Spectral evidence for NiPS3 as a Mott-Hubbard insulator

Yifeng Cao Department of Physics, Boston University, Boston, Massachusetts, 02215, USA. Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, California, 94720, USA.    Nicholas Russo Department of Physics, Boston University, Boston, Massachusetts, 02215, USA. Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, California, 94720, USA.    Qishuo Tan Department of Chemistry, Boston University, Boston, Massachusetts, 02215, USA.    Xi Ling Department of Chemistry, Boston University, Boston, Massachusetts, 02215, USA.    Jinghua Guo Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, California, 94720, USA.    Yi-de Chuang Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, California, 94720, USA.    Kevin E. Smith ksmith@bu.edu Department of Physics, Boston University, Boston, Massachusetts, 02215, USA.
Abstract

The layered van der Waals trichalcogenide NiPS3 has attracted widespread attention due to its unique optical, magnetic, and electronic properties. The complexity of NiPS3 itself, however, has also led to ongoing debates regarding its characteristics such as the existence of self-doped ligand holes. In this study, X-ray absorption spectroscopy and resonant inelastic X-ray scattering have been applied to investigate the electronic structure of NiPS3. With the aid of theoretical calculations using the charge-transfer multiplet model, we provide experimental evidence for NiPS3 being a Mott-Hubbard insulator rather than a charge-transfer insulator. Moreover, we explain why some previous XAS studies have concluded that NiPS3 is a charge-transfer insulator by comparing surface and bulk sensitive spectra.

I Introduction

Van der Waals (vdW) materials, including grapheneNovoselov et al. (2012), transition metal dichalcogenidesNicolosi et al. (2013); Xu et al. (2015), and boron nitrideWatanabe et al. (2004) are a promiment class of two-dimensional (2D) crystals. The reduced dimensionality of these systems often leads to exceptional physical properties, resulting from enhanced quantum effects and increased correlations due to the reduction of available phase space and diminished screeningAjayan et al. (2016). In this study, we focus on NiPS3, a type of 2D vdW metal phosphorus trichalcogenides (MPX3). NiPS3 is monoclinic, with space group C2/m𝐶2𝑚C2/mitalic_C 2 / italic_m; it can be visualized as layered NiS2 crystals with P-P pairs substituting for one third of the Ni sitesWang et al. (2018); Klingen et al. (1973) as shown in FIG. 1. NiPS3 exhibits intriguing properties in electronicsKim et al. (2018); Lane and Zhu (2020); Tan et al. (2022),spintronicsWang et al. (2021); Basnet et al. (2021); Scheie et al. (2023), magnetismWildes et al. (2015); Kim et al. (2019a); He et al. (2024), and opticsLiu et al. (2019); Wang et al. (2022). NiPS3 stands out among the 2D crystals due to its ultrasharp photoluminescence (PL) peak at 1.47 eV, induced by a spin-orbit-entangled exciton stateKang et al. (2020); Wang et al. (2021); Hwangbo et al. (2021). Understanding the properties of the exciton state requires an investigation of both the occupied and unoccupied states of NiPS3; although the exciton of NiPS3 has been extensively studiedBelvin et al. (2021); Ho et al. (2021); Klaproth et al. (2023); Kim et al. (2023), many fundamental electronic properties of NiPS3 have still remained controversial.

In X-ray absorption spectroscopy (XAS), a core electron is excited to an empty state, thereby probing the unoccupied states of the system, with information about the occupied states being inferred indirectlyFuggle et al. (1992). Meanwhile, in resonant inelastic X-ray scattering (RIXS), photons are scattered inelastically off matter, and we measure the excitations corresponding to the changes that occur during the scatteringAment et al. (2011). Both XASKim et al. (2018, 2019b); Kang et al. (2020); Yan et al. (2021) and RIXSKang et al. (2020); DiScala et al. (2023); He et al. (2024) have been applied to NiPS3, yet they have yielded differing results. The major controversy is whether NiPS3 is a charge transfer insulator or Mott-Hubbard insulator. A charge-transfer insulator is defined by a charge-transfer energy ΔΔ\Deltaroman_Δ (the energy cost of transferring an electron from the p𝑝pitalic_p level of a ligand to the metal ion) that is smaller than the Coulomb repulsion energy between two d𝑑ditalic_d electrons Uddsubscript𝑈𝑑𝑑U_{dd}italic_U start_POSTSUBSCRIPT italic_d italic_d end_POSTSUBSCRIPT (the energy cost of a d𝑑ditalic_d electron hopping to another already occupied d𝑑ditalic_d site); by contrast, when ΔΔ\Deltaroman_Δ is greater than Uddsubscript𝑈𝑑𝑑U_{dd}italic_U start_POSTSUBSCRIPT italic_d italic_d end_POSTSUBSCRIPT, the standard Hubbard model will be applied, resulting in a Mott-Hubbard insulatorKhomskii (2014). The simplified energy levels for these two types of insulatorsSahiner et al. (1996) are shown in FIG. 2: although both have a dominant ground state of d8superscript𝑑8d^{8}italic_d start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT, the Mott-Hubbard insulator’s lowest excited state is d9superscript𝑑9d^{9}italic_d start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT, while the charge-transfer insulator’s lowest excited state is d9L¯superscript𝑑9¯𝐿d^{9}\underline{L}italic_d start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT under¯ start_ARG italic_L end_ARG, where L¯¯𝐿\underline{L}under¯ start_ARG italic_L end_ARG represents a ligand hole. Kim et al. first applied XAS to the Ni L-edge of NiPS3 and compared with calculated spectra corresponding to different values of ΔΔ\Deltaroman_ΔKim et al. (2018). They found that NiPS3 is a self-doped charge transfer insulator with a negative value of ΔΔ\Deltaroman_Δ. With a negative ΔΔ\Deltaroman_Δ, the charge transfer from the ligand will occur spontaneously, resulting in a dominant ground state d9L¯superscript𝑑9¯𝐿d^{9}\underline{L}italic_d start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT under¯ start_ARG italic_L end_ARG. This claim, however, was directly challenged by a subsequent calculation which showed that NiPS3 would be a metal rather than an insulator with d9L¯superscript𝑑9¯𝐿d^{9}\underline{L}italic_d start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT under¯ start_ARG italic_L end_ARG ground stateKim et al. (2019b). They confirmed the presence of the paramagnetic Mott phase and the dominant ground state d8superscript𝑑8d^{8}italic_d start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT for NiPS3. Following XAS studiesKang et al. (2020); Yan et al. (2021), although challenging NiPS3 as having negative ΔΔ\Deltaroman_Δ with self-doped behavior, they still concluded that NiPS3 has a small but positive ΔΔ\Deltaroman_Δ and belongs to the category of charge-transfer insulator. The exciton of NiPS3 probed by RIXS is also a subject of debate. Initially, the exciton has been described as a Zhang-Rice singletKang et al. (2020). However, this does not explain the extremely narrow PL peak of the excitonHe et al. (2024) and its mismatch in response to an applied magnetic fieldJana et al. (2023). Instead of the Zhang-Rice mode, a most recent RIXS studies have revealed the dominant Hund’s character in NiPS3He et al. (2024).

In this study, we compared XAS measurements of an as-grown single crystal of NiPS3 to an exfoliated sample of NiPS3 in order to reconcile the controversy regarding the value of ΔΔ\Deltaroman_Δ for NiPS3. XAS was applied using both total electron yield (TEY) mode and total fluourescence yield (TFY) mode for surface sensitive (<10absent10<10< 10 nm) and bulk sensitive (>100absent100>100> 100 nm) measurements, respectivelyYang et al. (2021). Therefore, the surface and bulk features of NiPS3 were simultaneously collected, enabling us to capture the most subtle details of the electronic structure. By comparing the XAS results in TEY mode and TFY mode, we obtained different spectra from previous XAS studiesKim et al. (2018); Kang et al. (2020); Yan et al. (2021) on NiPS3. Furthermore, by comparison with calculations using the charge-transfer multiplet model we are led to conclude that, contrary to the previously held belief that NiPS3 has a very small value of ΔΔ\Deltaroman_Δ, it actually has a relatively large value of ΔΔ\Deltaroman_Δ. Further analysis was conducted to explain why previous XAS results concluded that NiPS3 belongs to the charge-transfer insulator. Moreover, we also applied RIXS to NiPS3 and found that dd𝑑𝑑dditalic_d italic_d excitations are dominant, while charge-transfer excitations were too weak to be observed. Combined with the conclusion from XAS that NiPS3 has a relatively large ΔΔ\Deltaroman_Δ, we provide strong evidence that NiPS3 is a Mott-Hubbard insulator.

II Methodology

NiPS3 single crystals were synthesized by the chemical vapor transport method. Pure elements, in a stoichiometric ratio of Ni:P:S === 1:1:3 (2 g in total), along with 40 mg iodine as transport agent, were enclosed in quartz ampules under vacuum of 1×\times×10--4 Torr. Subsequently, the ampules underwent heating in a two-zone furnace with the temperature range of 650 -- 600 C for 1 week, followed by cooling to room temperature. Bulk crystals were harvested from the lower temperature zone of the ampules. The as-grown NiPS3 single crystal was directly mounted on the sample holder with conductive copper tape. The exfoliated sample was prepared by removing the top layers using Scotch tape under an argon atmosphere in a glovebox, and was transferred to the analysis chamber using a vacuum suitcase to prevent oxidation.

The nickel L3,2-edge XAS spectra were measured at Advanced Light Source (ALS) beamline 7.3.1 at Lawrence Berkeley National Laboratory (LBNL). Both TEY and TFY modes were collected simultaneously. The XAS photon energy was calibrated using a Li(Ni1/3Mn1/3Co1/3)O2 reference sample. O K-edge XAS spectra with both TEY and TFY modes were also collected to determine the extent of O contamination in the sample.

The theoretically calculated XAS spectra were obtained using the Crispy software for the Quanty libraryRetegan (2019). A charge-transfer multiplet model is applied to the system, described by a semi-empirical Hamiltonian that includes atomic and crystal-field interactionsde Groot et al. (2021). In this study, nickel L3,2-edge XAS spectra were calculated by Crispy for various values of ΔΔ\Deltaroman_Δ. For simplicity, the symmetry is set to Ohsubscript𝑂O_{h}italic_O start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT. The applied parameters are: Fksubscript𝐹𝑘F_{k}italic_F start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT=0.8, Gksubscript𝐺𝑘G_{k}italic_G start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT=0.8, ζ𝜁\zetaitalic_ζ=1.0, U(3d,3d)𝑈3𝑑3𝑑U(3d,3d)italic_U ( 3 italic_d , 3 italic_d )=4.0 eV, U(2p,3d)𝑈2𝑝3𝑑U(2p,3d)italic_U ( 2 italic_p , 3 italic_d )=6.0 eV, 10Dq(3d)10𝐷𝑞3𝑑10Dq(3d)10 italic_D italic_q ( 3 italic_d )=1.0 eV, Veg(3d,L1)𝑉subscript𝑒𝑔3𝑑𝐿1Ve_{g}(3d,L1)italic_V italic_e start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( 3 italic_d , italic_L 1 )=1.4 eV, Vt2g(3d,L1)𝑉subscript𝑡2𝑔3𝑑𝐿1Vt_{2g}(3d,L1)italic_V italic_t start_POSTSUBSCRIPT 2 italic_g end_POSTSUBSCRIPT ( 3 italic_d , italic_L 1 )=1.0 eV. The broadening full width at half maximum parameters (eV) are set to be 0.48, 0.52 for Lorentzian and 0.3 for Gaussian. The values of ΔΔ\Deltaroman_Δ range from -1 eV to 5 eV.

The RIXS spectra of nickel L3-edge were obtained at beamline 8.0.1 at the ALS. The RIXS photon energy was calibrated by an undulator.

III Results and discussion

XAS was first applied to nickel L3,2-edge of as-grown NiPS3 single crystal as shown in FIG. 3(a). Surprisingly, the spectrum from TEY mode (blue solid line) and TFY mode (red solid line) exhibited different features. In TEY mode, both the L3-edge and L2-edge show a single peak feature with a satellite peak indicated by the blue arrow around 856.6 eV. This is precisely the feature corresponding to the small ΔΔ\Deltaroman_Δ observed in previous literature through experiments and calculations, indicating a charge-transfer insulator. However, as mentioned earlier, the TEY signal primarily reflects the surface electronic structure of a sample. In contrast, the TFY spectra reveal that the satellite splits into two peaks at 855.8 eV and 857.2 eV. Furthermore, distinguishable double peak features appear at both L3-edge (854.4 eV) and L2-edge (871.0 eV), as indicated by the red arrows.

Although many studies indicate that NiPS3 is stable in airLu et al. (2020); Fang et al. (2021), exposure to air may still cause surface oxidation and affect its TEY signalFrati et al. (2020). This can be inferred from the different features exhibited by the TEY and TFY signals in FIG. 3(a). To chracterize the surface oxidation, we also applied XAS in the oxygen K-edge region for the same sample used in FIG. 3(a) as presented in FIG. 4. As we previously inferred, the surface-sensitive TEY signal in the oxygen range exhibited very distinct signals, with characteristic peaks appearing in both the X-Ray absorption near edge structure (XANES) region (532 eV and 537 eV) and extended X-ray absorption fine structure (EXAFS) region (558 eV). This indicates that the oxygen on the surface of NiPS3 is not merely adsorbed free oxygen from the air, but has chemically bonded with the Ni cations. The spectrum in TFY mode, on the other hand, shows a very weak signal-to-noise ratio (SNR) in oxygen range compared to the spectrum in TEY mode. The appearance of an oxygen peak in the bulk-sensitive TFY mode does not mean that the entire NiPS3 sample is oxidized. Instead, it is because the TFY signal detection range covers the entire region to a depth of about 100 nanometers, including the few nanometers of the oxidized surface. It is precisely because the SNR of the TFY signal is significantly lower than that of the TEY signal that it proves the oxidation is only present on the surface of the NiPS3 sample (occupying only a small portion of the TFY detection depth). This is reasonable, as the bulk NiPS3 should not retain oxygen and should be air-stable.

Further investigation into the structure of the O K edge XAS has found that it is likely that nanoscaled NiO appear on the surface of NiPS3, evidenced by the similarity to the O K edge XAS from NiO nanowires as shown by Wu et al.Wu et al. (2005); these nanoscaled NiO exhibit very different features in the XAS O K-edge region compared to bulk NiO. By comparing with their XAS results, we find that the characteristic peaks in the TEY mode in FIG. 4 are perfectly consistent with those of NiO nanowires in both the XANES and EXAFS regions. The TEY probe depth shows that the nanoscaled NiO once again demonstrates that oxidation occurs only on the surface of the NiPS3 sample.

To eliminate the impact of surface oxidation on the experimental results, the most straightforward method is to directly remove the outermost few layers. As a vdW material, NiPS3 can be easily exfoliated using Scotch tape. The XAS spectrum was remeasured for the Ni L3,2-edge of exfoliated NiPS3 sample [FIG. 3(b)]. After exfoliation, we observed the following: (1) The TEY and TFY spectra exhibited excellent consistency, as indicated by the blue and red arrows, respectively; compared to the untreated sample, (2) the TEY signal showed significant changes, with the most notable being that both the Ni L2 peak and L3 peak split into a double peak structure, with more distinct features appearing in the satellite peak next to the L3 peak; (3) the TFY signal did not show obvious changes. These observations indicate that: (1) Exfoliation successfully removed the surface oxidation layers, making the TEY signal after exfoliation more likely to reflect the true sample signal; (2) The presence of the surface oxidation layers have a significant impact on the TEY signal; (3) The oxidation layers have a minimal impact on the TFY signal, allowing the TFY signal to remain reliable throughout the entire study.

To determine the ΔΔ\Deltaroman_Δ of NiPS3, we performed theoretical calculations using the charge-transfer multiplet model. The comparison between the experimental data for the exfoliated sample and the theoretical calculation is presented in FIG. 5. From the calculated results, it can be observed that with the increase in ΔΔ\Deltaroman_Δ, the most noticeable change is that the Ni L3 and L2 absorption edges transition from a single peak structure to a double peak structure, accompanied by a shift in the peak positions and changes in the satellite peaks. This trend is consistent with previous calculationsKim et al. (2018); Kang et al. (2020); Yan et al. (2021). When determining the ΔΔ\Deltaroman_Δ of NiPS3 by comparing the features with calculations, however, the main focus and point of contention has been around the L3 edge. By contrast, the most noticeable feature here, which differs from previous studies, is that the L2 peak also splits into a double peak structure. The best match to the experimental results is the calculated spectra with ΔΔ\Deltaroman_Δ=3 eV, yet the spectra with ΔΔ\Deltaroman_Δ=2.5 eV and ΔΔ\Deltaroman_Δ=3.5 eV also match some features of the experimental results very well. Therefore, we have obtained a relatively large ΔΔ\Deltaroman_Δ for NiPS3, suggesting it is highly likely that NiPS3 falls within the regime of a Mott-Hubbard insulator.

Thus far, we have determined the ΔΔ\Deltaroman_Δ of NiPS3 with XAS and theoretical calculation. To further confirm that such a ΔΔ\Deltaroman_Δ categorizes NiPS3 as a Mott-Hubbard insulator, we employed mapping of RIXS as shown in FIG. 6. It can be observed that parallel to the elastic peak (yellow solid line), two resonant X-ray Raman features (red dashed line) were found with energy-loss of 1.2 eV and 1.8 eV. Based on the scale of energy-lossAment et al. (2011) and previous optical spectra resultsKim et al. (2018), both resonant X-ray Raman features were assigned to dd𝑑𝑑dditalic_d italic_d excitations. Therefore, we found that the dd𝑑𝑑dditalic_d italic_d excitations are dominant, with the charge-transfer features being too weak to be observed. This indicates that standard Hubbard model applies to NiPS3, with transitions occurring only among d𝑑ditalic_d-electrons themselves, rather than the pd𝑝𝑑p-ditalic_p - italic_d model as in the case of charge-transfer insulator. Moreover, these energy-loss features match the previous RIXS resultHe et al. (2024), except that we did not find the ultrasharp PL peak due to the limitation of resolving power. With the aid of calculations, they determined the expectation values describing the NiPS3 wavefunction and revealed that NiPS3 has a dominant Hund’s character with a ground state d8superscript𝑑8d^{8}italic_d start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT, instead of d9L¯superscript𝑑9¯𝐿d^{9}\underline{L}italic_d start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT under¯ start_ARG italic_L end_ARG as in the Zhang-Rice scenario. This further proves that NiPS3 belongs to a Mott-Hubbard insulator rather than a charge-transfer insulator.

IV Conclusion

In this study, we used XAS and RIXS to investigate the electronic structure of NiPS3. By comparing experimental data from as-grown and exfoliated samples with theoretical calculations, we found that NiPS3 has a relatively large ΔΔ\Deltaroman_Δ (about 3 eV), indicating it is a Mott-Hubbard insulator. Additionally, we observed surface oxidation on as-grown samples, which may have contributed to the previous debates over the XAS results and was mitigated by exfoliation. RIXS data showed dominant dd𝑑𝑑dditalic_d italic_d excitations with weak charge-transfer features, further supporting this classification. These findings resolve previous controversies, contribute to the broader knowledge of NiPS3, and emphasize the importance of controlling surface adsorption in studies of the electronic structure of vdW materials.

V Acknowledgement

This research used resources of the Advanced Light Source, which is a Department of Energy (DOE) Office of Science User Facility under contract no. DE-AC02-05CH11231. Work done by Q.T. and X.L. was supported by National Science Foundation (NSF) under Grant No. 2216008 and No. 1945364, and the U.S. DOE, Office of Science, Basic Energy Science (BES) under Award Number DE-SC0021064. Q.T. also acknowledges the support of the Laursen Fellowship at Boston University.

References

  • Novoselov et al. (2012) K. S. Novoselov, L. Colombo, P. Gellert, M. Schwab, K. Kim, et al., nature 490, 192 (2012).
  • Nicolosi et al. (2013) V. Nicolosi, M. Chhowalla, M. G. Kanatzidis, M. S. Strano, and J. N. Coleman, Science 340, 1226419 (2013).
  • Xu et al. (2015) S.-Y. Xu, I. Belopolski, N. Alidoust, M. Neupane, G. Bian, C. Zhang, R. Sankar, G. Chang, Z. Yuan, C.-C. Lee, et al., Science 349, 613 (2015).
  • Watanabe et al. (2004) K. Watanabe, T. Taniguchi, and H. Kanda, Nature materials 3, 404 (2004).
  • Ajayan et al. (2016) P. Ajayan, P. Kim, and K. Banerjee, Physics Today 69, 38 (2016).
  • Wang et al. (2018) F. Wang, T. A. Shifa, P. Yu, P. He, Y. Liu, F. Wang, Z. Wang, X. Zhan, X. Lou, F. Xia, et al., Advanced Functional Materials 28, 1802151 (2018).
  • Klingen et al. (1973) W. Klingen, R. Ott, and H. Hahn, Zeitschrift für anorganische und allgemeine Chemie 396, 271 (1973).
  • Kim et al. (2018) S. Y. Kim, T. Y. Kim, L. J. Sandilands, S. Sinn, M.-C. Lee, J. Son, S. Lee, K.-Y. Choi, W. Kim, B.-G. Park, et al., Physical review letters 120, 136402 (2018).
  • Lane and Zhu (2020) C. Lane and J.-X. Zhu, Physical Review B 102, 075124 (2020).
  • Tan et al. (2022) Q. Tan, W. Luo, T. Li, J. Cao, H. Kitadai, X. Wang, and X. Ling, Applied Physics Reviews 9 (2022).
  • Wang et al. (2021) X. Wang, J. Cao, Z. Lu, A. Cohen, H. Kitadai, T. Li, Q. Tan, M. Wilson, C. H. Lui, D. Smirnov, et al., Nature Materials 20, 964 (2021).
  • Basnet et al. (2021) R. Basnet, A. Wegner, K. Pandey, S. Storment, and J. Hu, Physical Review Materials 5, 064413 (2021).
  • Scheie et al. (2023) A. Scheie, P. Park, J. W. Villanova, G. E. Granroth, C. L. Sarkis, H. Zhang, M. B. Stone, J.-G. Park, S. Okamoto, T. Berlijn, et al., Physical Review B 108, 104402 (2023).
  • Wildes et al. (2015) A. R. Wildes, V. Simonet, E. Ressouche, G. J. Mcintyre, M. Avdeev, E. Suard, S. A. Kimber, D. Lançon, G. Pepe, B. Moubaraki, et al., Physical Review B 92, 224408 (2015).
  • Kim et al. (2019a) K. Kim, S. Y. Lim, J.-U. Lee, S. Lee, T. Y. Kim, K. Park, G. S. Jeon, C.-H. Park, J.-G. Park, and H. Cheong, Nature communications 10, 345 (2019a).
  • He et al. (2024) W. He, Y. Shen, K. Wohlfeld, J. Sears, J. Li, J. Pelliciari, M. Walicki, S. Johnston, E. Baldini, V. Bisogni, M. Mitrano, and M. P. M. Dean, Nature Communications 15, 3496 (2024).
  • Liu et al. (2019) J. Liu, X. Li, Y. Xu, Y. Ge, Y. Wang, F. Zhang, Y. Wang, Y. Fang, F. Yang, C. Wang, et al., Nanoscale 11, 14383 (2019).
  • Wang et al. (2022) X. Wang, J. Cao, H. Li, Z. Lu, A. Cohen, A. Haldar, H. Kitadai, Q. Tan, K. S. Burch, D. Smirnov, et al., Science Advances 8, eabl7707 (2022).
  • Kang et al. (2020) S. Kang, K. Kim, B. H. Kim, J. Kim, K. I. Sim, J.-U. Lee, S. Lee, K. Park, S. Yun, T. Kim, et al., Nature 583, 785 (2020).
  • Hwangbo et al. (2021) K. Hwangbo, Q. Zhang, Q. Jiang, Y. Wang, J. Fonseca, C. Wang, G. M. Diederich, D. R. Gamelin, D. Xiao, J.-H. Chu, et al., Nature Nanotechnology 16, 655 (2021).
  • Belvin et al. (2021) C. A. Belvin, E. Baldini, I. O. Ozel, D. Mao, H. C. Po, C. J. Allington, S. Son, B. H. Kim, J. Kim, I. Hwang, et al., Nature communications 12, 4837 (2021).
  • Ho et al. (2021) C.-H. Ho, T.-Y. Hsu, and L. C. Muhimmah, npj 2D Materials and Applications 5, 8 (2021).
  • Klaproth et al. (2023) T. Klaproth, S. Aswartham, Y. Shemerliuk, S. Selter, O. Janson, J. van den Brink, B. Büchner, M. Knupfer, S. Pazek, D. Mikhailova, et al., Physical Review Letters 131, 256504 (2023).
  • Kim et al. (2023) D. S. Kim, D. Huang, C. Guo, K. Li, D. Rocca, F. Y. Gao, J. Choe, D. Lujan, T.-H. Wu, K.-H. Lin, et al., Advanced Materials 35, 2206585 (2023).
  • Fuggle et al. (1992) J. C. Fuggle, J. E. Inglesfield, and P. Andrews, Unoccupied electronic states: fundamentals for XANES, EELS, IPS and BIS, Vol. 69 (Springer, 1992).
  • Ament et al. (2011) L. J. Ament, M. Van Veenendaal, T. P. Devereaux, J. P. Hill, and J. Van Den Brink, Reviews of Modern Physics 83, 705 (2011).
  • Kim et al. (2019b) H.-S. Kim, K. Haule, and D. Vanderbilt, Physical review letters 123, 236401 (2019b).
  • Yan et al. (2021) M. Yan, Y. Jin, Z. Wu, A. Tsaturyan, A. Makarova, D. Smirnov, E. Voloshina, and Y. Dedkov, The Journal of Physical Chemistry Letters 12, 2400 (2021).
  • DiScala et al. (2023) M. DiScala, D. Staros, A. de la Torre, A. Lopez, D. Wong, C. Schulz, M. Bartkowiak, B. Rubenstein, and K. Plumb, arXiv preprint arXiv:2302.07910  (2023).
  • Khomskii (2014) D. Khomskii, Transition metal compounds (Cambridge University Press, 2014).
  • Sahiner et al. (1996) A. Sahiner, M. Croft, Z. Zhang, M. Greenblatt, I. Perez, P. Metcalf, H. Jhans, G. Liang, and Y. Jeon, Physical Review B 53, 9745 (1996).
  • Jana et al. (2023) D. Jana, P. Kapuscinski, I. Mohelsky, D. Vaclavkova, I. Breslavetz, M. Orlita, C. Faugeras, and M. Potemski, Physical Review B 108, 115149 (2023).
  • Yang et al. (2021) F. Yang, Y.-S. Liu, X. Feng, P.-A. Glans, J. Nasiatka, D. Voronov, J. Feng, Y.-D. Chuang, H. Padmore, and J. Guo, JSSRR 34, 306 (2021).
  • Retegan (2019) M. Retegan, Crispy: v0.7.4 (2019).
  • de Groot et al. (2021) F. M. de Groot, H. Elnaggar, F. Frati, R.-p. Wang, M. U. Delgado-Jaime, M. van Veenendaal, J. Fernandez-Rodriguez, M. W. Haverkort, R. J. Green, G. van Der Laan, et al., Journal of Electron Spectroscopy and Related Phenomena 249, 147061 (2021).
  • Lu et al. (2020) H. Lu, W. Wang, Y. Liu, L. Chen, Q. Xie, H. Yin, G. Cheng, and L. He, Applied Surface Science 504, 144405 (2020).
  • Fang et al. (2021) L. Fang, Y. Xie, P. Guo, J. Zhu, S. Xiao, S. Sun, W. Zi, and H. Zhao, Sustainable Energy & Fuels 5, 2537 (2021).
  • Frati et al. (2020) F. Frati, M. O. Hunault, and F. M. De Groot, Chemical reviews 120, 4056 (2020).
  • Wu et al. (2005) Z. Wu, C. Liu, L. Guo, R. Hu, M. Abbas, T. Hu, and H. Xu, The Journal of Physical Chemistry B 109, 2512 (2005).

VI Figures

Refer to caption
Figure 1: The crystal structure of NiPS3 seen along the c𝑐citalic_c axis (top) and a𝑎aitalic_a axis (bottom). Generated by VESTAKlingen et al. (1973).
Refer to caption
Figure 2: Simplified electronic energy levels for NiPS3 assuming it is a charge-transfer insulator (left) and assuming it is a Mott-Hubbard insulator (right)Sahiner et al. (1996).
Refer to caption
Figure 3: Nickel L3,2-edge XAS spectra with TEY mode (blue) and TFY mode (red) for (a) an untreated NiPS3 sample and (b) a NiPS3 sample after exfoliation of the surface layers in an air-isolated environment.
Refer to caption
Figure 4: Surface oxidation check: O K-edge XAS spectra with TEY mode (blue) and TFY mode (red) for the untreated NiPS3 sample used in FIG. 3(a).
Refer to caption
Figure 5: Experimental (Exp.) nickel L3,2-edge XAS spectra and calculated spectra from the charge-transfer multiplet model by Crispy softwareRetegan (2019). The top of the figure shows the experimental XAS data for the exfoliated NiPS3 sample with TEY signal (blue) and TFY signal (red). The subsequent spectra from top to bottom represent the calculated spectra corresponding to ΔΔ\Deltaroman_Δ decreasing from 5.0 eV to -1.0 eV.
Refer to caption
Figure 6: Nickel L3-edge mapping of RIXS of NiPS3. Yellow solid line: elastic peak. Red dashed line: energy lost features.