You are on page 1of 9

Journal of Physics and Chemistry of Solids 111 (2017) 25–33

Contents lists available at ScienceDirect

Journal of Physics and Chemistry of Solids


journal homepage: www.elsevier.com/locate/jpcs

3R and 2H polytypes of MoS2: DFT and DFPT calculations of structural,


optoelectronic, vibrational and thermodynamic properties
S.S. Coutinho a, M.S. Tavares a, b, C.A. Barboza c, N.F. Fraz~
ao d, E. Moreira b, David L. Azevedo e, f, *
a
Departamento de Física, Instituto Federal de Educaç~ ao, Ci^encia e Tecnologia do Maranh~ ao - IFMA, Campus Monte Castelo, 65030-005, S~ ao Luís, MA, Brazil
b
Departamento de Física, Centro de Ci^encias Tecnologicas, Universidade Estadual do Maranh~ ao - UEMA, Cidade Universit
aria Paulo VI, 65055-310, S~ao Luís, MA, Brazil
c
Departamento de Biofísica e Farmacologia, Universidade Federal do Rio Grande do Norte - UFRN, 59072-970, Natal, RN, Brazil
d
Centro de Educaç~ ao e Saúde, Universidade Federal de Campina Grande - UFCG, Campus Cuite, 58175-000, Cuite, PB, Brazil
e
Instituto de Física, Universidade de Brasília - UnB, Campus Universit ario Darcy Ribeiro - Asa Norte, 70919-970, Brasília, DF, Brazil
f
Faculdade UnB Planaltina - FUP, Universidade de Brasília - UnB, Area  Universit
aria 01, 73345-010, Planaltina, DF, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: We report the results of a theoretical study on the behavior of the structural, optoelectronic, vibrational, including
MoS2 infrared and Raman theoretical spectra, phonon spectrum, and thermodynamic properties of 3R- and 2H- poly-
DFT calculations types of molybdenum disulfide (MoS2) using density functional theory (DFT) considering both the local density
Optoelectronic properties
and generalized gradient approximation, LDA and GGA, respectively. Calculated lattice parameters are close to
Vibrational properties
the experimental measurements, and an indirect band gap E(A→KΓ)¼ 1:33 eV (0.68 eV) was obtained within the
Thermodynamic properties
GGA (LDA) level of calculation, considering the 3R-polytype, and for the 2H- polytype an indirect band gap
E(Γ→KΓ)¼ 1:30 eV (0.70 eV) was obtained within the GGA (LDA) approximation. The complex dielectric function
and absorption of 3R-MoS2 and 2H-MoS2 polytypes were shown to be sensitive to the plane of polarization of the
incident light. The phonon dispersion relation together with density of states (DOS) as well as theoretical peaks of
the infrared (IR) and Raman spectra in the frequency range of 0–800 cm1 was analyzed and assigned, consid-
ering the norm-conserved pseudopotentials. The thermodynamic potentials, the specific heat at constant volume
and Debye temperature of the 3R-MoS2 and 2H-MoS2 polytypes are also calculated, whose dependence on the
temperature are discussed.

1. Introduction vertically stacked, weakly interacting layers held together by Van der
Waals interactions (Fig. 1) [16] which justifies its use as lubricants. MoS2
The crystalline sulfide of molybdenum occurs as the mineral molyb- is a typical example from the layered transition-metal dichalcogenide
denite. It is the principal ore from which molybdenum metal is extracted family of materials. When MoS2 is synthesized it is possible to obtain
[1]. Molybdenum disulfide is relatively unreactive, being unaffected by hexagonal (2H), as well as, rhombohedral (3R) polytype layered crystals
dilute acids and oxygen. In its appearance and feel, molybdenum disul- which have quite different physical properties, and octahedral (1T)
fide is similar to graphite. Indeed, like graphite, it is widely used as a solid phase, it is metastable at room temperature, but metallic and of better
lubricant because of its ultra-low friction properties [2]. conductivity improving capacitances of supercapacitors [17] and it is an
Molybdenum disulfide is often a component of blends and composites effective contact electrode with record low-resistance [18]. Furthermore,
where low friction is sought [1], due to retaining their lubricity even in the two-dimensional materials are appealing for use in new-generation
cases of almost complete oil loss, being applied as aircraft engines and nanotechnology because of its excellent physical and chemical proper-
also used in ski wax to prevent static buildup in dry snow conditions and ties. Similar to graphene, but with distinct advantages in electronic ap-
to add glide when sliding in dirty snow. Indeed, MoS2 has attracted great plications [16], the transitions-metal dichalcogenides (TMD) brings a lot
interest for applications in optoelectronic, catalytic and lubricating of possible applications in nano-science and condensed matter physics, as
properties, such as hydrogen production, nanotribology, hydro- well as magneto-optical applications [19]. TMDs form a class of materials
desulphurization catalyst used for removing sulfur compounds from oil, with the formula TX2, where T is a transition Metal from groups IV-VI
solar cells, and photocatalysis [3–15]. Crystals of MoS2 are composed of (e.g. Mo, Ti, Nb), and X is a chalcogen atom (S, Se or Te). Widely

* Corresponding author. Instituto de Física, Universidade de Brasília - UnB, Campus Universitario Darcy Ribeiro - Asa Norte, 70919-970, Brasília, DF, Brazil.
E-mail address: david888azv@gmail.com (D.L. Azevedo).

http://dx.doi.org/10.1016/j.jpcs.2017.07.010
Received 10 May 2017; Received in revised form 19 June 2017; Accepted 12 July 2017
Available online 15 July 2017
0022-3697/© 2017 Elsevier Ltd. All rights reserved.
S.S. Coutinho et al. Journal of Physics and Chemistry of Solids 111 (2017) 25–33

Fig. 1. The MoS2 crystal unit cell. (a) The 2H-MoS2 polytype, and (b) the 3R-MoS2 polytype.

studied is the molybdenum disulfide sheet (MoS2) which has been suc- (DFT) formalism [34,35], as implemented in the CASTEP code [33]. Two
cessfully synthesized by chemical methods in the experiment [20]. Single exchange-correlation functionals were taken into account: the total local-
layers have 6.5 Å thickness and can be extracted using stotch tape density approximation (LDA) [35–37] and the generalized gradient
[21,22] or lithium based intercalation [23,24], furthermore the single- approximation [38,39]. The LDA functional uses standard parametriza-
layer MoS2 is a direct gap semiconductor [25,26] with a bandgap of tion [40,41], while the GGA functional is the one proposed by Perdew-
1.8 eV [25] whereas bulk MoS2 is semiconducting with an indirect Burke-Ernzerhof [42,43]. We have also adopted pseudopotentials to
bandgap of 1.2 eV [27]. Similar to carbon, MoS2 can form tubular, replace the core electrons in each atomic species. For the LDA case, ul-
fullerene [28,29] nanowires-likes nanostructures [30,31] and also show trasoft Vanderbilt-type pseudopotentials were used [44], while norm-
the influence of quantum mechanical confinement in their electronic and conserved pseudopotentials [45] were adopted in the GGA calculations
optical properties. with dispersion correction (GGA þ D) [46–51]. The integration over the
In this paper, we investigate by DFT and DFPT calculus of the struc- Brillouin zone was realized by means a k-points sampling using a 552
tural, vibrational, optoelectronic and thermodynamic properties for the Monkhorst-Pack grid [52]. This grid was sufficient to guarantee the
polytypes of MoS2 in the two forms 2H and 3R. In addition we study in convergence of the electronic structure. The atomic positions were
both polymorphic nanostructure the density-states (DOS), partial density optimized following the convergence thresholds for successive self-
states (PDOS), phonon dispersion relation (being compared with other consistent steps: total energy change smaller than 0.50105 eV/atom,
results), the complex dielectric function and absorption, infrared and maximum force over each atom below 0.01 eV/Å, pressure smaller than
Raman spectra. Our paper is organized as follows: in Section 2 we present 0.02 GPa, and maximum atomic displacement not exceeding
all computational procedures used to obtain the properties of systems. 0.50103 Å. The Broyden-Fletcher-Goldfarb-Shanon (BFGS) minimizer
Section 3 we show and discuss the all results find out to physical prop- [53] was employed to carry out the unit cell optimization. To represent
erties of 3R- and 2H- polytypes of MoS2. Finally, in Section 4 we present the Kohn-Sham orbitals, a plane-wave basis set with cutoff energy was
the conclusions. fixed at 1000.0 eV.
Upon obtained the optimized unit cell, the Kohn-Sham electronic
2. Computational methodology band structure and the density of states (total and partial, with contri-
butions per atom and per orbital in the last case) were evaluated for both
The initial lattice parameters for 3R- and 2H- polytypes of MoS2 were the optimized LDA-CAPZ and GGA-PBE unit cell, as well as the complex
obtained from the x-ray data provided by Sch€ onfeld et al. [32]. The dielectric function and absorption optical for polarized light. The IR and
structures of 3R- and 2H-MoS2 are shown in Fig. 1. For 3R-MoS2 poly- Raman spectra were evaluated for the optimized GGA unit cell as well as
type, the experimental lattice parameters are a ¼ 3:163 Å, b ¼ 3:163 Å, the frequencies of active modes, together with the calculated dielectric
and c ¼ 18:370 Å, and the space group is (C3v ) R3M (160), according to permittivity and polarisability tensors, following the recommendations
Table 1. For 2H-MoS2 polytype, the experimental lattice parameters are of Refs. [54] and [51].
a ¼ 3:161 Å, b ¼ 3:161 Å, and c ¼ 12:295 Å, and the space group is (D6h )
P63/MMC (194), according to Table 1, for the sake of comparison. 3. Results and discussion
All calculations were performed within the density functional theory
3.1. Geometry optimization
Table 1
Lattice parameters for 3R- and 2H- polytypes of MoS2 according to the DFT LDA-CAPZ and In Table 1 one can see the lattice parameters and unit cell volumes for
GGA-PBE approaches. Experimental data are also presented [32]. Lengths (a; b; c) are in Å, the 3R- and 2H- polytypes of MoS2 obtained from our LDA-CAPZ and
volumes (V) in Å3 and angles (α; β; γ) in degrees. GGA-PBE approaches, as well as the experimental values from Ref. [32].
a¼b c α¼β γ V The lattice parameters calculated within the GGA approach are consis-
3R-MoS2
tently bigger compared to those calculated using the LDA one and the
LDA 3.139(-0.7%) 17.942(-2.3%) 90 120 153.164(-3.7%) experimental data, which is coherent with the results found for other
GGA 3.199(þ1.1%) 20.139(þ8.7%) 90 120 178.567(þ10.8%) compounds and the well known underbinding effect for this kind func-
Exp. 3.163 18.370 90 120 159.162 tional. It is a very well-known fact that the LDA (GGA) exchange-
2H-MoS2
correlation functional tends to overestimate (underestimate) the inter-
LDA 3.134(-0.8%) 12.069(-1.8%) 90 120 102.720(-3.4%)
GGA 3.199(þ1.1%) 13.377(þ8.0%) 90 120 118.558(þ10.2%) atomic forces, thus predicting smaller (larger) bond lengths and lattice
Exp. 3.161 12.295 90 120 106.392 parameters in general. The calculated value for the 3R-MoS2 polytype

26
S.S. Coutinho et al. Journal of Physics and Chemistry of Solids 111 (2017) 25–33

lattice parameter using LDA-CAPZ exchange-correlation functional is


smaller than the X-ray value by 2.3% (c), and the unit cell volume is
smaller by 3.7%; and for the 2H-MoS2 polytype is smaller than the X-
ray value by 1.8% (c), and the unit cell volume is smaller by 3.4%.
The GGA-PBE lattice parameter, on the other hand, is larger than the
experimental measurement by 8.7% for the c parameter, for the 3R-MoS2
polytype, and the unit cell volume is bigger by 10.8%; whereas the 2H-
MoS2 polytype is larger than the experimental measurement by 8.0% for
the c parameter, and the unit cell volume is bigger by 10.2%. Notwith-
standing that, the GGA estimations using norm-conserved pseudopoten-
tials generated by the OPIUM code and van der Waals interactions have a
good agreement with available X-ray data for the a and b parameters.

3.2. Band structure and density states

The Kohn-Sham electronic band structure provides an illustration of


the electronic eigenenergies E as a function of the components of a wave
vector k in the first Brillouin zone (BZ). For a 3R-MoS2 and 2H-MoS2
polytypes of molybdenum disulfide, the paths in the BZ were selected
using straight segments connecting a set of high-symmetry points. These Fig. 3. Electronic Kohn-Sham band structure and density of states (DOS) of 2H-MoS2
calculated using LDA-CAPZ (dotted) and GGA-PBE (solid) exchange-correlation func-
points were chosen following the same sequence for the two polytypes: Γ tionals. The top valence bands have been aligned at the Γ point (zero energy), where their
(0.000, 0.000, 0.000), A (0.000, 0.000, 0.500), H (0.333, 0.667, 0.500), main energy band gaps are indirect (Γ to KΓ), where the Λ point depicts the bottom of
K (0.333, 0.667, 0.000), Γ (0.000, 0.000, 0.000), M (0.000, 0.500, conduction band.
0.000), L (0.000, 0.500, 0.500), H (0.333, 0.667, 0.500), and Γ (0.000,
0.000, 0.000). Figs. 2 and 3 show 3R-MoS2 and 2H-MoS2 Kohn-Sham
electronic band structures and density of states (DOS) calculated using valence band maximum occurs at the A (Γ) point in reciprocal space to
the LDA-CAPZ (dotted) and GGA-PBE (solid) exchange-correlation the 3R- (2H-) MoS2, with a secondary maximum at the Γ (A) point. The
functionals. The Fermi level (horizontal dotted line) represents the en- conduction band minimum, for both polytypes, is at about the KΓ middle
ergy of the highest occupied state and it was gauged to be zero in all plots point (it is depicted as Λ). Where the 3R-MoS2 DOS graph at the right side
at 0 K. of Fig. 2 indicates clearly at the point 0 eV the decrease of the LDA gap
In Fig. 2, one can see the LDA-CAPZ (GGA-PBE) band structures of 3R- (Eg ¼ 0:68 eV) compared to the GGA (Eg ¼ 1:33 eV). For the 2H-MoS2
MoS2 for the range of 5 eV to 5 eV; including a set of 18 (21) valence polytype, the energy band gaps depicted the same pattern, being 0.70 eV
bands from 5 eV to 0 eV, and a set of 12 (12) conduction bands between (LDA) and 1.30 eV (GGA).
0 eV and 5 eV. Already, the 2H-MoS2 band structures (Fig. 3), calculated Considering the full band structure, for both polytypes, the GGA es-
by LDA-CAPZ (GGA-PBE) approach, show 13 (12) valence bands and 12 timate is similar to the experimental band gap value (see Ref. [56]), but
(12) conduction bands between the range of 6 eV to 10 eV. Being that the LDA approximation predicted a smaller value. We remember that due
the valence bands of this last polytype are divided in two regions: lower to the approximate nature of exchange-correlation DFT functionals,
(0.70–4.63 eV) and upper (5.17–9.44 eV) giving rise to a gap of 0.54 eV. theoretically calculated energy gaps are inaccurate and could over- or
For both plots, LDA eigenenergies are smaller for the conduction bands underestimate the experimental data [57]. Besides that, the electronic
(CB) and approximately the same value for the valence bands (VB) at Γ eigenenergies estimation from DFT Kohn-Sham simulation do not match
point, resulting in smaller LDA band gaps in comparison to GGA. The the correct excitation energies, and a correct band gap prediction can be
made only by using the exact (but unknown) expression for the exchange-
correlation functional, as shown by Perdew and Levy [58]. As a matter of
fact, LDA calculations underestimate the main band gap of semi-
conductors and insulators by at about 40%, however GGA is better [59].
Despite this problem, there is evidence pointing that a rigid shift (up-
wards the LDA conduction bands) is enough to produce a reasonable
agreement with the more sophisticated quasi-particle GW approxima-
tion, which is able to predict optical excitation frequencies in semi-
conductors [60–63]. As the GGA band structure presented here looks
very similar to the LDA close to the valence and conduction bands limits,
we think that our analysis are valid to achieve an agreement with the true
3R- and 2H-MoS2 experimental band gap. According to our results, 3R-
MoS2 (AVB → KΓCB ) and 2H-MoS2 (ΓVB → KΓCB ) polytypes of molybdenum
disulfide are an indirect band gap semiconductor material, and it is in
complete agreement with the experimental measurement of the energy
gap reported in Ref. [56]. We have depicted high symmetry Λ point
(between K and Γ) as the bottom of conduction band for both phases.
Looking to the DOS curves in Fig. 2, we could see that LDA states are
lowered relative to GGA calculation, in the region under Fermi level
energy; being more evidence to the peak locates at 1.5 eV and 2.6 eV.
Now, in Fig. 3 this characteristic is more visible, where all peaks os states
Fig. 2. Full Kohn-Sham band structure and density of states (DOS) of the 3R-polytype of are displaced downward (with one exception to the states at about
molybdenum disulfide (MoS2) for full energy range using: LDA-CAPZ (dotted) and GGA-
2.9 eV), when it is compared with GGA approximation. On the other
PBE (solid) exchange-correlation functionals. The top valence bands have been aligned
at the A point with their main energy band gaps indirect (A to KΓ), where the Λ point hand, the DOS on the conduction band for the 3R- and 2H- polytypes are
depicts the bottom of conduction band. completely different, because the 2H-MoS2 (Fig. 3) states are well defined

27
S.S. Coutinho et al. Journal of Physics and Chemistry of Solids 111 (2017) 25–33

by two regions, which do not appear in Fig. 2. This is characterized by the appearance of an upper conduction bands in 2H-MoS2 polytype, char-
presence of the Mo-4d and S-3p (lower region), following by Mo-5s and acterized by a gap between the two conduction bands, but this does not
Mo-4p (upper region). affect the full band gap (see Fig. 3).
Fig. 4 (a) and 4(b) are a detailed plot of the partial density of states
(PDOS) of 3R-MoS2 per atomic species (Molybdenum and Sulfur) per 3.3. Optical properties
orbital, after the calculations within the LDA-CAPZ and GGA-PBE ex-
change-correlation functional, where in comparison the two functionals The description of the optical properties are directly connected to the
revealed a very similar pattern. The same property and functional results calculation of the real part ε1 ðωÞ and imaginary part ε2 ðωÞ of the complex
are shown in Fig. 5(a) and (a) for the 2H-MoS2 polytype. The Fermi level dielectric function as a function of the energy (in eV). Figs. 6 and 7 show
is taken to be 0.0 eV. At the top of Fig. 4(a) and (b), one can see the the behavior of these functions for the polymorphic phases 3R- (Rhom-
contribution to the electronic structure from the Mo atoms. Its most bohedral) and 2H- (Hexagonal) of the molybdenum disulphide (MoS2),
important feature for the valence bands arising from the d DOS band that respectively. These calculations were performed in different directions of
correspond from 5.0–5.0 eV, with maximums at 3.9 eV, 3.3 eV, polarization and polycrystalline (POLY) using the GGA-PBE approach. In
2.6 eV, 2.1 eV, 1.0 eV, and 0.6 eV; and to the conduction bands the Fig. 6, we find that the electronic static dielectric constant ε0 undergoes
contribution are given by Mo-4d bands with two maximum points at variations depending on the polarization direction adopted and the
1.8 eV and 2.2 eV. In comparison, the PDOS for Mo in 2H-MoS2 polytype polymorphic form. In the 2H- phase we measured ε0 ¼ 14:21 for the
has qualitatively the same features, considering the same energy ranges. directions [010, 100], and [110]; 8.05 in the directions [001, 101], and
However, the 2H- polytype has an expressive contribution of the 4p and [111]; and 12.04 for the direction POLY. While in phase 3R-the value
5s orbitals between 6.0 and 10.0 eV, with Mo-4p DOS band maxima determined for the plans [010, 100], and [110] is 13.59; in the orien-
smaller than 0.7 eV1 (GGA) and Mo-5s DOS band maxima smaller than tations [001, 101], and [111] the measured value is 7.04 and finally
0.5 eV1 (GGA). along direction POLY ε0 ¼ 11:35. We consider in our results that the
The S atoms in 3R-MoS2 polytype, on the other hand, contribute with values ε1 ðωÞ and ε2 ðωÞ depend directly on the polymorphic form, where
two 4p DOS bands: the first one between 5.0 and 0.0 eV and the ε1 ðωÞ always assume more intense values in 2H- compared to 3R-phase,
second between 1.2 and 4.8 eV, as it is depicted at the bottom of Fig. 4(a) which represents the greater polarization capacity of the material in the
and (b). Already, analyzing the 2H-MoS2 (Fig. 5(a) and (b)), it is easy to 2H- phase. Another important point is the dependence of the plasma
see that arising up a third contribution of the 5s orbital at the upper frequency of the crystal with the direction of polarization, presenting a
conduction bands between 5.7 and 9.1 eV, in comparison with S-GGA more intense value in the crystallographic directions [001, 101], and
and S-LDA of the 3R-MoS2. The 5s orbitais, in both polytypes, originate [111]. Moreover, in the 2H- phase (5.4 eV) the plasma frequency is
DOS bands with most pronounced maxima at: 3.9, 3.3, and 2.1 eV higher than plasma frequency in 3R-phase (5.3 eV), which guarantees the
(the dispersive bands observed at the valence bands); and 2.1, 3.3, and semiconductor a extension to its type-metal behavior [64], where the
3.9 eV (the dispersive bands observed at the conduction bands). Being absorption is smaller and consequently the transparency is greater in
that, the 5s contribution can be neglected, because the S-5s DOS band at percentages ranging from 2:1% to 2:6% depending on the direction of
the range 5.7 and 9.1 eV has height smaller than 0.1 eV1. Accordingly, polarization. The imaginary part ε2 ðωÞ is related to the real electronic
one may conclude also that the Mo atoms are responsible for the transitions between the occupied and unoccupied states (interband

Fig. 4. 3R-MoS2 PDOS as a function of energy (in eV) for the range 5 eV to 5 eV, near the Fermi level energy represented by a vertical dotted line at 0 eV. The plots show the contributions
from each atomic species (at the top is the Molybdenum (Mo) atom and at the bottom is the Sulfur (S) atom) calculated by (a) LDA-CAPZ and (b) GGA-PBE. Where s, p, and d orbital
contributions are represented by solid, dashed, and dotted lines, respectively.

28
S.S. Coutinho et al. Journal of Physics and Chemistry of Solids 111 (2017) 25–33

Fig. 5. 2H-MoS2 PDOS as a function of energy (in eV) for the range 6 eV to 10 eV. The Fermi level energy is depicted by a vertical dotted line at 0 eV. The plots show the contributions
from the Molybdenum (Mo) atom and from the Sulfur (S) atom calculated by (a) LDA-CAPZ and (b) GGA-PBE. Solid, dashed, and dotted lines correspond to s, p, and d contributions,
respectively.

Fig. 6. (a) Real (ε1 ) and (b) imaginary (ε2 ) parts of the GGA complex dielectric function Fig. 7. (a) Real (ε1 ) and (b) imaginary (ε2 ) parts of the GGA complex dielectric function
for the 3R-polytype of MoS2. Curves for incident light polarized along distinct crystalline for the 2H- polytype of MoS2. Curves for incident light polarized along distinct crystalline
planes and for light incident on a polycrystalline sample (Poly) are shown. planes and for light incident on a polycrystalline sample (Poly) are shown.

contribution), which is directly linked to the absorption. In addition, the This interdependence is clear in Figs. 6 and 7 when we observe the
dielectric constant describes a causal response to this effect and we can typical behavior of ε1 ðωÞ and ε2 ðωÞ for the same frequency values.
map the behavior of ε1 ðωÞ in ε2 ðωÞ from the Kramres-Kroning trans- Looking to Fig. 7, we find that for the polarization in the directions [010,
formations, given by Eqs. (1) and (2). 100], and [110], in both configurations, ε2 ðωÞ has a more pronounced
peak around 2.8 eV due to the electronic transition involving the valence
2 ∞ ω0 ε2 ðω0 Þ states Mo-4d character to S-3p conduction states according to the analysis
ε1 ðωÞ  1 ¼ P∫ 0 02 dω0 ; (1)
π ðω  ω2 Þ of Figs. 4 and 5. Looking now to Fig. 8, we can perform a direct com-
parison between the optical absorption for the 3R- and 2H- polytypes of
2 ∞ ω0 ε1 ðω0 Þ MoS2 using the GGA-PBE functional. The contribution of the imaginary
ε2 ðωÞ ¼  P∫ 0 02 dω0 : (2) part Im (εðωÞ) to the absorption phenomena is clear, since in the range
π ðω  ω2 Þ
between 2.0 eV and 7.0 eV, we have a considerable increase of the

29
S.S. Coutinho et al. Journal of Physics and Chemistry of Solids 111 (2017) 25–33

Fig. 8. Optical absorption of the structures (a) 3R-MoS2 and (b)2H-MoS2 near the main Fig. 10. Raman spectrum of 3R-MoS2 in the 250–500 cm1 range. The numbers corre-
energy band gap when the incident light is polarized along the crystal planes [001, 010, spond to the normal modes shown in Table 2. The smearing used to adjust the sharp peaks
100, 101, 110, 111] and for a polycrystalline sample (Poly), using the GGA exchange- was 2 cm1.
correlation functional. The maximum absorption occurs at near ultraviolet.

with the respective irreducible representations and the assignment of IR


absorption, region where we have the most intense peaks for ε2 ðωÞ.
and Raman active modes. The E and A1 modes are IR and Raman actives,
whereas the G is totally inactive (silent mode) for the 3R-polytype of
3.4. Vibrational properties MoS2. The E1u and A2u modes are IR active, the E2g , E1g and A1g are
Raman active, whereas the B2g , E2u and B1u are silent modes for the 2H-
After the optimization process of structures, we performed the cal- polytype of MoS2. Looking at Fig. 9, the most intense IR absorption peaks
culations from density functional perturbation theory (DFPT). The DFPT (two) occur at 372.34 cm1, indicated as 372 (391.2 cm1 [68]), both for
is a theory based in the linear response formalism [55,65], according as 3R- and 2H- polytypes of MoS2, is related to a in-plane E mode (3R-) and a
previous works [46,48,50,51]. All calculations were carried out from in-plane E1u mode (2H-), where sulfur atoms are moving in phase in the
GGA calculations within the norm-conserving potential scheme. opposite direction to the atom of molybdenum along the b-axis and a-
The convergence thresholds were: total energy convergence tolerance axis, and perpendicular to the c-axis, as in Ref. [68]. The other peaks
smaller than 5  106 eV/atom, maximum ionic force smaller than 102 (two) appear at 454.46 cm1 in both cases (3R-MoS2 and 2H-MoS2),
eV/Å, maximum ionic displacement tolerance of 5  104 Å, and indicated as 454 (469.4 cm1 [68]) is assigned to an A1 mode (3R-) and
maximum stress component smaller than 2  102 GPa. For the self- an A2u mode, where sulfur atoms are moving in the opposite direction to
consistent field calculations, the convergence criteria took into account the atom of molybdenum along the c-axis [68]. Fig. 10 presents the
a total energy per atom variation smaller than 5  107 eV and electronic calculated Raman scattering spectrum for the 3R-MoS2, in the
eigenenergy variation smaller than 1:667  107 eV. 250–500 cm1 range, with the most intense peak at 372.34 cm1, indi-
The IR and Raman spectra of the 3R- and 2H- polytypes of MoS2 are cated as 372 (387.8 [68]), a mode with irreducible representation E; the
shown in Figs. 9–11. Tables 2 and 3 present the predicted normal modes second most intense peak occurs at 396.54 cm1 (indicated as 396) with

Fig. 9. Infrared spectra of 3R-MoS2 (black: solid line) and 2H-MoS2 (red: dot line) in the
0–800 cm1. The numbers correspond to the normal modes shown in Tables 2 and 3. The Fig. 11. Raman spectrum of 2H-MoS2 in the 0–500 cm1 range. The numbers correspond
smearing used to adjust the sharp peaks was 2 cm1. (For interpretation of the references to the normal modes shown in Table 3. The smearing used to adjust the sharp peaks
to colour in this figure legend, the reader is referred to the web version of this article.) was 2 cm1.

30
S.S. Coutinho et al. Journal of Physics and Chemistry of Solids 111 (2017) 25–33

Table 2 ω ¼ 0 (ω ¼ ∞) are εxx ¼ 14:669 (14.329), εyy ¼ 14:669 (14.329), εzz ¼


Normal modes of 3R-MoS2 at k ¼ 0. Irreducible representations (Irrep) are indicated, as 4:654 (4.637), and εxy ¼ εxz ¼ εyz ¼ 0 (0). Whereas for 2H-MoS2 at ω ¼
well as the IR and Raman active modes.
0 (ω ¼ ∞) are εxx ¼ 14:692 (14.347), εyy ¼ 14:692 (14.347), εzz ¼ 4:704
N k (cm1) Irrep IR Raman (4.685), and εxy ¼ εxz ¼ εyz ¼ 0 (0). On the other hand, the polarisability
1 8.9 G N N tensor (in units of Å3) for 3R-MoS2 at ω ¼ 0-static (ω ¼ ∞-optical) has
2 8.9 G N N the following components: pxx ¼ 194:234 (189.402), pyy ¼ 194:234
3 8.9 G N N
4 8.9 G N N
(189.402), pzz ¼ 51:928 (51.676), and pxy ¼ pxz ¼ pyz ¼ 0 (0). For 2H-
5 20.1 E N N MoS2 at ω ¼ 0-static (ω ¼ ∞-optical) are pxx ¼ 129:177 (125.926), pyy ¼
6 20.1 E N N 129:177 (125.926), pzz ¼ 34:942 (34.764), and pxy ¼ pxz ¼ pyz ¼ 0 (0).
7 276.0 G N N
There is a clear optical anisotropy that comes from its hexagonal sym-
8 276.0 G N N
9 276.0 G N N
metries. For hexagonal symmetry the permittivity tensor has two main
10 276.0 G N N axes, hence the optical anisotropy. The tensor permittivity obtained for
11 276.5 E Y Y both phases characterizes uniaxial positive materials.
12 276.5 E Y Y The phonon's properties were studied by means of the harmonic
13 372.3 E Y Y
approximation implemented in CASTEP code [33]. In the CASTEP the
14 372.3 E Y Y
15 372.3 G N N lattice vibration are described by tensor:
16 372.3 G N N  
17 372.3 G N N ∂2 E
Dμν ðR  R0Þ ¼ ; (3)
18 372.3 G N N ∂uμ ðRÞ∂uν ðR0Þ u¼0
19 395.6 E N N
20 395.6 E N N
21 396.5 A1 Y Y such that u and E stand for displacement and total energy of a given atom,
22 454.4 A1 Y Y respectively.
23 455.8 E N N Thereby, the phonon density of states of the band n, can be ob-
24 455.8 E N N
tained as:

dk
Nn ðωÞ ¼ ∫ δ½ω  ωn ðkÞ; (4)
Table 3 4π 3
Normal modes of 2H-MoS2 at k ¼ 0. Irreducible representations (Irrep) are indicated, as
well as the IR and Raman active modes. being that ωn (k) stands for the dispersion of the band n. It's worth
N k (cm1
) Irrep IR Raman highlighting that Nn ðωÞ is obtained over the Brillouin zone.
The phonon dispersion curves along some high-symmetry points in
1 13.7 E2g N Y
the Brillouin zone (left panel) and the total density of phonon states mode
2 13.7 E2g N Y
3 25.0 B2g N N (right panel) for the 3R- and 2H- polytypes of MoS2 are summarized in
4 276.1 E2u N N Fig. 12 and practically identical, showing the frequency range from 0 to
5 276.1 E2u N N 500 cm1. It indicates that the phonon structure is stable since
6 276.5 E1g N Y throughout the Brillouin zone all phonon frequencies are positive. The
7 276.5 E1g N Y
main features of the dispersion curves is the presence of two well defined
8 372.3 E1u Y N
9 372.3 E1u Y N frequency regions corresponding to a region from 0 to 240 cm1, with
10 372.4 E2g N Y contribution of three acoustic modes, as in Ref. [68], which are linear
11 372.4 E2g N Y around Γ; and a second region from, approximately, 260–460 cm1, after
12 395.4 B1u N N a gap of 42 cm1, corresponding to the case of the A1g (out-of-plane) and
13 396.5 A1g N Y
E2g (in-plane) modes. In the left panel of Fig. 12, depicting the phonon
14 454.4 A2u Y N
15 455.8 B2g N N density of states spectrum, the peaks (gaps) are related to the corre-
sponding modes for each region.

3.5. Thermodynamic potentials


irreducible representation A1 ; the third peaks can be seen at 276.56,
indicated as 276 (288.7 cm1 [68]), corresponding to a E mode; and the
The DFPT calculations can be used to compute the enthalpy (H), free
smaller peaks at 454.46 cm1 is related to a A1 mode. Fig. 11 shows the
energy (F), the temperature times the entropy term TS ¼ U  F (U being
profiles of the calculated Raman scattering spectrum for the 2H-MoS2, in
the internal energy), and the lattice heat capacity (CV ) as functions of
the 0–500 cm1 range. The main difference is in the frequency
temperature. These potentials were obtained at O K, according to
13.71 cm1 (indicated as 13), that appears related to phase 2H-, assigned
Refs. [51,55].
to a E2g mode (33.7 cm1 [66] and 32 cm1 [67]) attributed to the vi-
The enthalpy (H) is calculated through formula,
bration of the adjoining rigid layers, as shown in Refs. [66,67], while the
other vibrational modes correspond to the movement of S-Mo-S groups ℏω
perpendicular to the c-axis, except for the A1g mode (along the c-axis). HðTÞ ¼ Etot þ Ezp þ ∫   NðωÞdω; (5)
exp ℏω 1
The intense peak at 372.42 cm1 (383 [66,67], and 387.8 cm1 [68]) is kT

related to a E2g mode, the peak at 396.58 cm1 (409 [66], 408 [67], and being Ezp the zero point vibrational energy, kB is Boltzmann's constant, ℏ
412 cm1 [68]) is assigned to an A1g mode, and the peak at 276.57 cm1 is reduced Planck's constant and NðωÞ is the phonon density of states. The
(287 [66], 286 [67], and 288.7 cm1 [68]) associated with a E1g mode. In vibrational contribution to the free energy, F, is given by:
practice we find the same eigenfrequencies between the infrared ab-
  
sorption and the Raman scattering, except for the E2g mode in Fig. 11, for ℏω
FðTÞ ¼ Etot þ Ezp þ kB T∫ NðωÞln 1  exp  dω; (6)
3R- and 2H- polytypes of MoS2, due to intermolecular (or interlayer) kB T
interactions known as Davydov splitting that is typically observed in
layered crystals [66,69]. while for the temperature times the entropy term, TS ¼ U  F, can
The optical permittivity tensor components calculated for 3R-MoS2 at be obtained:

31
S.S. Coutinho et al. Journal of Physics and Chemistry of Solids 111 (2017) 25–33

enthalpy for the 3R-MoS2 (black: solid line) assumes higher energy
values as a function of the temperature. The free energy (red: medium
dashed line) decreases more rapidly, up to 200 K, as the temperature
increases for the 3R-MoS2 comparing to 2H- polytype (green: short dash
dot line). The term TS (blue: small dashed line), for the 3R-MoS2, in-
creases more exponentially as a function of the temperature in compar-
ison with 2H- polytype. Therefore, one can observe that the structure 2H-
MoS2 becomes little more stable energetically than 3R-MoS2 as the
temperature increases. Besides, analyzing the entropy and enthalpy for
each phase, we conclude that the synthesis process for both phases are
non-spontaneous.
In Fig. 14 we represent the constant volume heat capacity CV , as a
function of the temperature (in K). We can observe that the heat capacity
CV increases as the temperature increases, reaching the Dulong-Petit
limit, at around 600 K more rapidly for the structure 2H-MoS2 than
3R-MoS2. The heat capacity CV for 2H-phase is lower than 3R-phase at
any temperature indicating that the 2H-MoS2 phase heats more easily
Fig. 12. Phonon dispersion curves (left panel) and the density of phonon states (DOS) than 3R-MoS2. Physical applications that require thermal insulation
mode/cm1 (right panel) of the structures 3R-MoS2 (black: solid line) and 2H-MoS2 (red: would be more appropriate to use the 3R-MoS2 phase. Using the right-
dot line) in the frequency range from 0 to 500 cm1, calculated using the GGA-PBE- hand scale of Fig. 14, we have also shown the behavior of the tempera-
OPIUM exchange-correlation functionals. (For interpretation of the references to colour
tures dependent ΘD ðTÞ (red solid and green short dash dot lines) which
in this figure legend, the reader is referred to the web version of this article.)
are equal. These results can be compared with the experimental data, or
used to predict the phase stability for different structural modifications.
8 9
< ℏω    =
ℏω 4. Conclusions
TSðTÞ ¼ kB T ∫  NðωÞdω  ∫ NðωÞ 1  exp 
kB T

: exp ℏω  1 kB T ;
kB T
In summary, we have performed the geometry optimization, opto-
(7) electronic and vibrational properties, dielectric permittivities and
We have preferred to use TSðTÞ instead of SðTÞ to emphasize the polarisability tensors, phonon spectrum and density of states, besides the
contribution of the thermodynamic potentials. Thermal expansion plays thermodynamic properties of 3R- and 2H- polytypes of molybdenum
an important role in a gas, less in a liquid, much less in a solid, as we are disulfide, MoS2, using DFT and DFPT calculations. The GGA approxi-
dealing with solids, in our calculations, we did not considered ther- mation overestimates the 3R-MoS2 and 2H-MoS2 lattice parameters,
mal expansion. while the LDA lattice parameters are underestimated. The 3R-polytype of
Fig. 13 depicts the profiles of the calculated thermodynamic poten- MoS2 crystal has a LDA (GGA) indirect band gap of 0.68 eV (1.33 eV)
tials' enthalpy, free energy and temperature times the entropy term, involving AVB →KΓCB transition from valence to conduction band, while
TS ¼ U  F, as a function of temperature (in K), for the 3R- and 2H- 2H- polytype of MoS2 crystal has a LDA (GGA) indirect band gap of
polytypes of MoS2, for the sake of comparison. From there one can see 0.70 eV (1.30 eV) involving ΓVB →KΓCB transition. For the two polytypes,
that the enthalpy (green: short dash dot line) has an almost linear the complex dielectric functions are anisotropic with respect to light
behavior as a function of the temperature for the 2H-MoS2, while the polarization, with its real and imaginary parts exhibiting more pro-
nounced variation with energy when the polarization plane is aligned
with the [010, 100], and [110] directions; and the same occurs for the
optical absorption, until 5.1 eV. An important aspect is the plasma

Fig. 13. Comparison between the profiles of the calculated thermodynamic potentials'
enthalpy (black: solid line), free energy (red: medium dashed line), T  entropy is tem-
perature multiplied by entropy in eV units (blue: small dashed line), as a function of the
temperature, for the 3R-MoS2 polytype; and the thermodynamic potentials for the 2H- Fig. 14. Constant volume heat capacity CV for the 3R-MoS2 and 2H-MoS2 polytypes, as a
MoS2 polytype - enthalpy (cyan: short dash dot lines), free energy (magenta: short dash function of the temperature (in K), for the sake of comparison. The red solid and green
dot lines), and T  entropy (green: short dash dot lines). (For interpretation of the ref- short dash dot lines, using the right-hand side scale, depict the temperature dependence of
erences to colour in this figure legend, the reader is referred to the web version of the Debye temperature ΘD ðTÞ. (For interpretation of the references to colour in this figure
this article.) legend, the reader is referred to the web version of this article.)

32
S.S. Coutinho et al. Journal of Physics and Chemistry of Solids 111 (2017) 25–33

frequency, that is higher in the 2H- phase (5.4 eV) than in the 3R-phase [19] F. Qu, A.C. Dias, J. Fu, L. Villegas-Lelovsky, D.L. Azevedo, Sci. Rep. 7 (2017) 41044.
[20] H.S.S. Ramakrishna Matte, A. Gomathi, A.K. Manna, D.J. Late, R. Datta, S.K. Pati,
(5.3 eV), which guarantees the semiconductor a extension to its type-
C.N.R. Rao, Angew. Chem. Int. Ed. 49 (2010) 4059.
metal behavior. We have also obtained the IR and Raman spectra, both [21] K.S. Novoselov, D. Jiang, F. Schedin, T.J. Booth, V.V. Khotkevich, S.V. Morozov,
exhibiting a very good agreement with other theoretical and experi- A.K. Geim, Proc. Natl. Acad. Sci. U. S. A. 102 (2005) 10451.
mental results, with the same eigenfrequencies between the IR absorp- [22] R.F. Frindt, J. Appl. Phys. 37 (1966) 1928.
[23] P. Joensen, R.F. Frindt, S.R. Morrison, Mater. Res. Bull. 21 (1986) 457.
tion and the Raman scattering, due to intermolecular interactions known [24] A. Schumacher, L. Scandella, N. Kruse, R. Prins, Surf. Sci. Lett. 289 (1993) L595.
as Davydov splitting, except for the E2g mode (13.71 cm1) in Raman [25] K.F. Mak, C. Lee, J. Hone, J. Shan, T.F. Heinz, Phys. Rev. Lett. 105 (2010) 136805.
spectrum for 2H-MoS2. In relation to the phonon spectrum, the main [26] A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.-Y. Chim, G. Galli, F. Wang, Nano
Lett. 10 (2010) 1271.
feature of its dispersion curves is the presence of two well defined fre- [27] K.K. Kam, B.A. Parkinson, J. Phys. Chem. 86 (1982) 463.
quency regions being evident that no imaginary phonon frequency exists [28] R. Tenne, L. Margulis, M. Genut, G. Hodes, Nature 360 (1992) 444.
in the whole Brillouin zone indicating the dynamical stability of our [29] L. Margulis, G. Salitra, R. Tenne, M. Talianker, Nature 365 (1993) 113.
[30] M. Remskar, A. Mrzel, Z. Skraba, A. Jesih, M. Ceh, J. Demsar, P. Stadelmann,
systems, in perfect agreement with the literature. Finally, the thermo- F. Levy, D. Mihailovic, Science 292 (2001) 479.
dynamic properties of these crystal systems are presented by its enthalpy, [31] F. Schwierz, Graphene transistors, Nat. Nanotech 5 (2010) 487.
entropy, free energy, heat capacity and Debye temperature. These results [32] B. Sch€onfeld, J.J. Huang, S.C. Moss, Acta Cryst. B 39 (1983) 404.
[33] M.D. Segall, P.L.D. Lindan, M.J. Probert, C.J. Pickard, P.J. Hasnip, S.J. Clark,
show that the structure 2H-MoS2 crystal become a little more stable M.C. Payne, J. Phys. Cond. Matt 14 (2002) 2717.
energetically than 3R-MoS2 crystal as the temperature increases, and for [34] P. Hohenberg, W. Kohn, Phys. Rev. 136 (1964) B864.
physical applications that require thermal insulation, would be more [35] W. Kohn, L.J. Sham, Phys. Rev. 140 (1965) A1133.
[36] R.O. Jones, O. Gunnarsson, Rev. Mod. Phys. 61 (1989) 689.
appropriate to use the 3R-MoS2 phase.
[37] O.V. Gritsenko, P.R. Schipper, E.J. Baerends, J. Chem. Phys. 107 (1997) 5007.
[38] A. Dal Corso, A. Pasquarello, A. Baldereschi, R. Car, Phys. Rev. B 53 (1996) 1180.
Acknowledgments [39] M. Fuchs, M. Bockstedte, E. Pehlke, M. Scheffler, Phys. Rev. B 57 (1998) 2134.
[40] D.M. Ceperley, B.J. Alder, Phys. Rev. Lett. 45 (1980) 566.
[41] J.P. Perdew, A. Zunger, Phys. Rev. B 23 (1981) 5048.
The authors thank the Maranh~ao Research Foundation–FAPEMA for [42] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865.
financial support from FAPEMA projects: Universal–00795/15 and Uni- [43] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh,
versal–00798/16. David L. Azevedo acknowledges the support by the C. Fiolhais, Phys. Rev. B 46 (1992) 6671.
[44] D. Vanderbilt, Phys. Rev. B 41 (1990) 7892.
Mato Grosso Research Foundation – FAPEMAT for financial support [45] J.S. Lin, A. Qteish, M.C. Payne, V. Heine, Phys. Rev. B 47 (1993) 4174.
through the Grant PRONEX CNPq/FAPEMAT–850109/2009. [46] E. Moreira, J.M. Henriques, D.L. Azevedo, E.W.S. Caetano, V.N. Freire,
E.L. Albuquerque, J. Solid State Chem. 184 (2011) 921.
[47] E. Moreira, J.M. Henriques, D.L. Azevedo, E.W.S. Caetano, V.N. Freire,
References E.L. Albuquerque, J. Solid State Chem. 187 (2012) 186.
[48] E. Moreira, J.M. Henriques, D.L. Azevedo, E.W.S. Caetano, V.N. Freire, U.L. Fulco,
[1] R.F. Sebenik, A.R. Burkin, R.R. Dorfler, J.M. Laferty, G. Leichtfried, H. Meyer- E.L. Albuquerque, J. Appl. Phys. 112 (2012) 043703.
Grünow, P.C.H. Mitchell, M.S. Vukasovich, D.A. Church, G.G. Van Riper, [49] J.E.F.S. Rodrigues, E. Moreira, D.M. Bezerra, A.P. Maciel, C.W.A. Paschoal, Mater.
J.C. Gilliland, S.A. Thielke, Molybdenum and Molybdenum Compounds, Ullmanns Res. Bull. 48 (2013) 3298.
Encyclopedia of Industrial Chemistry, Wiley-VCH, Weinheim, 2000. [50] J.M. Henriques, C.A. Barboza, E.L. Albuquerque, U.L. Fulco, E. Moreira, J. Phys.
[2] C. Donnet, J.M. Martin, Th. Le Mogne, M. Belin, Tribol. Int. 29 (1996) 123. Chem. Solids 76 (2015) 45.
[3] L. Rapoport, Y. Bilik, Y. Feldman, M. Homyonfer, S. Cohen, R. Tenne, Nature 387 [51] E. Moreira, C.A. Barboza, E.L. Albuquerque, U.L. Fulco, J.M. Henriques, A.I. Araújo,
(1997) 791. J. Phys. Chem. Solids 77 (2015) 85.
[4] C. Lee, Q. Li, W. Kalb, X.Z. Liu, H. Berger, R.W. Karpic, J. Hone, Science 328 (2010) [52] H.J. Monkhorst, J.D. Pack, Phys. Rev. B 13 (1976) 5188.
76. [53] B.G. Pfrommer, M. Cote, S.G. Louie, M.L. Cohen, J. Comput. Phys. 131 (1997) 133.
[5] B. Hinnemann, P. Moses, J. Bonde, K. Jorgensen, J. Nielsen, S. Horch, [54] Accelrys, Materials Studio CASTEP, Accelrys, San Diego, 2009.
I. Chorkendorff, J. Norskov, J. Am. Chem. Soc. 127 (2005) 5308. [55] S. Baroni, S. de Gironcoli, A. dal Corso, P. Giannozzi, Rev. Mod. Phys. 73 (2001)
[6] T.F. Jaramillo, K.P. Jorgensen, J. Bonde, J.H. Nielsen, S. Horch, I. Chorkendorff, 515.
Science 317 (2007) 100. [56] I. Song, C. Park, H.C. Choi, RSC Adv 10 (2015) 7495.
[7] T. Todorava, R. Prins, T. Weber, J. Catal. 246 (2007) 109. [57] H. Mizoguchi, P.M. Woodward, C.-H. Park, D.A. Keszler, J. Am. Chem. Soc. 126
[8] P. Raybaud, J. Hafner, G. Kresse, S. Kasztelan, H. Toulhoat, J. Catal. 189 (2005) (2004) 9796.
129. [58] J.P. Perdew, M. Levy, Phys. Rev. Lett. 51 (1983) 1884.
[9] P.G. Moses, B. Hinnemann, H. Topsoe, J.K. Norskov, J. Catal. 248 (2007) 188. [59] S.Q. Wang, H.Q. Ye, J. Phys. Condens. Matter 15 (2003) L197.
[10] J.V. Lauritsen, J. Kibsgaard, G.H. Olesen, P.G. Moses, B. Hinnemann, S. Helveg, [60] R.W. Godby, M. Schlter, L.J. Sham, Phys. Rev. B 37 (1988) 10159.
J.K. Norskov, B.S. Clausen, H. Topsoe, E. Laegsgaard, F. Besenbacher, J. Catal. 249 [61] Z.H. Levine, D.C. Allan, Phys. Rev. B 43 (1991) 4187.
(2007) 220. [62] U. Schonberger, F. Aryasetiawan, Phys. Rev. B 52 (1995) 8788.
[11] J. Lauritsen, J. Bollinger, E. Laegsgaard, K. Jacobsen, J.K. Norskov, B. Clausen, [63] A. Molina-Snchez, D. Sangalli, K. Hummer, A. Marini, L. Wirtz, Phys. Rev. B 88
H. Topsoe, F. Besenbacher, J. Catal. 221 (2004) 510. (2013) 045412.
[12] J. Lauritsen, S. Helveg, E. Laegsgaard, I. Stensgaard, B. Clausen, H. Topsoe, [64] M.S. Dresselhaus, SOLID STATE PHYSICS PART II Optical Properties of Solids,
E. Besenbacher, J. Catal. 197 (2001) 1. 2001. MIT Solid State Physics Course.
[13] G. Kline, K. Kam, R. Ziegler, B. Parkinson, Sol. Energy Mater 6 (1982) 337. [65] D. Porezag, M.R. Pederson, Phys. Rev. B 54 (1996) 7830.
[14] J. Wilcoxon, T. Thurston, J. Martin, J. Nanostruct. Mater. 12 (1999) 993. [66] A.V. Kolobov, J. Tominaga, Two-dimensional Transition-metal Dichalcogenides,
[15] A. Kumar, P.K. Ahluwalia, Mater. Chem. Phys. 135 (2012) 755. vol. 239, Springer Series in Materials Science, Switzerland, 2016.
[16] B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, A. Kis, Nat. Nanotechnol. 6 [67] J. Chen, C. Wang, Solid State Commun. 14 (9) (1974) 857.
(2011) 147. [68] A. Molina-Sanchez, L. Wirtz, Phys. Rev. B 84 (15) (2011) 155413.
[17] L. Jiang, S. Zhang, S.A. Kulinich, X. Song, J. Zhu, X. Wang, H. Zeng, Mater. Res. Lett. [69] A.S. Davydov, Zh. Exper. Teor. Fiz. 18 (1948) 210.
3 (2015) 177.
[18] R. Kappera, D. Voiry, S.E. Yalcin, B. Branch, G. Gupta, A.D. Mohite, M. Chhowalla,
Nat. Mater 13 (2014) 1128.

33

You might also like