Next Article in Journal
A Microwave Photonic Converter with High in-Band Spurs Suppression Based on Microwave Pre-Upconversion
Previous Article in Journal
Dual Polarization Simultaneous Optical Intensity Modulation in Single Birefringent LiNbO3 Mach–Zehnder Optical Modulator
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

First-Principles Study of the Optical Properties of TMDC/Graphene Heterostructures

Department of Electrical Engineering, National Chung Hsing University, Taichung 40227, Taiwan
*
Author to whom correspondence should be addressed.
Photonics 2022, 9(6), 387; https://doi.org/10.3390/photonics9060387
Submission received: 3 May 2022 / Revised: 23 May 2022 / Accepted: 26 May 2022 / Published: 30 May 2022
(This article belongs to the Section Optoelectronics and Optical Materials)

Abstract

:
The transition-metal dichalcogenide (TMDC) in the family of MX 2 ( M = Mo , W ; X = S , Se ) and the graphene ( Gr ) monolayer are an atomically thin semiconductor and a semimetal, respectively. The monolayer MX 2 has been discovered as a new class of semiconductors for electronics and optoelectronics applications. Because of the hexagonal lattice structure of both materials, MX 2 and Gr are often combined with each other to generate van der Waals heterostructures. Here, the MX 2 / Gr heterostructures are investigated theoretically based on density functional theory (DFT). The electronic structure and the optical properties of four different MX 2 / Gr heterostructures are computed. We systematically compare these MX 2 / Gr heterostructures for their complex permittivity, absorption coefficient, reflectivity and refractive index.

1. Introduction

Materials in confined geometries are hosts to novel phenomena. For example, in one dimension (1D), the fractional charge excitation or the soliton is found in the electronic properties of the organic polyacetylene [1], and the Haldane conjecture [2], which corresponds to the symmetry-protected topological phase, is experimentally found in the spin-1 magnetic chain, e.g., CsNiCl3, [Ni(HF2)(3-Clpyradine)4]BF4, Ni(C2H8N2)2NO2(ClO4) and Y2BaNiO5 [3,4,5,6,7,8]. Quantum states conversion through entanglement manipulations [9] as well as new types of quantum phase transitions with a negative central charge [10] are among the recently investigated 1D systems.
Examples of novel physical phenomena in two dimension (2D) include quantum Hall systems in semiconductors [11], the supersolid phase of atomic systems in the optical lattice [12,13] and Dirac electrons in graphene [14]. Graphene (Gr) is one of the most widely investigated 2D materials; it possesses a high mobility [15,16,17,18,19] and is driven by Dirac cone dispersions near the Fermi level at the K-point and the K’-point in the hexagonal Brillouin zone [20]. However, the gapless nature of Gr limits its applications in the semiconductor industry. Therefore, finding new 2D materials with a large band gap is an interesting topic. Note that a new allotropy of carbon, the 2D biphenylene monolayer, was discovered very recently [21,22,23,24].
In addition to the carbon, a number of 2D materials with an energy band gap have been synthesized, including the quantum spin Hall insulator HgTe [25,26], free-standing silicene [27] and the transition-metal dichalcogenide (TMDC) family. Monolayer TMDCs [28,29,30,31,32,33,34,35,36,37,38,39], such as molybdenum disulfide (MoS2) [40,41], molybdenum diselenide (MoSe2) [42], tungsten disulfide (WS2) [43] and tungsten diselenide (WSe2) [44,45], have been found to be direct gap semiconductors [46] and have emerged as new optically active materials for novel device applications. The TMDCs in the family of MX2, where M = Mo , W and X = S , Se , have a rather large energy band gap compared to HgTe and silicene; thus, these materials are more versatile as candidates for thin, flexible device applications and useful for a variety of applications including transistors [47], lubrication [48], lithium-ion batteries [49] and thermoelectric devices [50]. Other TMDCs, e.g., PbTaSe2, MoxW1−xTe2 and NiTe2, can exhibit more unusual physical properties such as topological superconductivity and Dirac and Weyl semimetal physics [51,52,53,54,55,56]. Two-dimensional hydrogenated NiTe2, PdS2, PdSe2, and PtSe2 monolayers also exhibit a quantum anomalous Hall effect [57].
One of the approaches used to broaden the application of 2D materials is to form a heterostructure in which similar lattice structures for different monolayers are usually required, e.g., the hexagonal lattices of TMDC, Gr, boron nitride [58], and boron phosphide (BP) [59]. It has been suggested that MX2/BP heterostructures have a good ability for optical absorption [59]. On the other hand, the MX2/Gr of the type-I van der Waals heterostructures have attracted much attention for their electronic properties. Research on MoS2/Gr [60,61,62], MoSe2/Gr [63,64,65], WS2/Gr [66] and WSe2/Gr [67,68] has revealed that MX2/Gr is favorable for electronics applications and field-effect transistors. However, there are relatively few studies on its optical properties [69,70,71]; therefore, comprehensive research on the optical properties of MX2/Gr is desired.
In this communication, we theoretically investigate the optical properties of MX2/Gr heterostructures. By means of a density functional theory simulation, we systematically compare the energy band structures, the density of states, complex permittivity, absorption coefficients, reflectivity, and refraction indexes.

2. Materials and Methods

The first-principles calculations are based on the local-density approximation (LDA) proposed by Kohn and Sham [72], which approximates the total energy of the multielectron system. The simulations were implemented using the software package Nanodcal and its accompanying software packages DeviceStudio and OpticCal [73,74]. A plane wave basis was set with a cutoff energy of 80 Hartree and a 5 × 5 × 1 Γ -centered k-points grid. The atomic structures were relaxed until the force was smaller than 0.03 eV/Å and the total energy convergence criterion was set as 10 4 eV. In order to avoid interactions in the vertical direction between neighboring slabs, a vacuum layer of 24 Å was added between different slabs.
The lattice constants of the MoS2, MoSe2, WS2, WSe2 and Gr monolayers were 3.166 Å, 3.288 Å, 3.153 Å, 3.282 Å and 2.47 Å, respectively [32,63]. For all the MX2/Gr heterostructures, lattice mismatch ratios less than 5% were achieved by choosing a 3 × 3 MX2 supercell and a 4 × 4 Gr supercell, as shown in Figure 1. The details of the lattice mismatch ratios are listed in Table 1. After the structure optimization, the lattice contents were 9.668 Å, 9.854 Å, 9.648 Å and 9.845 Å for MoS2/Gr, MoSe2/Gr, WS2/Gr and WSe2/Gr, respectively. The details of the atom–atom distances as well as the MX 2 layer–Gr layer distances are listed in Table 2. The total number of atoms in the simulations was 59 for MX2/Gr, 27 for the 3 × 3 monolayer of MX2, and 32 for the 4 × 4 monolayer of Gr.
The binding energies E b are calculated by the energy difference between the heterostructures and the monolayers:
E b = E MX 2 / Gr E MX 2 E Gr ,
where E MX 2 / Gr , E MX 2 and E Gr are the total energies of the heterostructures, isolated MX2 and Gr, respectively. It is interesting to compare other heterostructures, such as MX2/BP [59], as regards binding energies. The interface binding energies of the most stable configurations of MoS2/BP, MoSe2/BP, WS2/BP and WSe2/BP are 196 meV, 130 meV, 201 meV and 141 meV, respectively [59]. As listed in Table 1, all the binding energies for the MX2/Gr heterostructures in this study are negative, and the values are close to MX2/BP, demonstrating that the structural stability of MX2/Gr is similar to that of MX2/BP.

3. Results

3.1. Band Structures

The direct band gaps of the MoS2, MoSe2, WS2 and WSe2 monolayers were 1.82 eV, 1.56 eV, 1.95 eV and 1.64 eV, respectively, [36], where both the conduction band minimum (CBM) and the valence band maximum (VBM) were at the K-point in the Brillouin zone. The Dirac point of the Gr was located at the K-point and pinned to the Fermi level. All the results of the energy band structures for the MX2 monolayers are consistent with the existing literature [17,35]. Now we report the energy band structures and the electronic density of states (DOS) for the four MX2/Gr, as shown in Figure 2. We found that the band structure of MX2/Gr could basically be regarded as the overlap of the band structures of the MX2 and Gr monolayers. Due to van der Waals interactions, the Dirac point opens a small gap of 8.3 meV, 7.9 meV, 8.86 meV and 3.53 meV for MoS2/Gr, MoSe2/Gr, WS2/Gr and WSe2/Gr, respectively. The total and partial DOS values of MX2/Gr are also shown in Figure 2. The partial DOS is the relative contribution of a particular orbital to the total DOS. The d-orbital contributions mainly come from the W or Mo atoms, and the p-orbital contributions come from S or Se, and C atoms. Contributions from the s-orbital are minimal. These results reflect a fairly weak interfacial coupling at the interface between MX2 and Gr.
The Schottky contacts are formed between the semiconducting MX2 and the metallic Gr. The Schottky barrier is one of the most important characteristics of a semiconductor–metal junction and dominates the transport properties. Based on the Schottky–Mott model [75,76,77], at the interface of the metal and semiconductor, an n-type Schottky barrier height (SBH) Φ B n is defined as the energy difference between the Fermi level E F and the conduction band minimum E C , i.e., Φ B n = E C E F . Similarly, the p-type SBH Φ B p is defined as the energy difference between E F and the valence band maximum E V , i.e., Φ B p = E F E V . The n/p-type Schottky contacts are classified by the smaller SBH. The SBH are Φ B n = 0.29 eV and Φ B p = 1.23 eV for MoS2/Gr, Φ B n = 0.75 eV and Φ B p = 0.65 eV for MoSe2/Gr, Φ B n = 0.39 eV and Φ B p = 1.13 eV for WS2/Gr, and Φ B n = 1.00 eV and Φ B p = 0.61 eV for WSe2/Gr, respectively. Therefore, MoS2/Gr, MoSe2/Gr, WS2/Gr and WSe2/Gr are the n-type, p-type, n-type and p-type Schottky contacts, respectively [64,66,67].

3.2. Optical Properties

In order to understand the optical properties, the complex permittivity or the so-called dielectric function was computed under the long-wave approximation, i.e., q = 0 . The complex permittivity ε ( ω ) = ε real ( ω ) + ε imag ( ω ) as a function of incident photon energy is [78]
ε ( ω ) = 1 + 2 e 2 ε 0 m 2 1 V α β | ψ β | e ^ · p | ψ α | 2 ( E β E α ) 2 / 2 f ( E α ) f ( E β ) E β E α ω i η
where e is the electron charge, ε 0 is the vacuum permittivity, m is the electron mass, e ^ is the direction of the vector potential, p is the momentum operator, ω is the photon energy, and f ( E α ) and f ( E β ) are the Fermi–Dirac distribution functions. Since ε = n c 2 , i.e., ε real + i ε imag = ( n + i κ ) 2 , the refraction index n and the extinction coefficient κ are obtained:
n 2 = | ε | + ε real 2 ,
κ 2 = | ε | ε real 2 .
The real part ε real is caused by various kinds of displacement polarization inside the material and represents the energy storage term of the material. The imaginary part ε imag is related to the absorption of the material, including gain and loss. Therefore, the permittivity must be real in the absence of the incident photon energy, i.e., ε ( ω = 0 ) = ε real . It is expected that the smaller the energy gap of a material, the larger its ε ( 0 ) . On the other hand, since the DOS only appears in a finite range of energy in the numerical simulation, electron transitions by the absorption of photons do not occur if the photon energy is too large. Therefore, without photon absorption the computed ε imag is close to zero in the large limit of the incident photon energy, and the permittivity approaches a constant real value.
In Figure 3, we show the real part and the imaginary part of the complex permittivity for MX2/Gr heterostructures and MX2 and Gr monolayers. Since the energies corresponding to the peak positions of the DOS in the conduction and valence bands of the MX2/Gr were smaller than those of MX2, the peak positions of the real part ε real ( ω ) of MX2/Gr in Figure 3c had a red shift compared with the MX2 in Figure 3a. The highest peak corresponding energy of the imaginary part ε imag ( ω ) of the MX2/Gr in Figure 3d was about 0.8 eV to 0.9 eV, which corresponds to the position where the real part decreases the fastest.
After obtaining the dielectric function, the absorption coefficient α is computed by
α = 2 ω κ c ,
and the reflectivity R is
R = ( n 1 ) 2 + κ 2 ( n + 1 ) 2 + κ 2 .
Figure 4 compares MX2/Gr, MX2 and Gr for their absorption coefficient α , reflectivity R and refractive index n. Since MX2/Gr has a smaller optical band gap compared with the MX2 monolayer, MX2/Gr has a wider range of light absorption, from 0.6 eV to 4 eV. As shown in Figure 4b,e, MX2/Gr had a higher reflectivity than MX2 in the infrared area ( ω < 1.2 eV ), and had a higher reflectivity than Gr in the visible light area. We compare the real part of the permittivity of monolayer MX2 in Figure 3a with the refractive index of monolayer MX2 in Figure 4c, and compare the real part of the permittivity of the MX2/Gr heterostructure in Figure 3c with that of MX2/Gr in Figure 4f. It was found that the changing trends from Figure 3a to Figure 4c are similar to those from Figure 3c to Figure 4f. This means that the real part of the dielectric constant dominates the effect of the refractive index. The above simulation results suggest that MX2/Gr heterostructures are good candidate materials for optical applications.

4. Discussion

Two-dimensional heterostructures based on TMDCs exhibit the enhancement of electrical and optoelectrical properties, which are promising for next-generation optoelectronics devices. We systematically computed the complex permittivity ε ( ω ) , absorption coefficient α ( ω ) , reflectivity R ( ω ) and refractive index n ( ω ) for MX2/Gr heterostructures, where M = Mo, W; and X = S, Se. Our results qualitatively agree with those from previous studies on MoS2/Gr [69] and WSe2/Gr [70], where red shifts in the α ( ω ) , R ( ω ) and n ( ω ) were found compared with MoS2 and WSe2 monolayers. We extended the investigations to other MX2/Gr heterostructures, and found qualitatively similar behavior in their optical properties.
It is worth comparing our MX2/Gr results with the recent simulation results on the MX2/BP in terms of absorption abilities [59]. Although different types of van der Waals heterostructures were found in MX2/BP (type-I for MoS2/BP and WS2/BP; type-II for MoSe2/BP and WSe2/BP), all the materials of MX2/BP have excellent absorption abilities in the infrared and visible light range, i.e., α ( ω ) 0.01 nm 1 to 0.05 nm 1 for the wavelength 400 nm λ 1200 nm [59]. In our study, all the MX2/Gr were type-I heterostructures, and the values of the absorption coefficients were in the same range compared with MX2/BP in the infrared and visible light range. Therefore, MX2/Gr can be utilized as alternative materials for the applications of solar optoelectronics devices.

Author Contributions

Conceptualization, C.-H.Y.; computation and data curation, C.-H.Y.; computing support, S.-T.C.; analysis, C.-H.Y.; writing, C.-H.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding. The APC was funded by C.-H.Y.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Numerical data generated during the study are available from the corresponding author.

Acknowledgments

C.-H.Y. is grateful to Yu-Chin Tzeng for his invaluable discussions and helps.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
TMDCTransition metal dichalcogenide
GrGraphene
SBHSchottky barrier height
DFTDensity functional theory
DOSDensity of states
CBMConduction band minimum
VBMValence band maximum

References

  1. Su, W.P.; Schrieffer, J.R.; Heeger, A.J. Solitons in Polyacetylene. Phys. Rev. Lett. 1979, 42, 1698–1701. [Google Scholar] [CrossRef]
  2. Haldane, F.D.M. Nonlinear Field Theory of Large-Spin Heisenberg Antiferromagnets: Semiclassically Quantized Solitons of the One-Dimensional Easy-Axis Néel State. Phys. Rev. Lett. 1983, 50, 1153–1156. [Google Scholar] [CrossRef]
  3. Xu, G.; DiTusa, J.F.; Ito, T.; Oka, K.; Takagi, H.; Broholm, C.; Aeppli, G. Y2BaNiO5: A nearly ideal realization of the S = 1 Heisenberg chain with antiferromagnetic interactions. Phys. Rev. B 1996, 54, R6827–R6830. [Google Scholar] [CrossRef]
  4. Buyers, W.J.L.; Morra, R.M.; Armstrong, R.L.; Hogan, M.J.; Gerlach, P.; Hirakawa, K. Experimental evidence for the Haldane gap in a spin-1 nearly isotropic, antiferromagnetic chain. Phys. Rev. Lett. 1986, 56, 371–374. [Google Scholar] [CrossRef] [PubMed]
  5. Renard, J.P.; Verdaguer, M.; Regnault, L.P.; Erkelens, W.A.C.; Rossat-Mignod, J.; Stirling, W.G. Presumption for a Quantum Energy Gap in the Quasi-One-Dimensional S = 1 Heisenberg Antiferromagnet Ni(C2H8N2)2NO2(ClO4). Europhys. Lett. 1987, 3, 945–952. [Google Scholar] [CrossRef]
  6. Tzeng, Y.C. Parity quantum numbers in the density matrix renormalization group. Phys. Rev. B 2012, 86, 024403. [Google Scholar] [CrossRef] [Green Version]
  7. Tzeng, Y.C.; Onishi, H.; Okubo, T.; Kao, Y.J. Quantum phase transitions driven by rhombic-type single-ion anisotropy in the S = 1 Haldane chain. Phys. Rev. B 2017, 96, 060404. [Google Scholar] [CrossRef] [Green Version]
  8. Pajerowski, D.M.; Podlesnyak, A.P.; Herbrych, J.; Manson, J. High-pressure inelastic neutron scattering study of the anisotropic S = 1 spin chain [Ni(HF2)(3-Clpyradine)4]BF4. Phys. Rev. B 2022, 105, 134420. [Google Scholar] [CrossRef]
  9. Tzeng, Y.C.; Dai, L.; Chung, M.; Amico, L.; Kwek, L.C. Entanglement convertibility by sweeping through the quantum phases of the alternating bonds XXZ chain. Sci. Rep. 2016, 6, 26453. [Google Scholar] [CrossRef]
  10. Tu, Y.T.; Tzeng, Y.C.; Chang, P.Y. Rényi entropies and negative central charges in non-Hermitian quantum systems. arXiv 2022, arXiv:2107.13006. [Google Scholar]
  11. Klitzing, K.v.; Dorda, G.; Pepper, M. New Method for High-Accuracy Determination of the Fine-Structure Constant Based on Quantized Hall Resistance. Phys. Rev. Lett. 1980, 45, 494–497. [Google Scholar] [CrossRef] [Green Version]
  12. Ng, K.K.; Chen, Y.C.; Tzeng, Y.C. Quarter-filled supersolid and solid phases in the extended Bose-Hubbard model. J. Phys. Cond. Mat. 2010, 22, 185601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Chen, Y.C.; Yang, M.F. Two supersolid phases in hard-core extended Bose-Hubbard model. J. Phys. Commun. 2017, 1, 035009. [Google Scholar] [CrossRef] [Green Version]
  14. Geim, A.K. Graphene: Status and prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Bolotin, K.I.; Sikes, K.J.; Jiang, Z.; Klima, M.; Fudenberg, G.; Hone, J.; Kim, P.; Stormer, H.L. Ultrahigh electron mobility in suspended graphene. Solid State Commun. 2008, 146, 351–355. [Google Scholar] [CrossRef] [Green Version]
  16. Chan, K.T.; Neaton, J.B.; Cohen, M.L. First-principles study of metal adatom adsorption on graphene. Phys. Rev. B 2008, 77, 235430. [Google Scholar] [CrossRef] [Green Version]
  17. Torbatian, Z.; Asgari, R. Plasmonic physics of 2D crystalline materials. Appl. Sci. 2018, 8, 238. [Google Scholar] [CrossRef] [Green Version]
  18. Koukaras, E.N.; Kalosakas, G.; Galiotis, C.; Papagelis, K. Phonon properties of graphene derived from molecular dynamics simulations. Sci. Rep. 2015, 5, 12923. [Google Scholar] [CrossRef] [Green Version]
  19. Wendler, F.; Knorr, A.; Malic, E. Ultrafast carrier dynamics in Landau-quantized graphene. Nanophotonics 2015, 4, 224–249. [Google Scholar] [CrossRef]
  20. Castro Neto, A.H.; Guinea, F.; Peres, N.M.R.; Novoselov, K.S.; Geim, A.K. The electronic properties of graphene. Rev. Mod. Phys. 2009, 81, 109–162. [Google Scholar] [CrossRef] [Green Version]
  21. Fan, Q.; Yan, L.; Tripp, M.W.; Krejčí, O.; Dimosthenous, S.; Kachel, S.R.; Chen, M.; Foster, A.S.; Koert, U.; Liljeroth, P.; et al. Biphenylene network: A nonbenzenoid carbon allotrope. Science 2021, 372, 852–856. [Google Scholar] [CrossRef] [PubMed]
  22. Al-Jayyousi, H.K.; Sajjad, M.; Liao, K.; Singh, N. Two-dimensional biphenylene: A promising anchoring material for lithium-sulfur batteries. Sci. Rep. 2022, 12, 4653. [Google Scholar] [CrossRef] [PubMed]
  23. Bafekry, A.; Faraji, M.; Fadlallah, M.M.; Jappor, H.R.; Karbasizadeh, S.; Ghergherehchi, M.; Gogova, D. Biphenylene monolayer as a two-dimensional nonbenzenoid carbon allotrope: A first-principles study. J. Phys. Cond. Mat. 2021, 34, 015001. [Google Scholar] [CrossRef] [PubMed]
  24. Ren, K.; Shu, H.; Huo, W.; Cui, Z.; Xu, Y. Tuning electronic, magnetic and catalytic behaviors of biphenylene network by atomic doping. Nanotechnology 2022. [Google Scholar] [CrossRef]
  25. Bernevig, B.A.; Hughes, T.L.; Zhang, S.C. Quantum Spin Hall Effect and Topological Phase Transition in HgTe Quantum Wells. Science 2006, 314, 1757–1761. [Google Scholar] [CrossRef] [Green Version]
  26. König, M.; Wiedmann, S.; Brüne, C.; Roth, A.; Buhmann, H.; Molenkamp, L.W.; Qi, X.L.; Zhang, S.C. Quantum Spin Hall Insulator State in HgTe Quantum Wells. Science 2007, 318, 766–770. [Google Scholar] [CrossRef] [Green Version]
  27. Vogt, P.; De Padova, P.; Quaresima, C.; Avila, J.; Frantzeskakis, E.; Asensio, M.C.; Resta, A.; Ealet, B.; Le Lay, G. Silicene: Compelling Experimental Evidence for Graphenelike Two-Dimensional Silicon. Phys. Rev. Lett. 2012, 108, 155501. [Google Scholar] [CrossRef]
  28. Coleman, J.N.; Lotya, M.; O’Neill, A.; Bergin, S.D.; King, P.J.; Khan, U.; Young, K.; Gaucher, A.; De, S.; Smith, R.J.; et al. Two-Dimensional Nanosheets Produced by Liquid Exfoliation of Layered Materials. Science 2011, 331, 568–571. [Google Scholar] [CrossRef] [Green Version]
  29. Latzke, D.W.; Zhang, W.; Suslu, A.; Chang, T.R.; Lin, H.; Jeng, H.T.; Tongay, S.; Wu, J.; Bansil, A.; Lanzara, A. Electronic structure, spin–orbit coupling, and interlayer interaction in bulk MoS2 and WS2. Phys. Rev. B 2015, 91, 235202. [Google Scholar] [CrossRef] [Green Version]
  30. Hsu, W.T.; Lu, L.S.; Wang, D.; Huang, J.K.; Li, M.Y.; Chang, T.R.; Chou, Y.C.; Juang, Z.Y.; Jeng, H.T.; Li, L.J.; et al. Evidence of indirect gap in monolayer WSe2. Nat. Commun. 2017, 8, 929. [Google Scholar] [CrossRef] [Green Version]
  31. Zhang, Y.; Chang, T.R.; Zhou, B.; Cui, Y.T.; Yan, H.; Liu, Z.; Schmitt, F.; Lee, J.; Moore, R.; Chen, Y.; et al. Direct observation of the transition from indirect to direct bandgap in atomically thin epitaxial MoSe2. Nat. Nanotechnol. 2014, 9, 111–115. [Google Scholar] [CrossRef] [PubMed]
  32. Yang, C.H.; Chung, Y.F.; Su, Y.S.; Chen, K.T.; Huang, Y.S.; Chang, S.T. Band structure of molybdenum disulfide: From first principle to analytical band model. J. Comput. Electron. 2022, 21, 571–581. [Google Scholar] [CrossRef]
  33. Kumar, S.; Schwingenschlogl, U. Thermoelectric response of bulk and monolayer MoSe2 and WSe2. Chem. Mater. 2015, 27, 1278–1284. [Google Scholar] [CrossRef]
  34. Kormányos, A.; Burkard, G.; Gmitra, M.; Fabian, J.; Zólyomi, V.; Drummond, N.D.; Fal’ko, V. k·p theory for two-dimensional transition metal dichalcogenide semiconductors. 2D Mater. 2015, 2, 022001. [Google Scholar] [CrossRef]
  35. Cheiwchanchamnangij, T.; Lambrecht, W.R.L. Quasiparticle band structure calculation of monolayer, bilayer, and bulk MoS2. Phys. Rev. B 2012, 85, 205302. [Google Scholar] [CrossRef] [Green Version]
  36. Haldar, S.; Vovusha, H.; Yadav, M.K.; Eriksson, O.; Sanyal, B. Systematic study of structural, electronic, and optical properties of atomic-scale defects in the two-dimensional transition metal dichalcogenides MX2 (M = Mo,W; X = S, Se, Te). Phys. Rev. B 2015, 92, 235408. [Google Scholar] [CrossRef] [Green Version]
  37. Gusakova, J.; Wang, X.; Shiau, L.L.; Krivosheeva, A.; Shaposhnikov, V.; Borisenko, V.; Gusakov, V.; Tay, B.K. Electronic properties of bulk and monolayer TMDs: Theoretical study within DFT framework (GVJ-2e method). Phys. Status Solidi A 2017, 214, 1700218. [Google Scholar] [CrossRef]
  38. Molina-Sánchez, A.; Wirtz, L. Phonons in single-layer and few-layer MoS2 and WS2. Phys. Rev. B 2011, 84, 155413. [Google Scholar] [CrossRef] [Green Version]
  39. Gu, X.; Yang, R. Phonon transport in single-layer transition metal dichalcogenides: A first-principles study. Appl. Phys. Lett. 2014, 105, 131903. [Google Scholar] [CrossRef] [Green Version]
  40. Chang, T.R.; Lin, H.; Jeng, H.T.; Bansil, A. Thickness dependence of spin polarization and electronic structure of ultra-thin films of MoS2 and related transition-metal dichalcogenides. Sci. Rep. 2014, 4, 6270. [Google Scholar] [CrossRef] [Green Version]
  41. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.Y.; Galli, G.; Wang, F. Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef] [PubMed]
  42. Helmrich, S.; Sampson, K.; Huang, D.; Selig, M.; Hao, K.; Tran, K.; Achstein, A.; Young, C.; Knorr, A.; Malic, E.; et al. Phonon-Assisted Intervalley Scattering Determines Ultrafast Exciton Dynamics in MoSe2 Bilayers. Phys. Rev. Lett. 2021, 127, 157403. [Google Scholar] [CrossRef] [PubMed]
  43. Gutiérrez, H.R.; Perea-López, N.; Elías, A.L.; Berkdemir, A.; Wang, B.; Lv, R.; López-Urías, F.; Crespi, V.H.; Terrones, H.; Terrones, M. Extraordinary Room-Temperature Photoluminescence in Triangular WS2 Monolayers. Nano Lett. 2013, 13, 3447–3454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Zhou, H.; Wang, C.; Shaw, J.C.; Cheng, R.; Chen, Y.; Huang, X.; Liu, Y.; Weiss, N.O.; Lin, Z.; Huang, Y.; et al. Large area growth and electrical properties of p-type WSe2 atomic layers. Nano Lett. 2015, 15, 709–713. [Google Scholar] [CrossRef]
  45. Liu, R.Y.; Lin, M.K.; Chen, P.; Suzuki, T.; Clark, P.C.J.; Lewis, N.K.; Cacho, C.; Springate, E.; Chang, C.S.; Okazaki, K.; et al. Symmetry-breaking and spin-blockage effects on carrier dynamics in single-layer tungsten diselenide. Phys. Rev. B 2019, 100, 214309. [Google Scholar] [CrossRef] [Green Version]
  46. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef] [Green Version]
  47. Yoon, Y.; Ganapathi, K.; Salahuddin, S. How Good Can Monolayer MoS2 Transistors Be? Nano Lett. 2011, 11, 3768–3773. [Google Scholar] [CrossRef]
  48. Lee, C.; Li, Q.; Kalb, W.; Liu, X.Z.; Berger, H.; Carpick, R.W.; Hone, J. Frictional Characteristics of Atomically Thin Sheets. Science 2010, 328, 76–80. [Google Scholar] [CrossRef] [Green Version]
  49. Chang, K.; Chen, W. In situ synthesis of MoS2/graphene nanosheet composites with extraordinarily high electrochemical performance for lithium ion batteries. Chem. Commun. 2011, 47, 4252–4254. [Google Scholar] [CrossRef]
  50. Jobayr, M.R.; Salman, E.M.T. Investigation of the thermoelectric properties of one-layer transition metal dichalcogenides. Chin. J. Phys. 2021, 74, 270–278. [Google Scholar] [CrossRef]
  51. Chang, T.R.; Chen, P.J.; Bian, G.; Huang, S.M.; Zheng, H.; Neupert, T.; Sankar, R.; Xu, S.Y.; Belopolski, I.; Chang, G.; et al. Topological Dirac surface states and superconducting pairing correlations in PbTaSe2. Phys. Rev. B 2016, 93, 245130. [Google Scholar] [CrossRef] [Green Version]
  52. Bian, G.; Chang, T.R.; Sankar, R.; Xu, S.Y.; Zheng, H.; Neupert, T.; Chiu, C.K.; Huang, S.M.; Chang, G.; Belopolski, I.; et al. Topological nodal-line fermions in spin–orbit metal PbTaSe2. Nat. Commun. 2016, 7, 10556. [Google Scholar] [CrossRef] [PubMed]
  53. Chang, T.R.; Xu, S.Y.; Chang, G.; Lee, C.C.; Huang, S.M.; Wang, B.; Bian, G.; Zheng, H.; Sanchez, D.S.; Belopolski, I.; et al. Prediction of an arc-tunable Weyl Fermion metallic state in MoxW1−xTe2. Nat. Commun. 2016, 7, 10639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Tzeng, Y.C.; Yang, M.F. Fate of Fermi-arc states in gapped Weyl semimetals under long-range interactions. Phys. Rev. B 2020, 102, 035148. [Google Scholar] [CrossRef]
  55. Sufyan, A.; Macam, G.; Hsu, C.H.; Huang, Z.Q.; Huang, S.M.; Lin, H.; Chuang, F.C. Theoretical prediction of topological insulators in two-dimensional ternary transition metal chalcogenides (MM’X4, M = Ta, Nb, or V; M’ = Ir, Rh, or Co; X = Se or Te). Chin. J. Phys. 2021, 73, 95–102. [Google Scholar] [CrossRef]
  56. Hlevyack, J.A.; Feng, L.Y.; Lin, M.K.; Villaos, R.A.B.; Liu, R.Y.; Chen, P.; Li, Y.; Mo, S.K.; Chuang, F.C.; Chiang, T.C. Dimensional crossover and band topology evolution in ultrathin semimetallic NiTe2 films. npj 2D Mater. Appl. 2021, 5, 40. [Google Scholar] [CrossRef]
  57. Feng, L.Y.; Villaos, R.A.B.; Cruzado, H.N.; Huang, Z.Q.; Hsu, C.H.; Hsueh, H.C.; Lin, H.; Chuang, F.C. Magnetic and topological properties in hydrogenated transition metal dichalcogenide monolayers. Chin. J. Phys. 2020, 66, 15–23. [Google Scholar] [CrossRef]
  58. Yelgel, C.; Yelgel, Ö.C.; Gülseren, O. Structural and electronic properties of MoS2, WS2, and WS2/MoS2 heterostructures encapsulated with hexagonal boron nitride monolayers. J. Appl. Phys. 2017, 122, 065303. [Google Scholar] [CrossRef] [Green Version]
  59. Ren, K.; Sun, M.; Luo, Y.; Wang, S.; Yu, J.; Tang, W. First-principle study of electronic and optical properties of two-dimensional materials-based heterostructures based on transition metal dichalcogenides and boron phosphide. Appl. Surf. Sci. 2019, 476, 70–75. [Google Scholar] [CrossRef]
  60. Phuc, H.V.; Hieu, N.N.; Hoi, B.D.; Phuong, L.T.T.; Nguyen, C.V. First principle study on the electronic properties and Schottky contact of graphene adsorbed on MoS2 monolayer under applied out-plane strain. Surf. Sci. 2018, 668, 23–28. [Google Scholar] [CrossRef]
  61. Nguyen, C.V. Tuning the electronic properties and Schottky barrier height of the vertical graphene/MoS2 heterostructure by an electric gating. Superlattices Microstruct. 2018, 116, 79–87. [Google Scholar] [CrossRef]
  62. Ma, Y.; Dai, Y.; Guo, M.; Niu, C.; Huang, B. Graphene adhesion on MoS2 monolayer: An ab initio study. Nanoscale 2011, 3, 3883–3887. [Google Scholar] [CrossRef] [PubMed]
  63. Ma, Y.; Dai, Y.; Wei, W.; Niu, C.; Yu, L.; Huang, B. First-principles study of the graphene@MoSe2 heterobilayers. J. Phys. Chem. C 2011, 115, 20237–20241. [Google Scholar] [CrossRef]
  64. Zhang, W.; Hao, G.; Zhang, R.; Xu, J.; Ye, X.; Li, H. Effects of vertical strain and electrical field on electronic properties and Schottky contact of graphene/MoSe2 heterojunction. J. Phys. Chem. Solids 2021, 157, 110189. [Google Scholar] [CrossRef]
  65. Hao, G.-Q.; Zhang, R.; Zhang, W.-J.; Chen, N.; Ye, X.-J.; Li, H.-B. Regulation and control of Schottky barrier in graphene/MoSe2 heteojuinction by asymmetric oxygen doping. Acta Phys. Sin. 2022, 71, 017104. [Google Scholar] [CrossRef]
  66. Zheng, J.; Li, E.; Ma, D.; Cui, Z.; Peng, T.; Wang, X. Effect on Schottky barrier of graphene/WS2 heterostructure with vertical electric field and biaxial strain. Phys. Status Solidi B 2019, 256, 1900161. [Google Scholar] [CrossRef]
  67. Zhang, X.; Sun, Y.; Gao, W.; Lin, Y.; Zhao, X.; Wang, Q.; Yao, X.; He, M.; Ye, X.; Liu, Y. Sizable bandgaps of graphene in 3d transition metal intercalated defective graphene/WSe2 heterostructures. RSC Adv. 2019, 9, 18157–18164. [Google Scholar] [CrossRef] [Green Version]
  68. Ma, H.-H.; Zhang, X.-B.; Wei, X.-Y.; Cao, J.-M. Theoretical study on Schottky regulation of WSe2/graphene heterostructure doped with nonmetallic elements. Acta Phys. Sin. 2020, 69, 20200080. [Google Scholar] [CrossRef]
  69. Qiu, B.; Zhao, X.; Hu, G.; Yue, W.; Ren, J.; Yuan, X. Optical Properties of Graphene/MoS2 Heterostructure: First Principles Calculations. Nanomaterials 2018, 8, 962. [Google Scholar] [CrossRef] [Green Version]
  70. Qiu, B.; Zhao, X.; Hu, G.; Yue, W.; Yuan, X.; Ren, J. Tuning optical properties of Graphene/WSe2 heterostructure by introducing vacancy: First principles calculations. Physica E 2020, 116, 113729. [Google Scholar] [CrossRef]
  71. Sun, Z.; Chu, H.; Li, Y.; Zhao, S.; Li, G.; Li, D. Theoretical investigation on electronic and optical properties of the graphene-MoSe2-graphene sandwich heterostructure. Mater. Des. 2019, 183, 108129. [Google Scholar] [CrossRef]
  72. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef] [Green Version]
  73. Taylor, J.; Guo, H.; Wang, J. Ab initio modeling of quantum transport properties of molecular electronic devices. Phys. Rev. B 2001, 63, 245407. [Google Scholar] [CrossRef] [Green Version]
  74. Taylor, J.; Guo, H.; Wang, J. Ab initio modeling of open systems: Charge transfer, electron conduction, and molecular switching of a C60 device. Phys. Rev. B 2001, 63, 121104. [Google Scholar] [CrossRef] [Green Version]
  75. Bardeen, J. Surface States and Rectification at a Metal Semi-Conductor Contact. Phys. Rev. 1947, 71, 717–727. [Google Scholar] [CrossRef]
  76. Tung, R.T. Recent advances in Schottky barrier concepts. Mater. Sci. Eng. R Rep. 2001, 35, 1–138. [Google Scholar] [CrossRef]
  77. Spicer, W.E.; Chye, P.W.; Skeath, P.R.; Su, C.Y.; Lindau, I. New and unified model for Schottky barrier and III–V insulator interface states formation. J. Vac. Sci. Technol. 1979, 16, 1422–1433. [Google Scholar] [CrossRef]
  78. Grosso, G.; Parravicini, G.P. Solid State Physics; Academic Press: Cambridge, MA, USA, 2013. [Google Scholar]
Figure 1. The stacking of the 3 × 3 MoS2 supercell and the 4 × 4 Gr supercell. (a) The side view. (b) The top view. For the other MX2/Gr, the atom–atom distances are listed in Table 2.
Figure 1. The stacking of the 3 × 3 MoS2 supercell and the 4 × 4 Gr supercell. (a) The side view. (b) The top view. For the other MX2/Gr, the atom–atom distances are listed in Table 2.
Photonics 09 00387 g001
Figure 2. The energy band structures and the DOS for the MX2/Gr van der Waals heterostructures. The partial DOS values are labeled with s, p, or d, corresponding to the orbital contributions, and the total DOS is labeled with T. (a,b) MoS2, (c,d) MoSe2, (e,f) WS2, (g,h) WSe2.
Figure 2. The energy band structures and the DOS for the MX2/Gr van der Waals heterostructures. The partial DOS values are labeled with s, p, or d, corresponding to the orbital contributions, and the total DOS is labeled with T. (a,b) MoS2, (c,d) MoSe2, (e,f) WS2, (g,h) WSe2.
Photonics 09 00387 g002
Figure 3. The complex permittivities for MX2/Gr heterostructures and MX2 and Gr monolayers as functions of incident photon energy. (a,c) The real part of the permittivity ε real . (b,d) The imaginary part of the permittivity ε imag .
Figure 3. The complex permittivities for MX2/Gr heterostructures and MX2 and Gr monolayers as functions of incident photon energy. (a,c) The real part of the permittivity ε real . (b,d) The imaginary part of the permittivity ε imag .
Photonics 09 00387 g003
Figure 4. The comparison of MX2/Gr heterostructures and MX2 and Gr monolayers according to their (a,d) absorption coefficient α , (b,e) reflectivity R and (c,f) refractive index n. The rainbow bars represent the energy range of visible light, from 1.59 eV to 3.26 eV.
Figure 4. The comparison of MX2/Gr heterostructures and MX2 and Gr monolayers according to their (a,d) absorption coefficient α , (b,e) reflectivity R and (c,f) refractive index n. The rainbow bars represent the energy range of visible light, from 1.59 eV to 3.26 eV.
Photonics 09 00387 g004
Table 1. The lattice mismatch ratios and the biding energies of TMDCs/graphene with 3 × 3 supercell of MX 2 and 4 × 4 supercell of graphene.
Table 1. The lattice mismatch ratios and the biding energies of TMDCs/graphene with 3 × 3 supercell of MX 2 and 4 × 4 supercell of graphene.
MX 2 /Gr MoS 2 /Gr MoSe 2 /Gr WS 2 /Gr WSe 2 /Gr
mismatch ratio1.83%0.10%2.04%0.01%
E b 194  meV 146  meV 178  meV 246  meV
Table 2. The optimized atom–atom distances as well as the MX 2 layer–Gr layer distances.
Table 2. The optimized atom–atom distances as well as the MX 2 layer–Gr layer distances.
MoS 2 /GrMo-SMo-CS-CC-C MoS 2 -Gr
2.404 Å4.826 Å3.089 Å1.600 Å2.973 Å
MoSe 2 /GrMo-SeMo-CSe-CC-C MoSe 2 -Gr
2.349 Å4.920 Å3.270 Å1.626 Å3.161 Å
WS 2 /GrW-SW-CS-CC-C WS 2 -Gr
2.379 Å4.774 Å3.078 Å1.596 Å2.954 Å
WSe 2 /GrW-SeW-CSe-CC-C WSe 2 -Gr
2.354 Å4.940 Å3.286 Å1.625 Å3.178 Å
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, C.-H.; Chang, S.-T. First-Principles Study of the Optical Properties of TMDC/Graphene Heterostructures. Photonics 2022, 9, 387. https://doi.org/10.3390/photonics9060387

AMA Style

Yang C-H, Chang S-T. First-Principles Study of the Optical Properties of TMDC/Graphene Heterostructures. Photonics. 2022; 9(6):387. https://doi.org/10.3390/photonics9060387

Chicago/Turabian Style

Yang, Cheng-Hsien, and Shu-Tong Chang. 2022. "First-Principles Study of the Optical Properties of TMDC/Graphene Heterostructures" Photonics 9, no. 6: 387. https://doi.org/10.3390/photonics9060387

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop