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

Negative refraction in molybdenum disulfide

Open Access Open Access

Abstract

Recently, negative refractions have been demonstrated in uniaxial crystals with no necessary of negative permittivity and permeability. However, the small anisotropy parameterγin the uniaxial crystals limits the negative refraction occurrence only in a small range of the incident light angle, retarding its practical applications. In this paper, we report negative refraction induced by a pronounced anisotropic behavior in the bulk MoS2. Using the first-principles, the dielectric function and refractive index calculations confirm a uniaxial trait of MoS2 with a calculated anisotropy parameterγlarger than 2.5 in the entire range of visible wavelength. The critical incident angle to trigger a negative refraction in the bulk MoS2 is calculated up to 90°. The finite-difference time-domain simulations prove that the incident light with a density of 59.5% can be negatively refracted in a MoS2 slab with a thickness of 0.1 µm. Our results open up a new pathway for MoS2-like materials to a novel field of optical integration.

© 2015 Optical Society of America

1. Introduction

Negative refraction has attracted considerable interest because of its potential applications in realizing a “superlense” with a subwavelength resolution and observing a reversal of the Doppler shift and Vavilov-Cerenkov radiation [1]. Photonic crystals have been artificially fabricated and employed to meet the requirement of negative permittivity and permeability for the realization of the negative refraction phenomenon. In photonic crystal, the direction of the averaged Poynting vector is generally determined by group velocity while the wave vector direction is determined by Snell's law. Negative group index can be realized through a special dispersion relation of the band-gap edge, leading to a negative refraction in the photonic crystals [2]. However, the complex structural features and complicated nanofabrication processes hinder the photonic crystals in the applications of superlense [3], antenna technique [4], invisible weapons [5] as well as ultra-sensitive detection [6]. Recent research indicates that indefinite anisotropic materials with optical axes tilted to the interface can, for some directions of incidence, negatively refract the incoming wave. Negative refraction and total transmission refraction have been demonstrated at a unique type of interface in uniaxial crystals with twinning structures [7]. P. A. Belov et al. have showed the possibilities of negative refraction without backward waves by employing homogeneous transverse magnetic (TM) plane waves [8]. Liu et al. have calculated the angle size between the incident electromagnetic waves and the normal of the surface to achieve negative refraction in a uniaxial crystal [9]. It has been reported that the maximum incident angle and the bending negative refractive angle are only dependent on an anisotropy parameter [10]. The current uniaxial crystals utilized for negative refraction are generally with a small anisotropy parameter. Therefore, it results in the negative refraction occurring only in a small range of incident angles, limiting its applications.

MoS2 as the typical representative of two-dimensional layered materials has attracted considerable attentions due to their distinctive electronic [11–13] and optical [14–16] properties. Its monolayer structure consists of a layer Mo atoms sandwiched by two layers of S atoms. The bulk MoS2 is formed by stacking the sandwiched monolayer structures with weak interlayer van der Waals forces [17]. The strong interaction effect only exists in the basal plane of covalent bonds. As a result, bulk MoS2 exhibits strong anisotropic behavior. In this paper, we, for the first time, report negative refraction in the bulk MoS2 in a wide range of incident angles. Using the first-principles, we demonstrate the dielectric functions of bulk MoS2. Conditions for negative refraction in the interface between isotropic medium and anisotropic crystal have also been obtained. We confirm the negative refraction in bulk MoS2 by using the Finite-difference time-domain (FDTD) simulations and systematically examine the transmission coefficient of negatively refracted incident beam in the MoS2 slabs with different thickness.

2. Methods and simulation model

All the first-principle calculations are carried out with projector augmented wave potentials implemented in the Vienna ab initio Simulation Package (VASP) [18,19]. To optimize the structure of the bulk MoS2 (Fig. 1(a)), a well-converged Monkhorst-Pack k–point set (12 × 12 × 3) is utilized for the calculations. The exchange correlation contributions are treated with Perdew-Burke-Ernzerhof (PBE) function [20]. The plane wave cutoff energy of 500 eV is employed and a conjugate gradient scheme is used to optimize the geometries until the force on every atom is less than 1 meV/Å. However, as a ground-state theory, DFT has certain shortcomings in providing a quantitative description of optical and electronic excitations [21]. In order to obtain more accurate dielectric function and optical spectra in our calculations, we consider the effects of the electron-hole (e-h) interactions by employing a theory of single-particle transition. The electron-electron (e-e) repulsion is included by using the Green’s function (GW) [21] methods while both e-e and e-h interactions are taken into account with the GW-Bethe-Salpeter equation [22] (GW-BSE) formalism. We rely on the GW procedure together with the solution of the Bethe-Salpeter equation in the Tamm-Dancoff approximation. Strong excitonic effects [23,24] in MoS2 have been taken into account, influencing the optical properties of the MoS2. In addition, spin-orbit interaction is considered in our calculations to remove the degeneracy at the maximum of the valence band [25,26]. We started to calculate the optical spectra from the Kohn-Sham spinor wave functions, and the energies were calculated with the density-functional theory (DFT) in the general gradient approximation (GGA). The 360-unoccupied bands were used to ensure the accuracy of the dielectric function in the optical calculations. Through the Kohm-Sham wave functions and the quasiparticle energies, the optical-spectra were calculated on the level of the BSE [21]:

 figure: Fig. 1

Fig. 1 (a) The schematic illustration of the bulk MoS2 structure. (b) The calculated imaginary parts of MoS2 permittivity ϵ2(E) along x (black solid), y (red dash) and z (blue solid) axis as a function of incident photon energy. (c) The calculated real parts of MoS2 permittivity ϵ1(E) along x (black solid), y (red dash) and z (blue solid) axis as a function of incident photon energy. (d) The calculated frequency-dependent refractive indexes n along two principal dielectric axes. Red solid line represents the component parallel to the c axis and black solid line represents the component perpendicular to the c axis.

Download Full Size | PDF

(EckEvk)AvckS+k'v'c'vck|Keh|k'v'c'Av'c'k'S=ΩSAvckS,

The electronic excitations can be expressed in a basis of electron-hole pairs at a given K point, from a state in the valence band with quasiparticle energy Evk toconduction-band state with energy Eck. The Avcks is the expansion coefficients of the excitons in the electron-hole basis and the Ωs is the eigenenergy corresponding to the possible excitation energy of the system. The interaction kernel Keh describes the screened Coulomb interaction between electrons and holes, and the exchange interaction from a so-called local field effect. Under an illumination of an external electromagnetic field, the optical responses are generally governed by the Maxwell equations. The dielectric function, ε(E)=ε1(E)+i*ε2(E), bridges the electric displacements with the electric field in the constitutive relations of the Maxwell equations. Where ε1(E) is the real part, ε2(E) is the imaginary part, E is the photon energy. In the framework of BSE method, the imaginary part ε2(E) of the dielectric function can be obtained theoretically from the calculations of momentum matrix elements between the occupied and unoccupied states.

ε2(ω)ΣS|ΣcvkAvckSck|pi|vkξckξvk|δ(ΩSωΓ),

where ck|p|vk are the dipole matrix elements for electronic transitions from valence-band to conduction-band states. The broadening energy is expressed as Γ and P is the dipole polarization. The real part ϵ1(ω) can be obtained from the imaginary part ϵ2(ω) according to the Kramer-Kronig relation [27].

To confirm the negative refraction in bulk MoS2 crystals, the propagation of the electromagnetic wave was calculated by the FDTD method. The FDTD algorithm, which is direct solution of Maxwell’s time dependent curl equations, was firstly formulated by Yee [28] and further developed by Taflove and others [29,30]. Combing with the frequency-dependent dielectric function calculated by GW-BSE approach, we employed two dimensional (2D) FDTD to simulate the electromagnetic wave propagating from an isotropic medium to anisotropic medium MoS2. In our FDTD simulations, the boundary conditions were set with perfectly matched layer (PML) [31]. The dielectric function utilized for bulk MoS2 is based on our calculations. The computation domain size is 9.6 µm *12 µm. The width and the height of the middle layer (size of MoS2, Calcite and YVO4 crystals) along x and z axis is 2 and 20 µm, respectively. We set Yee’s cells with a mesh size of dx = dz = 0.0005µm as well as a Gaussian form pulse centered at 1.9 eV with an incidence angle of 45°. To study the efficiency of the negative refracted light in the bulk MoS2 for the further application in novel optical device, we also calculated the transmittance of negative refraction in the MoS2 slabs with different thickness.

3. Results and discussions

The optical property of the material is mainly dominated by ε(E) the dielectric function. To investigate the optical properties of the bulk MoS2, we calculated the dielectric function of the bulk MoS2 by using the GW-BSE approaches. The components of the imaginary part of dielectric function values ε2(E) along all three directions of x, y and z in bulk MoS2 crystal are demonstrated in the Fig. 1(b). The real part ε1(E) (Fig. 1(c)) was calculated from the imaginary part of the dielectric function by the Kramers-Kroning equations [27]. There is a huge hexagonal anisotropy existing between x (y) component and z-component for the incident photon energy from 1.3 to 3.8 eV, covering most of visible wavelength region. In brief, we found that in the bulk MoS2 crystals,εxx =εyyεzz, where εxxyy and εzz are the diagonal elements of the dielectric matrix εij. It indicates that the bulk MoS2 crystals exhibit a behavior of a uniaxial crystal. Here, we define the different components of the dielectric function perpendicular and parallel to the c axis by ε(E)=ε(E)xx+ε(E)yy2 and ε=ε(E)zz, respectively. Then the refraction index can be obtained from the frequency dependent dielectric function ε(E) through the GW-BSE approaches by a function of n=ε1+2. Figure 1(d) presents the refractive index components as a function of the incident photon energy in the bulk MoS2. To demonstrate the anisotropic property in the material, we define an anisotropy parameter as γ=none, where no=ε(E) and ne=ε(E) are ordinary and extraordinary refractive indexes, respectively. In the bulk MoS2, the calculated value of the γ is larger than 2.5 in the entire range of visible wavelength, much higher than the previous reported value in the traditional uniaxial crystals [10]. The large difference in the refractive indexes between the components perpendicular and parallel to the c axis gives a rise to a pronounced anisotropy. As a result, it could induce a strong modulation in propagation of the electromagnetic wave in the bulk MoS2.

To further evaluate the anisotropy-induced modulation of the incident electromagnetic wave, we have carried out an analysis of optical response to an external electromagnetic wave in the bulk MoS2. In our coordinate system, all the wave vectors and the optical axis are in the x-z plane, where the interface is in the x-z plane at z = 0 as shown in Fig. 2. θi is the incident angle and the angle of the optical axis φ is the angle between the optical axis of anisotropic media and the z direction. In xoz coordinate system, KtX and KtZ are the components of wave vectors of the transmitted light in the direction of x and z axis. XOZ is the coordinate system of optical axis, where KtX and KtZ are the components of wave vectors of the transmitted light in the direction of X and Z axis. The ordinary wave travels in the same propagating direction while the different travelling direction occurs for the extraordinary wave. Therefore we focus on the extraordinary wave effect in the following discussion. The extraordinary wave number vector can be written as:

{ktX=ktzcosφktxsinφktZ=ktzsinφktxcosφ
The dispersion relation of the extraordinary wave is:
(ktycosφktxsinφ)2no2+(ktysinφ+ktzcosφ)2ne2=k02
where k02=kix2+kiz2ni2, ni is the refractive index of the isotropic medium on the left-handed in Fig. 2, kix and kiz are the components of wave vectors of the incident light in the direction of x and z axis. The incident angle is calculated by the function of tanθi= kixkiz and the refractive angle of the wave vector θγ has a relation of tanθγ= ktxktz. Combining Eq. (3) and Eq. (4), we obtained the relation between incident angle θi and the refractive angle of the extraordinary wave vector θγ. Similarly, the relationship between incident angle θi and the refractive angle of the extraordinary time-averaged Poynting vector βγ was also achieved. Due to the anisotropy of uniaxial crystal, Poynting vector S and wave vector k travel in different directions. For a lossless medium, the direction of the time-averaged Poynting vector, defining the direction of the propagation of light, is normal to the k surface at any point. In Fig. 2, Stx=Etz × Hty and Stz=Etx × Hty are the components of Poynting vectors of the transmitted light in the direction of x and z axis; while βγ the refractive angle of Poynting vector expresses as tanβγ= StxStz. Since D=0, we can get the relationship between the components of wave vector and the components of Electric field EtX and EtZ in the XOZ coordinate system:
no2ktXEtX+ne2ktzEtZ=0
Where EtX and EtZ are the components of Electric field on XOZ coordinate system. The relationships between EtX, EtZ and EtX, EtZ can be described as:
{EtX=EtxsinφEtzcosφEtZ=EtxcosφEtzsinφ
Combing Eq. (5) and Eq. (6), the relation between incident angle θi and the refractive angle of the extraordinary time-averaged Poynting vector βγ can then be obtained. Based on the above analysis, the refractive angle βγ for energy flow and transmitted wave angle θγ as a function of incident angle θi were plotted in Fig. 3(a). As demonstrated in Fig. 1(d), the value of ne and no are 1.4 and 4.7, respectively at a photon energy of 1.8 eV. As long as the refractive angle βγ and transmitted wave angle θγ are both positive or negative, the extraordinary wave exhibits positive refraction as shown in Fig. 3(a); while the refractive angle βγ< 0 and transmitted wave vector angle θγ< 0, it leads to negative refraction. We attribute it to the large anisotropy parameterγexisting in the bulk MoS2 crystal. The negative refraction phenomenon can be very distinct as arranging the optical axial angle and anisotropic parameters appropriately. For eachγvalue and anisotropic refractive index, there is a difference between the corresponding optimal optical axis angle and maximum incident critical angle. The critical incident angle θic with the given direction of the optical axis can be solved under the condition of Stx  = 0,10
θic=arcsin(|no2ne2|sin2φ2ni(no2sin2φ+ne2cos2φ)1/2)
Substituting γ into Eq. (7), the maximum critical incident angle θic can be obtained as the optical axisφ is chosen by
φopt=arccos(1/(γ+1))
Therefore, the maximum incident angle θicmax is:

 figure: Fig. 2

Fig. 2 Schematic illustration for mechanism of negative refraction as an electromagnetic wave travelling from isotropic medium to uniaxial medium. It demonstrates the propagating path of time-averaged Poynting vectors Si, Sr and St together with the corresponding wave vectors Ki, Kγ and Kt.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 (a) Refractive angle βγ for energy flow (black line) and transmitted wave vector angle θγ (red line) as a function of incident angle θi through a uniaxial crystal of MoS2. (b) The dependence of the particular angle θic as a function of optical axis angle φ in the Calcite (black), YVO4 (red) and bulk MoS2 material (blue).

Download Full Size | PDF

θicmax=arcsin(1ni(non)e)

The maximum incident angle was calculated to be θicmax = 90°by the Eq. (9) to yield a negative refraction in the bulk MoS2 crystal. Figure 3(b) illustrates that the negative refraction occurs in an angle range of 0°to 90°, corresponding to the whole third quadrant of the schematic diagram shown in the Fig. 2. Compared to the bulk MoS2 (no = 4.7, ne = 1.4) with a γ value of 3.35 at a photo energy of 1.8 eV, we calculated the maximum incident angle of the Calcite (no = 1.658, ne = 1.486) and YVO4 (no = 1.993, ne = 2.215)10 with γof 1.12 and 0.90, respectively as shown in Fig. 3(b). In contrast with the maximum incident angles of 9.9° in Calcite and 12.9°in YVO4 crystals, the larger anisotropy behavior in bulk MoS2 crystal can yield negative refraction at up to a maximum incident angle of 90°. The range of incident angle triggering negative refraction in the bulk MoS2 is much larger than the previous uniaxial crystals.

To confirm our calculations of the negative refraction in bulk MoS2 crystal, we employed the 2D FDTD to simulate the electromagnetic wave propagating from an isotropic medium to such an anisotropic medium. The light intensity distribution in Fig. 4 illustrates that the refractive light stays at the same side as the incident light resulting from the negative refraction occurring in bulk MoS2 crystal; while in Calcite and YVO4 crystals, the refractive light is at different side of the incident side, indicating a positive refraction. Because most of light intensity is refracted, it is hard to demonstrate an obvious reflection phenomena at the interface. However, a simulation on an electromagnetic wave beam propagating from an isotropic medium to an anisotropic medium of bulk MoS2 derived as a function of time reveals both the reflection and refraction behaviors at the interface (See Visualization 1). There are zigzag features observed in the light intensity distribution in the middle layers. We attribute it to the standing waves resulted from the reaction between the incidence light and reflected light. For the further application of negative refraction in novel optical device, it is necessary to investigate the utilization rate of the negative refracted light. To study the efficiency of the negative refracted light in bulk MoS2, we calculated the transmittance of negative refraction in MoS2 slabs with different thickness as shown in Fig. 5(a). The intensity of the transmitted light was collected at a distance of 3.4 µm along x axis to assure the electromagnetic field totally passing through the bulk MoS2 crystal. With the increase of slab thickness, light intensity decreases. The incident light of ~59.5% is negatively refracted through a 0.1 µm thick slab as shown in our calculations. Figure 5(b) shows the negative refraction angle βγ as a function of incident angle θi in MoS2 with the slab thickness of 2 μm. It suggests that bulk MoS2 materials could provide an efficient way to realize negative refraction and be integrated with future optoelectronic devices. Considering the similarity between optical and acoustic crystal, the anisotropic density of mass and Lam'e coefficients could be utilized to modulate propagation of acoustic Wave [32], possibly finding a more efficient way in sound manipulation.

 figure: Fig. 4

Fig. 4 Light intensity distribution as an electromagnetic wave propagating from isotropic medium (n = 1) to anisotropic medium of bulk MoS2, Calcite and YVO4, respectively.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 (a) Transmittance as a function of the slab thickness of bulk MoS2 collected at a distance of 3.4 μm along the x axis. (b) Negative refraction angle βγ as a function of incident angle θi in MoS2 with the slab thickness of 2 μm.

Download Full Size | PDF

4. Summary

In summary, the inherent anisotropic optical properties of the bulk MoS2 were systematically studied by the first-principles, behaving as a uniaxial crystal with a large anisotropy parameter of γ. The region of incident angle inducing negative refraction has been achieved in our calculations, demonstrating the left-handed behavior in MoS2 crystals with no necessary of a negative permittivity and permeability. The FDTD simulations indicated that the incident light of 59.5% can be negatively refracted in a 0.1µm thick slab. Layered MoS2 provides an efficient way to realize negative refraction, leading to novel device concepts. It substantially extends the capabilities of MoS2 photonics in the future.

Acknowledgments

This research was supported by the National Natural Science Foundation of China (21373196, 11434009).

References and links

1. J. B. Pendry, “Negative Refraction Makes a Perfect Lens,” Phys. Rev. Lett. 85(18), 3966–3969 (2000). [CrossRef]   [PubMed]  

2. M. Notomi, “Negative refraction in photonic crystals,” Opt. Quantum Electron. 34(1-3), 133–143 (2002). [CrossRef]  

3. E. Cubukcu, K. Aydin, E. Ozbay, S. Foteinopoulou, and C. M. Soukoulis, “Subwavelength resolution in a two-dimensional photonic-crystal-based superlens,” Phys. Rev. Lett. 91(20), 207401 (2003). [CrossRef]   [PubMed]  

4. E. R. Brown, C. D. Parker, and E. Yablonovitch, “Radiation properties of a planar antenna on a photonic-crystal substrate,” J. Opt. Soc. Am. B 10(2), 404–407 (1993). [CrossRef]  

5. D. Schurig, J. J. Mock, B. J. Justice, S. A. Cummer, J. B. Pendry, A. F. Starr, and D. R. Smith, “Metamaterial Electromagnetic Cloak at Microwave Frequencies,” Science 314(5801), 977–980 (2006). [CrossRef]   [PubMed]  

6. G. Dolling, M. Wegener, C. M. Soukoulis, and S. Linden, “Negative-index metamaterial at 780 nm wavelength,” Opt. Lett. 32(1), 53–55 (2007). [CrossRef]   [PubMed]  

7. Y. Zhang, B. Fluegel, and A. Mascarenhas, “Total negative refraction in real crystals for ballistic electrons and light,” Phys. Rev. Lett. 91(15), 157404 (2003). [CrossRef]   [PubMed]  

8. P. A. Belov, “Backward waves and negative refraction in uniaxial dielectrics with negative dielectric permittivity along the anisotropy axis,” Microw. Opt. Technol. Lett. 37(4), 259–263 (2003). [CrossRef]  

9. Z. Liu, Z. F. Lin, and S. T. Chui, “Negative refraction and omnidirectional total transmission at a planar interface associated with a uniaxial medium,” Phys. Rev. B 69(11), 115402 (2004). [CrossRef]  

10. H. L. Luo, W. Hu, X. N. Yi, H. Y. Liu, and J. Zhu, “Amphoteric refraction at the interface between isotropic and anisotropic media,” Opt. Commun. 254(4–6), 353–360 (2006).

11. H. P. Komsa and A. V. Krasheninnikov, “Effects of confinement and environment on the electronic structure and exciton binding energy of MoS 2 from first principles,” Phys. Rev. B 86(24), 233–243 (2012). [CrossRef]  

12. E. S. Kadantsev and P. Hawrylak, “Electronic structure of a single MoS 2 monolayer,” Solid State Commun. 152(10), 909–913 (2012). [CrossRef]  

13. K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, “Atomically thin MoS60: a new direct-gap semiconductor,” Phys. Rev. B 105(13), 474–479 (2010).

14. H. Li, G. Lu, Z. Yin, Q. He, H. Li, Q. Zhang, and H. Zhang, “Optical identification of single- and few-layer MoS₂ sheets,” Small 8(5), 682–686 (2012). [CrossRef]   [PubMed]  

15. H. L. Shi, H. Pan, Y. W. Zhang, and B. I. Yakobson, “Quasiparticle band structures and optical properties of strained monolayer MoS2 and WS2,” Phys. Rev. B 87(15), 2095–2100 (2012).

16. D. Y. Qiu, F. H. da Jornada, and S. G. Louie, “Optical spectrum of MoS2: many-body effects and diversity of exciton states,” Phys. Rev. Lett. 111(21), 216805 (2013). [CrossRef]   [PubMed]  

17. S. W. Han, H. Kwon, S. K. Kim, S. Ryu, W. S. Yun, D. H. Kim, J. H. Hwang, J. S. Kang, J. Baik, H. J. Shin, and S. C. Hong, “Band-gap transition induced by interlayer van der Waals interaction in MoS2,” Phys. Rev. B 84(4), 2593–2598 (2011). [CrossRef]  

18. G. Kresse and J. Hafner, “Ab initio molecular dynamics for liquid metals,” Phys. Rev. B Condens. Matter 47(1), 558–561 (1993). [CrossRef]   [PubMed]  

19. G. Kresse and J. Furthmuller, “Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set,” Comput. Mater. Sci. 6(1), 15–50 (1996). [CrossRef]  

20. J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized gradient approximation made simple,” Phys. Rev. Lett. 78(7), 1396 (1997). [CrossRef]   [PubMed]  

21. F. Aryasetiawan and O. Gunnarson, “The GW method,” Rep. Prog. Phys. 61(3), 237–312 (1998). [CrossRef]  

22. C. Schwartz, “Solution of a Bethe-Salpeter equation,” Phys. Rev. 137(3B), 717–719 (1965). [CrossRef]  

23. A. Ramasubramaniam, “Large excitonic effects in monolayers of molybdenum and tungsten dichalcogenides,” Phys. Rev. B 86(11), 2757–2764 (2012). [CrossRef]  

24. S. Sim, J. Park, J. G. Song, C. In, Y. S. Lee, H. Kim, and H. Choi, “Exciton dynamics in atomically thin MoS2: Interexcitonic interaction and broadening kinetics,” Phys. Rev. B 88(7), 4521–4526 (2013). [CrossRef]  

25. A. Molina-Sánchez, D. Sangalli, K. Hummer, A. Marini, and L. Wirtz, “Effect of spin-orbit interaction on the optical spectra of single-layer, double-layer, and bulk MoS2,” Phys. Rev. B 88(4), 2358–2367 (2013). [CrossRef]  

26. T. Cheiwchanchamnangij and W. R. Lambrecht, “Quasiparticle band structure calculation of monolayer, bilayer, and bulk MoS2,” Phys. Rev. B 85(20), 1449–1454 (2012). [CrossRef]  

27. M. E. Orazem and B. Tribollet, “Electrochemical impedance spectroscopy,” Wiley &Sons. 44(24), 4345–4356 (2008).

28. K. S. Yee, “Numerical solution of initial boundary value problems involving maxwell’s equations in isotropic media,” IEEE Trans. Antennas and Propagation 14(3), 302–307 (1966). [CrossRef]  

29. G. Mur, “Absorbing boundary conditions for the finite-difference approximation of the time-domain electromagnetic-field equations,” IEEE Trans. Electromagnetic Compatibility 23(4), 377–382 (1981). [CrossRef]  

30. A. Taflove, “Application of the finite-difference time-domain method to sinusoidal steady-state electromagnetic-penetration problems,” IEEE Trans. Electromagnetic Compatibility 22(3), 191–202 (1980). [CrossRef]  

31. J. P. Berenger, “A perfectly matched layer for the absorption of electromagnetic waves,” J. Comput. Phys. 114(2), 185–200 (1994). [CrossRef]  

32. L. Feng, X. P. Liu, M. H. Lu, Y. B. Chen, Y. F. Chen, Y. W. Mao, J. Zi, Y. Y. Zhu, S. N. Zhu, and N. B. Ming, “Acoustic backward-wave negative refractions in the second band of a sonic crystal,” Phys. Rev. Lett. 96(1), 014301 (2006). [CrossRef]   [PubMed]  

Supplementary Material (1)

NameDescription
Visualization 1: AVI (794 KB)      Refraction angle as a function of incident angle

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 (a) The schematic illustration of the bulk MoS2 structure. (b) The calculated imaginary parts of MoS2 permittivity ϵ 2 ( E ) along x (black solid), y (red dash) and z (blue solid) axis as a function of incident photon energy. (c) The calculated real parts of MoS2 permittivity ϵ 1 ( E ) along x (black solid), y (red dash) and z (blue solid) axis as a function of incident photon energy. (d) The calculated frequency-dependent refractive indexes n along two principal dielectric axes. Red solid line represents the component parallel to the c axis and black solid line represents the component perpendicular to the c axis.
Fig. 2
Fig. 2 Schematic illustration for mechanism of negative refraction as an electromagnetic wave travelling from isotropic medium to uniaxial medium. It demonstrates the propagating path of time-averaged Poynting vectors Si, Sr and St together with the corresponding wave vectors Ki, Kγ and Kt.
Fig. 3
Fig. 3 (a) Refractive angle β γ for energy flow (black line) and transmitted wave vector angle θ γ (red line) as a function of incident angle θ i through a uniaxial crystal of MoS2. (b) The dependence of the particular angle θ ic as a function of optical axis angle φ in the Calcite (black), YVO4 (red) and bulk MoS2 material (blue).
Fig. 4
Fig. 4 Light intensity distribution as an electromagnetic wave propagating from isotropic medium (n = 1) to anisotropic medium of bulk MoS2, Calcite and YVO4, respectively.
Fig. 5
Fig. 5 (a) Transmittance as a function of the slab thickness of bulk MoS2 collected at a distance of 3.4 μm along the x axis. (b) Negative refraction angle βγ as a function of incident angle θi in MoS2 with the slab thickness of 2 μm.

Equations (9)

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

( E ck E vk ) A vck S + k ' v ' c ' vck| K eh | k ' v ' c ' A v ' c ' k ' S = Ω S A vck S ,
ε 2 ( ω ) Σ S | Σ cvk A vck S ck| p i |vk ξ ck ξ vk |δ( Ω S ωΓ ),
{ k tX = k tz cosφ k tx sinφ k tZ = k tz sinφ k tx cosφ
( k ty cosφ k tx sinφ) 2 n o 2 + ( k ty sinφ+ k tz cosφ) 2 n e 2 = k 0 2
n o 2 k tX E tX + n e 2 k tz E tZ =0
{ E tX = E tx sinφ E tz cosφ E tZ = E tx cosφ E tz sinφ
θ ic =arcsin( | n o 2 n e 2 |sin2φ 2 n i ( n o 2 sin 2 φ+ n e 2 cos 2 φ ) 1/2 )
φ opt =arccos( 1/( γ+1 ) )
θ ic max =arcsin( 1 n i ( n o n ) e )
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.