1/2NiO4. Phys. Rev. B71, 174412 (2005). [Google Scholar]
  • 50.Peng, Y. et al. Dispersion, damping, and intensity of spin excitations in the monolayer (Bi, Pb)2(Sr, La)2CuO6+δ cuprate superconductor family. Phys. Rev. B98, 144507 (2018). [Google Scholar]
  • 51.Lipscombe, O. J. et al. Spin waves in the (π,0) magnetically ordered iron chalcogenide Fe1.05Te. Phys. Rev. Lett.106, 057004 (2011). [DOI] [PubMed] [Google Scholar]
  • 52.Carlson, E. W., Yao, D. X. & Campbell, D. K. Spin waves in striped phases. Phys. Rev. B70, 064505 (2004). [Google Scholar]
  • 53.Scagnoli, V. et al. Role of magnetic and orbital ordering at the metal-insulator transition in NdNiO3. Phys. Rev. B73, 100409 (2006). [Google Scholar]
  • 54.Wilkins, S. et al. Comparison of stripe modulations in La1.875Ba0.125CuO4 and La1.48Nd0.4Sr0.12CuO4. Phys. Rev. B84, 195101 (2011). [Google Scholar]
  • 55.Hayden, S. M., Aeppli, G., Perring, T. G., Mook, H. A. & Doğan, F. High-frequency spin waves in YBa2Cu3O6.15. Phys. Rev. B54, R6905 (1996). [DOI] [PubMed] [Google Scholar]
  • 56.Reznik, D. et al. Direct observation of optical magnons in YBa2Cu3O6.2. Phys. Rev. B53, R14741 (1996). [DOI] [PubMed] [Google Scholar]
  • 57.Lin, J. et al. Strong superexchange in a d9−δ nickelate revealed by resonant inelastic X-ray scattering. Phys. Rev. Lett.126, 087001 (2021). [DOI] [PubMed] [Google Scholar]
  • 58.Liu, Z. et al. Electronic correlations and partial gap in the bilayer nickelate La3Ni2O7. Nat. Commun.15, 7570 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Kajimoto, R., Ishizaka, K., Yoshizawa, H. & Tokura, Y. Spontaneous rearrangement of the checkerboard charge order to stripe order in La1.5Sr0.5NiO4. Phys. Rev. B67, 014511 (2003). [Google Scholar]
  • 60.Sternlieb, B. et al. Charge and magnetic order in La0.5Sr1.5MnO4. Phys. Rev. Lett.76, 2169 (1996). [DOI] [PubMed] [Google Scholar]
  • 61.Zaliznyak, I., Hill, J., Tranquada, J., Erwin, R. & Moritomo, Y. Independent freezing of charge and spin dynamics in La1.5Sr0.5CoO4. Phys. Rev. Lett.85, 4353 (2000). [DOI] [PubMed] [Google Scholar]
  • 62.Wilkins, S., Hatton, P., Roper, M., Prabhakaran, D. & Boothroyd, A. Soft x-ray resonant magnetic diffraction. Phys. Rev. Lett.90, 187201 (2003). [DOI] [PubMed] [Google Scholar]
  • 63.Souliou, S. M. et al. Rapid suppression of the charge density wave in YBa2Cu3O6.6 under hydrostatic pressure. Phys. Rev. B97, 020503 (2018). [Google Scholar]
  • 64.Cyr-Choinière, O. et al. Sensitivity of Tc to pressure and magnetic field in the cuprate superconductor YBa2Cu3Oy: Evidence of charge-order suppression by pressure. Phys. Rev. B98, 064513 (2018). [Google Scholar]
  • 65.Zhou, K.-J. et al. I21: an advanced high-resolution resonant inelastic x-ray scattering beamline at diamond light source. J. Synchrotron Radiat.29, 563–580 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Cowan, R. D. The Theory of Atomic Structure and Spectra (Univ of California Press, 1981).
  • 67.Haverkort, M. W. Quanty for core level spectroscopy - excitons, resonances and band excitations in time and frequency domain. J. Phys. Conf. Ser.712, 012001 (2016). [Google Scholar]
  • 68.Kresse, G. & Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B54, 11169 (1996). [DOI] [PubMed] [Google Scholar]
  • 69.Kresse, G. & Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B59, 1758 (1999). [Google Scholar]
  • 70.Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett.77, 3865 (1996). [DOI] [PubMed] [Google Scholar]
  • 71.Wang, V., Xu, N., Liu, J.-C., Tang, G. & Geng, W.-T. Vaspkit: a user-friendly interface facilitating high-throughput computing and analysis using vasp code. Comput. Phys. Commun.267, 108033 (2021). [Google Scholar]
  • 72.Dudarev, S. L., Botton, G. A., Savrasov, S. Y., Humphreys, C. & Sutton, A. P. Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+ U study. Phys. Rev. B57, 1505 (1998). [Google Scholar]
  • 73.Haverkort, M. W. Theory of resonant inelastic X-ray scattering by collective magnetic excitations. Phys. Rev. Lett.105, 167404 (2010). [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Data Availability Statement

All data shown in the main text are available via Zenodo data repository (10.5281/zenodo.13955595).

All code used to perform the XAS and RIXS calculation is available from the corresponding authors upon reasonable request.


Articles from Nature Communications are provided here courtesy of Nature Publishing Group

RESOURCES