Introduction
In 2D magnets, enhanced fluctuations and lattice connectivity strike a balance from which collective states unobtainable in three dimensions may emerge. The ability to prepare isolated monolayers of van der Waals (vdW) magnets has enabled access to new magnetic phases and tests of fundamental theorems of magnetism,[1–6] and opens up possibilities for controlling or engineering unconventional states through stacking.[7] Much work has concentrated on vdW materials with a net ferromagnetism in the 2D limit.[5,8–10] However, antiferromagnets may offer more possibilities to explore complex magnetic order, topological spin textures, or quantum spin-liquids that arise from frustrated interactions and which are stabilized in 2D.[11–15]
NiPS3 stands as one of the few exfoliatable materials that exhibits both antiferromagnetic order and strong correlations.[16,17] Recent Raman scattering measurements suggest that the magnetic order in NiPS3 is highly sensitive to dimensionality and find that long-range order vanishes in the monolayer limit in favor of a fluctuating magnetic phase.[3] Based on the magnetic Hamiltonian that was determined by inelastic neutron scattering on bulk samples,[18] the thickness-dependent Raman data were associated with the proliferation of vorticies through a Berezinskii–Kosterlitz–Thouless phase transition in the 2D material. However, this explanation assumes that the few-layer magnetic Hamiltonian is identical to that in the bulk. More direct experimental access to the electronic energy scales and magnetic interactions is thus necessary to resolve the nature of the magnetic state in exfoliated NiPS3.
In this work, we demonstrate that Ni-S electronic energy scales are strongly altered by dimensionality in NiPS3 leading to a few-layer magnetic Hamiltonian that differs from that of the bulk. We use resonant Inelastic X-ray Scattering (RIXS) to directly interrogate correlated electronic states in exfoliated flakes of NiPS3 and reveal a systematic softening and broadening of NiS6 multiplet excitations with decreasing thickness that is reproduced by a multiplet ligand-field model. Overall, we find that decreased hopping integrals and charge transfer energy in 2D result in a more covalent character for Ni d-orbitals (Figure 1a). We additionally compute the relevant magnetic exchange integrals and find a decrease in the second- and third-nearest neighbor magnetic interaction strengths, and an increase in the first-nearest neighbor interaction strength. This change moves NiPS3 closer to the boundary between the stripy antiferromagnetic and spiral ordered phases of the honeycomb antiferromagnet. Lastly, we show how the change of electronic energy scales in thinner samples occurs due to decreased electronic vdW delocalization across layers in the 2D limit. Since this mechanism is not necessarily specific to NiPS3, its effect will be important to the properties of a broad class of strongly correlated few-layer vdW magnets.
[IMAGE OMITTED. SEE PDF]
Experimental Results
NiPS3 crystallizes in the monoclinic C2/m space group with Ni (S = 1) atoms on honeycomb lattices in the ab-planes. Each Ni atom is octahedrally coordinated by six S atoms, with P atoms situated between the 2D sheets of NiS6 that form the vdW gap along the c-axis with an interlayer spacing of ≈ 0.636 nm.[2,16] NiPS3 is known to magnetically order at a transition temperature of TN = 155 K; ferromagnetic zig-zag chains form with moments parallel to the a-axis and antiferromagnetically coupled along the b-axis, with a small out of plane component.[16,21]
Figure 1c,e show representative RIXS spectra for bulk and three-layer (3L) NiPS3, respectively. Spectra were collected at the peak of the Ni L3-edge XAS Ei = 853 eV, corresponding to 2p3/2 to 3d electronic transitions. We concentrate on the low energy region Eloss = 0.2 → 2.15 eV that contains excitations within the NiS6 multiplet. The bulk and three-layer (3L) spectra are qualitatively similar except for an overall energy broadening and softening that is readily visible in the 3L data (Figure 1b). While the qualitative similarity between bulk and 3L spectra is consistent with the fact that there are no drastic structural reconstructions upon exfoliation, the apparent broadening and softening indicates a change in the electronic structure of NiPS3 with thickness.
In order to elucidate the origin of this change, we first concentrate our analysis on the bulk spectra and identify all relevant features. We found that a minimum of six Gaussian modes were required to fit the bulk data, labeled A - F in Figure 1c, with minimum energy widths fixed by the experimental energy resolution (≈ 55meV FWHM). Each of these features can be identified as an excitation within the electronic multiplet of trigonally distorted NiS6 octahedra (Figure 2a; Figure SV, Supporting Information). The center of mass positions of the two bulk peaks are assigned to the t2g → eg (d-d) excitations of 3T2g and 3T1g symmetry, respectively, in good agreement with optical measurements.[17, 22] The trigonal distortion introduces a D3d symmetry which splits 3T2g → 3A1g + 3Eg (peaks B & C) and 3T1g → 3A2g + 3Eg (peaks E & F), in agreement with Raman and optical measurements.[22,23] Access to spin-flip (ΔS ≠ 0) excitations in the RIXS cross-section leads us to assign peak D 1Eg symmetry as the next highest excited state above 3T2g in a 3d8 system. Lastly, peak A is assigned to a charge transfer excitation with character, where denotes n ligand holes. The 800 meV energy scale of this peak indicates a small charge transfer energy in NiPS3. We verify these peak assignments through the application of the NiS6 multiplet ligand-field model described below. We note that since our incident energy was tuned to the peak of the Ni L3-edge XAS, our measurements were not sensitive to the sharp 1.47 eV peak reported in Ref. [24].
[IMAGE OMITTED. SEE PDF]
We now bring our attention to the spectra collected on the exfoliated samples, and in particular, the 3L sample. Following the fitting procedure from the bulk spectra, we again fit six Gaussian modes; however, Empirically fitting the 3L spectra to a minimum of six Gaussian peaks resulted in two scenarios of equally good fit quality. In order to address the width of peak A, we explore the two fitting scenarios for the 3L spectra. In scenario one, the widths of all peaks were held fixed at the experimental resolution; this fit converged with a systematic softening of all peaks between the bulk and 3L data sets. In scenario two, the FWHM of peak A was allowed to relax, while peaks B - D were fixed to the experimental resolution, and peaks E - F were fixed to the bulk fitted FWHM to experimental resolution ratio; this fit converged with minimal softening of all peaks, but systematic broadening and increased intensity attributed to peak A. We found that peak energies extracted from scenario one could only be reproduced within physically meaningful parameters using a negative charge transfer energy, while scenario two is reproduced with a small positive charge transfer energy.[25] A negative charge transfer energy for the 3L sample implies a zero-crossing of the charge transfer energy as a function of thickness between bulk and 2D exfoliated samples. We rule out such a transition based on the smooth evolution of thickness dependent RIXS data and Raman spectra.[23] The full details of each fitting scenario to the 3L spectrum, including details of the bulk spectrum fitting, can be found in the Supporting Information.[25]
Fits for scenario two are shown in Figure 1e, while Figure 1d summarizes the centroids of the fitted peaks for the various sample thicknesses measured. Minimal differences were found between the bulk and 60L, placing a lower limit on bulk behavior for exfoliated NiPS3 at ≈ 38 nm. From bulk to 3L, peaks B - F vary slightly in Eloss. We find a broadening in the FWHM of peak A by 360(50) meV over the bulk data that suggests a change in the charge transfer energy in few-layer samples. Furthermore, the observed systematic broadening of excitations signifies an electronic structure intricately connected to sample thickness in NiPS3.
Theoretical Analysis
Having identified a clear empirical trend, we perform a thorough fit of the experimental data for paramagnetic NiPS3 using a high-fidelity multiplet ligand-field model (MLFM) in the basis of symmetry-adapted linear combinations of ligand orbitals.[26] Crucially, we use rigorously converged parameter-free solutions of the Schrödinger equation to physically guide our search of the MLFM parameter space. Our MLFM model includes Slater-Condon parameters(, , ), covalent hopping integrals between S 3p- and Ni 3d- orbitals, pdσ and pdπ, S 3p- orbital level splitting, Tpp = ppσ − ppπ, cubic crystal field (10Dq) and trigonal distortion δ, Ni 3d-3d and 2p-3d on-site Coulomb interactions Udd and Upd = 1.2Udd, and charge transfer energy Δ. The values of the transition metal-ligand (pdσ, pdπ) and ligand–ligand (ppσ, ppπ) hopping integrals for bulk and monolayer geometries are calculated independently by self-consistently converging to the electronic ground state of the nonmagnetic configuration of NiPS3 using Density Functional Theory (DFT). We employ a substantial plane wave energy cutoff of more than 3,400 eV, sample the Brillouin zone such that all energies are converged to within less than 0.001 eV, and use correlation-consistent pseudopotentials obtained from fully many-body ab initio calculations.[38] We then generate maximally-localized Wannier functions (MLWF) which span a substantial subspace of the rigorously converged DFT ground states, and use the resulting tight-binding inter-orbital expectation values to solve for the hopping integrals within the two-center linear combination of atomic orbitals approximation.[25]
We carried out a search of the remaining parameter space for Δ, 10Dq, and by minimizing the difference between calculated energies peaks A - F while keeping F(G)pd fixed to 80% of their atomic Hartree-Fock values.[25,27]. For initial comparisons of this model to our data, we used physically meaningful parameters for octahedrally coordinated NiS6.[28–30] Figure 2a,c shows the calculated energy levels for a NiS6 cluster as a function of the charge transfer energy Δ using fixed parameters that give the best agreement between measured and calculated peak energies for bulk and 3L NiPS3, respectively.[25] Previous optical and X-ray absorption (XAS) studies classified NiPS3 as a negative charge transfer insulator,[17] while more recent RIXS and XAS measurements indicate a positive charge transfer gap.[24,31] We find that, for bulk NiPS3, a small positive Δ = 0.83 eV was necessary to give an accurate match to the data; for 3L NiPS3, the charge transfer gap decreases, and accordingly, the best match to the data is obtained with Δ = 0.37 eV.
In Figure 3a,b, we show RIXS spectra calculated using the open-source toolkit EDRIXS[32] compared to the experimental data. Intensities were normalized to the nominal 10Dq line. Covalent hopping integrals, pdσ and pdπ, as well as Tpp were fixed to those obtained from ab initio calculations of the nonmagnetic configuration (Table 1) while Δ was allowed to vary. To account for broadening of excitations not captured by our multiplet ligand-field model and facilitate better comparison with experimental data, the calculated spectra were broadened by increasing the final-state lifetime above 2 eV in Eloss. We can reproduce the observed broadening and softening of NiS6 multiplet excitations in the 3L spectrum by a decrease in charge transfer energy and transition metal-ligand hopping integrals, as parameterized by Δ/Tpd, and Δ/U, where Tpd = and on-site 3d Coulomb repulsion [26,33]: Δ/Tpd = 0.45 and Δ/U = 0.10 in bulk, and Δ/Tpd = 0.23 and Δ/U = 0.04 in 3L.
Table 1 Fixed values in eV of hopping integrals extracted from ab initio calculations of the nonmagnetic configuration. Charge transfer energy Δ, and intra-orbital Coulomb repulsion extracted from RIXS modeling.[26,33]
pdσ | pdπ | ppσ | ppπ | Tpp | Δ | U | |
Bulk | −1.07 | 0.67 | 0.89 | −0.09 | 0.98 | 0.83 | 8.3 |
3L | −0.93 | 0.46 | 0.62 | −0.01 | 0.63 | 0.37 | 8.3 |
[IMAGE OMITTED. SEE PDF]
We determine that the underlying mechanism responsible for the significant change in the RIXS signal with thickness is predominantly electronic rather than structural in origin, though the lattice constant is slightly overestimated in the PBE-optimized monolayer. We find that as NiPS3 gets thinner, metal-ligand π-hopping is reduced (pdπ decreases) due to the removal of π-like interlayer vdW interactions. The same effect also causes pdσ and Tpp to change significantly because of the mixed σ- and π-bonding character present in the sp3-hybridized phosphorus atoms that bridge the NiS6 clusters. In the context of our MLWFs, this is reflected in a change in the largest tight-binding energies used to solve for pdσ and pdπ (Figure 1b).[25]
The combination of RIXS measurements and ab initio calculations constrain the electronic ground state that underlies the magnetic properties of NiPS3. In Figure 4c, we investigate the change in the ground state character of extracted from our multiplet ligand-field model as a function of Tpp and Δ/Tpd. We find a negligible contribution from the state and nearly equal populations of the and states. In bulk, , and in 3L, , implying a small increase in the magnitude of the paramagnetic Ni moment. As shown in Figure 4c, this minimal change is accounted for by the dependence of the ground state character on both the hybridization Δ/Tpd and Tpp. While an increased transition metal-ligand hybridization tends to enhance the character, this is offset by the reduction in the ligand–ligand hybridization Tpp.
[IMAGE OMITTED. SEE PDF]
Despite the small change in ground state character, the large change in hybridization and hopping parameters influences the magnetic exchange interactions. Using the parameters obtained from our RIXS measurements and ab initio modeling, we compute the superexchange interactions up to the third-nearest neighbor within a sixth order cell-perturbation.[30,34,35] The first J1, second J2, and third J3 nearest neighbor expressions are given by the second order perturbation terms for the states, and fourth and sixth order terms for the states;[30] Figure 4a shows the first-nearest neighbor superexchange pathway for the configuration, while a detailed description of all these expressions is given in the Supporting Information.[25] In bulk NiPS3, we find meV, meV, and meV, in excellent agreement with recently reported values from inelastic neutron scattering.[18,21] The decrease in Tpp and Δ/Tpd leads to an overall enhancement of meV, a vanishing , and decrease in meV. In Figure 4b, we summarize the dependence of J3/J1 on Tpp and Δ/Tpd. We find that J1 is dominated by the state, while J2 and J3 are dominated by the state. Thus, the decrease in Tpp is directly responsible for an increased contribution to J1. As a consequence of the overall reduction in the average exchange interaction strength, the magnetic transition temperature is expected to be reduced in few-layer samples compared to bulk samples. Furthermore, the decrease in J3/J1 from -3.3 in bulk to -1.7 in 3L, positions 3L NiPS3 closer to a phase boundary between the stripy AFM phase and a spiral ordered phase.[36] It is likely that, in the 2D limit, NiPS3 is driven into a highly frustrated regime on this phase boundary.
Conclusion
In summary, we used RIXS to access the electronic ground state properties of an exfoliated, correlated antiferromagnet in the 2D limit. We found that the electronic energy scales associated with Ni-S hybridization, and consequently the magnetic exchange interactions, are altered in a non-trivial way though the modification of interlayer energy scales upon exfoliation of NiPS3 despite minimal structural changes. Our findings demonstrate that magnetic exchange parameters determined from measurements on bulk materials are not applicable in the 2D limit, as interlayer interactions, absent in 2D, affect intralayer ones. The underlying electronic mechanism we have identified points to the possibility of controlling magnetic interactions in strongly correlated van der Waals heterostructures by tuning interfacial energy scales toward the design of the next generation of 2D strongly correlated magnetic materials.
Experimental Section
Crystal Growth
Single crystal samples of NiPS3 were grown by vapor transport, following previously published methods.[16,18] Stoichiometric quantities of metallic Nickel (99.994% purity), crystalline Phorphorus (99.999% purity) and Sulfur (99.999% purity), were added to a quartz tube in an argon-filled glovebox. The quartz tube was then evacuated, sealed, and placed in a two-zone furnace and heated to 700 °C/750 °C over 6 h; the furnace was held at this temperature gradient for 2 days, and cooled to 670 °C/620 °C over 8 h and held for an additional 16 days. After a total of 18 days in the furnace, samples were cooled to room temperature over 8 h. The crystals formed were shiny gray metallic and had hexagonal motif characteristic of NiPS3.
Exfoliated Sample Preparation
Bulk NiPS3 was exfoliated using conventional scotch-tape methods[2] and deposited either onto a blank SiO2 substrate or onto a SiO2 substrate pre-treated with a patterned Copper (Cu) grid. Sample layer count was determined using a combination of atomic-force microscope thickness measurements and optical contrast with SiO2 substrate. See Supporting Information and Figure SII (Supporting Information) for more details.[25] The patterned Cu grid consisted of 100 µ m × µ m SiO2 cells separated by 200 µ m of 50 nm thickness Cu (Figure SIb,c, Supporting Information).[25] The 3L sample was deposited onto a blank SiO2 substrate and was later patterned with a Cu fiducial marker using electron-beam lithography (Figure SIa, Supporting Information), again with a Cu thickness of 50 nm.[25] In both cases, Cu was chosen as a material that could provide a fluorescence contrast to SiO2 in the soft X-ray regime. This fluorescence contrast proved invaluable in locating small samples whose signals were weak under an X-ray beam. The Cu grid proved a useful method for locating sample(s) as a unique grid scheme could be defined for each chip if the orientation of each chip remained consistent; however, the fiducial marker had the advantage of of being visible by eye, resulting in unequivocal sample location and removing the requirement of a grid scheme. Exfoliated samples were spin coated with a PMMA protective layer and stored in an Ar atmosphere to prevent degradation.
RIXS Measurements
Room temperature RIXS measurements on exfoliated flakes were carried out on the PEAXIS beamline at BESSY II.[37] The PMMA coatings were removed immediately prior to loading the samples into the RIXS vacuum chamber via washing with acetone and isopropyl alcohol. All samples chosen for measurement had lateral dimensions larger than, or comparable to, the 15 (H) µ m x 4 (V) µ m x-ray beam spot size at the PEAXIS beamline. A horizontal scattering geometry of 2θ = 90° was used with an ≈ 235 meV energy resolution (full width at half max, FWHM) using linear horizontal polarization and specular geometry. Spectra were collected in 30 min segments to minimize sample exposure to the X-ray beam. It was noted here that the 7 L spectra only was collected with marginally better instrumental resolution of ≈ 190 meV. Bulk NiPS3 RIXS measurements were carried out on the SIX beamline at NSLS II with identical scattering geometry, but with an ≈ 55 meV energy resolution FWHM. The main resonance peak at the Ni L3-edge was chosen via X-ray absorption spectroscopy (XAS) preformed on-site, prior to each RIXS measurements.
Acknowledgements
The authors thank Mark Dean for helpful discussions and comments on this manuscript. The authors also thank Naiyuan J. Zhang and Erin Morissette for their guidance and consultation on pattern fabrication. M.F.D. and K.W.P. were supported by the National Science Foundation under grant no. OMA-1936221. A.D.L.T. was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Award Number DE-SC0021. D.S. and B.R. were supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division, as part of the Computational Materials Sciences Program and the Center for Predictive Simulation of Functional Materials, while A.L. was supported by the Brown University Diversity Fellowship. RIXS measurements were carried out at the U41-PEAXIS beamline at the BESSY II electron storage ring operated by the Helmholtz-Zentrum Berlin für Materialien und Energie. This research was conducted using the computational resources of the National Energy Research Scientific Computing Center under Contract No. DE- AC0205CH11231, which is a U.S. Department of Energy Office of Science User Facility, as well as resources and services at the Center for Computation and Visualization, Brown University.
Conflict of Interest
The authors declare no conflict of interest.
Data Availability Statement
The data that support the findings of this study are openly available in Zenodo at , reference number 10125572.
J.‐G. Park, J. Phys.: Condens. Matter 2016, 28, [eLocator: 301001].
C.‐T. Kuo, M. Neumann, K. Balamurugan, H. J. Park, S. Kang, H. W. Shiu, J. H. Kang, B. H. Hong, M. Han, T. W. Noh, J.‐G. Park, Sci. Rep. 2016, 6, [eLocator: 20904].
K. Kim, S. Y. Lim, J.‐U. Lee, S. Lee, T. Y. Kim, K. Park, G. S. Jeon, C.‐H. Park, J.‐G. Park, H. Cheong, Nat. Commun. 2019, 10, 345.
N. Sivadas, S. Okamoto, X. Xu, Craig. J. Fennie, D. Xiao, Nano Lett. 2018, 18, 7658.
C. Gong, L. Li, Z. Li, H. Ji, A. Stern, Y. Xia, T. Cao, W. Bao, C. Wang, Y. Wang, Z. Q. Qiu, R. J. Cava, S. G. Louie, J. Xia, X. Zhang, Nature 2017, 546, 265.
Y. Tang, K. Su, L. Li, Y. Xu, S. Liu, K. Watanabe, T. Taniguchi, J. Hone, C.‐M. Jian, C. Xu, K. F. Mak, J. Shan, Nat. Nanotechnol. 2023, 18, 233.
F. Hellman, A. Hoffmann, Y. Tserkovnyak, G. S. D. Beach, E. E. Fullerton, C. Leighton, A. H. MacDonald, D. C. Ralph, D. A. Arena, H. A. Dürr, P. Fischer, J. Grollier, J. P. Heremans, T. Jungwirth, A. V. Kimel, B. Koopmans, I. N. Krivorotov, S. J. May, A. K. Petford‐Long, J. M. Rondinelli, N. Samarth, I. K. Schuller, A. N. Slavin, M. D. Stiles, O. Tchernyshyov, A. Thiaville, B. L. Zink, Rev. Mod. Phys. 2017, 89, [eLocator: 025006].
B. Huang, G. Clark, E. Navarro‐Moratalla, D. R. Klein, R. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A. McGuire, D. H. Cobden, W. Yao, D. Xiao, P. Jarillo‐Herrero, X. Xu, Nature 2017, 546, 270.
K. S. Burch, D. Mandrus, J.‐G. Park, Nature 2018, 563, 47.
M. Gibertini, M. Koperski, A. F. Morpurgo, K. S. Novoselov, Nat. Nanotechnol. 2019, 14, 408.
A. Kitaev, Annals of Physics 2006, 321, 2.
K. W. Plumb, J. P. Clancy, L. J. Sandilands, V. V. Shankar, Y. F. Hu, K. S. Burch, H.‐Y. Kee, Y.‐J. Kim, Phys. Rev. B 2014, 90, [eLocator: 041112(R)].
J. Nasu, J. Knolle, D. L. Kovrizhin, Y. Motome, R. Moessner, Nature Phys 2016, 12, 912.
K. H. Lee, S. B. Chung, K. Park, J.‐G. Park, Phys. Rev. B 2018, 97, [eLocator: 180401].
B. Zhou, Y. Wang, G. B. Osterhoudt, P. Lampen‐Kelley, D. Mandrus, R. He, K. S. Burch, E. A. Henriksen, J. Phys. Chem. Solids 2019, 128, 291.
A. R. Wildes, V. Simonet, E. Ressouche, G. J. McIntyre, M. Avdeev, E. Suard, S. A. J. Kimber, D. Lançon, G. Pepe, B. Moubaraki, T. J. Hicks, Phys. Rev. B 2015, 92, [eLocator: 224408].
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, C. Jeon, H.‐D. Kim, C.‐H. Park, J.‐G. Park, S. J. Moon, T. W. Noh, Phys. Rev. Lett. 2018, 120, [eLocator: 136402].
D. Lançon, R. A. Ewings, T. Guidi, F. Formisano, A. R. Wildes, Phys. Rev. B 2018, 98, [eLocator: 134414].
K. Momma, F. Izumi, J. Appl. Cryst. 2011, 44, 1272.
B. O. Community, Blender ‐ a 3D modelling and rendering package, Blender Foundation, Stichting Blender Foundation, Amsterdam 2018.
A. R. Wildes, J. R. Stewart, M. D. Le, R. A. Ewings, K. C. Rule, G. Deng, K. Anand, Phys. Rev. B 2022, 106, [eLocator: 174422].
D. Afanasiev, J. R. Hortensius, M. Matthiesen, S. Mañas‐Valero, M. Šiškins, M. Lee, E. Lesne, H. S. J. van der Zant, P. G. Steeneken, B. A. Ivanov, E. Coronado, A. D. Caviglia, Sci. Adv. 2021, 7, [eLocator: eabf3096].
X. Wang, J. Cao, H. Li, Z. Lu, A. Cohen, A. Haldar, H. Kitadai, Q. Tan, K. S. Burch, D. Smirnov, W. Xu, S. Sharifzadeh, L. Liang, X. Ling, Sci. Adv. 2022, 8, [eLocator: eabl7707].
S. Kang, K. Kim, B. H. Kim, J. Kim, K. I. Sim, J.‐U. Lee, S. Lee, K. Park, S. Yun, T. Kim, A. Nag, A. Walters, M. Garcia‐Fernandez, J. Li, L. Chapon, K.‐J. Zhou, Y.‐W. Son, J. H. Kim, H. Cheong, J.‐G. Park, Nature 2020, 583, 785.
See supporting information for details on sample preparation and preservation, additional details about the exact diagonalization calculations as well as further information about the analysis and fitting of the rixs spectra, details of the ab initio calculation methodology in obtaining mlwfs, and tm‐l and l‐l hopping parameters, and detailed description of the superexchange expressions.
M. W. Haverkort, M. Zwierzycki, O. K. Andersen, Phys. Rev. B 2012, 85, [eLocator: 165113].
G. Ghiringhelli, M. Matsubara, C. Dallera, F. Fracassi, R. Gusmeroli, A. Piazzalunga, A. Tagliaferri, N. B. Brookes, A. Kotani, L. Braicovich, J. Phys.: Condens. Matter 2005, 17, 5397.
A. E. Bocquet, T. Mizokawa, T. Saitoh, H. Namatame, A. Fujimori, Phys. Rev. B 1992, 46, 3771.
S. R. Krishnakumar, D. D. Sarma, Phys. Rev. B 2003, 68, [eLocator: 155110].
K. Takubo, T. Mizokawa, J.‐Y. Son, T. Nambu, S. Nakatsuji, Y. Maeno, Phys. Rev. Lett. 2007, 99, [eLocator: 037203].
M. Yan, Y. Jin, Z. Wu, A. Tsaturyan, A. Makarova, D. Smirnov, E. Voloshina, Y. Dedkov, J. Phys. Chem. Lett. 2021, 12, 2400.
Y. Wang, G. Fabbris, M. P. M. Dean, G. Kotliar, Comput. Phys. Commun. 2019, 243, 151.
Y. Shen, J. Sears, G. Fabbris, J. Li, J. Pelliciari, I. Jarrige, X. He, I. Bozovic, M. Mitrano, J. Zhang, J. F. Mitchell, A. S. Botana, V. Bisogni, M. R. Norman, S. Johnston, M. P. M. Dean, Phys. Rev. X 2022, 12, [eLocator: 011055].
J. H. Jefferson, H. Eskes, L. F. Feiner, Phys. Rev. B 1992, 45, 7959.
H. Eskes, J. H. Jefferson, Phys. Rev. B 1993, 48, 9788.
J. Fouet, P. Sindzingre, C. Lhuillier, Eur. Phys. J. B 2001, 20, 241.
C. Schulz, K. Lieutenant, J. Xiao, T. Hofmann, D. Wong, K. Habicht, J. Synchrotron Rad. 2020, 27, 238.
A. Annaberdiyev, G. Wang, C. A. Melton, M. C. Bennett, L. Shulenburger, L. Mitas, The Journal of Chemical Physics, 2018 149. [DOI: https://dx.doi.org/10.1063/1.5040472]
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2024. This work is published under http://creativecommons.org/licenses/by/4.0/ (the "License"). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
The sustained interest in investigating magnetism in the 2D limit of insulating antiferromagnets is driven by the possibilities of discovering, or engineering, novel magnetic phases through layer stacking. However, due to the difficulty of directly measuring magnetic interactions in 2D antiferromagnets, it is not yet understood how intralayer magnetic interactions in insulating, strongly correlated, materials can be modified through layer proximity. Herein, the impact of reduced dimensionality in the model van der Waals antiferromagnet NiPS3 is explored by measuring electronic excitations in exfoliated samples using Resonant Inelastic X‐ray Scattering (RIXS). The resulting spectra shows systematic broadening of NiS6 multiplet excitations with decreasing layer count from bulk down to three atomic layers (3L). It is shown that these trends originate from a decrease in transition metal‐ligand and ligand–ligand hopping integrals, and by charge‐transfer energy evolving from Δ = 0.83 eV in the bulk to 0.37 eV in 3L NiPS3. Relevant intralayer magnetic exchange integrals computed from the electronic parameters exhibit a decrease in the average interaction strength with thickness. This study underscores the influence of interlayer electronic interactions on intralayer ones in insulating magnets, indicating that magnetic Hamiltonians in few‐layer insulating magnets can greatly deviate from their bulk counterparts.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details




1 Department of Physics, Brown University, Providence, RI, USA
2 Department of Chemistry, Brown University, Providence, RI, USA
3 Department of Dynamics and Transport in Quantum Materials, Helmholtz‐Zentrum Berlin für Materialen und Energie, Berlin, Germany
4 National Synchrotron Light Source II, Brookhaven National Laboratory, Upton, NY, USA