Skip to content
BY 4.0 license Open Access Published by De Gruyter October 28, 2020

High refractive index and extreme biaxial optical anisotropy of rhenium diselenide for applications in all-dielectric nanophotonics

  • Anton A. Shubnic , Roman G. Polozkov , Ivan A. Shelykh and Ivan V. Iorsh ORCID logo EMAIL logo
From the journal Nanophotonics

Abstract

We establish a simple quantitative criterium for the search of new dielectric materials with high values of refractive index in the visible range. It is demonstrated, that for light frequencies below the bandgap, the latter is determined by the dimensionless parameter η calculated as the ratio of the sum of the widths of conduction and valence bands and the bandgap. Small values of this parameter, which can be achieved in materials with almost flat bands, lead to dramatic increase of the refractive index. We illustrate this rule with a particular example of rhenium dichalcogenides, for which we perform ab initio calculations of the band structure and optical susceptibility and predict the values of the refractive index n>5 in a wide frequency range around 1 eV together with comparatively low losses. Our findings open new perspectives in search for the new high-index/low-loss materials for all-dielectric nanophotonics.

1 Introduction

All-dielectric photonics [1], [2] is arguably the most rapidly evolving field of modern nano-optics. The basic component in all-dielectric photonics, a dielectric nanoantenna, supports optical Mie resonances [3], with properties that can be flexibly controlled by the geometry of a nanoantenna. In particular, the variation of its shape allows changing the nature of the lowest energy optical resonance from electric dipole to magnetic dipole [4], [5], the functionality inaccessible in plasmonic devices. With the current stage of technology, the fabrication of specifically designed arrays of resonant antennae with finely controlled shape and lattice geometry became a routine task, which paved the way toward unprecedented control over linear [6], [7], [8] and nonlinear [9], [10] light manipulation.

The key factor defining the functionality of resonant all-dielectric nanostructures, besides their shape, is the value of the refractive index of the material forming resonant nanoantennas. Indeed, the quality factor of the fundamental Mie resonance scales is n2, and the ratio of the nanoantenna size to the resonant wavelength is 1/n. Thus, the increase in the refractive index would allow for both more sharp and more deep subwavelength resonances in all-dielectric nanostructures, provided that the extinction coefficient, related to the imaginary part of the refractive index, is kept small.

While virtually purely real refractive indices n>10 can be easily achieved in the wide frequency range from terahertz to mid-IR, in the near-IR and visible ranges, the values of refractive index are currently limited by a value of approximately 5. Around a characteristic photon energy of 1 eV, standard high-index reference materials are crystalline silicon (n4), gallium arsenide (n3.7), and germanium (n4.5). It is clear that the broadband large refractive index in this frequency range is provided by strong interband polarization. However, the questions of what are the other parameters which can contribute to high values of the off-resonant refractive index, are there any fundamental upper limit for this quantity, and whether it can be substantially increased with respect to the values characterizing current high-index materials remain unexplored to our knowledge.

In this work, we show the sum of the widths of the conduction Δc and valence Δv bands to the bandgap.

(1)η=Δc+ΔvEG

is the key parameter which defines the value of the refractive index for the frequencies slightly below the bandgap. We derive a simplified estimation for the susceptibility and demonstrate that the refractive index can be substantially increased if η becomes small, as it happens in materials with flattened valence and conduction bands. This situation takes place in ReSe2, for which we perform the ab initio analysis of the optical properties and predict that for the photon energies around 1 eV, the real part of the refractive index n>5, which is the current record high value, and its imaginary part remains small. We also predict extreme biaxial optical anisotropy in this material.

2 Theoretical model

We start from considering a simplified model of interband polarization in bulk material and show that the condition of the weak dependence of the bandgap on the wave vector in the whole Brillouin zone (BZ) (the extreme case being a material with flat bands) results in a high value of optical susceptibility close to the absorption edge.

We start from the simplest expression for the susceptibility tensor given by the Kubo formula

(2)χα,β(ω)=ωΩNkk,m,nf(ϵnk)f(ϵmk)ϵnkϵmknk|jα|mkmk|jβ|nkω(ϵnkϵmk)+iγ,

where α,β=x,y,z,Ω is the unit cell volume, Nk is the number of points in the BZ, over which we make the summation, f is the Fermi–Dirac distributions, |nk is the Bloch eigenfunctions for the corresponding band, and jα is the current operators given by:

(3)jˆ=ekHk,

where Hk is the hamiltonian in the basis of Bloch eigenfunctions.

Let us consider the simplest case of a cubic unit cell, for which Ω=D3, where D is the lattice period. We limit ourselves to the case of single conduction and valence bands separated by the bandgap EG, assuming that EGkT and putting fc=0,fv=1. We then can rewrite the expression for the susceptibility as:

(4)χα,β(ω)=α(k0D)(2m0D2)21NkkD2Pα(k)Pβ(k)[ϵc(k)ϵv(k)][ω(ϵc(k)ϵv(k))+iγ]

where α1/137 is the fine structure constant, k0=ω/c, m0 is the free electron mass, and Pα(k)=d3ruc,k*(r)i(/rαuv,k(r) is the matrix element of interband polarization. The presence of the prefactor α/(k0D) allows understanding why one usually observes large permittivities at longer wavelengths. Indeed for the frequency 2πω=1GHz and typical unit cell size D5.7×1010m (lattice constant of germanium), the term α/(k0D) can be as large as 106. On the other hand, in the range of optical frequencies ω1eV, α/(k0D)2.5. The characteristic energy 2/(m0D2)0.5 eV is comparable with the bandgap width.

At the first sight, it seems that the recipe for the large value of susceptibility is straightforward: if the material has ϵc(k)ϵv(k)const(k)=ϵ˜, then tuning the frequency slightly below ϵ˜, we can achieve arbitrary large susceptibility with vanishing losses. Unfortunately, this would not work so directly because conduction and valence band dispersions and matrix elements P are directly related to each other. This can be easily seen if we use kp perturbation theory. As a very rough approximation, we can assume that the dispersion in the whole BZ is parabolic and isotropic (which of course is not the case in real materials) and write

(5)ϵc(k)ϵv(k)=EG+2k2(12mc12mv)=EG+2k22μ=EG+2k22m0(42|P|2m0EG),

where the last equality follows from two band kp perturbation theory. We can see that the constant energy difference between the two bands can be achieved only if |P|2 vanishes, but this means that optical transitions between valence and conduction bands are forbidden, which corresponds to zero net susceptibility. The dependence of χ on the bandwidths is therefore determined by the competition between the numerator and denominator in Eq. (4).

If we set ω=EGδ,δEG and go from summation to the integration over the BZ, we can derive simple analytical expression for susceptibility:

(6)χ(ω)α(kEGD)1π2tan1(πη)212tan1(πη/2)η,

where kEG=EG/(c) and η=2/(μD2EG)=(Δc+Δv)/EG. This simplified expression substantially underestimates the dielectric constant for real semiconductors (e.g., for GaAs, it gives the estimate for ϵ7 instead of the experimental value ϵ13.5). This is mainly due to the effects of nonparabolicity of the electron dispersion and neglect of the interband polarization corresponding to the higher bands. Nevertheless, Eq. (6) allows determining qualitatively the dependence of the susceptibility on η, as it contains the universal decaying function of this parameter. Therefore, although the matrix element of the interband transition decays with decrease in η, maximal values of the susceptibility should be still expected for materials with small η, i.e., for those having large gap and narrow conduction and valence bands.

Fortunately, one can find representatives of this class of the materials, the example being bulk ReSe2 [11]. This layered material belonging to the rhenium dichalcogenide family has been gaining recently increasing attention, mainly owing to its pronounced in-plane anisotropy and suppressed interlayer van der Waals coupling [12]. To check the results of our qualitative analysis, in the next section, we provide the data of ab initio modeling of the linear optical response of bulk ReSe2.

3 Results of the ab initio modeling

All ab initio calculations were performed using QUANTUM ESPRESSO package [13]. The analysis of the optical response was performed in three steps.

In the first step, we determined the equilibrium positions of the ions in the lattice by full self-consistent geometry optimization within density functional theory (DFT). The top and side views of the resulting structure are presented in Figure 1. The obtained lattice parameters and their comparison to experimental XRD data are given in Supplementary.

Figure 1: The unit cell, side and top views, of the ReX2${\text{ReX}}_{2}$ bulk structure. The a, b, and c are crystallographic axes marking the unit cell. Rhenium and sulfur (or selenium) atoms are indicated in blue and yellow, respectively. It can be seen that the set of the elementary cells forms a quasi–two-dimensional bulk material. The x, y, z, axes are marking the cartesian coordinates system.
Figure 1:

The unit cell, side and top views, of the ReX2 bulk structure. The a, b, and c are crystallographic axes marking the unit cell. Rhenium and sulfur (or selenium) atoms are indicated in blue and yellow, respectively. It can be seen that the set of the elementary cells forms a quasi–two-dimensional bulk material. The x, y, z, axes are marking the cartesian coordinates system.

In the second step, we employed two different DFT approaches (LDA and GGA) to calculate the three-dimensional (3D) band structure.

In the third step, we calculated the diagonal components of the unit cell polarizability tensor αii via time-dependent DFT (TDDFT) (see details in Supplementary).

Most of the previous calculations of the band structure of the considered materials focused only on the dispersion along the special sets of directions, example being the 2D projection into the layers plane [14], [15]. In the other cases, highly symmetric paths were chosen in the 3D BZ connecting 3D special points, without exploring the structure of the whole BZ [16], [17] and thus missing the true band edges, which are crucially important for the determination of the optical response. The possible drawbacks of this approach were analyzed in detail in the recent work [11], wherein the necessity of the calculation of the band energies everywhere within the BZ and determination of the constant energy surfaces was stressed. In the present paper, we used the k-path, developed in the study by Gunasekara et al. [11], and calculated the band structure using the following path: –CBM–CBM′–ZΓ–VBM–VBM′–ZΓ, where CBM is the conduction band minimum (LUMO), VBM is the valence band maximum (HOMO), and CBM′ and VBM′ are the intersection points of the boundary surface of the BZ and the straight lines connecting Γ and CBM or VBM points, respectively. The paths from CBM′ or VBM′ points to the Z point run along the surface of the BZ containing the Z point. To find the k-points corresponding to the VBM and CBM, we used the different grids in BZ–6 × 6 × 6 and 7 × 7 × 7. The calculated values of the direct and indirect bandgap are presented in Table 1 and compared with the previously reported data [11], [16].

Table 1:

Bandgap values calculated within LDA and GGA.

Type/MaterialLDA/GGALDA [11]/GGA [11]GGA + GW [16]
Indirect ReSe21.0043/1.04630.87/0.99
Direct (Z) ReSe21.0714/1.09090.97/1.001.38
Direct (Γ) ReSe21.3262/1.3421
Indirect ReSe21.1424/1.1790
Direct (Z) ReSe21.1908/1.21471.60
Direct (Γ) ReSe21.6809/1.74411.88
  1. CBM, conduction band minimum; VBM, valence band maximum.

  2. The grid 6 × 6 × 6 was used to define the VBM point for ReSe2 and CBM point for ReS2, and the grid 7 × 7 × 7 was used to define the CBM point for ReSe2 and VBM point for ReS2. This choice was dictated by the reasons of grid parity and symmetry of the BZ. All energies are given in electron volt.

The obtained band structures of ReSe2 and ReS2 are presented in Figure 2. It can be seen that the energy of the lowest conduction and the highest valence bands depend only weakly on the wavevector, and the bands are close to the flat ones.

Figure 2: The band structure of the bulk ReSe2 and ReS2 calculated by the GGA level of DFT for the path going through Γ${\Gamma}$ and Z points and the band extrema (CBM and VBM). The Fermi energy is set to zero. The CBM′ and VBM′ points are the intersection points of the boundary surface of the BZ and the straight lines connecting the Γ${\Gamma}$ point and CBM or VBM points, respectively. The bands are flattened in a wide range of the wave vectors. CBM and VBM points correspond to the different k-values, so this is an indirect semiconductor. CBM, conduction band minimum; VBM, valence band maximum; DFT, density functional theory.
Figure 2:

The band structure of the bulk ReSe2 and ReS2 calculated by the GGA level of DFT for the path going through Γ and Z points and the band extrema (CBM and VBM). The Fermi energy is set to zero. The CBM′ and VBM′ points are the intersection points of the boundary surface of the BZ and the straight lines connecting the Γ point and CBM or VBM points, respectively. The bands are flattened in a wide range of the wave vectors. CBM and VBM points correspond to the different k-values, so this is an indirect semiconductor. CBM, conduction band minimum; VBM, valence band maximum; DFT, density functional theory.

In order to get the the optical response, we employ the TDDFT approach, calculating the diagonal terms of the unit cell polarizability αii. The average dielectric susceptibility χii(ω) was found as the ratio of the polarization and the unit cell volume V, χii=αii/V. The dielectric permittivity was then found as ϵii=1+4πχii. The spectra of the dielectric permittivities are shown in Figure 3.

Figure 3: Real and imaginary parts of ε of the bulk ReSe2 (left) and ReS2 (right) for different light polarizations calculated within TDDFT (see details of calculations in Supplementary). It can be seen that the peaks of the imaginary part of the permittivity, and hence the absorption edges, are 1.32 eV for ReSe2 and 1.7 eV for ReS2. TDDFT, time-dependent density functional theory.
Figure 3:

Real and imaginary parts of ε of the bulk ReSe2 (left) and ReS2 (right) for different light polarizations calculated within TDDFT (see details of calculations in Supplementary). It can be seen that the peaks of the imaginary part of the permittivity, and hence the absorption edges, are 1.32 eV for ReSe2 and 1.7 eV for ReS2. TDDFT, time-dependent density functional theory.

It can be seen that for both ReSe2 and ReS2, the absorption sharply increases only when the photon frequency approaches the bandgap in Γ point, 1.32 and 1.7 eV, respectively. Furthermore, we observe strong biaxial anisotropy for both crystals, stemming both from the layered structure of the material and reduced in-plane symmetry of the rhenium dichalcogenides [18]. Noteworthily, large refractive index contrast and strong uniaxial anisotropy have been recently experimentally measured in bulk transition metal dichalcogenides [19], [20], [21] and hexagonal boron nitride [22]. The pronounced biaxial anisotropy may be used for the the observation of the Dyakonov surface waves [23] in the rhenium dichalcogenide waveguides.

One can see that the real parts of the in-plane components of the permittivity of ReSe2 exceed 25 in the broad frequency range, which should correspond to the high refractive index. Since for biaxial crystals, the refractive index depends on the light propagation direction, we limit ourselves to the case when light travels along one of the three principal axes. In this case, the refractive index for each of the two polarizations is just a square root of the corresponding permittivity. The spectra of the refractive index components are shown in Figure 4. One sees that in the broad wavelength range, the refractive index exceeds 5, which is substantially larger than the index of the modern state-of-the-art high-index materials such as c-Si [24], GaP [25], and Ge [26]. The refractive index of ReS2 is smaller than that of ReSe2, but is still quite large. It should be noted that the imaginary part of the refractive index in this wavelength range is negligibly small.

Figure 4: The calculated dependence of the the real and imaginary parts of refractive indexes of ReSe2 (left) and ReS2 (right) on the wavelength for different cases of light polarizations. The known experimental data for the modern state-of-the-art high-index materials such as c-Si [24], GaP [25], and Ge [26] are shown for comparison.
Figure 4:

The calculated dependence of the the real and imaginary parts of refractive indexes of ReSe2 (left) and ReS2 (right) on the wavelength for different cases of light polarizations. The known experimental data for the modern state-of-the-art high-index materials such as c-Si [24], GaP [25], and Ge [26] are shown for comparison.

In our analysis, we have not accounted for the excitonic contribution to the dielectric permittivity, which can be substantial and would lead to the additional absorption losses at the frequencies slightly below the bandgap. Nevertheless, since the range of the large refractive index covers more than 0.5 eV below the bandgap, it should be possible to detune from the exciton absorption lines to probe the lossless large refractive index.

4 Conclusions

We established a simple quantitative criterium for the search of high-refractive-index dielectric materials, expressed in terms of the single dimensionless parameter calculated as the ratio of the sum of the widths of conduction and valence bands and the bandgap. With the use of it, we found that ReSe2 is low-loss material, possessing a record high refractive index n>5 in a wide frequency range around 1 eV. Our conclusion is supported by the results of the DFT simulation of the optical response. Since currently there exist vast material databases containing the data on their band structures, the proposed criterion can be used for the automated search of the perspective candidates for the novel high-index materials for all-dielectric photonics.


Corresponding author: Ivan V. Iorsh, Department of Physics and Engineering, ITMO University, Saint-Petersburg, Russia, E-mail:

Funding source: Russian Science Foundation

Award Identifier / Grant number: 20-12-00224

Funding source: Ministry of Education and Science of the Russian Federation 10.13039/501100003443

Award Identifier / Grant number: 14.Y26.31.0015

Funding source: Icelandic research fund

Award Identifier / Grant number: 163082-051

  1. Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: The authors acknowledge the support from the mega-grant No. 14.Y26.31.0015 of the http://dx.doi.org/10.13039/501100003443, “Ministry of Education and Science of the Russian Federation.” I.A.S. acknowledges the support from the Icelandic research fund, grant No. 163082-051.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

[1] A. I. Kuznetsov, A. E. Miroshnichenko, M. L. Brongersma, Y. S. Kivshar, and B. Luk’yanchuk, “Optically resonant dielectric nanostructures,” Science, vol. 354, no. 6314, p. aag2472, 2016, https://doi.org/10.1126/science.aag2472.10.1126/science.aag2472Search in Google Scholar PubMed

[2] Y. Kivshar, “All-dielectric meta-optics and non-linear nanophotonics,” Natl. Sci. Rev., vol. 5, no. 2, pp. 144–158, 2018, https://doi.org/10.1093/nsr/nwy017.10.1093/nsr/nwy017Search in Google Scholar

[3] G. Mie, “Beiträge zur optik trüber medien, speziell kolloidaler metallösungen,” Ann. Phys., vol. 330, no. 3, pp. 377–445, 1908, https://doi.org/10.1002/andp.19083300302.10.1002/andp.19083300302Search in Google Scholar

[4] A. B. Evlyukhin, S. M. Novikov, U. Zywietz, et al., “Demonstration of magnetic dipole resonances of dielectric nanospheres in the visible region,” Nano Lett., vol. 12, no. 7, pp. 3749–3755, 2012, https://doi.org/10.1021/nl301594s.10.1021/nl301594sSearch in Google Scholar PubMed

[5] A. I. Kuznetsov, A. E. Miroshnichenko, Y. H. Fu, J. Zhang, and B. Luk’Yanchuk, “Magnetic light,” Sci. Rep., vol. 2, p. 492, 2012, https://doi.org/10.1038/srep00492.10.1038/srep00492Search in Google Scholar PubMed PubMed Central

[6] M. Decker, I. Staude, M. Falkner, et al., “High-efficiency dielectric huygens’ surfaces,” Adv. Opt. Mater., vol. 3, no. 6, pp. 813–820, 2015, https://doi.org/10.1002/adom.201400584.10.1002/adom.201400584Search in Google Scholar

[7] A. Arbabi, Y. Horie, M. Bagheri, and A. Faraon, “Dielectric metasurfaces for complete control of phase and polarization with subwavelength spatial resolution and high transmission,” Nat. Nanotechnol., vol. 10, no. 11, pp. 937–943, 2015, https://doi.org/10.1038/nnano.2015.186.10.1038/nnano.2015.186Search in Google Scholar PubMed

[8] S. Jahani and Z. Jacob, “All-dielectric metamaterials,” Nat. Nanotechnol., vol. 11, no. 1, pp. 23–36, 2016, https://doi.org/10.1038/nnano.2015.304.10.1038/nnano.2015.304Search in Google Scholar PubMed

[9] M. R. Shcherbakov, P. P. Vabishchevich, A. S. Shorokhov, et al., “Ultrafast all-optical switching with magnetic resonances in nonlinear dielectric nanostructures,” Nano Lett., vol. 15, no. 10, pp. 6985–6990, 2015, https://doi.org/10.1021/acs.nanolett.5b02989.10.1021/acs.nanolett.5b02989Search in Google Scholar PubMed

[10] C. Schlickriede, S. S. Kruk, L. Wang, B. Sain, Y. Kivshar, and T. Zentgraf, “Nonlinear imaging with all-dielectric metasurfaces,” Nano Lett., vol. 20, no. 6, pp. 4370–4376, 2020, https://doi.org/10.1021/acs.nanolett.0c01105.10.1021/acs.nanolett.0c01105Search in Google Scholar PubMed

[11] S. M. Gunasekera, D. Wolverson, L. S. Hart, and M. Mucha-Kruczynski, “Electronic band structure of rhenium dichalcogenides,” J. Electron. Mater., vol. 47, pp. 4314–4320, 2018, https://doi.org/10.1007/s11664-018-6239-0.10.1007/s11664-018-6239-0Search in Google Scholar

[12] H. Zhao, J. Wu, H. Zhong, et al., “Interlayer interactions in anisotropic atomically thin rhenium diselenide,” Nano Res., vol. 8, no. 11, pp. 3651–3661, 2015, https://doi.org/10.1007/s12274-015-0865-0.10.1007/s12274-015-0865-0Search in Google Scholar

[13] P. Giannozzi, S. Baroni, N. Bonini, et al., “Quantum espresso: A modular and open-source software project for quantum simulations of materials,” J. Phys. Condens. Matter, vol. 21, no. 39, p. 395502, 2009, https://doi.org/10.1088/0953-8984/21/39/395502.10.1088/0953-8984/21/39/395502Search in Google Scholar PubMed

[14] K. Dileep, R. Sahu, S. Sarkar, S. C. Peter, and R. Datta, “Layer specific optical band gap measurement at nanoscale in mos2 and res2 van der waals compounds by high resolution electron energy loss spectroscopy,” J. Appl. Phys., vol. 119, no. 11, p. 114309, 2016, https://doi.org/10.1063/1.4944431.10.1063/1.4944431Search in Google Scholar

[15] W. Wen, Y. Zhu, X. Liu, et al., “Anisotropic spectroscopy and electrical properties of 2d res2(1–x)se2x alloys with distorted 1t structure,” Small, vol. 13, no. 12, p. 1603788, 2017, https://doi.org/10.1002/smll.201603788.10.1002/smll.201603788Search in Google Scholar PubMed

[16] J. P. Echeverry and I. C. Gerber, “Theoretical investigations of the anisotropic optical properties of distorted 1tres2 and rese2 monolayers, bilayers, and in the bulk limit,” Phys. Rev. B, vol. 97, p. 075123, 2018, https://doi.org/10.1103/physrevb.97.075123.10.1103/PhysRevB.97.075123Search in Google Scholar

[17] A. Arora, J. Noky, M. Drüppel, et al., “Highly anisotropic in-plane excitons in atomically thin and bulklike 1t’-rese2,” Nano Lett., vol. 17, no. 5, pp. 3202–3207, 2017, https://doi.org/10.1021/acs.nanolett.7b00765.10.1021/acs.nanolett.7b00765Search in Google Scholar PubMed

[18] H. Tian, J. Tice, R. Fei, et al., “Low-symmetry two-dimensional materials for electronic and photonic applications,” Nano Today, vol. 11, no. 6, pp. 763–777, 2016, https://doi.org/10.1016/j.nantod.2016.10.003.10.1016/j.nantod.2016.10.003Search in Google Scholar

[19] R. Verre, D. G. Baranov, B. Munkhbat, J. Cuadra, M. Käll, and T. Shegai, “Transition metal dichalcogenide nanodisks as high-index dielectric mie nanoresonators,” Nat. Nanotechnol., vol. 14, no. 7, pp. 679–683, 2019, https://doi.org/10.1038/s41565-019-0442-x.10.1038/s41565-019-0442-xSearch in Google Scholar PubMed

[20] S. Busschaert, R. Reimann, M. Cavigelli, R. Khelifa, A. Jain, and L. Novotny, “Tmdc resonators for second harmonic signal enhancement,” arXiv:2001.08959v3 preprint, 2020. https://arxiv.org/pdf/2001.08959.pdf.Search in Google Scholar

[21] G. A. Ermolaev, D. V. Grudinin, Y. V. Stebunov, et al., “Giant optical anisotropy in transition metal dichalcogenides for next-generation photonics,” arXiv:2006.00884 preprint, 2020. https://arxiv.org/ftp/arxiv/papers/2006/2006.00884.pdf.10.1038/s41467-021-21139-xSearch in Google Scholar PubMed PubMed Central

[22] A. Segura, L. Artús, R. Cuscó, T. Taniguchi, G. Cassabois, and B. Gil, “Natural optical anisotropy of h-bn: highest giant birefringence in a bulk crystal through the mid-infrared to ultraviolet range,” Phys. Rev. Mater., vol. 2, p. 024001, 2018, https://doi.org/10.1103/physrevmaterials.2.024001.10.1103/PhysRevMaterials.2.024001Search in Google Scholar

[23] M. I. D’yakonov, “New type of electromagnetic wave propagating at an interface,” Sov. Phys. JETP, vol. 67, no. 4, pp. 714–716, 1988.Search in Google Scholar

[24] H. R. Philipp and E. A. Taft, “Optical constants of silicon in the region 1 to 10 ev,” Phys. Rev., vol. 120, pp. 37–38, 1960, https://doi.org/10.1103/physrev.120.37.10.1103/PhysRev.120.37Search in Google Scholar

[25] S. Adachi, “Optical dispersion relations for gap, gaas, gasb, inp, inas, insb, alxga1-xas, and in1-xgaxasyp1-y,” J. Appl. Phys., vol. 66, no. 12, pp. 6030–6040, 1989, https://doi.org/10.1063/1.343580.10.1063/1.343580Search in Google Scholar

[26] G. E. JellisonJr, “Optical functions of gaas, gap, and ge determined by two-channel polarization modulation ellipsometry,” Opt. Mater., vol. 1, no. 3, pp. 151–160, 1992, https://doi.org/10.1016/0925-3467(92)90022-f.10.1016/0925-3467(92)90022-FSearch in Google Scholar


Supplementary Material

The online version of this article offers supplementary material (https://doi.org/10.1515/nanoph-2020-0416).


Received: 2020-07-23
Accepted: 2020-09-29
Published Online: 2020-10-28

© 2020 Anton A. Shubnic et al., published by De Gruyter, Berlin/Boston

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

Downloaded on 23.4.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2020-0416/html
Scroll to top button