Skip to content
BY 4.0 license Open Access Published by De Gruyter September 29, 2018

Measuring the optical permittivity of two-dimensional materials without a priori knowledge of electronic transitions

  • Gwang-Hun Jung , SeokJae Yoo ORCID logo and Q-Han Park EMAIL logo
From the journal Nanophotonics

Abstract

We propose a deterministic method to measure the optical permittivity of two-dimensional (2D) materials without a priori knowledge of the electronic transitions over the spectral window of interest. Using the thin-film approximation, we show that the ratio of reflection coefficients for s and p polarization can give a unique solution to the permittivity of 2D materials within the measured spectral window. The uniqueness and completeness of our permittivity measurement method do not require a priori knowledge of the electronic transitions of a given material. We experimentally demonstrate that the permittivity of monolayers of MoS2, WS2, and WSe2 in the visible frequency range can be accurately obtained by our method. We believe that our method can provide fast and reliable measurement of the optical permittivity of newly discovered 2D materials.

1 Introduction

Atomically thin two-dimensional (2D) materials exhibit various distinct electronic and optical properties. At the early stage of 2D materials research, a single atomic layer of carbon, namely graphene, was intensively studied because this material is not only of fundamental importance to quantum electrodynamics but also has promising applications, such as in transparent electrodes [1], optical modulators [2], and infrared photodetectors [3]. In recent years, it was shown that the family of transition-metal dichalcogenides (TMDCs) can be obtained in atomically thin layers [4]. 2D TMDCs have a direct bandgap in the visible frequency region, while graphene does not. The energy bands in 2D TMDCs are also degenerate in the momentum space, namely valleys, and electrons can have the degree of freedom to populate each valley [5], [6]. The transitions at the bandgap in 2D TMDCs also give strong photoluminescence, and the polarization state of the emitted photons indicates which valleys give rise to the transitions [5], [6]. 2D TMDC-based optoelectronic devices have also been intensively studied because of the huge potential in valley-contrasting physics [7].

The permittivity of materials plays a pivotal role in the design of optoelectronic devices. For the family of 2D TMDCs, including MoS2, MoSe2, WS2, and WSe2, permittivity has been measured by ellipsometric spectroscopy [8], [9], [10], [11]. To accurately obtain the permittivity of a given material using ellipsometric spectroscopy, we need a priori knowledge of their electronic transitions outside the spectral range of interest because conventional analysis methods for ellipsometric spectroscopy data fits the parameters of known spectral functions such as Lorentzian functions. For example, for permittivity measurements of 2D TMDCs in the visible frequency range, we have to know the location of their electronic transitions in the ultraviolet (UV) wavelength range because strong electronic transitions in the UV range can affect the permittivity in the visible range [9]. Therefore, we have to define the Lorentzian functions outside the spectral range of interest in the conventional measurement of the permittivity. To obtain the permittivity accurately without a priori knowledge on the electronic transitions, we have to measure the samples in the UV range, but this is often impossible because the high-energy UV light can damage the sample.

The field of 2D materials is emerging, and novel 2D materials are being discovered daily. Furthermore, the significant changes in their optoelectronic properties brought about by phase changes in their crystal structure [12], surface functionalization [13], [14], dielectric screening effects [15], external magnetic fields [16], and optical pumping [17] are being actively reported. However, despite these findings indicating that the properties of 2D materials are potentially highly tunable, in practice this tunability often makes it hard to measure their permittivity because of the lack of a priori knowledge of the electronic transitions of novel 2D materials over the wider spectral range including high-energy bands up to tens of electronvolts (eV).

In this paper, we propose a deterministic method for measuring the electric permittivity of 2D materials. In striking contrast to conventional measurement methods in ellipsometric spectroscopy, our method does not require a priori knowledge of the electronic transitions of materials outside the measured spectral window. Given the measured data, but without a priori knowledge of the 2D material, our measurement method yields unique and robust permittivity results, while the conventional methods based on parameter-fitting may predict permittivity with a large degree of error. We demonstrate that our method can provide accurate permittivity measurements for MoS2, WS2, and WSe2 monolayers.

2 Results

2.1 Deterministic measurement of the refractive index

We suggest a method to directly measure the refractive index of 2D materials using conventional spectroscopic ellipsometers. These conventional instruments measure the complex ratio of the reflection coefficients for p- and s-polarized light, i.e. ρ=r(p)/r(s), within a finite spectral range. Figure 1 shows a schematic of the measurement procedures. Suppose that a thin film of 2D material of thickness df and refractive index nf is prepared on a dielectric substrate of refractive index nsub. The light beam impinges on the thin film from the upper half-plane with the index n0. We measure the complex reflection ratio of both the thin 2D film material on the substrate (ρf) and the bare substrate alone (ρsub). Since the thin 2D film material is atomically thin, we can approximate the complex reflection ratio by

Figure 1: Principle of the deterministic permittivity measurement.(A) Schematic of the deterministic permittivity measurement of 2D materials. (B) Complex reflection ratio of the 2D thin film material of thickness df on the substrate (ρf) and of the bare substrate (ρsub). Ef is the reflected field when the incident field Einc impinges on the sample with the angle of incidence θ0. Reflectivity can be characterized by the reflection amplitude ratio tan(ψ) and the phase difference Δ.
Figure 1:

Principle of the deterministic permittivity measurement.

(A) Schematic of the deterministic permittivity measurement of 2D materials. (B) Complex reflection ratio of the 2D thin film material of thickness df on the substrate (ρf) and of the bare substrate (ρsub). Ef is the reflected field when the incident field Einc impinges on the sample with the angle of incidence θ0. Reflectivity can be characterized by the reflection amplitude ratio tan(ψ) and the phase difference Δ.

(1)ρf(ϕ)=ρf|ϕ=0+dρfdϕ|ϕ=0ϕ=ρsub{1+(1rf(p)drf(p)dϕ1rf(s)drf(s)dϕ)|ϕ=0ϕ},

using the Taylor expansion for the optical path length ϕ=nfk0dfcosθf in the 2D thin film material, with the vacuum wavenumber k0=2π/λ, the wavelength of light λ, the propagating angle of light inside the 2D material θF, and the reflection coefficients of the thin 2D film material on the substrate for p-polarized (s-polarized) light rf(p)(rf(s)). The angle θF is determined by Snell’s law, nFsinθF=n0sinθ0, with the incidence angle of light θ0. Note that we used the relation ρsubs=(rf(p)/rf(s))|ϕ=0 to derive Eq. (1). The reflection coefficients rf(p,s) are given by

(2)rf(p,s)=r(p,s)(n0,nf,θ0)eiϕ+r(p,s)(nf,nsub,θf)eiϕeiϕ+r(p,s)(n0,nf,θ0)r(p,s)(nf,nsub,θf)eiϕ,

Then, the terms in the round bracket in Eq. (1) evaluated at ϕ=0 can be written as

(3)1rf(p,s)drf(p,s)dϕ|ϕ=0=2i{r(p,s)(n0,nsub,θ0)r(p,s)(n0,nf,θ0)}{1r(p,s)(n0,nsub,θ0)r(p,s)(n0,nf,θ0)}r(p,s)(n0,nsub,θ0){1r(p,s)(n0,nf,θ0)2},

where the reflection coefficients at the interface of the two media with the refractive indices n0 and n are given by

(4)r(p)(n0,n,θ0)=n0n2n02sin2θ0n2cosθ0n0n2n02sin2θ0+n2cosθ0,
(5)r(s)(n0,n,θ0)=n0cosθ0n2n02sin2θ0n0cosθ0+n2n02sin2θ0.

Plugging Eqs. (3–5) into Eq. (1), we can obtain the quadratic equations for the refractive index of the 2D thin film material as

(6)(nf2nsub2)+εinc(nsub2nf21)=1α(ρfρsub1),

where the function α is given by

(7)α=4ik0dfn0nsub2cos(θ0)sin2(θ0)(n02nsub2){nsub2n02+(nsub2+n02)cos(2θ0)}.

The solution of Eq. (6) can be expressed in the form

(8)nf2=12{n02+nsub2+(δρ/α)}±12{n02+nsub2+(δρ/α)}24n02nsub2,

where δρ=(ρfρsub)/ρsub is the fractional change in the complex reflection ratio.

The ambiguity of the sign in Eq. (8) can be resolved by imposing the passivity condition Im(nf)>0. The expression for the refractive index of 2D materials, Eq. (8), is the key result of this paper. We can directly obtain the refractive index nf from the measured quantities ρf and ρsub. It requires neither the fitting procedures for the spectral line shape functions nor a priori knowledge of the electronic transitions of the 2D materials. Using our method, the deterministic solution for the permittivity can be obtained from the given measurement data.

2.2 Refractive indices of monolayer TMDCs

We apply our method to monolayer TMDCs, which are promising 2D materials possessing semiconductor properties, with their direct bandgap in visible wavelengths and their spin-valley properties [5], [18]. Large-area, uniform films of monolayer MoS2, WS2, and WSe2 were purchased from SixCarbon and prepared on sapphire (Al2O3) wafers. We measured their complex reflection ratios and the permittivity of the sapphire substrate using a commercial spectroscopic ellipsometer (alpha-SE, JA Woollam), which is used throughout this paper.

Figure 2A–C shows the Raman spectra of MoS2, WS2, and WSe2 monolayers. In Figure 2A, the MoS2 monolayer presents two sharp peaks at 385.48 (E2g1) and 402.90 cm−1 (A1g), which correspond to the first-order in-plane and out-of-plane modes at the Brillouin zone center, respectively [19]. The frequency difference between the E2g1 and A1g modes is Δω=17.42 cm−1, which is indicative of the MoS2 monolayer [6]. In Figure 2B, the WS2 monolayer presents a peak at 348.93 cm−1 (A1g) with a shoulder at 330.32 cm−1 (2LA). The shoulder originates from the second-order longitudinal acoustic (LA) phonon mode, and is indicative of the WS2 monolayer [20]. In Figure 2C, the WSe2 monolayer presents a peak emerging at 250.53 cm−1 due to the first-order out-of-plane mode (A1g). It is known that this strikingly sharp peak disappears when the WSe2 layer increases in thickness to a few layers more than a bilayer [19] and that the emergence of the sharp A1g peak is indicative of a WSe2 monolayer.

Figure 2: Confirmation of monolayer TMDCs.Raman spectra of monolayers of (A) MoS2, (B) WS2, and (C) WSe2. The optical contrast of monolayer (D) MoS2, (E) WS2, and (F) WSe2 along the black lines in the optical images in the insets.
Figure 2:

Confirmation of monolayer TMDCs.

Raman spectra of monolayers of (A) MoS2, (B) WS2, and (C) WSe2. The optical contrast of monolayer (D) MoS2, (E) WS2, and (F) WSe2 along the black lines in the optical images in the insets.

We also use the optical contrast of the monolayers on the Si/SiO2 wafer to confirm that the TMDC monolayers are really monolayers. The optical contrast is defined as (IIsub)/Isub using the image intensities I (Isub) of the 2D TMDCs (the bare substrate). The interference of light from the SiO2 spacer layer on the Si substrate allows the number of layers of 2D TMDCs to be confirmed by optical contrast [21]. When the optical contrast of the 2D TMDC flakes lies between 0.25 and 0.4, we can conclude that they are composed of a monolayer [21]. First, the samples of monolayers MoS2, WS2, and WSe2 were transferred to Si/SiO2 wafers with an oxide thickness of 280 nm, and optical images were then recorded using a charge-coupled device (CCD) camera. Figure 2D–F shows the optical contrast of the monolayers of MoS2, WS2, and WSe2 along with the black lines of the optical images in the insets. The optical contrasts of our samples lie between 0.25 and 0.4, indicating that they are monolayers [21].

Figure 3 shows the permittivity spectra of monolayers of MoS2, WS2, and WSe2 in the visible spectral range, obtained using Eq. (8). For all monolayers TMDCs, the A and B exciton bands can be found using the peaks of the imaginary part of the measured permittivity near the energy of 2.0 eV. At the higher energy band near 3.0 eV, the strong interband transition (peak I) and its fine structure (peak If) are observed [8]. For monolayers WS2 and WSe2 (Figure 3B,C), their interband transitions and fine structure are clearly resolved, while for monolayer MoS2 the fine structure is barely visible (Figure 2A). By resolving these peaks, we demonstrate that the permittivity of 2D TMDCs can be obtained without a priori knowledge of the electronic transitions in the visible frequency range. In the next section, we will confirm that our method is not influenced by knowledge of the strong electronic transitions outside the spectral window.

Figure 3: The permittivity of monolayer TMDCs.Real (left column) and the imaginary part (right column) of the permittivity of monolayers of (A, B) MoS2, (C, D) WS2, and (E, F) WSe2. Solid and dashed lines, respectively, correspond to the data obtained by our method and by Tauc–Lorentz fitting with a limited number of poles (see Supporting Information for the tabulated data of the permittivity and the refractive index plotted in Figure 3).
Figure 3:

The permittivity of monolayer TMDCs.

Real (left column) and the imaginary part (right column) of the permittivity of monolayers of (A, B) MoS2, (C, D) WS2, and (E, F) WSe2. Solid and dashed lines, respectively, correspond to the data obtained by our method and by Tauc–Lorentz fitting with a limited number of poles (see Supporting Information for the tabulated data of the permittivity and the refractive index plotted in Figure 3).

2.3 Effect of a priori knowledge of high-energy electronic transitions on permittivity measurements

Conventional spectroscopic ellipsometry methods for semiconductor materials use Tauc–Lorentzian (TL) spectral line shape functions to describe the permittivity of these materials [9], [22]. The imaginary permittivity can be written as the sum of the N TL functions, i.e. Imε(E)=k=1NImεk(E) with

(9)Imεk(E)=Ak(E0)kCk(EEb)2(E2E0,k2)2+Ck2E21E  for  E>Eb,k,=0  for  EEb,k,

with the amplitude A, the broadening C, the photon energy E, the peak transition energy E0, and the bandgap Eb. The subscript k in Eq. (9) denotes the parameters of the kth pole. Each TL function corresponds to the respective electric transitions of the material. The real part of the permittivity can be obtained by the Kramers–Kronig relation [23]

(10)Reε(E)=1+2πPEbξImε(ξ)ξ2E2dξ.

Equation (9) automatically gives the constraints to the obtained permittivity satisfying the Kramers–Kronig relation, resulting in a causal description of the optical response of the materials. However, this approach requires a priori knowledge of the materials because electronic transitions outside the spectral range of interest can affect the analysis of the permittivity in the finite spectral range. For example, high-energy electronic transitions up to 30 eV should be taken into account for the permittivity of monolayer TMDCs within the narrow visible spectral range of 1.5–3.0 eV [9]. Higher energy electronic transitions are generally very strong, and their off-resonant tails can reach the visible spectral range. If these high-energy electronic transitions are neglected in the analysis of the permittivity, it may lead to multiple results satisfying the same reflectivity spectrum.

In Figure 4, we demonstrate how, when using the N TL spectral functions, high-energy electronic transitions outside the spectral range of interest influence the permittivity analysis, i.e. Eq. (9). We numerically find the TL parameters in Eq. (9) that fit the experimentally measured spectra of the reflection ratio, i.e. ρ=r(p)/r(s)=tan(Ψ)eiΔ, with the reflection amplitude ratio tan(ψ) and the phase difference Δ. The accuracy of the spectral fitting is determined for a given number of TL functions N. In contrast, our deterministic index analysis method, based on Eq. (8), does not depend on N. In Figure 4, we compare the errors made by these two permittivity analysis methods. The error can be evaluated by δψ=ψexp–ψmod and δΔ=Δexp–Δmod, where ψexp and ψmodexp and Δmod) are the reflection amplitude ratio ψ (the phase difference Δ) of the experimental results and the model results, respectively. Without a priori knowledge of the electronic transitions in the material, the function number N is determined by the number of peaks in the visible frequency range, i.e. Nvis=3 (Nvis=4) for monolayer MoS2 (WS2 and WSe2). As shown in Figure 4, TL function analysis using the function number N=Nvis does not provide a perfect fit for the experimentally measured ψexp and Δexp. Note that the error made by the TL function analysis is significant even for the A and B exciton bands, which are far from the higher energy bands with interband transitions. This implies that the optical response of monolayer TMDCs across the entire visible frequency range is also influenced by higher energy electronic transitions such as the fine structures of the interband transition, partial plasma resonances, and the full plasma excitation involving all valence electrons [8]. On the other hand, our deterministic analysis method shows reduced error compared to the TL function analysis with the function number N=Nvis. We also note that the errors δψ and δΔ in our method tend to increase in the high-energy bands near 3.0 eV because the thin-film approximation starts to fail, but this does not significantly affect the visible frequency range in which the A and B exciton bands are located. Table 1 summarizes the spectral average of the errors given by δΨ={ω1ω2δΨ(ω)dω}/(ω2ω1) and δΔ={ω1ω2δΔ(ω)dω}/(ω2ω1) for the two methods. It shows that the spectral averages of the errors of our analysis method are smaller than those of the TL analysis method.

Figure 4: The errors compared to the conventional method.(A) Reflection amplitude ratio (ψ) and (B) phase difference Δ. Solid lines: experimental results; white squares: results of our method; black squares: results of TL fitting with N poles. (C) Errors in the amplitude ratio δψ and (D) errors in the phase difference δΔ of monolayers of MoS2 (left column), WS2 (middle column), and WSe2 (right column). Solid lines with white squares show the errors in the results of our method and solid lines with black squares are the errors in the results of the TL fitting method with N poles.
Figure 4:

The errors compared to the conventional method.

(A) Reflection amplitude ratio (ψ) and (B) phase difference Δ. Solid lines: experimental results; white squares: results of our method; black squares: results of TL fitting with N poles. (C) Errors in the amplitude ratio δψ and (D) errors in the phase difference δΔ of monolayers of MoS2 (left column), WS2 (middle column), and WSe2 (right column). Solid lines with white squares show the errors in the results of our method and solid lines with black squares are the errors in the results of the TL fitting method with N poles.

Table 1:

Summary of the spectral averaged errors δψ and δΔ.

MaterialsOur deterministic analysisTauc–Lorentzian analysis
δψδΔδψδΔ
Monolayer MoS20.321.880.302.22
Monolayer WS20.231.270.272.77
Monolayer WSe20.130.880.312.63

2.4 Effect of anisotropy of the 2D materials

It is also worth noting that the permittivity of the 2D materials can be described in the tensor form

(11)ε=(ε000ε000ε)

with the in-plane permittivity ε|| and the out-of-plane permittivity ε because of material anisotropy. Using the anisotropic permittivity tensor, we can relate the in-plane and out-of-plane permittivities with the complex reflection ratio. In the same manner as when deriving Eq. (6), the relation is given by

(12)(εnsub2)+εinc(nsub2ε1)=1α(ρfρsub1).

Equation (12) shows that the in-plane permittivity ε|| and the out-of-plane permittivity ε cannot be simultaneously determined by our method. However, we can find that the effect of the out-of-plane permittivity ε becomes significant when ε is small. The zero out-of-plane permittivity, i.e. ε=0, falls into the singularity of the term nsub2/ε in Eq. (12), but only for metallic materials whose permittivity decreases to the point of becoming negative. However, for most 2D materials, the out-of-plane permittivity is unlikely to become metallic [24]. Therefore, we can speculate that the anisotropic effect on permittivity is insignificant for the optical response of 2D materials.

3 Conclusion

Although our method allows the determination of the permittivity of 2D materials without a priori knowledge, it can result in significant errors in two specific cases: (1) in high-frequency regions and (2) for large refractive index of 2D materials, because cases of both high frequencies ω=ck0 and large refractive indices nf increase the optical path length ϕ=nfk0d. In our method, a large optical path length ϕ results in the breakdown of the first-order approximation of the optical path length ϕ [Eq. (1)]. Therefore, we would like to emphasize that for these two cases careful consideration is required.

To sum up, we have provided a method to measure and analyze the permittivity of 2D materials. We showed that the first-order approximation of the optical path length ϕ for the complex reflection ratio ρ in ellipsometric spectroscopy measurements can provide the deterministic permittivity of 2D materials without a priori knowledge of the 2D material or parameter-fitting for a certain spectral function. Using monolayer MoS2, WS2, and WSe2 as test cases, we showed that our method can provide robust permittivity measurements. We expect that our method can provide a fast, robust, and accurate way to determine the optical properties of newly discovered 2D materials.

Acknowledgments

This work was supported by the Samsung Science and Technology Foundation under Project No. SSTF-BA1401-05. SeokJae Yoo was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (2017R1A6A3A11034238).

References

[1] Wang X, Zhi L, Müllen K. Transparent, conductive graphene electrodes for dye-sensitized solar cells. Nano Lett 2008;8: 323–7.10.1021/nl072838rSearch in Google Scholar PubMed

[2] Liu M, Yin X, Ulin-Avila E, et al. A graphene-based broadband optical modulator. Nature 2011;474:64–7.10.1038/nature10067Search in Google Scholar PubMed

[3] Yu T, Wang F, Xu T, Ma L, Pi X, Yang D. Graphene coupled with silicon quantum dots for high-performance bulk-silicon-based schottky-junction photodetectors. Adv Mater 2016;28:4912–9.10.1002/adma.201506140Search in Google Scholar PubMed

[4] Splendiani A, Sun L, Zhang Y, et al. Emerging photoluminescence in monolayer MoS2. Nano Lett 2010;10:1271–5.10.1021/nl903868wSearch in Google Scholar PubMed

[5] Mak KF, He K, Shan J, Heinz TF. Control of valley polarization in monolayer MoS2 by optical helicity. Nat Nanotechnol 2012;7:494–8.10.1038/nnano.2012.96Search in Google Scholar PubMed

[6] Zeng H, Dai J, Yao W, Xiao D, Cui X. Valley polarization in MoS2 monolayers by optical pumping. Nat Nanotechnol 2012;7:490–3.10.1038/nnano.2012.95Search in Google Scholar PubMed

[7] Schaibley JR, Yu H, Clark G, et al. Valleytronics in 2D materials. Nat Rev Mater 2016;1:16055.10.1038/natrevmats.2016.55Search in Google Scholar

[8] Beal AR, Hughes HP. Kramers-Kronig analysis of the reflectivity spectra of 2H-MoS2, 2H-MoSe2 and 2H-MoTe2. J Phys C Solid State Phys 1979;12:881–90.10.1088/0022-3719/12/5/017Search in Google Scholar

[9] Li Y, Chernikov A, Zhang X, et al. Measurement of the optical dielectric function of monolayer transition-metal dichalcogenides: MoS2, MoSe2, WS2 and WSe2. Phys Rev B 2014;90:1–6.10.1007/978-3-319-25376-3_5Search in Google Scholar

[10] Zhang H, Ma Y, Wan Y, et al. Measuring the refractive index of highly crystalline monolayer MoS2 with high confidence. Sci Rep 2015;5:1–7.10.1038/srep08440Search in Google Scholar PubMed PubMed Central

[11] Liu HL, Shen CC, Su SH, et al. Optical properties of monolayer transition metal dichalcogenides probed by spectroscopic ellipsometry. Appl Phys Lett 2014;105:201905.10.1063/1.4901836Search in Google Scholar

[12] Kang Y, Najmaei S, Liu Z, et al. Plasmonic hot electron induced structural phase transition in a MoS2 monolayer. Adv Mater 2014;26:6467–71.10.1002/adma.201401802Search in Google Scholar PubMed

[13] Chen X, McDonald AR. Functionalization of two-dimensional transition-metal dichalcogenides. Adv Mater 2016;28:5738–46.10.1002/adma.201505345Search in Google Scholar PubMed

[14] Nguyen EP, Carey BJ, Ou JZ, et al. Electronic tuning of 2D MoS2 through surface functionalization. Adv Mater 2015;27: 6225–9.10.1002/adma.201503163Search in Google Scholar PubMed

[15] Hüser F, Olsen T, Thygesen KS. How dielectric screening in two-dimensional crystals affects the convergence of excited-state calculations: monolayer MoS2. Phys Rev B 2013;88:245–309.10.1103/PhysRevB.88.245309Search in Google Scholar

[16] Srivastava A, Sidler M, Allain AV, Lembke DS, Kis A, Imamoglu A. Valley Zeeman effect in elementary optical excitations of monolayer WSe2. Nat Phys 2015;11:141–7.10.1038/nphys3203Search in Google Scholar

[17] Kim J, Jin C, Chen B, et al. Observation of ultralong valley lifetime in WSe2/MoS2 heterostructures. Sci Adv 2017;3:e1700518.10.1126/sciadv.1700518Search in Google Scholar PubMed PubMed Central

[18] Xiao D, Bin Liu G, Feng W, Xu X, Yao W. Coupled spin and valley physics in monolayers of MoS2 and other group-VI dichalcogenides. Phys Rev Lett 2012;108:1–5.10.1103/PhysRevLett.108.196802Search in Google Scholar PubMed

[19] Tonndorf P, Schmidt R, Böttger P, et al. Photoluminescence emission and Raman response of monolayer MoS2, MoSe2, and WSe2. Opt Exp 2013;21:4908.10.1364/OE.21.004908Search in Google Scholar PubMed

[20] Berkdemir A, Gutiérrez HR, Botello-Méndez AR, et al. Identification of individual and few layers of WS2 using Raman spectroscopy. Sci Rep 2013;3:1755.10.1038/srep01755Search in Google Scholar

[21] Benameur MM, Radisavljevic B, Héron JS, Sahoo S, Berger H, Kis A. Visibility of dichalcogenide nanolayers. Nanotechnology 2011;22:125706.10.1088/0957-4484/22/12/125706Search in Google Scholar PubMed

[22] Liu HL, Shen CC, Su SH, Hsu CL, Li MY, Li LJ. Investigation of the optical properties of MoS2 thin films using spectroscopic ellipsometry. Appl Phys Lett 2014;104:103114.10.1063/1.4868108Search in Google Scholar

[23] Jackson JD. Classical electrodynamics. Weinheim, Germany: Wiley-VCH Verlag GmbH & Co. KGaA, 1998.Search in Google Scholar

[24] Laturia A, Van de Put ML, Vandenberghe WG. Dielectric properties of hexagonal boron nitride and transition metal dichalcogenides: from monolayer to bulk. NPJ 2D Mater Appl 2018;2:6.10.1038/s41699-018-0050-xSearch in Google Scholar


Supplementary Material

The online version of this article offers supplementary material (https://doi.org/10.1515/nanoph-2018-0120).


Received: 2018-08-12
Revised: 2018-09-14
Accepted: 2018-09-18
Published Online: 2018-09-29

©2019 Q-Han Park et al., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 Public License.

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2018-0120/html
Scroll to top button