research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767

Structural, elastic, electronic, optical and vibrational properties of single-layered, bilayered and bulk molybdenite MoS2-2H

crossmark logo

aBiological, Geological and Environmental Sciences, University of Bologna, P. Porta San Donato 1, Bologna, Emilia-Romagna 40127, Italy, and bInterdisciplinary Centre of Biomineralogy, Crystallography and Biomaterials, University of Bologna, P. Porta San Donato 1, Bologna, Emilia-Romagna 40127, Italy
*Correspondence e-mail: gianfranco.ulian2@unibo.it, giovanni.valdre@unibo.it

Edited by H. Brand, Australian Synchrotron, ANSTO, Australia (Received 7 October 2022; accepted 17 March 2023; online 25 April 2023)

In recent years, transition metal dichalcogenides have received great attention since they can be prepared as two-dimensional semiconductors, presenting heterodesmic structures incorporating strong in-plane covalent bonds and weak out-of-plane interactions, with an easy cleavage/exfoliation in single or multiple layers. In this context, molybdenite, the mineralogical name of molybdenum disulfide, MoS2, has drawn much attention because of its very promising physical properties for optoelectronic applications, in particular a band gap that can be tailored with the material's thickness, optical absorption in the visible region and strong light–matter interactions due to the planar exciton confinement effect. Despite this wide interest and the numerous experimental and theoretical articles in the literature, these report on just one or two specific features of bulk and layered MoS2 and sometimes provide conflicting results. For these reasons, presented here is a thorough theoretical analysis of the different aspects of bulk, monolayer and bilayer MoS2 within the density functional theory (DFT) framework and with the DFT-D3 correction to account for long-range interactions. The crystal chemistry, stiffness, and electronic, dielectric/optical and phonon properties of single-layered, bilayered and bulk molybdenite have been investigated, to obtain a consistent and detailed set of data and to assess the variations and cross correlation from the bulk to single- and double-layer units. The simulations show the indirect–direct transition of the band gap (KK′ in the first Brillouin zone) from the bulk to the single-layer structure, which however reverts to an indirect transition when a bilayer is considered. In general, the optical properties are in good agreement with previous experimental measurements using spectroscopic ellipsometry and reflectivity, and with preliminary theoretical simulations.

1. Introduction

Molybdenite (MoS2) is a sulfide mineral that crystallizes in the hexagonal crystal system (space group P63/mmc) under standard pressure and temperature conditions. The unit cell of molybdenite is shown graphically in Fig. 1[link], presenting two formula units (Z = 2) of molybdenum disulfide, each one forming a layer where the atoms are linked by mixed covalent/ionic bonds. The molybdenum atom is sixfold coordinated, with a trigonal prism arrangement of the S2− ions. These layers are extended indefinitely along the a and b axes and stacked along the c-axis direction. The whole structure is held along the [001] direction by weak long-range (van der Waals) interactions. Because the crystal structure is made up of stacked layers, molybdenite is subject to polytypism, and the structural model in Fig. 1[link] represents the MoS2-2H polymorph stable under standard conditions of pressure and temperature (1 atm, 298.15 K; 1 atm = 101 325 Pa).

[Figure 1]
Figure 1
(Left and middle) Ball-and-stick models of bulk molybdenite MoS2-2H viewed along the (left) [110] and (middle) [001] directions. The black lines represent the hexagonal unit cell of the mineral. The S—Mo—S bond angle Θ, the interlayer distance dI and the z parameter (i.e. the atomic shift of sulfur with respect to molybdenum atoms, scaling with the length of the c axis) are reported. (Right) The electronic band structure of MoS2-2H, highlighting the indirect band gap (I, green line) and the excitonic KK′ (used in experimental work) transitions 1 (from valence band v1 to conduction band c1, black line) and 2 (v2 → c1, blue line).

Molybdenite is a widely employed mineral phase in several and manifold geological, industrial and technological applications. It is the main ore mineral for the extraction of metallic molybdenum (Hess, 1924[Hess, F. L. (1924). Molybdenum Deposits: A Short Review. Bulletin 761. United States Geological Survey. Washington, DC: Department of the Interior.]) and an element used in alloy steels, superalloys and pigments, just to cite a few examples (Kropschot, 2010[Kropschot, S. J. (2010). Molybdenum - A Key Component of Metal Alloys. Fact Sheet 2009-3106. Reston: US Geological Survey.]). Bulk molybdenite can be used as a dry lubricant because of its very low friction coefficient (Martin et al., 1993[Martin, J. M., Donnet, C., Le Mogne, T. & Epicier, T. (1993). Phys. Rev. B, 48, 10583-10586.], 1994[Martin, J. M., Pascal, H., Donnet, C., Le Mogne, T., Loubet, J. L. & Epicier, T. (1994). Surf. Coat. Technol. 68-69, 427-432.]), even under high pressure (Wu et al., 2022[Wu, S., Meng, Z., Tao, X. & Wang, Z. (2022). Friction, 10, 209-216.]).

In addition, molybdenite was the first discovered semiconductor material in the early 20th century (Aminoff & Broome, 1935[Aminoff, G. & Broome, B. (1935). Z. Kristallogr. 91, 77-94. ]), and it has drawn much attention in the past decade for its use as a two-dimensional (2D) material in field-emission transistors (FETs) and other electronic devices, such as solar cells and light-emitting diodes (Radisavljevic et al., 2011[Radisavljevic, B., Radenovic, A., Brivio, J., Giacometti, V. & Kis, A. (2011). Nat. Nanotech. 6, 147-150.]; Fontana et al., 2013[Fontana, M., Deppe, T., Boyd, A. K., Rinzan, M., Liu, A. Y., Paranjape, M. & Barbara, P. (2013). Sci. Rep. 3, 1634.]; Sebastian et al., 2021[Sebastian, A., Pendurthi, R., Choudhury, T. H., Redwing, J. M. & Das, S. (2021). Nat. Commun. 12, 693.]; Wang et al., 2021[Wang, Y., Tang, H., Xie, Y., Chen, X., Ma, S., Sun, Z., Sun, Q., Chen, L., Zhu, H., Wan, J., Xu, Z., Zhang, D. W., Zhou, P. & Bao, W. (2021). Nat. Commun. 12, 3347.]). Thanks to the weak interactions along the [001] direction, molybdenite is easily cleavable, and a monolayer of molybdenite was exfoliated by Joensen et al. (1986[Joensen, P., Frindt, R. F. & Morrison, S. R. (1986). Mater. Res. Bull. 21, 457-461.]) using intercalated lithium. It was shown that a single layer of MoS2 has superior properties to graphene for the fabrication of FETs and optoelectronic devices because it presents a gap in the band structure (Radisavljevic et al., 2011[Radisavljevic, B., Radenovic, A., Brivio, J., Giacometti, V. & Kis, A. (2011). Nat. Nanotech. 6, 147-150.]). This band gap is tunable by increasing or decreasing the number of layers of the material (Yang & Li, 2020[Yang, X. & Li, B. (2020). Nanophotonics, 9, 1557-1577.]), and even changes from an indirect gap of about 1.23 eV in bulk MoS2-2H (Kam & Parkinson, 1982[Kam, K. K. & Parkinson, B. A. (1982). J. Phys. Chem. 86, 463-467.]) to a direct gap of about 1.8 eV in the monolayer (Kuc et al., 2011[Kuc, A., Zibouche, N. & Heine, T. (2011). Phys. Rev. B, 83, 245213.]). This indirect–direct band-gap transition is fundamental to enhancing light emission and optical absorption, and to inducing photoluminescence in the monolayer of molybdenite (Mak & Shan, 2016[Mak, K. F. & Shan, J. (2016). Nat. Photon. 10, 216-226.]).

Several experimental and theoretical reports have investigated the structural, electronic and vibrational properties of bulk and layered molybdenite to provide a better understanding of the quantum confinement effects originating from `layering' the mineral and to drive the development of MoS2-based materials for photonics and optoelectronic applications. However, most of them focused only on specific features, such as the band gap (Kuc et al., 2011[Kuc, A., Zibouche, N. & Heine, T. (2011). Phys. Rev. B, 83, 245213.]; Kam & Parkinson, 1982[Kam, K. K. & Parkinson, B. A. (1982). J. Phys. Chem. 86, 463-467.]), the dielectric properties (Beal & Hughes, 1979[Beal, A. R. & Hughes, H. P. (1979). J. Phys. C Solid State Phys. 12, 881-890.]; Kumar & Ahluwalia, 2012[Kumar, A. & Ahluwalia, P. K. (2012). Mater. Chem. Phys. 135, 755-761.]), the elasticity (Feldman, 1976[Feldman, J. L. (1976). J. Phys. Chem. Solids, 37, 1141-1144.]; Peelaers & Van de Walle, 2014[Peelaers, H. & Van de Walle, C. G. (2014). J. Phys. Chem. C, 118, 12073-12076.]) and the vibrational properties (Ataca et al., 2011[Ataca, C., Topsakal, M., Aktürk, E. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 16354-16361.]; Funke et al., 2016[Funke, S., Miller, B., Parzinger, E., Thiesen, P., Holleitner, A. W. & Wurstbauer, U. (2016). J. Phys. Condens. Matter, 28, 385301.]; Wieting & Verble, 1971[Wieting, T. J. & Verble, J. L. (1971). Phys. Rev. B, 3, 4286-4292.]). Some of these reports also showed conflicting results, such as a shift of some Raman modes when MoS2 is prepared in layers.

Because of the relevance of this material in the next generation of devices, and to provide a detailed and comprehensive cross-correlated analysis of the different properties of molybdenite and its layered derivatives, we have performed a series of ab initio simulations at the density functional theory (DFT) level of the structures of the MoS2-2H bulk, the monolayer (here labelled as MoS2-1L) and the bilayer (MoS2-2L). This choice is dictated by previous theoretical analyses (Kuc et al., 2011[Kuc, A., Zibouche, N. & Heine, T. (2011). Phys. Rev. B, 83, 245213.]), which highlighted the occurrence of the direct band gap only in the monolayer, with bilayer MoS2 behaving at the electronic level similarly to the bulk mineral. In the following, an analysis of the structures and elastic, electronic, dielectric and phonon properties is provided, cross-correlated, and discussed against previous results reported in the scientific literature. Most important in our work is that all the simulations were performed with a correction to include the van der Waals interactions in the physical treatment, a non-trivial choice when dealing with layered materials such as molybdenite, other transition metal dichalcogenides and phyllosilicates (Moro et al., 2016[Moro, D., Ulian, G. & Valdrè, G. (2016). Appl. Clay Sci. 131, 175-181.]).

2. Computational methods

The present simulations have been conducted by means of the Vienna ab initio Software Package (VASP) code (Kresse & Furthmüller, 1996[Kresse, G. & Furthmüller, J. (1996). Comput. Mater. Sci. 6, 15-50.]; Kresse & Hafner, 1993[Kresse, G. & Hafner, J. (1993). Phys. Rev. B, 48, 13115-13118.]) using the density functional PBE (Perdew et al., 1996[Perdew, J. P., Burke, K. & Ernzerhof, M. (1996). Phys. Rev. Lett. 77, 3865-3868.]). The projector augmented-wave (PAW) basis set was expanded using a cut-off of 600 eV throughout the different calculations (Kresse & Joubert, 1999[Kresse, G. & Joubert, D. (1999). Phys. Rev. B, 59, 1758-1775.]). The contribution of the weak van der Waals forces was included according to the DFT-D3 approach (Grimme et al., 2011[Grimme, S., Ehrlich, S. & Goerigk, L. (2011). J. Comput. Chem. 32, 1456-1465.]),

[E_{\rm disp} = - {1 \over 2} \sum\limits_{i=1}^N \sum\limits_{j=1}^N \sum\limits_L \left [ f_{{\rm d},6} \left ( r_{ij,L} \right ) {{C_{6ij}} \over {r_{ij,L}^6}} + f_{{\rm d},8} \left ( r_{ij,L} \right ) {{C_{8ij}} \over {r_{ij,L}^8}} \right ] , \eqno(1)]

where the first two summations are over all N atoms in the cell and the third one is over the translations of the unit cell L = (l1, l2, l3). C6ij and C8ij are the dispersion coefficients of the atom pair ij, and rij,L is the distance between atom i in the reference cell (L = 0) and atom j in the cell L. The function fd,n is the Becke–Jonson damping function, which is given by the expression

[f_{{\rm d},n} = {{s_n r_{ij}^n} \over {r_{ij}^n + \left ( a_1 R_{0ij} + a_2 \right )^n}} , \eqno(2)]

with s6 = 1 and a1, a2 and s8 as variable parameters depending on the geometry of the system.

Geometry optimization of the bulk (MoS2-2H) was conducted by means of an iterative procedure involving the successive minimization of the a lattice parameter and c/a ratio, starting from the experimentally refined unit cell (Schönfeld et al., 1983[Schönfeld, B., Huang, J. J. & Moss, S. C. (1983). Acta Cryst. B39, 404-407.]). Within this procedure, the tolerance on forces was set at 10−4 eV Å−1. The molybdenite monolayer (MoS2-1L) and bilayer (MoS2-2L) structures were modelled as slabs containing just one and two structural layers, respectively, placed in a unit cell with the a axis equal to the refined a parameter of the bulk MoS2-2H. A vacuum region along the [001] direction of 20 and 30 Å thickness for MoS2-1L and MoS2-2L, respectively, was employed to avoid interaction between replicas of layers in neighbouring cells along the c axis. The sampling in the first Brillouin zone for the bulk mineral was conducted on a Monkhorst–Pack Γ-centred grid (Monkhorst & Pack, 1976[Monkhorst, H. J. & Pack, J. D. (1976). Phys. Rev. B, 13, 5188-5192.]) with 17 × 17 × 9 k-points for the geometry optimization procedure, whereas a finer mesh of 33 × 33 × 9 k-points was employed for the calculation of the electronic band structure, density of states and dielectric function. For the layered systems MoS2-1L and MoS2-2L, due to the increased c lattice parameter, the grids were reduced along the [001] direction to just one k-point, leading to 33 × 33 × 1 and 33 × 33 × 1 k-point meshes. For all the calculations, the threshold controlling the convergence on the total energy between two consecutive self-consistent field steps was set to 10−8 eV. Electronic properties (band structure and density of states) at the GW level of theory were obtained through generating maximally localized Wannier functions, which were calculated using the Wannier90 code (Pizzi et al., 2020[Pizzi, G., Vitale, V., Arita, R., Blügel, S., Freimuth, F., Géranton, G., Gibertini, M., Gresch, D., Johnson, C., Koretsune, T., Ibañez-Azpiroz, J., Lee, H., Lihm, J. M., Marchand, D., Marrazzo, A., Mokrousov, Y., Mustafa, J. I., Nohara, Y., Nomura, Y., Paulatto, L., Poncé, S., Ponweiser, T., Qiao, J. F., Thöle, F., Tsirkin, S. S., Wierzbowska, M., Marzari, N., Vanderbilt, D., Souza, I., Mostofi, A. A. & Yates, J. R. (2020). J. Phys. Condens. Matter, 32, 165902.]). The frequency-dependent dielectric function was calculated with zero momentum transfer [q = 0; q = (4π/λ)sinθ, where θ is half the scattering angle and λ is the wavelength of the incident radiation] using both the independent particle approximation (without local field effects) (Gajdoš et al., 2006[Gajdoš, M., Hummer, K., Kresse, G., Furthmüller, J. & Bechstedt, F. (2006). Phys. Rev. B, 73, 045112.]) and the random phase approximation, the latter performed on top of a single-shot G0W0 calculation (Shishkin & Kresse, 2006[Shishkin, M. & Kresse, G. (2006). Phys. Rev. B, 74, 035101.], 2007[Shishkin, M. & Kresse, G. (2007). Phys. Rev. B, 75, 235102.]). The frequency-dependent dielectric function was then compared with known measures and models reported in the scientific literature. Phonon properties were calculated using the Phonopy package (Togo & Tanaka, 2015[Togo, A. & Tanaka, I. (2015). Scr. Mater. 108, 1-5.]), calculating the Γ-point (q = 0) frequency and the phonon dispersion relations (q ≠ 0) using finite displacements and the modified Parlinski–Li–Kawazoe method (Parlinski et al., 1997[Parlinski, K., Li, Z. Q. & Kawazoe, Y. (1997). Phys. Rev. Lett. 78, 4063-4066.]) on 3 × 3 × 3 and 3 × 3 × 1 supercells for the bulk and layered molybdenite models, respectively.

3. Results and discussions

3.1. Molybdenite structure

Geometry optimization is the first step necessary prior to any further analysis to assess the quality of the simulation approach. The structural results for the different molybdenite models (bulk, monolayer and bilayer) are reported in Table 1[link], alongside previous experimental and theoretical determinations. For the bulk mineral, our simulation at the PBE-D3 level of theory is correctly able to reproduce the structural features observed from the X-ray diffraction (XRD) refinements of Bronsema et al. (1986[Bronsema, K. D., De Boer, J. L. & Jellinek, F. (1986). Z. Anorg. Allge. Chem. 540, 15-17.]). In detail, we observed a small underestimation of the a and c lattice parameters of about −0.35 and −1.74%, respectively, which is an expected result at 0 K. The bond lengths are in good agreement with the XRD results, with a slight underestimation of about −1.88 and −0.46% for the S—S and Mo—S bonds, respectively. The structural data are also in good agreement with the experimental lattice parameters obtained via scanning tunnelling microscopy, a = 3.160 Å and c = 12.294 Å (Schönfeld et al., 1983[Schönfeld, B., Huang, J. J. & Moss, S. C. (1983). Acta Cryst. B39, 404-407.]). Furthermore, our theoretical Mo—S and S—S bond lengths (2.399 and 3.149 Å, respectively, at the DFT-D3 level) are in good agreement with those measured experimentally (2.41 and 3.19 Å, respectively; Schönfeld et al., 1983[Schönfeld, B., Huang, J. J. & Moss, S. C. (1983). Acta Cryst. B39, 404-407.]).

Table 1
Calculated lattice parameters a and c (Å), S—S and Mo—S bond distances (Å), S—Mo—S bond angle Θ (°), interlayer distance dI (Å), internal parameter z (dimensionless), bulk modulus (K0, GPa), interlayer binding energy ΔEbind (meV per atom) and cohesive energy ΔEcoh (eV per atom) for bulk (MoS2-2H) and layered (MoS2-1L and MoS2-2L) structures, compared with previous theoretical and experimental (Exp.) results

The asterisk (*) marks fixed parameters in the present theoretical simulations.

System Method a c S—S Mo—S Θ dI z K0 ΔEbind ΔEcoh
Bulk (MoS2-2H) PBE-D3a 3.149 12.080 3.130 2.399 81.46 2.910 0.130 42 (2) 47 5.438
LDAb 3.125 12.137 3.12 2.39         55  
PW91b 3.215 15.540             3  
PW91-D2b 3.220 12.411 3.150 2.436 80.56     44 53 5.105
PBEc 3.18 14.68         0.143 2   5.12
PBE-D2c 3.19 12.42         0.125 39   5.37
vdW-DFd 3.23 12.6           39 60  
Exp. 3.160e 12.294e 3.190e 2.410e     0.121e 53.4f   5.18g
                       
Monolayer (MoS2-1L) PBE-D3a 3.149 20* 3.137 2.401 81.58   0.078     5.344
LDAh 3.11   3.11 2.37 81.62         6.35
PW91b 3.20   3.13 2.42 80.69         5.18
PW91-D2b 3.220   3.153 2.437 80.62         5.052
Exp. 3.20i                  
                       
Bilayer (MoS2-2L) PBE-D3a 3.149 30* 3.134 2.400 81.53 2.929 0.052   22 5.389
                     
References: (a) present work, (b) Ataca et al. (2011[Ataca, C., Topsakal, M., Aktürk, E. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 16354-16361.]), (c) Bučko et al. (2010[Bučko, T., Hafner, J., Lebègue, S. & Ángyán, J. G. (2010). J. Phys. Chem. A, 114, 11814-11824.]), (d) Rydberg et al. (2003[Rydberg, H., Dion, M., Jacobson, N., Schröder, E., Hyldgaard, P., Simak, S. I., Langreth, D. C. & Lundqvist, B. I. (2003). Phys. Rev. Lett. 91, 126402.]), (e) Bronsema et al. (1986[Bronsema, K. D., De Boer, J. L. & Jellinek, F. (1986). Z. Anorg. Allge. Chem. 540, 15-17.]), (f) Aksoy et al. (2006[Aksoy, R., Ma, Y., Selvi, E., Chyu, M. C., Ertas, A. & White, A. (2006). J. Phys. Chem. Solids, 67, 1914-1917.]), (g) Raybaud et al. (1997[Raybaud, P., Kresse, G., Hafner, J. & Toulhoat, H. (1997). J. Phys. Condens. Matter, 9, 11085-11106.]), (h) Ataca & Ciraci (2011[Ataca, C. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 13303-13311.]) and (i) Joensen et al. (1986[Joensen, P., Frindt, R. F. & Morrison, S. R. (1986). Mater. Res. Bull. 21, 457-461.]).

The binding energy between the MoS2 layers was calculated according to the following formula:

[\Delta E_{\rm bind} = 2 E_{1{\rm L}} - E_{\rm bulk} , \eqno(3)]

where Ebulk is the total energy of the bulk mineral and E1L is the energy of a single layer (the factor 2 considers the presence of two layers in the bulk unit cell). Our theoretical value (ΔEbind = 47 meV per atom = 0.456 J m−2) is in very good agreement with both the ΔEbind = 0.52 (4) J m−2 obtained by Weiss & Phillips (1976[Weiss, K. & Phillips, J. M. (1976). Phys. Rev. B, 14, 5392-5395.]) from the experimental specific surface energy of molybdenite and the binding energy value of 0.55 (13) J m−2 measured by Fang et al. (2020[Fang, Z., Li, X., Shi, W., Li, Z., Guo, Y., Chen, Q., Peng, L. & Wei, X. (2020). J. Phys. Chem. C, 124, 23419-23425.]) using an in situ peeling-to-fracture method. In the latter study, the authors performed a theoretical simulation of the mechanical exfoliation of molybdenite with the VASP code using the optB86b-vdw kernel, which is a modified version of the vdW-DF approach. The results provided ΔEbind = 0.422 J m−2, which is slightly lower than the value obtained from our simulations. We infer that this small deviation is due to the different computational settings employed by Fang et al. (2020[Fang, Z., Li, X., Shi, W., Li, Z., Guo, Y., Chen, Q., Peng, L. & Wei, X. (2020). J. Phys. Chem. C, 124, 23419-23425.]), especially the different description of the van der Waals long-range interactions.

We also calculated the cohesive energy of bulk molybdenite using the formula

[\Delta E_{\rm coh} = \left ( 2E_A^{\rm Mo} + 4E_A^{\rm S} \right ) - E_{\rm bulk} \eqno(4)]

and of the layered structures using

[\Delta E_{\rm coh} = \left ( nE_A^{\rm Mo} + 2nE_A^{\rm S} \right ) - E_{nL} , \eqno(5)]

where [E_A^{\rm Mo}] and [E_A^{\rm S}] are the atomic energy of the isolated Mo and S atoms, respectively, n is the number of layers, and EnL is the energy of the layered system. The cohesive energy (5.438 eV per atom) is in good agreement with the experimental value of 5.18 eV per atom reported by Raybaud et al. (1997[Raybaud, P., Kresse, G., Hafner, J. & Toulhoat, H. (1997). J. Phys. Condens. Matter, 9, 11085-11106.]). These results are also in line with those of Bučko et al. (2013[Bučko, T., Lebègue, S., Hafner, J. & Ángyán, J. G. (2013). J. Chem. Theory Comput. 9, 4293-4299.]), who obtained ΔEcoh = 5.37 eV per atom. As explained by those authors, this slight overestimation is mainly related to a high contribution from the correction for the dispersive forces, since the sole PBE functional provided a cohesive energy closer to the experimental value. The ΔEcoh value that we obtained from the DFT-D3 approach seems to confirm this hypothesis.

Comparison with the results using the standard PBE functional (Bučko et al., 2010[Bučko, T., Hafner, J., Lebègue, S. & Ángyán, J. G. (2010). J. Phys. Chem. A, 114, 11814-11824.]) indicates that the inclusion of dispersive forces, at least via a posteriori corrections, plays a fundamental role in determining the physical–chemical properties of MoS2-2H and other layered materials, in agreement with previous statements on the topic (Cutini et al., 2020[Cutini, M., Maschio, L. & Ugliengo, P. (2020). J. Chem. Theory Comput. 16, 5244-5252.]; Ulian et al., 2013[Ulian, G., Tosoni, S. & Valdrè, G. (2013). J. Chem. Phys. 139, 204101.], 2021[Ulian, G., Moro, D. & Valdrè, G. (2021). Phys. Chem. Chem. Phys. 23, 18899-18907.]). Indeed, the lack of an appropriate treatment results in an overly overestimated c lattice parameter and extremely low bulk modulus, as shown by Ataca et al. (2011[Ataca, C., Topsakal, M., Aktürk, E. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 16354-16361.]). The uncorrected PW91 functional provided a good description of the structural information directly related to the MoS2 layers, i.e. the a lattice parameter, but the interlayer binding energy dropped to the negligible value of 3 meV per atom (−94%), resulting in a dramatic increase of the c lattice parameter (25% with respect to the PW91-D2 results). Hence, since most of a mineral's properties (optical, electronic and so on) are strictly related to its structure, it is of the utmost importance to use a theoretical approach that considers all the relevant physical forces.

One of the first attempts to include dispersive forces in the physical description of layered structures (e.g. graphite and molybdenite) was provided by Rydberg et al. (2003[Rydberg, H., Dion, M., Jacobson, N., Schröder, E., Hyldgaard, P., Simak, S. I., Langreth, D. C. & Lundqvist, B. I. (2003). Phys. Rev. Lett. 91, 126402.]), who included these kinds of forces in a nonlocal correlation density functional called vdW-DF, coded in the PWscf program. The structural results (lattice parameters and bulk modulus) provided were in reasonably good agreement with the available experimental data (Aksoy et al., 2006[Aksoy, R., Ma, Y., Selvi, E., Chyu, M. C., Ertas, A. & White, A. (2006). J. Phys. Chem. Solids, 67, 1914-1917.]; Bronsema et al., 1986[Bronsema, K. D., De Boer, J. L. & Jellinek, F. (1986). Z. Anorg. Allge. Chem. 540, 15-17.]). Compared with the results of the present study, the vdW-DF approach provided a slightly overestimated unit-cell volume due to larger a and c parameters. More recently, in the work of Bučko et al. (2013[Bučko, T., Lebègue, S., Hafner, J. & Ángyán, J. G. (2013). J. Chem. Theory Comput. 9, 4293-4299.]), who employed the VASP code and a DFT-D2 correction for the dispersive forces, the calculated lattice parameters of the unit cell were a = 3.19 Å and c = 12.42 Å, which are slightly larger than those obtained in the present study, in particular for the c axis. This discrepancy is mainly related to the different correction for the dispersive forces, because the original D2 scheme of Grimme (2006[Grimme, S. (2006). J. Comput. Chem. 27, 1787-1799.]) is more empirical than the D3, where the C6 parameters are adjusted for each atom according to the local chemical environment (charge density) in which it is located. Also, the 8 × 8 × 8 k-point mesh employed by Bučko et al. (2013[Bučko, T., Lebègue, S., Hafner, J. & Ángyán, J. G. (2013). J. Chem. Theory Comput. 9, 4293-4299.]) has fewer sampling points than the one adopted here (17 × 17 × 9), although the authors considered a very large kinetic energy cut-off (i.e. quality of the basis set) of 1500 eV.

Our simulation results for bulk MoS2-2H are also in good agreement with previous theoretical results at the DFT level reported by Ataca et al. (2011[Ataca, C., Topsakal, M., Aktürk, E. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 16354-16361.]). In that work, the authors employed the PWscf package in QuantumEspresso (Giannozzi et al., 2009[Giannozzi, P., Baroni, S., Bonini, N., Calandra, M., Car, R., Cavazzoni, C., Ceresoli, D., Chiarotti, G. L., Cococcioni, M., Dabo, I., Dal Corso, A., de Gironcoli, S., Fabris, S., Fratesi, G., Gebauer, R., Gerstmann, U., Gougoussis, C., Kokalj, A., Lazzeri, M., Martin-Samos, L., Marzari, N., Mauri, F., Mazzarello, R., Paolini, S., Pasquarello, A., Paulatto, L., Sbraccia, C., Scandolo, S., Sclauzero, G., Seitsonen, A. P., Smogunov, A., Umari, P. & Wentzcovitch, R. M. (2009). J. Phys. Condens. Matter, 21, 395502.]), plane-wave basis sets with ultrasoft pseudo­potentials (Vanderbilt, 1990[Vanderbilt, D. (1990). Phys. Rev. B, 41, 7892-7895.]) and the PW91 density functional (Perdew et al., 1992[Perdew, J. P., Chevary, J. A., Vosko, S. H., Jackson, K. A., Pederson, M. R., Singh, D. J. & Fiolhais, C. (1992). Phys. Rev. B, 46, 6671-6687.]), corrected with the DFT-D2 scheme. In particular, the unit-cell lattice and internal geometry of bulk molybdenite obtained with our approach were within about 1% of those previously simulated, and no difference in the binding energy ΔEbind was observed. We note that the formation energy from our simulation is about 6% more than that calculated at the PW91-D2 level, which may be due to the different choice of the density functional and basis sets.

Compared with the bulk mineral, the molybdenite monolayer (MoS2-1L) and bilayer (MoS2-2L) structures showed very small variations regarding their internal geometry, with differences not higher than 0.1%. This is a typical trend in layered minerals and in materials whose structural features do not vary significantly when simulated in layers (Ulian et al., 2018[Ulian, G., Moro, D. & Valdrè, G. (2018). Compos. Struct. 202, 551-558.]). This can also be evinced from the cohesive energy of both the mono- and bilayer models, which were within 1% of that of bulk MoS2-2H. It is interesting that the difference between the cohesive energy of MoS2-1L and MoS2-2H corresponds to the binding energy per MoS2 unit, which is twice the ΔEbind per atom. The same applies to the bilayer slab of molybdenite. Finally, and as expected, the binding energy between the two layers of the molybdenite bilayer structure is 22 meV per atom = 0.213 J m−2, which is half the ΔEbind value of the bulk mineral because in the latter each MoS2 sheet interacts with one above and one below, whereas there is only a single layer-to-layer interaction in the bilayer model. The above observations and discussion are in line with the results reported by Ataca et al. (2011[Ataca, C., Topsakal, M., Aktürk, E. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 16354-16361.]), who compared the structural properties of bulk and monolayer molybdenite. However, those authors did not consider a possible bilayer structure in their simulations.

3.2. Elasticity

To provide a further assessment of the quality of our simulation approach, we calculated the second-order elastic moduli of bulk molybdenite. The elasticity of bulk molyb­denite was obtained using a finite strain approach, calculating the stiffness moduli from the stress–strain relationship (Yu et al., 2010[Yu, R., Zhu, J. & Ye, H. Q. (2010). Comput. Phys. Commun. 181, 671-675.]; Nye, 1957[Nye, J. F. (1957). Physical Properties of Crystals. Oxford University Press.]). For a hexagonal structure there are five independent elastic moduli that, according to the 6 × 6 matrix notation of Voigt, can be expressed as

[C = \left ( \matrix{C_{11} & C_{12} & C_{13} & \cdot & \cdot & \cdot \cr & C_{11} & C_{13} & \cdot & \cdot & \cdot \cr & & C_{33} & \cdot & \cdot & \cdot \cr & & & C_{44} & \cdot & \cdot \cr & & & & C_{44} & \cdot \cr & & & & & C_{66}} \right ) , \eqno(6)]

where C66 = (C11C12)/2 and the dots are null moduli. Three strain patterns were necessary to obtain the elastic constants and, for each of them, five equally spaced strain magnitudes between −0.015 and 0.015 were employed. For the calculation of the elastic moduli, the unit cell vectors a and c were oriented parallel to the x and z Cartesian axes, respectively, which is the standard crystallographic orientation of a hexagonal unit cell proposed by the Institute of Radio Engineering (Brainerd, 1949[Brainerd, J. G. (1949). Proc. IRE, 37, 1378-1395.]). The simulation results are presented in Table 2[link] and compared with previous experimental and theoretical data. Note the strong anisotropy between the C11 = 231.82 GPa and C33 = 63.47 GPa stiffness matrix components, which are related to elastic deformations along the a and c axes, respectively. As previously mentioned, this is a consequence of the different bonding scheme in molybdenite, where the Mo and S atoms are linked by strong covalent bonds and the MoS2 layers, stacked along the [001] direction, are held together by van der Waals interactions. A similar anisotropic behaviour has been observed for other layered phases as well (Gatta et al., 2015[Gatta, G. D., Lotti, P., Merlini, M., Liermann, H.-P., Lausi, A., Valdrè, G. & Pavese, A. (2015). Phys. Chem. Miner. 42, 309-318.]; Ulian et al., 2014[Ulian, G., Tosoni, S. & Valdrè, G. (2014). Phys. Chem. Miner. 41, 639-650.]).

Table 2
Elastic moduli of MoS2-2H polytype (in GPa), as obtained from different theoretical and experimental (Exp.) approaches

  PBE-D3a Exp.b HSE06-D2c PBEd Hartree–Focke
C11 231.82 238 238 211 255
C12 55.20 −54 64 49 −38
C13 10.63 23 12 3 17
C33 63.47 52 57 37 35
C44 21.35 19 18 30 15
References: (a) present work, (b) Feldman (1976[Feldman, J. L. (1976). J. Phys. Chem. Solids, 37, 1141-1144.]), (c) Peelaers & Van de Walle (2014[Peelaers, H. & Van de Walle, C. G. (2014). J. Phys. Chem. C, 118, 12073-12076.]), (d) Wei et al. (2010[Wei, L., Jun-fang, C., Qinyu, H. & Teng, W. (2010). Physica B, 405, 2498-2502.]) and (e) Alexiev et al. (2000[Alexiev, V., Prins, R. & Weber, T. (2000). Phys. Chem. Chem. Phys. 2, 1815-1827.]).

These theoretical findings are in line with the approximate values calculated from different experimental measurements (neutron dispersion curves and compressibility data from XRD refinements) by Feldman (1976[Feldman, J. L. (1976). J. Phys. Chem. Solids, 37, 1141-1144.]). Our results show an underestimation of about −2.6% for the C11 modulus and an overestimation of the C13 stiffness component of about 22%, but they nevertheless fall within the large uncertainties associated with the experimental data reported and discussed by Feldman (1976[Feldman, J. L. (1976). J. Phys. Chem. Solids, 37, 1141-1144.]). In addition, we are aware that the C12 stiffness component is positive in our work and negative in the cited study, a discrepancy caused by the different orientation of the Cartesian axes relative to the unit cell of the mineral (see above).

3.3. Electronic properties

The electronic band structure and density of states for molybdenite bulk (MoS2-2H), calculated along the KGMKHLAH path in the first Brillouin zone, are reported in Fig. 2[link], and the same results are presented in Fig. 3[link] for the MoS2-1L (monolayer) and MoS2-2L (bilayer) structures. It is possible to note from these results that there are four groups of bands in the calculated structure and density of states. The first group is related to valence bands located at low energy, between −12 and −15 eV, related mainly to the 3s orbital of the sulfur atom. The second group of valence bands, below the Fermi energy and separated from the first one by a large gap of about 7 eV, is the result of a visible hybridization of the 3p and 4d orbitals of S and Mo, respectively, with a small contribution from the 5s orbital of molybdenum around −5 eV. The third group is above the Fermi energy, representing the first range of conduction bands of molybdenite that are predominantly the result of the 4d orbitals of Mo and, to lesser extent, the 3p orbitals of S. The last group of (conduction) bands is located above ca 5 eV, given by a mixed contribution (hybridization) of the 5s and 5d states of molybdenum and the 3p orbitals of sulfur. Since the d orbitals are the main contributor to the bands located near the band gap, the bands are almost flat.

[Figure 2]
Figure 2
(a) Band structure and density of states (total and projected on the Mo and S atoms) of bulk molybdenite MoS2-2H. (b) Orbital-level details of the density of states, showing the contributions of the 3s (purple line) and 3p (green line) states of sulfur and of the 5s (red line) and 4d (blue line) states of molybdenum to the total density of states (DOS, black line).
[Figure 3]
Figure 3
(a), (c) Band structures and densities of states (total and projected on the Mo and S atoms) of molybdenite (a) monolayer MoS2-1L and (c) bilayer MoS2-2L. (b), (d) Orbital-level details of the density of states, showing the contributions of the 3s (purple line) and 3p (green line) states of sulfur and of the 5s (red line) and 4d (blue line) states of molybdenum to the total density of states (DOS, black line) of (b) monolayer MoS2-1L and (d) bilayer MoS2-2L.

While these features are shared between the different structures (bulk and monolayer), there are important differences arising from the thickness of the materials. For instance, in the bulk MoS2-2H, the maximum energy of the highest occupied valence band is located on the Γ point and the minimum energy of the lowest occupied conduction band is on a point in the KΓ path. This means that bulk molybdenite is a semiconductor with an indirect band gap of 0.79 eV. The band gap becomes a wider direct one, Eg = 1.90 eV, located on the K point for a monolayer of the mineral. However, even the bilayer MoS2 structure reverts the band gap from the direct KK′ to the indirect one observed for the bulk mineral, albeit the gap is slightly higher (1.22 eV).

The present results are in very good agreement with previous theoretical ones reported in the literature. For example, Kumar & Ahluwalia (2012[Kumar, A. & Ahluwalia, P. K. (2012). Mater. Chem. Phys. 135, 755-761.]), using Troullier–Martin norm-conserving pseudopotentials as coded in the SIESTA software, calculated a band gap for bulk molybdenite of 0.75 eV by means of the local density approximation (LDA) functional, which increased to 1.05 eV when the authors employed the DFT functional of Perdew et al. (1996[Perdew, J. P., Burke, K. & Ernzerhof, M. (1996). Phys. Rev. Lett. 77, 3865-3868.]). For the MoS2-1L monolayer the obtained values were higher, namely 1.89 and 1.55 eV at the LDA and PBE levels of theory, respectively. Ataca & Ciraci (2011[Ataca, C. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 13303-13311.]) performed similar simulations by means of the projector augmented-wave basis set, calculating the band gap of bulk MoS2 using LDA (or the generalized gradient approximation) as 0.72 eV (0.85 eV), whereas they obtained 1.87 eV (1.58 eV) for the molybdenite monolayer.

However, as also reported by Kumar & Ahluwalia (2012[Kumar, A. & Ahluwalia, P. K. (2012). Mater. Chem. Phys. 135, 755-761.]), all theoretical simulations on bulk MoS2-2H underestimated the band gap, as the experimental investigation by photocurrent spectroscopy assessed this value at 1.23 eV (Kam & Parkinson, 1982[Kam, K. K. & Parkinson, B. A. (1982). J. Phys. Chem. 86, 463-467.]). This is a common issue related to both LDA and GGA (generalized gradient approximation) DFT functionals. To overcome this drawback, in the present work the GW approximation was employed, which allows us to describe the quasiparticle electronic properties. Fig. 4[link] shows the band structures and densities of state for the MoS2-2H bulk [Figs. 4[link](a) and 4[link](d)], the MoS2-1L monolayer [Figs. 4[link](b) and 4[link](e)] and the MoS2-2L bilayer [Figs. 4[link](c) and 4[link](f)]. The calculated band gaps are 1.14 eV (indirect), 2.57 eV (direct) and 1.88 eV (indirect) for the bulk mineral, monolayer and bilayer MoS2, respectively. Hence, the GW approach provides a better description of the band gaps of the MoS2-2H bulk and the layered structures. The results for the monolayer are also in very good agreement with the theoretical results of Zibouche et al. (2021[Zibouche, N., Schlipf, M. & Giustino, F. (2021). Phys. Rev. B, 103, 125401.]), who calculated the band gap (2.68 eV) using the Sternheimer equation (Umari et al., 2009[Umari, P., Stenuit, G. & Baroni, S. (2009). Phys. Rev. B, 79, 201104.]).

[Figure 4]
Figure 4
Band structures of (a) bulk MoS2-2H, (b) monolayer MoS2-1L and (c) bilayer MoS2-2L calculated at the GW level. Details of the band gap region are reported for the three structures in panels (d), (e) and (f), respectively. The blue and red dots in the highlighted regions are the maxima of the valence and minima of the conduction bands, respectively.

3.4. Optical response function

The optical properties are due to the electronic transition from the occupied to the unoccupied state, in other words they are two-particle excitations. At the DFT level, the key quantity that is calculated is the frequency-dependent complex dielectric function ɛ(ω) = ɛ1(ω) + iɛ2(ω), with ɛ1 and ɛ2 the real and imaginary parts, respectively. This knowledge is very useful because the real component of ɛ(ω) provides information on the transmission of electromagnetic waves through the medium, whereas the ɛ2(ω) component is related to interband electronic transitions, i.e. single-particle excitations (Raether, 1980[Raether, H. (1980). Excitation of Plasmons and Interband Transitions by Electrons. Berlin, Heidelberg: Springer.]). The imaginary part of ɛ is calculated as a summation over empty states, according to the equations reported by Gajdoš et al. (2006[Gajdoš, M., Hummer, K., Kresse, G., Furthmüller, J. & Bechstedt, F. (2006). Phys. Rev. B, 73, 045112.]), whereas ɛ1(ω) was obtained using the Kramers–Kronig relation. In addition, the dielectric function can also be calculated with the electric vector oscillating either parallel or perpendicular to the c axis, because each considered phase belongs to the hexagonal system. This is calculated with the following formulae:

[\varepsilon ^\parallel (\omega) = \varepsilon_{zz} (\omega) \eqno(7)]

and

[\varepsilon ^\bot (\omega) = {{\varepsilon_{xx} (\omega) + \varepsilon_{yy} (\omega)} \over 2} . \eqno(8)]

Other properties of interest can be derived from the complex dielectric function, such as the refractive index n(ω), the extinction coefficient κ(ω), the absorption coefficient α(ω), the energy-loss function EELS(ω) and the reflectivity R(ω), according to

[n (\omega) = \left [ {{\left ( \varepsilon_1^2 + \varepsilon_2^2 \right )^{1/2} + \varepsilon_1} \over 2} \right ]^{1/2} , \eqno(9)]

[\kappa (\omega) = \left [ {{\left ( \varepsilon_1^2 + \varepsilon_2^2 \right )^{1/2} - \varepsilon_1} \over 2} \right ]^{1/2} , \eqno(10)]

[\alpha (\omega) = {{\omega (2^{1/2})} \over c} \left [ \left ( \varepsilon_1^2 + \varepsilon_2^2 \right )^{1/2} - \varepsilon_1 \right]^{1/2} , \eqno(11)]

[{\rm EELS} (\omega) = {\rm Im} \left [ - {1 \over {\varepsilon (\omega)}} \right ] = {{\varepsilon_2} \over {\varepsilon_1^2 + \varepsilon_2^2}} , \eqno(12)]

[R (\omega) = {{(n - 1)^2 + k^2} \over {(n + 1)^2 + k^2}} . \eqno(13)]

These optical properties were evaluated for each molybdenite model for the MoS2-2H, MoS2-1L and MoS2-2L structures, considering a high number of unoccupied bands to increase the accuracy of the results. Fig. 5[link] reports the dielectric function calculated in the photon energy range 0–30 eV, subdivided into real (ɛ1) and imaginary (ɛ2) parts. Bulk molybdenite [Figs. 5[link](a) and 5[link](b)] shows excellent agreement with the data measured by Beal & Hughes (1979[Beal, A. R. & Hughes, H. P. (1979). J. Phys. C Solid State Phys. 12, 881-890.]) from the reflectivity spectra of MoS2-2H. As a consequence of the experimental setup, the comparison basis is the dielectric function calculated for the electric field oscillating perpendicular to the (001) plane. In this case, the calculated optical properties were not shifted to give a better match to the experimental findings, as was done in previous work (Kumar & Ahluwalia, 2012[Kumar, A. & Ahluwalia, P. K. (2012). Mater. Chem. Phys. 135, 755-761.]).

[Figure 5]
Figure 5
Imaginary (ɛ2) and real (ɛ1) components of the complex dielectric function, as obtained from the theoretical simulations at the PBE-D3 level for (a), (b) MoS2-2H, (c), (d) MoS2-1L and (e), (f) MoS2-2L. Red and blue lines refer to the on-plane (xx, E [\bot] c) and out-of-plane (xx, E || c) electric field oscillations, respectively. Experimental data from Beal & Hughes (1979[Beal, A. R. & Hughes, H. P. (1979). J. Phys. C Solid State Phys. 12, 881-890.]) are shown for direct comparison.

Three main peaks can be observed in the low-energy region (1–5 eV) of the imaginary part ɛ2, labelled A (2.76 eV), B (4.21 eV) and C (5.31 eV) according to the nomenclature proposed by Kumar & Ahluwalia (2012[Kumar, A. & Ahluwalia, P. K. (2012). Mater. Chem. Phys. 135, 755-761.]). The A peak corresponds to the C exciton (Funke et al., 2016[Funke, S., Miller, B., Parzinger, E., Thiesen, P., Holleitner, A. W. & Wurstbauer, U. (2016). J. Phys. Condens. Matter, 28, 385301.]). There is good agreement with the theoretical data reported by Kumar and Ahluwalia, who calculated A = 2.9 eV, B = 4.5 eV and C = 5.1 eV, but our computational setting allows us to detect the exciton at 1.72 eV, corresponding to the KK′ transition from the topmost valence (v1) band to the first conduction band (c1), marked with a black arrow in Fig. 5[link](a). However, the intensity of the signal is lower than the experimental one, and the second expected excitonic transition c2–v1 is not visible because of the overlap with the A peak. The present data are also in line with thermal ellipsometry data at 35 K measured by Le et al. (2019[Le, V. L., Kim, T. J., Park, H. G., Nguyen, H. T., Nguyen, X. A. & Kim, Y. D. (2019). Curr. Appl. Phys. 19, 182-187.]), who measured A = 2.91 eV and the KK′ transition at 1.96 eV.

MoS2-1L presents four peaks in the imaginary part of the complex dielectric function, labelled A (2.82 eV), B (3.70 eV), C (4.39 eV) and D (5.54 eV) in Fig. 5[link](c), and they are in agreement with the results of Kumar & Ahluwalia (2012[Kumar, A. & Ahluwalia, P. K. (2012). Mater. Chem. Phys. 135, 755-761.]), i.e. A = 2.9 eV, B = 3.8 eV, C = 4.5 eV and D = 5.5 eV. In this case, the B peak is a new feature that appears because of the single-layer structure, whereas the other signals (A, C and D) are slightly shifted at higher energy with respect to the A, B and C peaks of the mineral bulk. A single excitonic KK′ transition is also clearly visible at about 1.94 eV, in excellent agreement with that found at 1.9 eV by Funke et al. (2016[Funke, S., Miller, B., Parzinger, E., Thiesen, P., Holleitner, A. W. & Wurstbauer, U. (2016). J. Phys. Condens. Matter, 28, 385301.]), which was derived from spectroscopic ellipsometry measurements. A second peak at 2.05 eV was also measured by the same authors, which is related to spin–orbit coupling, i.e. a splitting of the topmost valence band in the single-layered MoS2 unit at about 140–150 meV. An example of this effect on the bands of 2D molybdenite can be seen in the experimental and theoretical work of Song et al. (2019[Song, B., Gu, H., Fang, M., Chen, X., Jiang, H., Wang, R., Zhai, T., Ho, Y. T. & Liu, S. (2019). Adv. Opt. Mater. 7, 1801250.]) and Moynihan et al. (2020[Moynihan, E., Rost, S., O'Connell, E., Ramasse, Q., Friedrich, C. & Bangert, U. (2020). J. Microsc. 279, 256-264.]). Since spin–orbit coupling was not included in the present simulations, this second peak (KK′ transition) is not visible in the dielectric function. Funke et al. (2016[Funke, S., Miller, B., Parzinger, E., Thiesen, P., Holleitner, A. W. & Wurstbauer, U. (2016). J. Phys. Condens. Matter, 28, 385301.]) found at about 3 eV the excitonic transition here labelled as A, in line with our calculations.

When the molybdenite structure is made of two layers, the shape of the dielectric function reverts to a form similar to that of bulk MoS2-2H [see Fig. 5[link](e)], presenting only three major peaks, A (2.73 eV), B (4.32 eV) and C (5.36 eV), in the range 0–5 eV.

In general, all the observed optical transitions occur in an energy range corresponding to electronic transitions between the p valence (occupied) bands of sulfur and the d conduction (unoccupied) bands of molybdenum (see Fig. 3[link]), as also observed in previous work (Beal & Hughes, 1979[Beal, A. R. & Hughes, H. P. (1979). J. Phys. C Solid State Phys. 12, 881-890.]; Kumar & Ahluwalia, 2012[Kumar, A. & Ahluwalia, P. K. (2012). Mater. Chem. Phys. 135, 755-761.]). According to the literature, the A peak in Figs. 5[link](a), 5[link](c) and 5[link](d) is mainly due to nearly parallel bands in the MΓ path, i.e. band nesting. The low-energy excitons are instead related to electronic transitions from the Mo dz orbital, which is split at the K point in the Brillouin zone because of spin–orbit interaction.

The real part of the dielectric function of the MoS2-2H structure is also in very good agreement with the experimental findings (Beal & Hughes, 1979[Beal, A. R. & Hughes, H. P. (1979). J. Phys. C Solid State Phys. 12, 881-890.]; Funke et al., 2016[Funke, S., Miller, B., Parzinger, E., Thiesen, P., Holleitner, A. W. & Wurstbauer, U. (2016). J. Phys. Condens. Matter, 28, 385301.]), with values at zero energy of [\varepsilon_1^\parallel] = 10.42 eV and [\varepsilon_1^\bot] = 16.43 eV that are in line with those measured from experiments, e.g. [\varepsilon_1^\parallel] = 10.42 eV and [\varepsilon_1^\bot] = 16.8 eV (Beal & Hughes, 1979[Beal, A. R. & Hughes, H. P. (1979). J. Phys. C Solid State Phys. 12, 881-890.]). These values are reduced when going from bulk to the monolayer ([\varepsilon_1^\parallel] = 3.29 eV and [\varepsilon_1^\bot] = 5.49 eV) and the bilayer ([\varepsilon_1^\parallel] = 4.53 eV and [\varepsilon_1^\bot] = 7.25 eV), showing a trend that increases with the number of MoS2 units in the structure. Our results are also in line with the theoretical simulations of Kumar & Ahluwalia (2012[Kumar, A. & Ahluwalia, P. K. (2012). Mater. Chem. Phys. 135, 755-761.]), who calculated [\varepsilon_1^\parallel] = 8.9 eV and [\varepsilon_1^\bot] = 12.8 eV for the MoS2-2H bulk, and [\varepsilon_1^\parallel] = 3.0 eV and [\varepsilon_1^\bot] = 4.8 eV for the molybdenite monolayer. A better agreement can be observed by comparing our results with the theoretical ones of Ben Amara et al. (2016[Ben Amara, I., Ben Salem, E. & Jaziri, S. (2016). J. Appl. Phys. 120, 051707.]), who obtained [\varepsilon_1^\parallel] = 8.3 eV and [\varepsilon_1^\bot] = 15.4 eV from PBE simulations.

The calculated electron energy-loss spectroscopy (EELS) spectra (Fig. 6[link]) provide further information on the plasmon modes, i.e. the collective oscillation of valence or conduction electrons in a material. These modes are related to transitions from the π and σ bonding states to the respective antibonding states π* and σ*, which occur in the EELS spectra as low-energy π and high-energy π+σ plasmons (interband plasmons). Such plasmon modes are present in the EELS spectrum when the real part of the dielectric function ɛ1(ω) crosses zero with a positive slope. In bulk molybdenite (MoS2-2H), the π plasmon falls at 8.52 eV, whereas the π+σ one can be seen at 22.55 eV for E [\bot] c. For an electric field oscillating parallel to the c axis, the π+σ plasmon presents two peaks at 22.98 and 23.17 eV, and no other well defined signals are visible in the spectrum. A similar trend is observed for layered MoS2, but both the monolayer and bilayer systems show a red shift of these plasmon signals (for E [\bot] c), at 8.06 eV (π) and 16.17 eV (π+σ) for MoS2-1L, and 8.31 eV (π) and 17.44 eV (π+σ) for the MoS2-2L model. For the electric field parallel to the c axis, the π+σ plasmon falls at 15.55 and 17.01 eV for MoS2-1L and MoS2-2L, respectively. These theoretical results are in very good agreement with the experimental EELS measurements conducted via scanning transmission electron microscopy (STEM) by Moynihan et al. (2020[Moynihan, E., Rost, S., O'Connell, E., Ramasse, Q., Friedrich, C. & Bangert, U. (2020). J. Microsc. 279, 256-264.]), who obtained for MoS2-2H 8.6 and 23 eV for the π and π+σ plasmons, respectively. Compared with the previous DFT simulations at the PBE level (Kumar & Ahluwalia, 2012[Kumar, A. & Ahluwalia, P. K. (2012). Mater. Chem. Phys. 135, 755-761.]), our results are in general agreement, with the π plasmon falling at lower energies than previously reported, i.e. 9.2 and 8.6 eV for the bulk and monolayer structures, respectively.

[Figure 6]
Figure 6
Calculated EELS and optical absorption α for (a), (b) MoS2-2H, (c), (d) MoS2-1L and (e), (f) MoS2-2L (ɛ2). Red and blue lines refer to the on-plane (xx, E [\bot] c) and out-of-plane (xx, E || c) electric field oscillations, respectively. Experimental data from Beal & Hughes (1979[Beal, A. R. & Hughes, H. P. (1979). J. Phys. C Solid State Phys. 12, 881-890.]) are shown for direct comparison. The green dashed boxes in panels (b), (d) and (f) show the part of the absorption spectrum in the visible region, which is reported in the upper right inset.

The optical absorption α is also reported in Fig. 6[link], in the photon energy range 0–30 eV. The simulation results for bulk molybdenite [Fig. 6[link](b)] are in excellent agreement with the experimental ones of Beal & Hughes (1979[Beal, A. R. & Hughes, H. P. (1979). J. Phys. C Solid State Phys. 12, 881-890.]) up to about 15 eV, and then there is some deviation between the curves due to the data extrapolation performed by those authors. In the region of the visible spectrum [Vis, 380–780 nm, see the inset in Fig. 6[link](b)], MoS2-2H strongly absorbs violet light, with two peaks between 400 and 450 nm. Experimentally, two absorption peaks were found at 614 and 670 nm related to the KK′ transitions, whereas a single almost flat signal centred at 662 nm was obtained from the simulations. Similar general figures were found for the MoS2-1L [Fig. 6[link](d)] and MoS2-2L [Fig. 6[link](f)] systems, with absorption peaks in the red (ca 650 nm) and violet regions (about 400 nm) of the visible spectrum.

3.5. Phonon properties

Bulk molybdenite (space group P63/mmc, point group D6h4) has six atoms in the unit cell, resulting in 18 degrees of freedom that, in the Γ point and from group-theory analysis, have the following irreducible representations (irreps):

[\Gamma = A_{1g} \otimes 2A_{2u} \otimes 2B_{1g} \otimes B_{1u} \otimes E_{2u} \otimes 2E_{2g} \otimes 2E_{1u} \otimes E_{1g} , \eqno(14)]

with Γa = [A_{2u} \otimes E_{1u}] and Γo = [A_{1g} \otimes A_{2u}] [\otimes \ 2B_{1g}] [\otimes \ B_{1u}] [\otimes \ E_{2u}] [\otimes \ 2E_{2g}] [\otimes \ E_{1u}] [\otimes \ E_{1g}] being the acoustic and optical modes, respectively. Modes with character A or B are non-degenerate, whereas the E modes are doubly degenerate. Vibrational motions that are symmetric (gerade, g) and anti-symmetric (ungerade, u) with respect to the inversion centre of the crystal are active in Raman and infrared spectroscopies, respectively.

The single layer of MoS2 ([P {\overline 6}m2] space group, D3h point group) has only nine degrees of freedom and different zone-centre irreps,

[\Gamma = A'_1 \otimes 2A''_2 \otimes 2E' \otimes E'' , \eqno(15)]

with Γa = [A''_2 \otimes E'] and Γo = [A'_1 \otimes A''_2 \otimes E' \otimes E'']. If we consider a correlation with the D6h point group, it is possible to associate the E′′ and [A'_1] modes of the monolayer with the E1g and A1g modes of the bulk, respectively. Instead, while the bilayer MoS2-2L has the same degrees of freedom as the mineral bulk, the [P {\overline 3}m1] space group ( D3d3 point group) leads to different irreps,

[\Gamma = 3A_{1g} \otimes 3A_{2u} \otimes 3E_u \otimes 3E_g , \eqno(16)]

where Γa = [A_{2u} \otimes E_u] and Γo = [3A_{1g} \otimes 2A_{2u} \otimes 2E_u \otimes 3E_g]. Again, gerade (A1g and Eg) and ungerade modes (A2u and Eu) are active in Raman and IR, respectively.

Γ-point frequencies are reported in Table 3[link], alongside selected results from both experiments (Funke et al., 2016[Funke, S., Miller, B., Parzinger, E., Thiesen, P., Holleitner, A. W. & Wurstbauer, U. (2016). J. Phys. Condens. Matter, 28, 385301.]; Wieting & Verble, 1971[Wieting, T. J. & Verble, J. L. (1971). Phys. Rev. B, 3, 4286-4292.]) and theoretical simulations (Ataca et al., 2011[Ataca, C., Topsakal, M., Aktürk, E. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 16354-16361.]; Jiménez Sandoval et al., 1991[Jiménez Sandoval, S., Yang, D., Frindt, R. F. & Irwin, J. C. (1991). Phys. Rev. B, 44, 3955-3962.]). The phonon dispersion curves for each model along the KΓMKH path in the first Brillouin zone are shown in Fig. 7[link]. No negative value of the phonon branch was found, meaning that the bulk and the free-standing layered structures are stable. In general, there is very good agreement between the different results, with mean absolute differences less than 3 cm−1. Two Raman modes are quite sensitive to the number of layers in the structure, i.e. the No. 7 E2g and No. 10 A1g modes in MoS2-2H, the former decreasing and the latter increasing with the number of layers of the 2D material, in agreement with previous experimental observations (Funke et al., 2016[Funke, S., Miller, B., Parzinger, E., Thiesen, P., Holleitner, A. W. & Wurstbauer, U. (2016). J. Phys. Condens. Matter, 28, 385301.]; Lee et al., 2010[Lee, C., Yan, H., Brus, L. E., Heinz, T. F., Hone, J. & Ryu, S. (2010). ACS Nano, 4, 2695-2700.]). This is an anomalous behaviour that deviates from the classical model of coupled harmonic oscillators, where it is expected that the cited E2g and A1g modes should both stiffen on increasing the thickness of the molybdenite. This is due to the stacking of the MoS2 layers along the c axis, which affects the intralayer bonding (see Table 1[link]) and hence the force constants of the vibrational motion of the atoms. Interestingly, according to Wieting & Verble (1971[Wieting, T. J. & Verble, J. L. (1971). Phys. Rev. B, 3, 4286-4292.]), the frequency difference between the E1u and E2g modes in the bulk is associated with the interlayer interaction, which is quite low in our simulations (no more than 3 cm−1). This extends to the bilayer model MoS2-2L, considering the Eu and Eg vibrations. However, while the absolute frequency shift of the A1g[A_1 '] mode is quite in line with the experimental one (ca 4 cm−1) of Lee et al. (2010[Lee, C., Yan, H., Brus, L. E., Heinz, T. F., Hone, J. & Ryu, S. (2010). ACS Nano, 4, 2695-2700.]), the E2gE′ mode is much less affected by the thickness of the material, in agreement with previous theoretical observations (Ataca et al., 2011[Ataca, C., Topsakal, M., Aktürk, E. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 16354-16361.]). We suppose that this is due to the absence of any interaction of the MoS2 monolayer or bilayer with the support (e.g. SiO2 or sapphire) that is commonly employed in experimental measurements.

Table 3
Zone centre vibrational frequencies ν (cm−1) of the bulk (MoS2-2H), monolayer (MoS2-1L) and bilayer (MoS2-2L) models, as obtained from PBE-D3 simulations, compared with experimental results (Exp.) where these are available

Activity in infrared and/or Raman spectroscopy is indicated with IR or R, respectively.

System Mode Irrepa νa IR/R Exp.b Exp.c VFFd PW91+De
MoS2-2H 1 E1u 0.0          
2 A2u 0.0          
3 E2g 36.0 R   32    
4 B1g 62.4 R        
5 E2u 282.1 IR       285.0
6 E1g 285.3 R   287   286.6
7 E2g 381.0 R 383.5 383   378.5
8 E1u 381.3 IR       378.8
9 B2u 402.5 IR       395.4
10 A1g 411.7 R 408.9 409   400.2
11 A2u 464.5 IR       456.9
12 B1g 470.1 R       460.3
                 
MoS2-1L 1 A2 " 0.0          
2 E 0.0          
3 E′′ 281.9 R     280 287.1
4 E 381.8 IR, R 384.4   384 380.2
5 [A_1 '] 402.3 R 403.1   407 406.1
6 A2 " 467.7 IR     481 465.0
                 
MoS2-2L 1 Eu 0.0          
2 A2u 0.0          
3 Eg 24.8 R        
4 A1g 42.6 R        
5 Eu 281.9 IR        
6 Eg 283.5 R        
7 Eg 381.3 R 383.9      
8 Eu 381.4 IR        
9 A2u 402.3 IR        
10 A1g 406.5 R 404.6      
11 A2u 466.8 IR        
12 A1g 468.8 R        
References: (a) present work, (b) Funke et al. (2016[Funke, S., Miller, B., Parzinger, E., Thiesen, P., Holleitner, A. W. & Wurstbauer, U. (2016). J. Phys. Condens. Matter, 28, 385301.]), (c) Wieting & Verble (1971[Wieting, T. J. & Verble, J. L. (1971). Phys. Rev. B, 3, 4286-4292.]), (d) Jiménez Sandoval et al. (1991[Jiménez Sandoval, S., Yang, D., Frindt, R. F. & Irwin, J. C. (1991). Phys. Rev. B, 44, 3955-3962.]) using the valence force field (VFF) model and (e) Ataca et al. (2011[Ataca, C., Topsakal, M., Aktürk, E. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 16354-16361.]) using the DFT/PW91-D approach.
[Figure 7]
Figure 7
Phonon band structures and atom-projected phonon densities of states for (a) bulk MoS2-2H, (b) single-layer MoS2-1L and (c) bilayer MoS2-2L structures.

Regarding the phonon dispersion (Fig. 7[link]), there is a satisfactory agreement with previous inelastic neutron scattering data measured by Wakabayashi et al. (1975[Wakabayashi, N., Smith, H. G. & Nicklow, R. M. (1975). Phys. Rev. B, 12, 659-663.]) along the ΓM path ([100] direction) for MoS2-2H. No significant differences in the phonon density of states were observed in our simulations by changing the number of layers in the system.

Quite good agreement was found by comparing the present results with those of Ataca et al. (2011[Ataca, C., Topsakal, M., Aktürk, E. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 16354-16361.]), who performed PW91-D simulations for the bulk and monolayer of molybdenite for both the phonon branches and the density of states. Their GGA+D approach resulted in E1g = 286.6 cm−1, E2g = 378.5 cm−1 and A1g = 400.2 cm−1 for MoS2-2H, and E′ = 380.2 cm−1 and [A'_1] = 406.1 cm−1 for MoS2-1L. It is worth noting the opposite behaviour of the A1g[A_1 '] frequency shift, which is explained by the different unit-cell parameters between theory and experiment.

4. Conclusions

In this work, with the density functional theory method, using plane-wave basis sets, the PBE functional and the DFT-D3 scheme to include van der Waals interactions, we have investigated different properties of bulk molybdenite MoS2-2H and its monolayer (MoS2-1L) and bilayer (MoS2-2L) structures. The scope was to provide a detailed cross-correlated analysis of these materials on the atomic scale, both to increase the knowledge of them and as a comparison basis for future work.

We have noted subtle variations in the crystal chemistry of MoS2 when simulated in layers, in particular the bond distances and angles, with binding energy values in agreement with experimental measurements that employed the peeling-to-fracture method. The selected PBE-D3 approach also described the structure of bulk MoS2-2H, with the expected unit-cell volume at 0 K smaller than that from room-temperature XRD refinements. The stiffness of bulk MoS2-2H was also in line with the only experimental work reported in the literature, with a slight overestimation of the C33 elastic modulus and an underestimation of the C13 one as a result of the approximations introduced in the simulations, chiefly the lack of thermal effects (athermal results at 0 K without any zero-point contribution).

The band gaps of bulk molybdenite, MoS2-1L and MoS2-2L calculated at the GGA (GW) level were an indirect 0.79 eV (1.14 eV), a direct 1.90 eV (2.57 eV) and an indirect 1.22 eV (1.88 eV), respectively, in line with both experimental data and theoretical simulations that considered quasi-particle excitations. We observed from the atom-projected density of states that the 3p orbitals of S hybridize with the 4d ones of Mo at the top of the valence band and at the bottom of the conduction band, whereas the core states are due to the 3s orbitals of the sulfide ions. We expect that hybrid functionals, such as HSE06, could perform similarly but with much higher computational costs, especially when using plane-wave basis sets.

The calculated complex dielectric function showed the direct (zero-momentum q) electronic transitions (imaginary part) and the plasmonic resonances (EELS spectrum) in the photon energy range 0–30 eV, which agree with recent experimental measurements. Different absorption peaks in the visible portion of the light spectrum were observed according to the number of layers of molybdenite.

Finally, we studied the IR and Raman vibrations at the zone centre and the phonon dispersion relations of the three MoS2 models (bulk, monolayer and bilayer). Our simulations confirm the experimental evidence that the No. 7 E2g mode decreases and the No. 10 A1g mode increases with the number of layers of the bidimensional material. However, the extent of the former vibration is less than expected because in our setting the mono- and bilayer are simulated in a vacuum, without any interaction with typical substrates used in experimental work, such as silica or sapphire, that could affect the force constant of these modes.

Funding information

Funding for this research was provided by Università di Bologna.

References

First citationAksoy, R., Ma, Y., Selvi, E., Chyu, M. C., Ertas, A. & White, A. (2006). J. Phys. Chem. Solids, 67, 1914–1917.  Web of Science CrossRef CAS Google Scholar
First citationAlexiev, V., Prins, R. & Weber, T. (2000). Phys. Chem. Chem. Phys. 2, 1815–1827.  Web of Science CrossRef CAS Google Scholar
First citationAminoff, G. & Broome, B. (1935). Z. Kristallogr. 91, 77–94.   CrossRef CAS Google Scholar
First citationAtaca, C. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 13303–13311.  Web of Science CrossRef CAS Google Scholar
First citationAtaca, C., Topsakal, M., Aktürk, E. & Ciraci, S. (2011). J. Phys. Chem. C, 115, 16354–16361.  Web of Science CrossRef CAS Google Scholar
First citationBeal, A. R. & Hughes, H. P. (1979). J. Phys. C Solid State Phys. 12, 881–890.  CrossRef CAS Web of Science Google Scholar
First citationBen Amara, I., Ben Salem, E. & Jaziri, S. (2016). J. Appl. Phys. 120, 051707.  Web of Science CrossRef Google Scholar
First citationBrainerd, J. G. (1949). Proc. IRE, 37, 1378–1395.  Google Scholar
First citationBronsema, K. D., De Boer, J. L. & Jellinek, F. (1986). Z. Anorg. Allge. Chem. 540, 15–17.  CrossRef ICSD Google Scholar
First citationBučko, T., Hafner, J., Lebègue, S. & Ángyán, J. G. (2010). J. Phys. Chem. A, 114, 11814–11824.  Web of Science PubMed Google Scholar
First citationBučko, T., Lebègue, S., Hafner, J. & Ángyán, J. G. (2013). J. Chem. Theory Comput. 9, 4293–4299.  Web of Science PubMed Google Scholar
First citationCutini, M., Maschio, L. & Ugliengo, P. (2020). J. Chem. Theory Comput. 16, 5244–5252.  Web of Science CrossRef CAS PubMed Google Scholar
First citationFang, Z., Li, X., Shi, W., Li, Z., Guo, Y., Chen, Q., Peng, L. & Wei, X. (2020). J. Phys. Chem. C, 124, 23419–23425.  Web of Science CrossRef CAS Google Scholar
First citationFeldman, J. L. (1976). J. Phys. Chem. Solids, 37, 1141–1144.  CrossRef CAS Web of Science Google Scholar
First citationFontana, M., Deppe, T., Boyd, A. K., Rinzan, M., Liu, A. Y., Paranjape, M. & Barbara, P. (2013). Sci. Rep. 3, 1634.  Web of Science CrossRef PubMed Google Scholar
First citationFunke, S., Miller, B., Parzinger, E., Thiesen, P., Holleitner, A. W. & Wurstbauer, U. (2016). J. Phys. Condens. Matter, 28, 385301.  Web of Science CrossRef PubMed Google Scholar
First citationGajdoš, M., Hummer, K., Kresse, G., Furthmüller, J. & Bechstedt, F. (2006). Phys. Rev. B, 73, 045112.  Google Scholar
First citationGatta, G. D., Lotti, P., Merlini, M., Liermann, H.-P., Lausi, A., Valdrè, G. & Pavese, A. (2015). Phys. Chem. Miner. 42, 309–318.  Web of Science CrossRef CAS Google Scholar
First citationGiannozzi, P., Baroni, S., Bonini, N., Calandra, M., Car, R., Cavazzoni, C., Ceresoli, D., Chiarotti, G. L., Cococcioni, M., Dabo, I., Dal Corso, A., de Gironcoli, S., Fabris, S., Fratesi, G., Gebauer, R., Gerstmann, U., Gougoussis, C., Kokalj, A., Lazzeri, M., Martin-Samos, L., Marzari, N., Mauri, F., Mazzarello, R., Paolini, S., Pasquarello, A., Paulatto, L., Sbraccia, C., Scandolo, S., Sclauzero, G., Seitsonen, A. P., Smogunov, A., Umari, P. & Wentzcovitch, R. M. (2009). J. Phys. Condens. Matter, 21, 395502.  Web of Science CrossRef PubMed Google Scholar
First citationGrimme, S. (2006). J. Comput. Chem. 27, 1787–1799.  Web of Science CrossRef PubMed CAS Google Scholar
First citationGrimme, S., Ehrlich, S. & Goerigk, L. (2011). J. Comput. Chem. 32, 1456–1465.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHess, F. L. (1924). Molybdenum Deposits: A Short Review. Bulletin 761. United States Geological Survey. Washington, DC: Department of the Interior.  Google Scholar
First citationJiménez Sandoval, S., Yang, D., Frindt, R. F. & Irwin, J. C. (1991). Phys. Rev. B, 44, 3955–3962.  Google Scholar
First citationJoensen, P., Frindt, R. F. & Morrison, S. R. (1986). Mater. Res. Bull. 21, 457–461.  CrossRef CAS Web of Science Google Scholar
First citationKam, K. K. & Parkinson, B. A. (1982). J. Phys. Chem. 86, 463–467.  CrossRef CAS Web of Science Google Scholar
First citationKresse, G. & Furthmüller, J. (1996). Comput. Mater. Sci. 6, 15–50.  CrossRef CAS Web of Science Google Scholar
First citationKresse, G. & Hafner, J. (1993). Phys. Rev. B, 48, 13115–13118.  CrossRef CAS Web of Science Google Scholar
First citationKresse, G. & Joubert, D. (1999). Phys. Rev. B, 59, 1758–1775.  Web of Science CrossRef CAS Google Scholar
First citationKropschot, S. J. (2010). Molybdenum – A Key Component of Metal Alloys. Fact Sheet 2009-3106. Reston: US Geological Survey.  Google Scholar
First citationKuc, A., Zibouche, N. & Heine, T. (2011). Phys. Rev. B, 83, 245213.  Web of Science CrossRef Google Scholar
First citationKumar, A. & Ahluwalia, P. K. (2012). Mater. Chem. Phys. 135, 755–761.  Web of Science CrossRef CAS Google Scholar
First citationLe, V. L., Kim, T. J., Park, H. G., Nguyen, H. T., Nguyen, X. A. & Kim, Y. D. (2019). Curr. Appl. Phys. 19, 182–187.  Web of Science CrossRef Google Scholar
First citationLee, C., Yan, H., Brus, L. E., Heinz, T. F., Hone, J. & Ryu, S. (2010). ACS Nano, 4, 2695–2700.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMak, K. F. & Shan, J. (2016). Nat. Photon. 10, 216–226.  Web of Science CrossRef CAS Google Scholar
First citationMartin, J. M., Donnet, C., Le Mogne, T. & Epicier, T. (1993). Phys. Rev. B, 48, 10583–10586.  CrossRef CAS Web of Science Google Scholar
First citationMartin, J. M., Pascal, H., Donnet, C., Le Mogne, T., Loubet, J. L. & Epicier, T. (1994). Surf. Coat. Technol. 68–69, 427–432.  CrossRef Web of Science Google Scholar
First citationMonkhorst, H. J. & Pack, J. D. (1976). Phys. Rev. B, 13, 5188–5192.  CrossRef Web of Science Google Scholar
First citationMoro, D., Ulian, G. & Valdrè, G. (2016). Appl. Clay Sci. 131, 175–181.  Web of Science CrossRef CAS Google Scholar
First citationMoynihan, E., Rost, S., O'Connell, E., Ramasse, Q., Friedrich, C. & Bangert, U. (2020). J. Microsc. 279, 256–264.  Web of Science CrossRef CAS PubMed Google Scholar
First citationNye, J. F. (1957). Physical Properties of Crystals. Oxford University Press.  Google Scholar
First citationParlinski, K., Li, Z. Q. & Kawazoe, Y. (1997). Phys. Rev. Lett. 78, 4063–4066.  CrossRef CAS Web of Science Google Scholar
First citationPeelaers, H. & Van de Walle, C. G. (2014). J. Phys. Chem. C, 118, 12073–12076.  Web of Science CrossRef CAS Google Scholar
First citationPerdew, J. P., Burke, K. & Ernzerhof, M. (1996). Phys. Rev. Lett. 77, 3865–3868.  CrossRef PubMed CAS Web of Science Google Scholar
First citationPerdew, J. P., Chevary, J. A., Vosko, S. H., Jackson, K. A., Pederson, M. R., Singh, D. J. & Fiolhais, C. (1992). Phys. Rev. B, 46, 6671–6687.  CrossRef CAS Web of Science Google Scholar
First citationPizzi, G., Vitale, V., Arita, R., Blügel, S., Freimuth, F., Géranton, G., Gibertini, M., Gresch, D., Johnson, C., Koretsune, T., Ibañez-Azpiroz, J., Lee, H., Lihm, J. M., Marchand, D., Marrazzo, A., Mokrousov, Y., Mustafa, J. I., Nohara, Y., Nomura, Y., Paulatto, L., Poncé, S., Ponweiser, T., Qiao, J. F., Thöle, F., Tsirkin, S. S., Wierzbowska, M., Marzari, N., Vanderbilt, D., Souza, I., Mostofi, A. A. & Yates, J. R. (2020). J. Phys. Condens. Matter, 32, 165902.  Web of Science CrossRef PubMed Google Scholar
First citationRadisavljevic, B., Radenovic, A., Brivio, J., Giacometti, V. & Kis, A. (2011). Nat. Nanotech. 6, 147–150.  Web of Science CrossRef CAS Google Scholar
First citationRaether, H. (1980). Excitation of Plasmons and Interband Transitions by Electrons. Berlin, Heidelberg: Springer.  Google Scholar
First citationRaybaud, P., Kresse, G., Hafner, J. & Toulhoat, H. (1997). J. Phys. Condens. Matter, 9, 11085–11106.  Web of Science CrossRef CAS Google Scholar
First citationRydberg, H., Dion, M., Jacobson, N., Schröder, E., Hyldgaard, P., Simak, S. I., Langreth, D. C. & Lundqvist, B. I. (2003). Phys. Rev. Lett. 91, 126402.  Web of Science CrossRef PubMed Google Scholar
First citationSchönfeld, B., Huang, J. J. & Moss, S. C. (1983). Acta Cryst. B39, 404–407.  CrossRef ICSD Web of Science IUCr Journals Google Scholar
First citationSebastian, A., Pendurthi, R., Choudhury, T. H., Redwing, J. M. & Das, S. (2021). Nat. Commun. 12, 693.  Web of Science CrossRef PubMed Google Scholar
First citationShishkin, M. & Kresse, G. (2006). Phys. Rev. B, 74, 035101.  Web of Science CrossRef Google Scholar
First citationShishkin, M. & Kresse, G. (2007). Phys. Rev. B, 75, 235102.  Web of Science CrossRef Google Scholar
First citationSong, B., Gu, H., Fang, M., Chen, X., Jiang, H., Wang, R., Zhai, T., Ho, Y. T. & Liu, S. (2019). Adv. Opt. Mater. 7, 1801250.  Web of Science CrossRef Google Scholar
First citationTogo, A. & Tanaka, I. (2015). Scr. Mater. 108, 1–5.  Web of Science CrossRef CAS Google Scholar
First citationUlian, G., Moro, D. & Valdrè, G. (2018). Compos. Struct. 202, 551–558.  Web of Science CrossRef Google Scholar
First citationUlian, G., Moro, D. & Valdrè, G. (2021). Phys. Chem. Chem. Phys. 23, 18899–18907.  Web of Science CrossRef CAS PubMed Google Scholar
First citationUlian, G., Tosoni, S. & Valdrè, G. (2013). J. Chem. Phys. 139, 204101.  Web of Science CrossRef PubMed Google Scholar
First citationUlian, G., Tosoni, S. & Valdrè, G. (2014). Phys. Chem. Miner. 41, 639–650.  Web of Science CrossRef ICSD CAS Google Scholar
First citationUmari, P., Stenuit, G. & Baroni, S. (2009). Phys. Rev. B, 79, 201104.  Web of Science CrossRef Google Scholar
First citationVanderbilt, D. (1990). Phys. Rev. B, 41, 7892–7895.  CrossRef CAS Web of Science Google Scholar
First citationWakabayashi, N., Smith, H. G. & Nicklow, R. M. (1975). Phys. Rev. B, 12, 659–663.  CrossRef CAS Web of Science Google Scholar
First citationWang, Y., Tang, H., Xie, Y., Chen, X., Ma, S., Sun, Z., Sun, Q., Chen, L., Zhu, H., Wan, J., Xu, Z., Zhang, D. W., Zhou, P. & Bao, W. (2021). Nat. Commun. 12, 3347.  Web of Science CrossRef PubMed Google Scholar
First citationWei, L., Jun-fang, C., Qinyu, H. & Teng, W. (2010). Physica B, 405, 2498–2502.  Web of Science CrossRef Google Scholar
First citationWeiss, K. & Phillips, J. M. (1976). Phys. Rev. B, 14, 5392–5395.  CrossRef CAS Web of Science Google Scholar
First citationWieting, T. J. & Verble, J. L. (1971). Phys. Rev. B, 3, 4286–4292.  CrossRef Web of Science Google Scholar
First citationWu, S., Meng, Z., Tao, X. & Wang, Z. (2022). Friction, 10, 209–216.  Web of Science CrossRef CAS Google Scholar
First citationYang, X. & Li, B. (2020). Nanophotonics, 9, 1557–1577.  Web of Science CrossRef CAS Google Scholar
First citationYu, R., Zhu, J. & Ye, H. Q. (2010). Comput. Phys. Commun. 181, 671–675.  Web of Science CrossRef CAS Google Scholar
First citationZibouche, N., Schlipf, M. & Giustino, F. (2021). Phys. Rev. B, 103, 125401.  Web of Science CrossRef Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767
Follow J. Appl. Cryst.
Sign up for e-alerts
Follow J. Appl. Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds