Next Article in Journal
Mechanisms of Grain Structure Evolution in a Quenched Medium Carbon Steel during Warm Deformation
Next Article in Special Issue
Ideal Photonic Weyl Nodes Stabilized by Screw Rotation Symmetry in Space Group 19
Previous Article in Journal
Effect of Deposition Power on the Thermoelectric Performance of Bismuth Telluride Prepared by RF Sputtering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Infrared Optical Conductivity of Bulk Bi2Te2Se

by
Elena S. Zhukova
1,2,*,
Hongbin Zhang
3,
Victor P. Martovitskiy
4,
Yurii G. Selivanov
4,
Boris P. Gorshunov
1,2 and
Martin Dressel
1,2,*
1
Moscow Institute of Physics and Technology, National Research University, 141701 Dolgoprudny, Moscow Region, Russia
2
1. Physikalisches Institut, Universität Stuttgart, Pfaffenwaldring 57, 70569 Stuttgart, Germany
3
Institute of Materials Science, Technische Universität Darmstadt, Jovanka-Bontschits-Str. 2, 64287 Darmstadt, Germany
4
P. N. Lebedev Physical Institute of the RAS, 119991 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Crystals 2020, 10(7), 553; https://doi.org/10.3390/cryst10070553
Submission received: 8 June 2020 / Revised: 21 June 2020 / Accepted: 24 June 2020 / Published: 28 June 2020
(This article belongs to the Special Issue Advances in Topological Materials)

Abstract

:
Mid- and near-infrared measurements reveal that the optical conductivity of the three-dimensional topological insulator, Bi2Te2Se, is dominated by bulk carriers and shows a linear-in-frequency increase at 0.5 to 0.8 eV. This linearity might be interpreted as a signature of three-dimensional (bulk) Dirac bands; however, band-structure calculations show that transitions between bands with complex dispersion contribute instead to the inter-band optical conductivity at these frequencies and, hence, the observed linearity is accidental. These results warn against the oversimplified interpretations of optical-conductivity measurements in different Dirac materials.

1. Introduction

Spin-orbit coupling often leads to the formation of linear bands in solids. Electrons in such bands (the Dirac electrons) manifest themselves in special ways in different experiments [1,2,3,4,5]. One of these manifestations is in their optical response: the contribution of a d-dimensional Dirac band to the inter-band optical conductivity, which is calculated to follow a simple power–law frequency dependence [6,7]:
σ(ω) ∝ ωd−2.
Such optical-conductivity behavior—unusual for conventional materials—has indeed been confirmed for (quasi)-2D electrons in graphene, graphite, and the line-node semimetal ZrSiS, where σ(ω) ≈ const(ω) was reported [8,9,10]. In turn, the 3D Dirac electrons in Dirac and Weyl semimetals, such as ZrTe5, Cd3As2, and TaAs, provide the inter-band optical conductivity to be proportional to frequency, σ(ω) ∝ ω [11,12,13]. The linearity in σ(ω) over a broad frequency range in a 3D electron system is often considered as a ’’smoking gun’’ for Dirac physics. For example, Timusk et al. [14] suggested the presence of 3D Dirac fermions in a number of quasicrystals, based entirely on the observation of a linear σ(ω) in these materials.
Besides, enormous efforts have been made to investigate the symmetry-protected surface states of topological insulators [2,3]. However, the dominant physics of the bulk often obscures the surface properties and hence is generally considered as an obstacle for experiments targeting the surface states. Achieving dissipationless surface spin currents may be of primary importance for potential applications of topological insulators, nevertheless, investigations into bulk electronic properties are essential for understanding the complete picture of the topological-state formation [15].
Our experiments reveal that the bulk optical conductivity of Bi2Te2Se follows a linear frequency dependence in an appreciably broad spectral range. Based on band-structure calculations, we argue that this linearity is not due to transitions within (a) particular 3D linear band(s), but instead a result of contributions from the transitions between the bands with complex dispersion.

2. Materials and Methods

Bi2Te2Se bulk crystals were synthesized by a modified Bridgman method [16]. Highly purified (99.9999%) elemental starting materials (Bi, Te, and Se) (Chimmed, Moscow, Russia) were loaded in quartz ampules inside an inert-gas glove box in the stoichiometric ratio 2:2:1. The sealed evacuated ampules were kept at 850 °C for 24 h with periodic stirring to ensure the homogeneity of the melt, followed by a cooldown to 520 °C with a rate of 5 °C/h. The crystals were then annealed at 520 °C for six days. The typical crystal sizes obtained in this way were in the centimeter range. The crystals were cut into appropriate pieces for X-ray, Hall, and optical measurements (and kept in vacuum until the measurements).
Utilizing an X’Pert Pro Extended MRD X-ray diffractometer (PANalytical, Almelo, the Netherlands) we have confirmed the high structural quality of the crystals, see Figure 1. The free-carrier concentration and mobility were measured in a standard Hall geometry. Indium-soldered contacts were applied to razor-cut Hall bars with typical dimensions of 2 × 0.5 × 0.2 mm3. For all samples, the conduction was by n-type carriers. The properties of the sample, used in our infrared studies, are listed in Table 1.
Optical reflectivity was measured from the (001) plane on freshly cleaved surfaces. The room-temperature experiments were performed in the mid- and near-infrared spectral ranges (600–8000 cm−1, 75 meV–1 eV) with a Bruker Vertex 80v Fourier-transform infrared spectrometer (Bruker Corporation, Billerica, MA, USA). Freshly evaporated gold mirrors served for reference measurements. We used unpolarized light, because Bi2Te2Se possesses C3 rotational symmetry along the [001] direction and hence the (001)-plane response, expressed via a second-rank tensor, such as optical conductivity, is isotropic.

3. Results and Discussion

In the top panel of Figure 2, we plot the raw reflectivity data recorded at 300 K. The reflectivity is very flat between 4000 and 8000 cm−1. In order to obtain the optical conductivity from the reflectivity data, we first tried to fit the measured spectra using a standard Drude–Lorentz procedure [18]. However, we found that such flat reflectivity is impossible to fit in an acceptable way with a physically meaningful number of Lorentzians. In an alternative approach, we used Kuzmenko’s variational dielectric function method [19], which produces optical functions with an accuracy equivalent to Kramers–Kronig. For the sake of convenience, the variable part of the dielectric response function was described by a large number of Lorentzians. Justification and details of this approach can be found in [20]. Similar to the Kramers–Kronig analysis, this method gives less accurate results near the edges of the experimental window. Thus, the results below approximately 2000 cm−1 and above 7000 cm−1 cannot be considered as accurate.
The real part of the optical conductivity obtained from this fit is plotted in the bottom panel of Figure 2. The eye-catching feature of the figure is the linear increase in σ(ω) at 4000 to 7000 cm−1 (~0.5–0.8 eV).
Let us first argue that the observed optical conductivity originates from the bulk of Bi2Te2Se. In Bi2Te2Se, the surface Dirac point lies inside the bulk band gap [21,22] and metallic surface states have been experimentally confirmed [21,22,23,24,25,26,27]. Nevertheless, Bi2Te2Se samples usually possess a significant concentration of bulk charges due to the basically unavoidable presence of defects, the so-called self-doping [26,27,28,29,30]. This is also the case for our sample—its bulk carrier concentration is rather large, as shown in Table 1. Furthermore, the skin depth, calculated from the complex optical conductivity, is above 30 nm at any measurement frequency, while the thickness of the topologically non-trivial surface layer is believed to be around 1 nm [3]. Hence, the response detected by our optical measurements is due to the bulk.
Let us also note that, at elevated temperatures, the optical detection of surface carriers in Bi2Te2Se, as well as in similar compounds, such as Bi2Te3 and Bi2Se3, remains so far elusive, while bulk carriers clearly manifest themselves in the optical response of Bi2Te2Se [16,28,29,30] and related compounds [31,32,33,34,35]. Reijnders et al. have reported on a mixed (surface plus bulk) optical response in Bi2Te2Se for low frequencies at temperatures below some 40 K [30]. However, at room temperature, as well as at frequencies above 2000 cm−1, their data are perfectly reconciled with entirely bulk response.
Coming back to the linear σ(ω), it is tempting to interpret it in terms of Equation (1), namely, as a signature of a 3D Dirac band (because our σ(ω) reflects the bulk response). Such a band, however, is not expected to appear in the bulk of Bi2Te2Se [36]. We would like to point out that all the available optical conductivity spectra (ours and those previously reported in [16,28,29,30]) are rather similar to each other, although the linearity of σ(ω) is most apparent in our data. The deviations between the data sets can be assigned, for example, to the abovementioned difference in the exact Fermi-level position in different samples of Bi2Te2Se. In order to check the origin of the linear frequency increase in σ(ω), we performed band-structure calculations for Bi2Te2Se and then calculated its inter-band optical conductivity.
The band-structure and optical-conductivity calculations were performed using the full potential linear augmented plane-wave method, as implemented in the WIEN2k code [37]. The exchange-correlation functional is parameterized using the GGA approximation [38]. The self-consistent charge-densities and optical-conductivity calculations were done with 400 and 2000 k-points in the whole Brillouin zone, respectively. The results of the calculations are shown in Figure 2 and Figure 3. The obtained band structure is basically identical to the one reported in [36]. In order to be reconciled with the bulk electron concentration (the self-doping problem mentioned above), the Fermi level needs to be shifted upwards, as compared to the undoped situation, as shown in Figure 3. From the figure, it is apparent that there is no truly Dirac band in the bulk of Bi2Te2Se.
The calculated optical conductivity is shown as a dashed line in Figure 2. Taking into account the generally poor reproducibility of the experimental infrared optical conductivity by first-principles calculations (cf., e.g., in [39,40]), the agreement between theory and experiment can be considered as fairly good. Further, we should point out that the computed σ(ω) has no intra-band (free-carrier) contribution. Thus, it is not surprising that the low-frequency experimental σ(ω) is larger than the theoretical line. Additionally, the effect of temperature broadening is absent in the calculations. Such broadening would make the smooth step at around 3000 cm−1 even broader [10].Taking into account the mentioned issues in the computations of σ(ω) is outside of our capacity and beyond the scope of the paper. The important result of our computations is that the linear σ(ω) is nicely reproduced at 4000 to 6000 cm−1 (~0.5–0.75 eV). Thus, we can conclude that this linearity comes as a cumulative effect of transitions between the bands, which do not have a simple linear dispersion. We note that recent measurements of BaCoS2 and GdPtBi provide other examples of linear σ(ω) not due to a simple 3D Dirac band [41,42].

4. Conclusions

We have experimentally found that the bulk optical conductivity of Bi2Te2Se is linear in frequency at 4000 to 7000 cm−1 (~0.5–0.8 eV). Our computations demonstrate that this linearity is not due to transitions within a 3D Dirac band, but emerges as a cumulative effect of transitions between the bands with complex dispersion. Obviously, similar situations can appear in other systems and, thus, suggestions for Dirac physics based on optical-conductivity measurements have to be made cautiously.

Author Contributions

Infrared measurements, E.S.Z. and B.P.G.; sample preparation, V.P.M. and Y.G.S.; data analysis, E.S.Z.; writing—original draft preparation, E.S.Z.; writing—review and editing, H.Z., M.D., and B.P.G.; project administration, B.P.G.; funding acquisition, Y.G.S. M.D., and B.P.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Higher Education of the Russian Federation (Program "5 top 100"), by the Russian Science Foundation (grant No. 17-12-01544) and by the Deutsche Forschungsgemeinschaft (DFG) via grants No. DR228/51 and No. ZH559/2-1.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Neto, A.C.; Guinea, F.; Peres, N.; Novoselov, K.S.; Geim, A.K. The electronic properties of graphene. Rev. Mod. Phys. 2009, 81, 109. [Google Scholar] [CrossRef] [Green Version]
  2. Hasan, M.Z.; Kane, C.L. Colloquium: Topological insulators. Rev. Mod. Phys. 2010, 82, 3045. [Google Scholar] [CrossRef] [Green Version]
  3. Qi, X.L.; Zhang, S.C. Topological insulators and superconductors. Rev. Mod. Phys. 2011, 83, 1057. [Google Scholar] [CrossRef] [Green Version]
  4. Wehling, T.O.; Black-Schaffer, A.M.; Balatsky, A.V. Dirac materials. Adv. Phys. 2014, 63, 1. [Google Scholar] [CrossRef] [Green Version]
  5. Armitage, N.P.; Mele, E.J.; Vishwanath, A. Weyl and Dirac semimetals in three-dimensional solids. Rev. Mod. Phys. 2018, 90, 015001. [Google Scholar] [CrossRef] [Green Version]
  6. Hosur, P.; Parameswaran, S.A.; Vishwanath, A. Charge transport in Weyl semimetals. Phys. Rev. Lett. 2012, 108, 046602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Bacsi, A.; Virosztek, A. Low-frequency optical conductivity in graphene and in other scale-invariant two-band systems. Phys. Rev. B 2013, 87, 125425. [Google Scholar] [CrossRef] [Green Version]
  8. Kuzmenko, A.B.; van Heumen, E.; Carbone, F.; van der Marel, D. Universal Optical conductance of graphite. Phys. Rev. Lett. 2008, 100, 117401. [Google Scholar] [CrossRef] [Green Version]
  9. Mak, K.F.; Sfeir, M.Y.; Wu, Y.; Lui, C.H.; Misewich, J.A.; Heinz, T.F. Measurement of the Optical conductivity of graphene. Phys. Rev. Lett. 2008, 101, 196405. [Google Scholar] [CrossRef]
  10. Schilling, M.B.; Schoop, L.M.; Lotsch, B.V.; Dressel, M.; Pronin, A.V. Flat Optical Conductivity in ZrSiS due to two-dimensional Dirac bands. Phys. Rev. Lett. 2017, 119, 187401. [Google Scholar] [CrossRef] [Green Version]
  11. Chen, R.Y.; Zhang, S.J.; Schneeloch, J.A.; Zhang, C.; Li, Q.; Gu, G.D.; Wang, N.L. Optical spectroscopy study of the three-dimensional Dirac semimetal ZrTe5. Phys. Rev. B 2015, 92, 075107. [Google Scholar] [CrossRef] [Green Version]
  12. Neubauer, D.; Carbotte, J.P.; Nateprov, A.A.; Löhle, A.; Dressel, M.; Pronin, A.V. Interband optical conductivity of the [001]-oriented Dirac semimetal Cd3As2. Phys. Rev. B 2016, 93, 121202. [Google Scholar] [CrossRef] [Green Version]
  13. Xu, B.; Dai, Y.M.; Zhao, L.X.; Wang, K.; Yang, R.; Zhang, W.; Liu, J.Y.; Xiao, H.; Chen, G.F.; Taylor, A.J.; et al. Optical spectroscopy of the Weyl semimetal TaAs. Phys. Rev. B 2016, 93, 121110. [Google Scholar] [CrossRef] [Green Version]
  14. Timusk, T.; Carbotte, J.P.; Homes, C.C.; Basov, D.N.; Sharapov, S.G. Three-dimensional Dirac fermions in quasicrystals as seen via optical conductivity. Phys. Rev. B 2013, 87, 235121. [Google Scholar] [CrossRef] [Green Version]
  15. Xu, S.-Y.; Xia, Y.; Wray, L.A.; Jia, S.; Meier, F.; Dil, J.H.; Osterwalder, J.; Slomski, B.; Bansil, A.; Lin, H.; et al. Topological phase transition and texture inversion in a tunable topological insulator. Science 2011, 332, 560. [Google Scholar] [CrossRef] [Green Version]
  16. Aleshchenko, Y.A.; Muratov, A.V.; Pavlova, V.V.; Selivanov, Y.G.; Chizhevskii, E.G. Infrared spectroscopy of Bi2Te2Se. JETP Lett. 2014, 99, 187. [Google Scholar] [CrossRef]
  17. Jia, S.; Ji, H.; Climent-Pascual, E.; Fuccillo, M.K.; Charles, M.E.; Xiong, J.; Ong, N.P.; Cava, R.J. Low-carrier-concentration crystals of the topological insulator Bi2Te2Se. Phys. Rev. B 2011, 84, 235206. [Google Scholar] [CrossRef] [Green Version]
  18. Dressel, M.; Grüner, G. Electrodynamics of Solids; Cambridge University Press: Cambridge, UK, 2002. [Google Scholar]
  19. Kuzmenko, A.B. Kramers–Kronig constrained variational analysis of optical spectra. Rev. Sci. Instrum. 2005, 76, 083108. [Google Scholar] [CrossRef] [Green Version]
  20. Chanda, G.; Lobo, R.P.S.M.; Schachinger, E.; Wosnitza, J.; Naito, M.; Pronin, A.V. Optical study of superconducting Pr2CuOx with x≃4. Phys. Rev. B 2014, 90, 024503. [Google Scholar] [CrossRef] [Green Version]
  21. Xu, S.-Y.; Wray, L.A.; Xia, Y.; Shankar, R.; Petersen, A.; Fedorov, A.; Lin, H.; Bansil, A.; Hor, Y.S.; Grauer, D.; et al. Discovery of several large families of topological insulator classes with backscattering-suppressed spin-polarized single-Dirac-cone on the surface. arXiv 2010, arXiv:1007.5111. [Google Scholar]
  22. Ren, Z.; Taskin, A.A.; Sasaki, S.; Segawa, K.; Ando, Y. Large bulk resistivity and surface quantum oscillations in the topological insulator Bi2Te2Se. Phys. Rev. B 2010, 82, 241306. [Google Scholar] [CrossRef] [Green Version]
  23. Xiong, J.; Petersen, A.C.; Qu, D.; Hor, Y.S.; Cava, R.J.; Ong, N.P. Quantum oscillations in a topological insulator Bi2Te2Se with large bulk resistivity (6 Ωcm). Physica E 2012, 44, 917. [Google Scholar] [CrossRef] [Green Version]
  24. Li, Z.; Chen, T.; Pan, H.; Song, F.; Wang, B.; Han, J.; Qin, Y.; Wang, X.; Zhang, R.; Wan, J.; et al. Two-dimensional universal conductance fluctuations and the electron-phonon interaction of surface states in Bi2Te2Se microflakes. Sci. Rep. 2012, 2, 595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Tian, J.; Miotkowski, I.; Hong, S.; Chen, Y.P. Electrical injection and detection of spin-polarized currents in topological insulator Bi2Te2Se. Sci. Rep. 2015, 5, 14293. [Google Scholar] [CrossRef]
  26. Arakane, T.; Sato, T.; Souma, S.; Kosaka, K.; Nakayama, K.; Komatsu, M.; Takahashi, T.; Ren, Z.; Segawa, K.; Ando, Y. Tunable Dirac cone in the topological insulator Bi2-xSbxTe3-ySey. Nat. Commun. 2011, 3, 636. [Google Scholar] [CrossRef] [Green Version]
  27. Neupane, M.; Xu, S.-Y.; Wray, L.A.; Petersen, A.; Shankar, R.; Alidoust, N.; Liu, C.; Fedorov, A.; Ji, H.; Allred, J.M.; et al. Topological surface states and Dirac point tuning in ternary topological insulators. Phys. Rev. B 2012, 85, 235406. [Google Scholar] [CrossRef] [Green Version]
  28. di Pietro, P.; Vitucci, F.M.; Nicoletti, D.; Baldassarre, L.; Calvani, P.; Cava, R.; Hor, Y.S.; Schade, U.; Lupi, S. Ab initio studying of topological insulator Bi2Se3 under the stress. Phys. Rev. B 2012, 86, 045439. [Google Scholar] [CrossRef] [Green Version]
  29. Akrap, A.; Tran, M.; Ubaldini, A.; Teyssier, J.; Giannini, E.; van der Marel, D.; Lerch, P.; Homes, C.C. Optical properties of Bi2Te2Se at ambient and high pressures. Phys. Rev. B 2012, 86, 235207. [Google Scholar] [CrossRef] [Green Version]
  30. Reijnders, A.A.; Tian, Y.; Sandilands, L.J.; Pohl, G.; Kivlichan, I.D.; Zhao, S.Y.F.; Jia, S.; Charles, M.E.; Cava, R.J.; Alidoust, N.; et al. Optical evidence of surface state suppression in Bi-based topological insulators. Phys. Rev. B 2014, 89, 075138. [Google Scholar] [CrossRef] [Green Version]
  31. LaForge, A.D.; Frenzel, A.; Pursley, B.C.; Lin, T.; Liu, X.; Shi, J.; Basov, D.N. Optical characterization of Bi2Se3 in a magnetic field: Infrared evidence for magnetoelectric coupling in a topological insulator material. Phys. Rev. B 2010, 81, 125120. [Google Scholar] [CrossRef] [Green Version]
  32. Sushkov, A.B.; Jenkins, G.S.; Schmadel, D.C.; Butch, N.P.; Paglione, J.; Drew, H.D. Far-infrared cyclotron resonance and Faraday effect in Bi2Se3. Phys. Rev. B 2010, 82, 125110. [Google Scholar] [CrossRef] [Green Version]
  33. Dordevic, S.V.; Wolf, M.S.; Stojilovic, N.; Lei, H.; Petrovic, C. Signatures of charge inhomogeneities in the infrared spectra of topological insulators Bi2Se3, Bi2Te3 and Sb2Te3. J. Phys. Condens. Matter 2013, 25, 075501. [Google Scholar] [CrossRef] [PubMed]
  34. Chapler, B.C.; Post, K.W.; Richardella, A.R.; Lee, J.S.; Tao, J.; Samarth, N.; Basov, D.N. Infrared electrodynamics and ferromagnetism in the topological semiconductors Bi2Te3 and Mn-doped Bi2Te3. Phys. Rev. B 2014, 89, 235308. [Google Scholar] [CrossRef] [Green Version]
  35. Post, K.W.; Lee, Y.S.; Chapler, B.C.; Schafgans, A.A.; Novak, M.; Taskin, A.A.; Segawa, K.; Goldflam, M.D.; Stinson, H.T.; Ando, Y.; et al. Infrared probe of the bulk insulating response in Bi2−xSbxTe3−ySey topological insulator alloys. Phys. Rev. B 2015, 91, 165202. [Google Scholar] [CrossRef] [Green Version]
  36. Wang, L.-L.; Johnson, D.D. Ternary tetradymite compounds as topological insulators. Phys. Rev. B 2011, 83, 241309. [Google Scholar] [CrossRef] [Green Version]
  37. Available online: http://www.wien2k.at (accessed on 6 June 2020).
  38. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  39. Frenzel, A.J.; Homes, C.C.; Gibson, Q.D.; Shao, Y.M.; Post, K.W.; Charnukha, A.; Cava, R.J.; Basov, D.N. Anisotropic electrodynamics of type-II Weyl semimetal candidate WTe2. Phys. Rev. B 2017, 95, 245140. [Google Scholar] [CrossRef] [Green Version]
  40. Neubauer, D.; Yaresko, A.; Li, W.; Löhle, A.; Hübner, R.; Schilling, M.B.; Shekhar, C.; Felser, C.; Dressel, M.; Pronin, A.V. Optical conductivity of the Weyl semimetal NbP. Phys. Rev. B 2018, 98, 195203. [Google Scholar] [CrossRef] [Green Version]
  41. Santos-Cottin, D.; Klein, Y.; Werner, P.; Miyake, T.; Medici, L.d.; Gauzzi, A.; Lobo, R.P.S.M.; Casula, M. Linear behavior of the optical conductivity and incoherent charge transport in BaCoS2. Phys. Rev. Materials 2018, 2, 105001. [Google Scholar] [CrossRef] [Green Version]
  42. Hütt, F.; Yaresko, A.; Schilling, M.B.; Shekhar, C.; Felser, C.; Dressel, M.; Pronin, A.V. Linear-in-frequency optical conductivity in GdPtBi due to transitions near the triple points. Phys. Rev. Lett. 2018, 121, 176601. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Bi2Te2Se X-ray diffraction pattern. Inset: Rocking curve for the (0 0 15) reflection peak.
Figure 1. Bi2Te2Se X-ray diffraction pattern. Inset: Rocking curve for the (0 0 15) reflection peak.
Crystals 10 00553 g001
Figure 2. Top panel: [001]-plane reflectivity of Bi2Te2Se at 300 K: measurements (black line) and fit (red line). Bottom panel: bulk optical conductivity (real part) of Bi2Te2Se, as obtained from the reflectivity fit (black straight line) and the inter-band portion of optical conductivity, computed from the band structure of Figure 3 at 0 K (dashed line), as detailed in the text. The thin orange line is to mimic a linear increase in frequency.
Figure 2. Top panel: [001]-plane reflectivity of Bi2Te2Se at 300 K: measurements (black line) and fit (red line). Bottom panel: bulk optical conductivity (real part) of Bi2Te2Se, as obtained from the reflectivity fit (black straight line) and the inter-band portion of optical conductivity, computed from the band structure of Figure 3 at 0 K (dashed line), as detailed in the text. The thin orange line is to mimic a linear increase in frequency.
Crystals 10 00553 g002
Figure 3. Band structure of Bi2Te2Se. Black dashed (red solid) horizontal line indicates the original (shifted) Fermi energy.
Figure 3. Band structure of Bi2Te2Se. Black dashed (red solid) horizontal line indicates the original (shifted) Fermi energy.
Crystals 10 00553 g003
Table 1. Room-temperature properties of the single-crystalline Bi2Te2Se sample used for the optical measurements. The mobility value is typical for the samples with such electron densities [17].
Table 1. Room-temperature properties of the single-crystalline Bi2Te2Se sample used for the optical measurements. The mobility value is typical for the samples with such electron densities [17].
Lateral DimensionsThicknessBulk electron DensityMobilityLattice Constant
5 × 5 mm2350 μm1.0 × 1018 cm−3330 cm2/Vs29.766 A

Share and Cite

MDPI and ACS Style

Zhukova, E.S.; Zhang, H.; Martovitskiy, V.P.; Selivanov, Y.G.; Gorshunov, B.P.; Dressel, M. Infrared Optical Conductivity of Bulk Bi2Te2Se. Crystals 2020, 10, 553. https://doi.org/10.3390/cryst10070553

AMA Style

Zhukova ES, Zhang H, Martovitskiy VP, Selivanov YG, Gorshunov BP, Dressel M. Infrared Optical Conductivity of Bulk Bi2Te2Se. Crystals. 2020; 10(7):553. https://doi.org/10.3390/cryst10070553

Chicago/Turabian Style

Zhukova, Elena S., Hongbin Zhang, Victor P. Martovitskiy, Yurii G. Selivanov, Boris P. Gorshunov, and Martin Dressel. 2020. "Infrared Optical Conductivity of Bulk Bi2Te2Se" Crystals 10, no. 7: 553. https://doi.org/10.3390/cryst10070553

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

Article Metrics

Back to TopTop