Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Self-consistent dielectric constant determination for monolayer WSe2

Open Access Open Access

Abstract

Frequency-dependent dielectric constant dispersion of monolayer WSe2, ε(ω)=ε1(ω)+i ε2(ω), was obtained from simultaneously measured transmittance and reflectance spectra. Optical transitions of the trion as well as A-, B-, and C-excitons are clearly resolved in the  ε2 spectrum. A consistent Kramers-Kronig transformation between the ε1 and  ε2 spectra support the validity of the applied analysis. It is found that the A- and B-exciton splitting in the case of the double-layer WSe2 can be attributed to the spin-orbit coupling, which is larger than that in the monolayer WSe2. In addition, the temperature-induced evolution of the A-exciton energy and its width are explained by model equations with electron-phonon interactions.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

In contrast to graphene and boron nitride that can be categorized as a gapless Dirac material and wide gap insulator, respectively [1,2], transition-metal dichalcogenides (TMDCs) such as MoS2, MoSe2, WS2, and WSe2, are semiconducting layered materials. Owing to the unique electronic and optical properties that TMDCs possess, they have attracted great interest from many fields [3–5]. Because of their finite energy gap, TMDCs can outperform graphene-based transistors in terms of the on/off ratio [6]. The large spin-orbit splitting in the case of MoS2 or WSe2 has led to long sustaining spin properties enabling novel functionalities in the field of valleytronics [7,8].

Because monolayer MoS2 or WSe2 are direct gap semiconductors with good radiative recombination efficiency, they can be utilized as light sources of atomic length scale. Due to large exciton binding energy light emission in TMDCs is featured by exciton states even at room temperature [4,9]. In the case of bulk crystals, the exciton binding energy can be determined using the following equation: Eb=μ*ε2×13.6 eV, where μ* is the reduced effective mass of electron and hole quasi-particles and ε is the dielectric constant [10]. However, in case of monolayers, the substantial field intrusion out of the host material leads to a complicated change in the binding energy; it has been shown that a series of exciton sublevels in monolayer MoS2 do not follow the Rydberg-type energy separation of the hydrogen atom, but they can be explained by field intrusion into the substrate [11]. Nevertheless, the dielectric constant and effective mass are still the most deterministic factors for the binding energy of two-dimensional (2D) monolayer excitons. While the effective mass can be obtained by measuring the energy band dispersion using angle-resolved photoemission spectroscopy [12,13], the dielectric constant can be obtained through optical characterizations.

The imaginary part of the dielectric constant, ε2(ω), which is proportional to the optical transition probability, provides valuable information about the density of energy states, including the excitons. Though the ellipsometry technique is used to obtain complex dielectric constants of monolayer MoS2 and WSe2 [14–16], it not only requires dispersion modeling, but also has difficulty in performing microscopic measurements. It has been proposed that the absorbance is proportional to the reflectance contrast in the case of atomically thin layers of graphene [17], In the opaque spectral range, however, the complex dielectric constant cannot be completely determined based on the reflectance or transmittance individually, but values of both quantities are necessary for singular extractions. Several works have reported on the determination of the dielectric constant of monolayer TMDCs based on their reflectance and transmittance [18,19]. In particular, Morozov et al. reported on the dielectric constants of monolayer TMDCs obtained using concerted transmittance and reflectance measurements at room temperature [19].

In this study, we report on the frequency-dependent dielectric constants of monolayer WSe2. By using the transfer matrix method (TMM), both real and imaginary components of the dielectric constant are obtained from the simultaneously measured reflectance and transmittance spectra. In addition, we study the temperature-induced evolution of the A-exciton and compare the dielectric constant of monolayer with that of the double layer WSe2.

2. Sample preparation and methods

Monolayer WSe2 was synthesized on a SiO2 (300 nm)/Si substrate using the chemical vapor deposition (CVD) method. The synthesized monolayer film was then transferred onto a quartz plate via the poly(methyl methacrylate) (PMMA) support method in order to perform the transmittance measurement. The flake size of the monolayer was typically ~100 μm ×100 μm, but, it depended on the growth conditions. As indicated by the optical image in Fig. 1(a), in addition to pure monolayers, some double-layers were formed as well.

 figure: Fig. 1

Fig. 1 Schematic diagram of the experimental setup for simultaneous reflection and transmission measurements. White light from the supercontinuum laser passes the monochromator and is focused onto WSe2 layers using an objective lens. Transmitted or reflected beams are simultaneously measured using photodiodes. The optical microscope image of the WSe2 layers is included in the figure.

Download Full Size | PDF

A broadband white light source from a supercontinuum laser (Compact, NKT photonics) was passed through a monochromator, and the filtered beam was then focused using an objective lens onto the monolayer or double layer region of WSe2. Both the transmitted (T) and reflected (R) intensities were simultaneously measured using silicon photodiodes. In addition, the Rq and Tq at adjacent quartz regions were measured, using which, the normalized reflectance (Rn) and transmittance (Tn), defined by Rn = R/Rq and Tn = T/Tq, respectively, were obtained. The sample was mounted inside an optical cryostat in vacuum and the lock-in technique referenced by an optical chopper was applied to suppress the detection noise.

3. Results and discussion

Figure 2 shows the Rn and Tn spectra of monolayer WSe2 measured at a temperature of 10 K. While Tn is less than unity because of the absorption by the WSe2 layer, Rn is greater than one over the entire range. In previous works, it has been demonstrated that the reflectance contrast, defined as RC=(RRq)/Rq, is positive and proportional to the absorbance in the case of graphene [17], indicating that Rn is more than one in opaque regions. In many studies on monolayer TMDCs, the reflectance was analyzed to represent the optical transition probability or the absorbance. This can be attributed, in part, to the fact that transmittance could not be measured in samples with a silicon substrate. However, as shown in Fig. 2, there is a difference in the spectral shape of the Rn and Tn spectra. Furthermore, as shown in the inset of Fig. 2, the A-exciton peak near 1.75 eV in Tn is slightly differentiated compared with that in Rn, which indicates that the reflectance alone cannot provide the precise position and shape of the exciton resonance, especially when the resonance is spectrally narrow.

 figure: Fig. 2

Fig. 2 Normalized reflectance (Rn) and transmittance (T) obtained in monolayer WSe2 at 80 K. The inset shows the comparison of Rn and Tn spectra around the sharp peak at 1.75 eV.

Download Full Size | PDF

In an optical system under study, the reflectance or transmittance is dependent on the dielectric constants of the constituent layers. In the opaque spectral region, the dielectric constant cannot be set based on reflectance alone. However, it can be unambiguously determined using both the Rn and Tn spectra. As previously mentioned, in order to obtain the dielectric constant from the measured Rn and Tn spectra, we applied the TMM approach [20–22]. According to the TMM approach, the matrix elements ABCD representing the light propagation from air to quartz through the WSe2 layer are presented as the multiples of the three sub matrices in Eq. (1), which correspond to the electric field transfer at the air/WSe2 interface, through the WSe2 layer, and at the WSe2/quartz interface, respectively.

[ABCD]=[ε+12ε12ε12ε+12]×[exp(iεωdc)11exp(iεωdc)]×[nq+ε2εnqε2εnqε2εnq+ε2ε]

Here, ω is the angular frequency and c is the speed of light. In addition, ε and nq are the dielectric constants of WSe2 and refractive index of quartz, respectively. With the ABCD elements known, the reflectance and transmittance can be obtained using R=|C/A|2 and T=|1/A|2×1/nq. In this study, based on the measured Rn and Tn spectra, the TMM approach was used in reverse to obtain the dielectric constant of the WSe2 layers. According to previous literature, d = 0.65 nm was assumed as the monolayer thickness [23]. We prepared a 2D array of (ε1,ε2), covering a broad range of real (ε1) and imaginary (ε2) components of the dielectric constant, and calculated the (Rnc, Tnc) for each set. Then, for each photon energy, a set of (ε1,ε2)was selected for which the deviation of (Rnc, Tnc) from the measured values or (RnRnc)2+(TnTnc)2 was the smallest.

In Fig. 3, the extracted dielectric constant of ε1 and ε2 are plotted for the monolayer WSe2 at 10 K. The ε2 spectrum clearly exhibits resonant transitions at the A-, B-, and C-excitons. The A-exciton at 1.748 eV is quite sharp with an FWHM linewidth of 15 meV and a large peak value of ε2=58. In contrast to the A-exciton, the B- and C-excitons are much broader with linewidths of ~35 and ~110 meV, respectively. In addition to neutral excitons, the signature of the trion or charged exciton is observed at left side of the A-exciton with an energy separation of ~30 meV consistent with the literature [9,24].

 figure: Fig. 3

Fig. 3 Real (a) and imaginary (b) spectra of the complex dielectric constant as a function of the photon energy for monolayer WSe2 at 80 K with the plots in the right highlighting the region near the A-exciton position. The notations A, B, and C in ε2(ω)  indicate positions of A-, B-, and C-excitons, respectively. The Kramers-Kronig transformation of the ε2(ω) , ε1KKR(ω) in (a), is included for comparison with ε1(ω). (c) ε1KKR(ω) spectrum in the low energy region below a photon energy of 1.4 eV obtained from the Kramers-Kronig transformation of ε2(ω) , and model fitting based on the Sellmeier single oscillator formula.

Download Full Size | PDF

While the ε2 spectrum is dominated by optical transitions at excitons, the ε1 spectrum is closely correlated by those resonances through the Kramers-Kronig relation (KKR) [25]. A closer comparison of the ε1 and ε2 spectra near the A-exciton, as shown in Fig. 3(b), indicates an abrupt change of the ε1 spectrum at the exciton peak, which is a typical signature of the KKR. In order to verify this correlation in more detail, we generated the ε1KKR(ω) through the Kramers-Kronig transformation of ε1(ω). In this transformation, in order to compensate for the high-energy optical transitions in WSe2 beyond the detection range, a single oscillator at 3.56 eV was added to ε2(ω). The consistency of the ε1KKR(ω) with ε1(ω) presented in Fig. 3(a), thus satisfying the KKR, clarifies the validity of the (ε1, ε2) extraction procedure. In the low energy region below the A-exciton energy, where no electronic transitions are allowed, ε1(ω) can be produced through the Kramers-Kronig transformation of the measured plus compensational ε2(ω) at 3.56 eV. Figure 3(c) shows that transformed ε1(ω) attains a value of ~21.1 in the zero-frequency limit. We find that the low-energy ε1 spectrum, which is indicated using a red line in Fig. 3(c), can be modeled by the Sellmeier single oscillator formula of ε1(E)=1+A0/(E2E02) with A0=183.5 eV2 and E0=3.02 eV as a function of photon energy E in the eV unit [26]. Here, the energy of the Sellmeier oscillator is differentiated from the compensational oscillator because the optical transitions within our detection range contributes as well in the Kramers-Kronig transformation.

According to Eq. (1) the extracted dielectric constant depends on the choice of the monolayer thickness d. As we varied the layer thickness from 0.5 to 0.8 nm it was found that the value of ε×d did not change noticeably. We note that this multiplied value, through the relation of σ2D=iωεε0 d with ε0 being the vacuum permittivity, corresponds to the 2D optical conductivity, which are widely used in order to explain the optical response of atomically thin monolayers [19,27,28].

The (Rn, Tn) measurement as well as the (ε1,ε2) extraction were repeated on the double-layer WSe2 region assuming a thickness of 2d. Figure 4 shows a comparison of the (ε1,ε2) spectra of the double-layer WSe2 with those of the monolayer WSe2 at 10 K. It can be observed that, for the double-layer WSe2, the energies of the A- and B-excitons are 36 meV and 8 meV lower than those of the monolayer WSe2, respectively. The energy separation between the A- and B-excitons, which is larger in the case of the double-layer WSe2, originates from the spin-orbit splitting of the valence band. It should be noted that the larger A-B splitting in the case of the double-layer over the monolayer was previously explained by the effect of interlayer interactions [29,30]. In contrast to the A- and B-excitons, the C-exciton experiences a greater energy shift of ~160 meV as the WSe2 changes from monolayer to double-layer. There have been some controversial theories about the origin of the C-excitons–whether they are due to the transition between parallel bands or Van Hove singularities at saddle points in the band structure [31,32]. We believe that the layer number dependence of C-exciton energy in combination with the theoretical interpretation might help in addressing this issue. Though the ε1 spectrum of the double-layer in Fig. 4 is accordingly differentiated from the monolayer, especially near the exciton resonances, the overall shape, including the low energy value of ε1=~26 at 1.5 eV, is similar to the monolayer.

 figure: Fig. 4

Fig. 4 ε1 and ε2 spectra of the double-layer WSe2 at 80 K compared with those of the monolayer WSe2 plotted with solid and dotted lines, respectively.

Download Full Size | PDF

As the temperature is increased, the electron-phonon interaction is accelerated and results in spectral broadening of the exciton resonances. We investigated the temperature dependence of excitons by measuring the dielectric constant at different temperatures. As is expected, the evolution of the ε2 spectrum of the monolayer WSe2 in Fig. 5 reveals a red shift and broadening of the A-, B-, and C-excitons with an increase in temperature. We plotted the temperature induced variation of A-exciton energy and width in the inset. From data fitting, we find that the energy shift follows the Varshni formula of EA(T)=EA0αT2T+β with EA0=1.749 eV, α=3.9×104 eV/K and β=158 K [33]. The temperature induced broadening, on the other hand, is well described by the Rudin’s relation of γ(T)=γ0+σT+γ'exp(ELO/kBT)1 with the Boltzmann constant kB and fitting parameters of γ0=5.1 meV, σ=0.038 meV/K, and γ'=55.4 meV [34], that takes into account the exciton interaction with longitudinal optical phonons (ELO=31.25 meV) and acoustic phonons.

 figure: Fig. 5

Fig. 5 ε2 spectrum at different temperatures from 10 to 290 K. Inset shows the temperature dependence of the energy and width of A-exciton resonance, where the dotted lines represent the model fitting of the energy and width according to the Varshni formula and Rudin’s relation, respectively.

Download Full Size | PDF

4. Conclusion

In this study, by performing TMM analysis of simultaneously measured transmittance and reflectance, frequency-dependent dielectric constant of monolayer WSe2 was obtained with self-consistency. The ε2 spectrum exhibits clear optical transitions at A-, B-, and C-excitons as well as the trion resonance. The Kramers-Kronig transformation leads to a dispersion model ε1(E)=1+A0/(E2E02) with A0=183.5 eV2 and E0=3.02 eV in the transparent low-energy region below 1.4 eV. In addition, thermal shift and broadening of A-exciton resonance acquired from temperature-dependent measurements demonstrates significant exciton relaxation because of electron-phonon interactions. We expect that self-consistent dielectric constant extraction performed in this study will be useful in characterizing other ultrathin layered materials and provide valuable optical data for TMDC-based optoelectronic devices.

Funding

National Research Foundation of Korea (NRF-2016R1A2B4009816); Research Fund of 2017 Chungnam National University; Institute for Basic Science (IBS-R011-D1).

References

1. E. Doni and G. P. Parravicini, “Energy bands and optical properties of hexagonal boron nitride and graphite,” Il Nuovo Cimento B 64(1), 117–144 (1969). [CrossRef]  

2. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos, and A. A. Firsov, “Two-dimensional gas of massless Dirac fermions in graphene,” Nature 438(7065), 197–200 (2005). [CrossRef]   [PubMed]  

3. K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, “Atomically thin MoS2: a new direct-gap semiconductor,” Phys. Rev. Lett. 105(13), 136805 (2010). [CrossRef]   [PubMed]  

4. A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C. Y. Chim, G. Galli, and F. Wang, “Emerging photoluminescence in monolayer MoS2,” Nano Lett. 10(4), 1271–1275 (2010). [CrossRef]   [PubMed]  

5. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, “Electronics and optoelectronics of two-dimensional transition metal dichalcogenides,” Nat. Nanotechnol. 7(11), 699–712 (2012). [CrossRef]   [PubMed]  

6. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, “Single-layer MoS2 transistors,” Nat. Nanotechnol. 6(3), 147–150 (2011). [CrossRef]   [PubMed]  

7. D. Xiao, G. B. Liu, W. Feng, X. Xu, and W. Yao, “Coupled spin and valley physics in monolayers of MoS2 and other group-VI dichalcogenides,” Phys. Rev. Lett. 108(19), 196802 (2012). [CrossRef]   [PubMed]  

8. Z. Zhu, Y. Cheng, and U. Schwingenschlögl, “Giant spin-orbit-induced spin splitting in two-dimensional transition-metal dichalcogenide semiconductors,” Phys. Rev. B 84(15), 153402 (2011). [CrossRef]  

9. K. He, N. Kumar, L. Zhao, Z. Wang, K. F. Mak, H. Zhao, and J. Shan, “Tightly bound excitons in monolayer WSe2.,” Phys. Rev. Lett. 113(2), 026803 (2014). [CrossRef]   [PubMed]  

10. R. S. Knox, Theory of Excitons (Academic, 1963).

11. Y. Li, A. Chernikov, X. Zhang, A. Rigosi, H. M. Hill, A. M. van der Zande, D. A. Chenet, E. M. Shih, J. Hone, and T. F. Heinz, “Measurement of the optical dielectric function of monolayer transition-metal dichalcogenides: MoS2, MoSe2, WS2, and WSe2,” Phys. Rev. B 90(20), 205422 (2014). [CrossRef]  

12. N. R. Wilson, P. V. Nguyen, K. Seyler, P. Rivera, A. J. Marsden, Z. P. Laker, G. C. Constantinescu, V. Kandyba, A. Barinov, N. D. Hine, X. Xu, and D. H. Cobden, “Determination of band offsets, hybridization, and exciton binding in 2D semiconductor heterostructures,” Sci. Adv. 3(2), e1601832 (2017). [CrossRef]   [PubMed]  

13. D. Le, A. Barinov, E. Preciado, M. Isarraraz, I. Tanabe, T. Komesu, C. Troha, L. Bartels, T. S. Rahman, and P. A. Dowben, “Spin-orbit coupling in the band structure of monolayer WSe2.,” J. Phys. Condens. Matter 27(18), 182201 (2015). [CrossRef]   [PubMed]  

14. C. Yim, M. O’Brien, N. McEvoy, S. Winters, I. Mirza, J. G. Lunney, and G. S. Duesberg, “Investigation of the optical properties of MoS2 thin films using spectroscopic ellipsometry,” Appl. Phys. Lett. 104(10), 103114 (2014). [CrossRef]  

15. S. M. Eichfeld, C. M. Eichfeld, Y. C. Lin, L. Hossain, and J. A. Robinson, “Rapid, non-destructive evaluation of ultrathin WSe2 using spectroscopic ellipsometry,” APL Mater. 2(9), 092508 (2014). [CrossRef]  

16. Y. Yu, Y. Yu, Y. Cai, W. Li, A. Gurarslan, H. Peelaers, D. E. Aspnes, C. G. Van de Walle, N. V. Nguyen, Y. W. Zhang, and L. Cao, “Exciton-dominated dielectric function of atomically thin MoS2 films,” Sci. Rep. 5(1), 16996 (2015). [CrossRef]   [PubMed]  

17. K. F. Mak, M. Y. Sfeir, Y. Wu, C. H. Lui, J. A. Misewich, and T. F. Heinz, “Measurement of the optical conductivity of graphene,” Phys. Rev. Lett. 101(19), 196405 (2008). [CrossRef]   [PubMed]  

18. B. Mukherjee, F. Tseng, D. Gunlycke, K. K. Amara, G. Eda, and E. Simsek, “Complex electrical permittivity of the monolayer molybdenum disulfide (MoS2) in near UV and visible,” Opt. Mater. Express 5(2), 447–455 (2015). [CrossRef]  

19. Y. V. Morozov and M. Kuno, “Optical constants and dynamic conductivities of single layer MoS2, MoSe2, and WSe2,” Appl. Phys. Lett. 107(8), 083103 (2015). [CrossRef]  

20. M. Born and E. Wolf, Principles of Optics: Electromagnetic Theory of Propagation, Interference and Diffraction of Light (Elsevier, 2013).

21. G. Scuri, Y. Zhou, A. A. High, D. S. Wild, C. Shu, K. De Greve, L. A. Jauregui, T. Taniguchi, K. Watanabe, P. Kim, M. D. Lukin, and H. Park, “Large excitonic reflectivity of monolayer MoSe2 encapsulated in hexagonal boron nitride,” Phys. Rev. Lett. 120(3), 037402 (2018). [CrossRef]   [PubMed]  

22. T. Zhan, X. Shi, Y. Dai, X. Liu, and J. Zi, “Transfer matrix method for optics in graphene layers,” J. Phys. Condens. Matter 25(21), 215301 (2013). [CrossRef]   [PubMed]  

23. J. A. Wilson and A. Yoffe, “The transition metal dichalcogenides discussion and interpretation of the observed optical, electrical and structural properties,” Adv. Phys. 18(73), 193–335 (1969). [CrossRef]  

24. J. S. Ross, P. Klement, A. M. Jones, N. J. Ghimire, J. Yan, D. G. Mandrus, T. Taniguchi, K. Watanabe, K. Kitamura, W. Yao, D. H. Cobden, and X. Xu, “Electrically tunable excitonic light-emitting diodes based on monolayer WSe2 p-n junctions,” Nat. Nanotechnol. 9(4), 268–272 (2014). [CrossRef]   [PubMed]  

25. M. Cardona and Y. Y. Peter, Fundamentals of Semiconductors (Springer, 2005).

26. O. N. Stavroudis and L. E. Sutton, “Rapid method for interpolating refractive index measurements,” J. Opt. Soc. Am. 51(3), 368–370 (1961). [CrossRef]  

27. H. Wang, C. Zhang, and F. Rana, “Ultrafast dynamics of defect-assisted electron-hole recombination in monolayer MoS2.,” Nano Lett. 15(1), 339–345 (2015). [CrossRef]   [PubMed]  

28. M. Merano, “Fresnel coefficients of a two-dimensional atomic crystal,” Phys. Rev. A 93(1), 013832 (2016). [CrossRef]  

29. T. Cheiwchanchamnangij and W. R. L. Lambrecht, “Quasiparticle band structure calculation of monolayer, bilayer, and bulk MoS2,” Phys. Rev. B 85(20), 205302 (2012). [CrossRef]  

30. T. Y. Jeong, B. M. Jin, S. H. Rhim, L. Debbichi, J. Park, Y. D. Jang, H. R. Lee, D. H. Chae, D. Lee, Y. H. Kim, S. Jung, and K. J. Yee, “Coherent lattice vibrations in mono- and few-layer WSe2,” ACS Nano 10(5), 5560–5566 (2016). [CrossRef]   [PubMed]  

31. A. R. Klots, A. K. Newaz, B. Wang, D. Prasai, H. Krzyzanowska, J. Lin, D. Caudel, N. J. Ghimire, J. Yan, B. L. Ivanov, K. A. Velizhanin, A. Burger, D. G. Mandrus, N. H. Tolk, S. T. Pantelides, and K. I. Bolotin, “Probing excitonic states in suspended two-dimensional semiconductors by photocurrent spectroscopy,” Sci. Rep. 4(1), 6608 (2015). [CrossRef]   [PubMed]  

32. A. Kormányos, G. Burkard, M. Gmitra, J. Fabian, V. Zólyomi, N. D. Drummond, and V. Fal’ko, “k· p theory for two-dimensional transition metal dichalcogenide semiconductors,” 2D Materials 2, 022001 (2015).

33. Y. P. Varshni, “Temperature dependence of the energy gap in semiconductors,” Physica 34(1), 149–154 (1967). [CrossRef]  

34. S. Rudin, T. L. Reinecke, and B. Segall, “Temperature-dependent exciton linewidths in semiconductors,” Phys. Rev. B Condens. Matter 42(17), 11218–11231 (1990). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 Schematic diagram of the experimental setup for simultaneous reflection and transmission measurements. White light from the supercontinuum laser passes the monochromator and is focused onto WSe2 layers using an objective lens. Transmitted or reflected beams are simultaneously measured using photodiodes. The optical microscope image of the WSe2 layers is included in the figure.
Fig. 2
Fig. 2 Normalized reflectance ( R n ) and transmittance (T) obtained in monolayer WSe2 at 80 K. The inset shows the comparison of R n and T n spectra around the sharp peak at 1.75 eV.
Fig. 3
Fig. 3 Real (a) and imaginary (b) spectra of the complex dielectric constant as a function of the photon energy for monolayer WSe2 at 80 K with the plots in the right highlighting the region near the A-exciton position. The notations A, B, and C in ε 2 ( ω )  indicate positions of A-, B-, and C-excitons, respectively. The Kramers-Kronig transformation of the ε 2 ( ω ) , ε 1KKR ( ω ) in (a), is included for comparison with ε 1 ( ω ). (c) ε 1KKR ( ω ) spectrum in the low energy region below a photon energy of 1.4 eV obtained from the Kramers-Kronig transformation of ε 2 ( ω ) , and model fitting based on the Sellmeier single oscillator formula.
Fig. 4
Fig. 4 ε 1 and ε 2 spectra of the double-layer WSe2 at 80 K compared with those of the monolayer WSe2 plotted with solid and dotted lines, respectively.
Fig. 5
Fig. 5 ε 2 spectrum at different temperatures from 10 to 290 K. Inset shows the temperature dependence of the energy and width of A-exciton resonance, where the dotted lines represent the model fitting of the energy and width according to the Varshni formula and Rudin’s relation, respectively.

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

[ A B C D ]=[ ε +1 2 ε 1 2 ε 1 2 ε +1 2 ]×[ exp( i ε ωd c ) 1 1 exp( i ε ωd c ) ]×[ n q + ε 2 ε n q ε 2 ε n q ε 2 ε n q + ε 2 ε ]
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.