Introduction

The discovery of the high-Tc superconducting cuprates is one of the greatest surprises in quantum materials. Due to the complexity stemming from multiple intertwined orders, the pairing mechanism of cuprates remains an enigma1. Many theories have proposed that the unconventional superconductivity is likely associated with the strong correlation nature of the d-orbital electrons in transition-metal oxides (TMOs)2, which can be characterized by the Mott insulating parent state at the filling of one hole per unit cell. This picture has motivated the study of superconductivity in several TMOs and, more recently, magic-angle twisted bilayer graphene3. However, the anticipated analogy materials, such as iridates and ruthenates, do not show high Tc as cuprates. Theoretically, it remains a question whether the single-band Hubbard model, as a prototype of correlated electron systems, indeed gives long-range ordered superconductivity4,5.

The recent discovery of superconducting nickelates provides a new opportunity to shed light on the puzzle6,7,8,9,10,11. Due to nickelates’ structural and chemical similarities with cuprates, a comparative study of these two materials may help establish the pairing mechanism and guide us to design materials with higher Tc12,13,14,15,16,17,18,19,20,21,22,23. Experiments have shown infinite-layer nickelates exhibiting similar magnetic excitations with cuprates24. However, nickelates show a much lower Tc. A possible reason is that the nickelates RNiO2 (R = La, Pr, and Nd) are Mott-Hubbard insulators18,22,25,26, while cuprates are charge-transfer insulators with more covalent-bond nature near the parent compound. The charge-transfer property of the latter is characterized by a prominent pre-peak in the oxygen K-edge x-ray absorption spectrum (XAS), usually referred to as the Zhang-Rice peak27. Given that the charge-transfer gap was also found to anti-correlate (and thus are indirectly proportional) with Tc in various cuprate compounds28,29, it is conceivable that the oxygen-orbital components of dopants may play a crucial role in pairing. A critical step to assess this is to manipulate the charge-transfer extent of nickelates and demonstrate the possibility of enhancing superconductivity30,31,32.

Here, we report an enhancement of Tc by growing the Pr0.8Sr0.2NiO2 film on a compressive substrate (LaAlO3)0.3(Sr2AlTaO6)0.7 (LSAT). Combining first-principles and many-body simulations, we reveal that the in-plane compressive strain strengthens the p − d hybridization between oxygen and nickel orbitals. Thus, a considerable number of doped carriers distribute into oxygen orbitals, increasing the pre-peak in the oxygen K-edge XAS. This comparison reflects the dominant role of the charge-transfer nature in unconventional superconductivity, including but not restricted to nickelates and cuprates. This comparative study of the impact of microscopic ingredients on Tc further motivates a systematic and efficient approach to designing superconductivity in TMOs.

Results and discussion

Substrate-induced strain and enhanced T c

As illustrated in Fig. 1a, b, we prepared Pr0.8Sr0.2NiO2 films of equal thickness (8 nm) on two different substrates, SrTiO3 and LSAT, using pulsed laser deposition (PLD) followed by a CaH2-assisted chemical reduction method (see Supplementary Note 1 for details). Figure 1c, d shows the high-angle annular dark-field (HAADF) scanning transmission electron microscopy (STEM) images of the two superconducting Pr0.8Sr0.2NiO2 films on the two substrates, respectively. The contrast difference of diffraction spots reveals the uniform growth of the films with an ordered structure of a coherent infinite layer, displaying no evidence of minority phases. The well-defined infinite-layer phase of the Pr0.8Sr0.2NiO2 films is further disclosed in the annular bright-field (ABF) STEM images in Fig. 1e, f. The enlarged views of the areas inside the rectangles in Fig. 1e, f are shown in Fig.1g, h for better visibility of the lattice structures of the films and the substrates near the interface. The clear and sharp interfaces between the films and the substrates confirm the former epitaxially strained on the latter. Besides, the reciprocal space maps (RSM) around the (103) reflection demonstrate that the films are strained to the substrates (see Supplementary Note 2 for details). Based on both measurements, the in-plane lattice constants of the nickelate films match those of the substrates, i.e., a = 3.868 Å, yielding −1% compressive in-plane strain in the films grown on LSAT compared to SrTiO310.

Fig. 1: Cross-sectional scanning transmission electron microscopy (STEM) images of two types of superconducting Pr0.8Sr0.2NiO2 films.
figure 1

a, b The schematic illustrations of heterostructures of Pr0.8Sr0.2NiO2 films on two different substrates SrTiO3 and (LaAlO3)0.3(Sr2AlTaO6)0.7 (LSAT) respectively. c, d The high-angle annular dark field (HAADF) STEM images of films on SrTiO3 and LSAT substrates, respectively. e, f The annular bright field (ABF) STEM images of film on SrTiO3 and LSAT substrates, respectively. g, h The enlarged views of the areas inside the rectangles in (e, f), respectively.

With the nickelate films on distinct substrates, we first examine their transport properties using a Quantum Design Physical Property Measurement System (PPMS) with a standard four-probe configuration. Figure 2a shows the temperature-dependent resistivity of these two Pr0.8Sr0.2NiO2 films on SrTiO3 and LSAT substrates, respectively. Both films exhibit a sudden decrease in resistivity, characteristic of the superconducting transition (see Supplementary Note 3 for similar measurements on more samples). Here, we employ the onset temperature to quantify the superconductivity critical temperature Tc, resulting in \({T}_{c}^{(onset)}\) = 9 K and 15 K (highlighted with arrows) for samples grown on SrTiO3 and LSAT, respectively. To further confirm the superconducting behavior, we investigate the transport properties with a magnetic field (up to 9.0-tesla) applied perpendicular to the ab-plane of the film. As shown in Fig. 2c, d, the field dependence exhibits similar behaviors: In the normal state, the films display negligible variation, whereas, below the superconducting transition, the conductance is suppressed significantly with an increasing magnetic field.

Fig. 2: Transport properties of superconducting Pr0.8Sr0.2NiO2 films on two different substrates.
figure 2

a Temperature-dependent resistivity of the Pr0.8Sr0.2NiO2 on SrTiO3 and (LaAlO3)0.3(Sr2AlTaO6)0.7 (LSAT) substrates, respectively. b The phase diagram of Pr1−xSrxNiO2 films adapted from Ref. 10 The Tc varies from 7 K to 12 K with thickness for Pr0.8Sr0.2NiO2 films adapted from Ref. 9 (pink bar). The \({T}_{c}^{(onset)}\)≈ 9 K (red diamond) and \({T}_{c}^{(onset)}\)≈ 15 K (blue diamond) of our 8 nm-thick uncapped Pr0.8Sr0.2NiO2/SrTiO3 and Pr0.8Sr0.2NiO2/LSAT. c, d Magnetic-field response of superconducting Pr0.8Sr0.2NiO2 film on SrTiO3 and LSAT substrates, at a varying magnetic field perpendicular to the (a, b) plane. The onset of superconducting transition at 9 K for the film on SrTiO3 is smaller than that of the film at 15 K on LSAT substrate. Zero-resistance is obtained at 2.5 K and 2.0 K for the films on LSAT and SrTiO3, respectively.

We summarize the critical temperatures of various nickelate experiments by the gray area in Fig. 2b to compare with previous experiments. Considering the sensitivity to the thickness, we further mark the Tc varying from 7 K to 12 K for existing Pr0.8Sr0.2NiO2 studies9. The \({T}_{c}^{(onset)}\) ≈ 9 K (denoted by the red diamond) of our 8 nm-thick uncapped Pr0.8Sr0.2NiO2 film grown on SrTiO3 is lower than \({T}_{c}^{(onset)}\) ≈ 12 K from the previous studies on the film with equal thickness. Interestingly, when substrated by LSAT, the same uncapped 8 nm-thick Pr0.8Sr0.2NiO2 film exhibits a considerably higher \({T}_{c}^{(onset)}\) ≈ 15 K, beyond the record of any (capped and uncapped) STO-substrated samples for the same doping. The uncapped films on both substrates in our study show a more significant broadening of superconducting transition, which is probably due to the detailed topotactic reduction and common in various superconducting nickelate films11,33. Nevertheless, the zero resistance develops at 2.0 K and 2.5 K in zero-field for the films on SrTiO3 and LSAT, respectively. It is noticed that such an observation of enhanced Tc was also reported in Nd1-xSrxNiO2 films on LSAT compared to those on SrTiO3 for different doping levels34. While we cannot completely rule out some other extrinsic effects such as crystalline disorder that might partially affect Tc, it seems natural to attribute this \({T}_{c}^{(onset)}\) enhancement in the comparative study to the crystal structure of Pr0.8Sr0.2NiO2, which is compressively strained by the LSAT substrate. We were aware that an even higher \({T}_{c}^{(onset)}\) ≈ 17 K was reported in Pr0.82Sr0.18NiO2/STO35, which does not necessarily conflict with our attribution of the enhancement to strain since it corresponds to a different doping level that expects to result in a higher \({T}_{c}^{(onset)}\) than that of Pr0.80Sr0.20NiO2/STO10. Besides, the capping of STO layers may help improve the sample quality.

Pre-edge peak in x-ray absorption spectra

To unveil the connection between the strain and the enhancement of Tc in Pr0.8Sr0.2NiO2/LSAT, we further examine their electronic structure using XAS. The XAS at the Ni L3,2 and Pr M5,4 absorption edges have been collected for Pr0.8Sr0.2NiO2 films to characterize the ionic valence states (see Supplementary Note 4 for further details). We select the O K-edge to reveal the nature of dopant carriers since its characteristic peaks have been regarded as a fingerprint in cuprates27. Considering that the top layer of the films can be oxidized due to air exposure as well as that the substrates are insulators, we have employed the total fluorescence yield (TFY) mode to collect the XAS at the oxygen K absorption edge. Similar to other recent studies, we find that the K-edge XAS of Pr0.8Sr0.2NiO2/SrTiO3 does not exhibit any pre-edge peak22,36, suggesting that the parent compound of superconducting nickelates (Pr0.8Sr0.2NiO2/SrTiO3) falls into the Mott-Hubbard insulator regime. This property is in sharp contrast to cuprates, where an extra pre-edge mobile carrier peak (MCP) arises with hole doping and carries intensity comparable with the doping concentration. We illustrate the difference in XAS features via the partial density of states (PDOS) distribution sketched in Fig. 3c. As charge-transfer insulators, cuprates have a relatively strong on-site Coulomb interaction Ud (primarily for the Cu 3d electrons) compared to the charge-transfer energy h in the hole language. Therefore, a doped hole primarily occupies the oxygen 2p orbital, which hybridizes with the Cu 3d orbital as the Zhang-Rice singlet band. XAS selectively probes the unoccupied PDOS on oxygen, resulting in the coexistence of the pre-edge MCP and the upper Hubbard band (UHB) peaks. Nevertheless, the h is slightly larger than the Ud in nickelates, leading to tiny dopant holes on oxygen orbitals and, accordingly, a main absorption peak with a shoulder instead of a pre-edge peak. It is important to note that the distinction between charge-transfer and Mott-Hubbard insulators occurs suddenly at h ~ Ud only in the atomic limit25, while the presence of orbital hybridization leads to a crossover in the intensity of the pre-edge peak (see Supplementary Note 5 for detailed discussions).

Fig. 3: Total fluorescence-yield x-ray absorption spectra (XAS) of two superconducting nickelate films.
figure 3

a, b The XAS at the O K edge of Pr0.8Sr0.2NiO2 thin films with the linear polarization perpendicular to the films’ c axis. The inset in (b) shows the fitted spectral features near the pre-edge peak (the regime inside of the dashed square). The mobile carrier peak (MCP) is absent in the film on SrTiO3 substrate of (a). c Schematic density of states (DOS) of transition-metal oxides (TMOs) and their corresponding XAS features (i): a charge-transfer insulator, represented by cuprates, distributes doped holes primarily to oxygen orbitals, resulting in a prominent pre-edge peak in XAS; (ii): a Mott-Hubbard insulator, represented by nickelates on STO substrates, distributes doped holes primarily to transition-metal 3d orbitals, resulting in negligible pre-edge peaks; (iii): strained nickelates exhibit stronger hybridization between 3d and 2p orbitals, producing a pre-edge peak. UHB and LHB represent the upper and lower Hubbard bands, respectively.

Unlike the STO-substrated nickelates, the XAS of our Pr0.8Sr0.2NiO2/LSAT sample exhibits an evident pre-peak to the O K main edge at 530.2 eV, reminiscent of the pre-peaks in cuprates27. The energy position of 530.2 eV for the pre-peak is nearly equal to the observed hybridization peak in La4Ni3O8, which belongs to another family of nickelate materials (Rn+1NinO2n+2)37. Therefore, we can attribute this pre-edge peak in Pr0.8Sr0.2NiO2/LSAT to the MCP. Since the charge-transfer nature appears to be one of the most significant differences between cuprates and nickelates, the emergence of the pre-edge XAS peak in LSAT-substrated nickelates provides insight for its enhanced Tc, consistent with previous results28,29,38.

Numerical explanation of the electronic structure’s change

To reveal the impact of the strained lattice on nickelates’ electronic structure, we simulate the doped carriers and spectral evolution upon straining using density functional theory (DFT) and the extracted many-body model. We first calculate the electronic structure of the pristine and −1% strained PrNiO2 using Perdew-Burke-Ernzerhof (PBE) functional (see Methods). As shown in Fig. 4, the simulated PDOS reflects that the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular (LUMO) orbitals primarily consist of Ni 3dx2−y2 orbitals. As pointed out in previous studies22, the rare-earth band also resides at low energy, leading to a “self-doping” effect. To address the experimental features observed in the oxygen K-edge XAS, we focus on the oxygen components of the electronic wavefunction. By projecting the simulated Kohn-Sham orbitals to single-particle Wannier orbitals, we extracted two oxygen orbitals (2px and 2py) which hybridize with the active Ni 3dx2−y2 orbital [see the inset of Fig. 4a]. We employ these three orbitals to construct a tight-binding model, the first four terms in Eq. (1) in the Methods section, where the hopping parameters and site energies are extracted from the Kohn-Sham Wannier orbitals. For the pristine PrNiO2, we obtain tpd = 1.36 eV, tpp = 0.57 eV, and Ed - Ep = 3.90 eV, consistent with previous simulations22 for the −1% strained PrNiO2, tpd are increased into 1.41 eV, while the changes of other parameters are less than 2%.

Fig. 4: Theoretical analysis of strain effects on the electronic structures and x-ray absorption spectra (XAS) of Praseodymium Nickelates.
figure 4

a Band structures of unstrained PrNiO2 from density functional theory (DFT) and Wannier tight-binding model. Nickel’s 3dx2−y2 and oxygen’s px and py orbitals are shown at the top, where gray, purple, and green spheres represent Pr, Ni, and O, respectively. The site energies of Ed and Ep, as well as the hopping parameters tpd (hopping between Ni: 3dx2−y2 and O: px, py) and tpp (hopping between O: px and py) in the Wannier tight-binding model. These band parameters are employed in exact diagonalization simulations. b Exact diagonalization XAS with a charge-transfer energy h = 8 eV and a varying tpd from 1.0 eV to 1.5 eV. The dashed and solid lines represent the spectra for 0 and 12.5 % hole-doping, respectively. The pre-peak area (illustrated by the shade) increases with an increasing tpd. The pre-peak area (c) and the average hole number per px,y orbital (d) as a function of h and tpd are shown by the intensity plots. Both increase with an increasing tpd and a decreasing h. The blue arrows indicate the change of tpd from 1.36 to 1.41 eV when 1 % compressive epitaxial strain is applied.

We then involve the strong Coulomb interactions, doping, and x-ray core-hole interactions on top of the DFT-extracted model to simulate the XAS spectra and analyze correlation effects. Based on the estimations from previous studies, we set the Hubbard-type Coulomb interactions on the Ni 3d and O 2p orbitals as Ud = 6 eV and Up = 2 eV, respectively. We also set the core-hole attractive interaction Uc = 3 eV to match the experimental separation between the pre-edge and main absorption peaks. The XAS spectral intensity is simulated using Eq. (2) in the Methods section. As shown in Fig. 4b, we manipulate the strength of tpd and fix other parameters, mimicking the primary impact on the electronic structure induced by the strain effect. Considering possible ambiguities of band parameters extracted from minimal Wannier-folded orbitals, we examine the spectral properties for a wide range of tpd values, from 1.0 to 1.5 eV. The simulated XAS results suggest that increased tpd, induced by the compressive strain effect, enhances the pre-edge peak, consistent with our experimental observations in Fig. 3a, b. As illustrated in Fig. 3c, although the strain does not switch the relative size of Ud to h in the atomic representation, the increased tpd mixes oxygen and nickel PDOS at the top of the lower Hubbard band (LHB). Therefore, a larger portion of the doped holes resides in oxygen, resulting in the rise of the pre-edge XAS peak.

To further demonstrate the aforementioned mechanism, we calculate the average hole concentrations ‹np› on the oxygen 2px/y orbitals. By comparing the undoped and doped XAS results, the pre-edge peak reflects most dopant carrier concentrations in oxygen. Therefore, we evaluate the pre-peak intensity and ‹np› for a wide range of charge-transfer energy h and tpd of the three-band Hubbard model. As shown in Fig. 4c, d, the parameter dependencies of these two quantities are consistent in the entire htpd parameter space. We also highlight the DFT-extracted model parameters for the pristine and strained nickelates to guide the eye. In this context, the pre-edge peak intensity (Fig. 4c) rises by 2.8%, and the oxygen-hole concentration (Fig. 4d) increases by 3.1%. This increase is smooth in the DFT-extracted model, in contrast to the sharp transition in the atomic model (see Supplementary Note 5), but the correlation between the spectral and carrier properties is the same. Due to the simplicity of the downfolded three-band, our simulation has neglected contributions from other oxygen orbitals in nickelates and substrates, underestimating the main-peak intensity. Despite the mismatch of this main peak, the rise of the pre-edge peak reflects that the nickelates in LSAT substrate distribute more dopant carriers into oxygen orbitals.

The fact that Pr0.8Sr0.2NiO2/LSAT displays a much higher Tc compared with Pr0.8Sr0.2NiO2/STO suggests the strain enabled via substrate design is an efficient route to control, and especially enhance, the superconductivity in nickelates. More importantly, the comparison between strained and unstrained nickelates, as well as nickelates and cuprates, provides significant insights into the long-standing puzzle – the pairing mechanism of unconventional superconductivity in correlated materials. Our experimental evidence for the concurrent rise of Tc and the emergence of the pre-edge peak in oxygen K-edge XAS reflects the intrinsic connection between the TM-oxygen hybridization in the pairing mechanism. Narrowing the dominant factor into the hybridization, one can further interpret its microscopic origin. A possible explanation is the enhanced spin fluctuations due to these Ni-O hybridizations, similar to the recently observed strained-enhanced magnons in monolayer cuprates39. According to our first-principles extracted three-band models, we simulate the dynamical spin structure factors of the pristine and strained PrNiO2. The top of magnon excitations rises from 178 meV into 192 meV with the −1% compressive strain (see Supplementary Note 6 for further details). Large-momentum spin fluctuations have been popular candidates for the d-wave pairing glue of unconventional superconductivity. Moreover, such a 7.8% increase of spin excitations out of the 1% strain indicates the outsized nonlinear impact of the lattice structure. Thus, an alternative but non-exclusive explanation for the 60% enhancements lies in the possible contribution of electron-phonon to unconventional superconductivity40,41,42,43,44,45. As a key difference from nickelates and other TMOs26, the unique charge-transfer insulating parent compounds of cuprates cause doped holes primarily to reside on oxygen, whose vibrations constitute most phonon modes. The present experimental and simulation results suggest that the strain-enhanced hybridization between O and Ni orbitals also leads to the similar charge transfer of doped carriers, which may further help pairing.

It is still challenging to examine if the magnetic or phonon-assisted mechanism, or their combined effect, may account for the more than 60% enhancement of Tc in Pr0.8Sr0.2NiO2/LSAT. The quantitative assessment of their contributions may identify the general pairing mechanism in correlated materials and motivates future experiments. For example, resonant inelastic x-ray scattering has characterized both the magnon and phonon dispersions in STO-substrated Nd1–xSrxNiO224. Comparative RIXS studies of nickelates on these two substrates may provide more insight into individual contributions from these excitations. Thus, our results, together with the further investigations of pairing mechanisms, should stimulate systematic engineering and design of high-Tc superconductivity.

Methods

Sample preparation and experimental characterization

The perovskite precursor Pr0.8Sr0.2NiO3 films were prepared by using pulsed laser deposition (PLD). The corresponding infinite-layer phase was acquired by the soft-chemistry reduction method. The superconducting transition temperature was confirmed by transport measurements using a Quantum Design Physical Property Measurement System (PPMS) with a standard four-probe configuration. Samples for the cross-sectional scanning transmission electron microscopy (STEM) were prepared by focused ion beam (FIB, Helios 600i) techniques. The high-angle annular dark field (HAADF) and annular bright field (ABF) STEM images were acquired on the ARM-200F (JEOL, Japan) operated at 200 kV with a CEOS Cs corrector (CEOS GmbH, Germany). The x-ray absorption spectroscopy (XAS) measurements were performed at beamline 29-ID IEX at the Advanced Photon Source, Argonne National Laboratory. The beamline uses an electromagnetic undulator providing a soft x-ray from 250 eV to 3000 eV. The spectra were normalized by the incident x-ray intensity (I0) using a drain current from a gold mesh upstream of the sample. All spectra were collected at 30 K.

Density functional theory calculations

The electronic band structure of PrNiO2 is calculated using the Quantum Espresso package46 with Perdew-Burke-Ernzerhof47 exchange-correlation functional, where the pseudopotential is based on the projected augmented wave (PAW) method48. We adopt the Monkhorst-Pack sampling49 with a Γ-centered k-mesh of 13 × 13 × 15 and a plane-wave cutoff energy of 40 rydberg is used to expand the wave function. The convergence criteria of structure relaxation and self-consistent field are set to 10−4 rydberg/bohr and 10−7 rydberg, respectively. For the −1% strained structure, we then change the lengths of the a- and b-axis directly while the c-axis remains unchanged.

To extract the site energies and hopping parameters, we employ the Wannier90 package50 to fit our DFT results and construct the tight-binding models, where a total number of 16 Wannier functions: five Pr’s d orbitals, five Ni’s d orbitals, and six O’s p orbitals, are considered. In our calculations, the disentanglement procedure is employed with an energy window from −10 to +0.8 eV.

Exact diagonalization simulation for XAS spectra

We employ the three-band Hubbard model extracted from the DFT Wannier orbitals as a minimal description of the charge-transfer physics in PSNO. The Hamiltonian reads as:

$$Hv = - \mathop{\sum}\limits_{\langle i,j \rangle \alpha \sigma}{t}_{pd}^{ij}({d}_{i \sigma }^{+}{p}_{j \alpha \sigma } + h.c.)\\ - \mathop{\sum} \limits_{\langle j,j^{\prime} \rangle \sigma }{t}_{pp}^{jj^{\prime}}({p}_{jx\sigma }^{+}{p}_{j^{\prime}y\sigma }+h.c.)+\mathop{\sum} \limits_{i\sigma }{E}_{d}{n}_{i\sigma }^{d}+\mathop{\sum} \limits_{j\alpha \sigma }{E}_{p}{n}_{j\alpha \sigma }^{p}\\ + \mathop{\sum} \limits_{i}{U}_{d}{n}_{i\uparrow }^{d}{n}_{i\downarrow }^{d}+\mathop{\sum} \limits_{j\alpha }{U}_{p}{n}_{j\alpha \uparrow }^{p}{n}_{j\alpha \downarrow }^{p},$$
(1)

Here, \({d}_{i\sigma }^{+}({p}_{j\alpha \sigma }^{+})\) creates an electron with spin σ at a Ni site i (O site j), and \({n}_{i\sigma }^{d}({n}_{j\sigma }^{p})\) is the Ni (O) electron number. The flavor α = {x, y} labels the 2px,y orbitals. The tpd and tpp determine the nearest-neighbor hopping amplitudes, and Ed (Ep) is the electronic site energy. These single-particle parameters are extracted from the Wannier orbitals of the DFT simulation. The last two terms are the on-site Hubbard interactions, whose strengths are set as ad hoc parameters Ud = 6 eV, Up = 2 eV, comparable with previous studies22,26. The charge-transfer energy in the hole representation is defined as h = Ed − Ep + Ud − Up. At the pristine PSNO, the DFT simulation gives h = 8 eV, |tpd | = 1.36 eV, |tpp | = 0.57 eV, while in the −1% strained PSNO, the extracted parameters of |tpd| and |tpp| are = 1.41 eV and 0.58 eV. Beyond the single-particle Kohn-Sham band parameters, we perform many-body simulations for the XAS spectrum of the three-orbital Hubbard model on an 8-site cluster for two different doping levels: 0% (half-filling) and 12.5% (underdoped).

The many-body ground state \(|G\rangle\) of the three-band valence-electron Hamiltonian Eq. (1) is solved by the exact diagonalization (ED). With this ground state, the zero-temperature XAS spectrum is calculated using

$${I}_{XAS}(\omega ,T)=-\frac{1}{\pi }{{\mbox{Im}}}\langle G|{D}^{+}\frac{1}{\omega +{E}_{G}-H+i\delta }D|G\rangle ,$$
(2)

where EG is the ground-state energy and δ is the broadening corresponding to the inverse lifetime of the intermediate state. For the convenience of extracting the integrated peak intensity, we employed a small broadening δ = 0.1 eV in the simulation, where realistic values should be much larger. Different from the three-band valence-electron Hamiltonian in Eq. (1), the intermediate-state Hamiltonian H in Eq. (2) describes an x-ray excited state and is defined in an extended Hilbert space consisting of both the valence electrons and a single core-level band resonant to the x-ray edge. Thus, the dipole operators D define the electron transition between the oxygen 1 s (core-level) and the 2p orbitals (valence level) selected by the x-ray K-edge edge

$$D=\mathop{\sum} \limits_{\alpha \sigma }{p}_{j\alpha \sigma }^{+}{c}_{j\sigma },$$
(3)

We ignore the matrix elements since we do not compare absolute intensities among different edges. In addition to the valence-band tight-binding model Hv, the intermediate state in Eq. (2) contains an extra core hole and its attractive interaction

$${H}_{c}=\mathop{\sum} \limits_{i\sigma }{E}_{edge}(1-{n}_{i\sigma }^{c})-{U}_{c}\mathop{\sum} \limits_{i\sigma \sigma ^{\prime}}{n}_{i\sigma }^{p}(1-{n}_{i\sigma ^{\prime}}^{c}),$$
(4)

where Eedge is the edge energy Eedge = 530 eV and Uc = 3 eV the core-hole potential.