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

Surface plasmon polaritons in topological insulator nano-films and superlattices

Open Access Open Access

Abstract

We investigate the propagation of surface plasmon polaritons (SPPs) in thin films of topological insulators. Cases of single films and multilayered stacks are analyzed. The materials considered are second generation three dimensional topological insulators Bi2Se3, Bi2Te3, and Sb2Te3. Dispersion relations and propagation lengths of SPPs are estimated numerically, taking into account the variation of bulk dielectric functions of topological insulators, as well as substrate, using the Drude-Lorentz model. The key factors affecting propagation length are identified and experimental modifications for tuning the dispersion relations are proposed. The apparent discrepancy between the experimental data and previously considered theory is resolved.

© 2016 Optical Society of America

1. Introduction

Since their discovery, three dimensional topological insulators (TIs) [1–3] have attracted enormous interest, both in theory [4–6] and in experiment [7–9], owing to the unconventional character of gapless topological surface states hosting “massless” helical electron liquid [10]. These surfaces are envisioned as a potentially disruptive platform for a wide range of frontier technological applications, from spintronics [11] and fault-tolerant quantum computing based on Majorana fermions [12], to terahertz optics and plasmonics [13–18]. Visible range plasmonics is also being actively explored [19, 20].

Here we numerically study the propagation of surface plasmon polaritons (SPPs) – coupled oscillations of surface charges and electromagnetic field [21] – in nano-meter thin films of topological insulators Bi2Se3, Bi2Te3, and Sb2Te3. Motivated by recent experimental observation of SPPs in Bi2Se3 in the far-infrared range [22], we resolve several outstanding issues. (i) We find the dispersion relations and propagation lengths of SPPs in Bi2Se3, Bi2Te3, and Sb2Te3 in the far-IR, using realistic material parameters allowing the comparison of these materials’ potential in plasmonics; (ii) identify key parameters determining the propagation length and demonstrate that the latter can be enhanced by two orders of magnitude in some cases – the finding which can be crucial for practical applications; (iii) revisit the problem advanced by Stauber et al. [23] regarding the proper analysis of the experimental data in [22] and propose a simple solution to it; (iv) analyze the effect of stacking of TI films and dielectric into 1D superlattice with the goal to modify the dispersion relation of SPPs.

Since optical response of materials depends on their bulk dielectric function ε(ω), it is imperative to account for its possible variation. To properly address the questions (i–iv) we approximate ε(ω) using the Drude-Lorentz model combined with available experimental data on far-IR optical properties of the materials under investigation. This methodology is the main and essential difference between our approach and the usual take on SPPs in TI films.

2. Background

Experimentally SPPs in TIs were first studied by Di Pietro et al. [22] in ribbons of Bi2Se3, where standing SPP waves with resonant frequencies defined by the ribbons width were formed. The idea of the experiment is illustrated in Fig. 1. As the size and spacing W of ribbons was varied the spectral position of SPP resonances shifted, allowing to map the dispersion curve ω(q) for the structures. The correspondence between the experiment and theoretical prediction, ω(q)q, based on the Dirac fermions description (see e.g. [24]) was remarkable. However, this result was later critically analyzed by Stauber et al. [23], who noted that when the long range Coulomb interaction between top and bottom surfaces is properly taken into account, the experiment should have shown the sensitivity to the thickness of the films as well as to the bulk dielectric function of Bi2Se3.

 figure: Fig. 1

Fig. 1 Schematics of the experiment for observing SPP resonances in Bi2Se3 (see [22]). Thin film, MBE grown on sapphire, is etched into a periodic array of ribbons of the width W, separated by a gap of the same width. Incident linearly polarized far-IR light excites standing waves of collective oscillations of surface charges. The resonant excitation for given ribbon width W is observed in the drop of the transmitted light intensity T(ν). The surface current density j, plotted on the vertical axis, describes standing waves with the period 2W.

Download Full Size | PDF

More accurate analytical expression for ω(q) was derived in [23] in the long-wavelength limit (qd ≪ 1):

ω2=vFkFe2ε0hqεT+εB+qdεTI

Here εT and εB are bulk dielectric functions of the media on top and bottom of the TI film. When qdεTIεT + εB) the well-known dispersion relation for two dimensional plasmons is recovered with ωq. This simplified dispersion relation was used to fit the experimental results in [22]. However, Stauber et al. pointed out that for εT = 1 (air), εB = 10 (sapphire), and εTI = 100 the term qdεTI in Eq. (1) can not be neglected. Indeed, the requirement qdεTI ≪ (εT + εB) leads to qd ≪ 0.1, which is not satisfied for the majority of data points in [22]: qd = 0.02, 0.05, 0.05, 0.08, 0.2 for W = 2.0, 2.5, 4.0, 8.0 and 20 microns, respectively. In order to satisfactorily fit the experiment with a more appropriate dispersion relation (1), [23] assumed additional contribution to the optical response from two dimensional spin-degenerate electron gas close to the surface. Although not excluding this possibility, we demonstrate that such an assumption is not necessary. Instead we reconcile the experiment [22] with the theory [23] by calculating the dispersion relation without the restriction ε = const for both substrate and the TI bulk. It must be noted, that the possibility of such approach has been mentioned in [23].

For sapphire, which is often used as a growth substrate, the bulk dielectric function in far-IR (below 400 cm1) can be approximated by

ε(ω,cm1)=n02+(n021)(λω)2+iγ(n021)(λω),
where n0 = 3.2, λ = 20.4 × 104 cm, and γ = 0.036 are experimentally determined parameters [25]. Although sapphire is optically anisotropic, this expression gives the refractive index for ordinary and extraordinary rays with 10% accuracy. Above 400 cm1 the bulk dielectric function of sapphire begins to change drastically due to the presence of IR-active modes of lattice vibrations [26] and then more careful modeling is required.

The bulk optical properties of Bi2Se3, Bi2Te3, and Sb2Te3 in the far and mid-IR are also relatively well known from reflectance measurements [27, 28]. The analysis of experimental data suggests that isotropic Drude-Lorentz model with 3 or 4 oscillators can quite satisfactorily describe the overall features of reflectance spectra in the range from 50 cm1 to 1000 cm1 for all three materials [27]. The bulk dielectric function in the far-IR range of interest can therefore be approximated with the expression

ε(ω,cm1)=εωD2ω2+iωγD+j=1j=3,4ωpj2ω0j2ω2iωγj.

Parameters for Bi2Se3 at room temperature [27] are shown in Table 1. Table 2 shows similar numbers for thin films of Bi2Se3 used in [22]. Fig. 2 shows the contribution of each term in the Drude-Lorentz model into the bulk ε(ω) for this material, while Fig. 3 demonstrates that ε(ω) varies significantly over the spectral range 10 cm1 to 200 cm1. We note that the contribution to ε(ω) from free electrons in the bulk, represented by the Drude term in Eq. (3), is substantial. It may be argued that at low frequencies the optical response is dominated by the surface states and the bulk contribution is small [29], but if the exact contribution of the bulk Drude term is not known it must be set as a free parameter. Indeed, considering that 1) Bi2Se3, Bi2Te3, and Sb2Te3 all have relatively narrow band gaps, 2) the Fermi level of many thin films lies in the bulk conduction or valence band, and 3) intrinsic bulk defects increase the free carrier concentration, it is important to fully consider the bulk Drude term.

Tables Icon

Table 1. Parameters for the Drude-Lorentz model with ε = 1, extracted from the reflectance measurements on Bi2Se3 at room temperatures in far and mid-IR [27].

Tables Icon

Table 2. Parameters for the Drude-Lorentz model for the unpatterned Bi2Se3 thin films at 300 K, see Supplementary Material to [22].

 figure: Fig. 2

Fig. 2 Contributions of various terms of the Drude-Lorentz model into the real part of the far-IR bulk dielectric function of Bi2Se3. Two IR-active phonon modes (Lorentz-α with ω ≈ 61 cm1 and Lorentz-β with ω ≈ 133 cm1) correspond to the first two terms in the Drude-Lorentz model (3). The oscillator Lorentz-Ω accounts for higher frequency absorption, in this case at ω ≈ 2029.5 cm1 = 252 meV which is close to the band-gap energy of Bi2Se3 bulk [28].

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Real ε′ and imaginary ε″ parts of the bulk dielectric function of Bi2Se3, according to the Drude-Lorentz model. Solid lines show the bulk dielectric function without the Drude term, e.g. in a perfect crystal at low temperature. Dashed lines reproduce the bulk dielectric function with the Drude term present. The effect of Drude term is very strong in this spectral range.

Download Full Size | PDF

According to [27], the Drude-Lorentz model for Bi2Te3 at room temperature requires the Drude term with ωp = 5651.5 cm1, γ = 111.86 cm1 and a single Lorentz oscillator with ω0 = 8386.6 cm1, ωp = 66024 cm1, γ = 10260 cm1 (ε = 1). Sb2Te3 needs only the Drude term with ωp = 6906.7 cm1, γ = 183.69 cm1 (ε = 51). These numbers, compared to the data for Bi2Se3, indicate that at room temperature both Bi2Te3 and Sb2Te3 exhibit highly metallic behavior in their optical response. This is in correspondence with the fact that the bulk band gap of Bi2Te3 and Sb2Te3 is smaller than the band gap of Bi2Se3 [5]. Room temperature optical responses of Bi2Te3 and Sb2Te3 therefore seem to be dominated by free electrons in the bulk.

Low temperature reflectance measurements on Bi2Te3 and Sb2Te3 in the far-IR reveal the presence of IR active vibrational modes with frequencies in the range 20 cm1 to 200 cm1. For Bi2Te3 see, for example, Fig. 6 in [28]. Such modes are characteristic to the whole family of rhombohedral V2–VI3 compounds and are represented by the oscillators “Lorentz α” and “Lorentz β” in the Drude-Lorentz model.

We emphasize the importance of taking these modes into account in order to adequately estimate the bulk ε(ω) of Bi2Se3, Bi2Te3, and Sb2Te3 in the far-IR. By doing so we are able to determine more realistic dispersion relations for SPPs and their propagation lengths in these materials, which is essential for potential applications.

3. Method

We now can calculate dispersion relations and estimate propagation lengths for SPPs in thin TI films, following the approach of [23]. A film of thickness d supports two SPP modes (see Refs. [24, 30]) and in the following we will focus on the higher-frequency optical mode. In general the dispersion relation for this mode can be found by solving the non-linear equation det[1 − v(q, d)χ0(ω, q)] = 0, where the matrix v(q, d) incorporates intra- and interlayer Coulomb interactions for the top and bottom surface of TI film, and χ0(ω, q) is charge density and transverse spin susceptibility tensor (see Eqs. (2)–(7) in [23]).

In many cases of interest the criterion for the long-wavelength limit, qd ≪ 1, is satisfied and the dispersion relation (1) can be used. Indeed, consider how the SPPs are excited using the etched grating made of TI ribbons with period 2W [22]. The wave-vector imposed by the grating is given by q = 2π/(2W) and the criterion qd ≪ 1 leads to the requirement Wπd. For TI films with thicknesses d = 10–100 nm the necessary limit is satisfied if W ≫ 30 300 nm. For the gratings with the widths of several microns one can safely use the Eq. (1).

To account for the variation of the bulk dielectric functions εTI(ω), εB(ω), and εT (ω), it is convenient to invert Eq. (1):

q(ω)=Aω2[εT(ω)+εB(ω)]1Aω2dεTI(ω),Aε0hνFkFe2.

Substituting the expressions for εB and εTI with appropriate material parameters (Eqs. (2) and (3), respectively), one can find the wave-vector q(ω) = q′(ω) + iq″(ω) for any frequency. The real part q′(ω) yields the dispersion relations ω(q′), and the imaginary part q″(ω) determines the effective propagation length L ≡ 1/(2q″(ω)) – the parameter important for applications of propagating SPPs.

In addition to bulk dielectric functions εT, εTI, and εB, the results of calculations depend on film thickness, the Fermi velocity, and the Fermi level. Below we present calculations for the representative values of these parameters. The Fermi velocity is set to vF = (5 ± 1) × 105 m/s [5, 31]. The thicknesses d = 15 nm, 30 nm, 60 nm, 120 nm and the Fermi energies 50 meV and 500 meV will be considered, corresponding to the Fermi wave-vectors kF = 0.15 × 109 m1 and 1.52 × 109 m1, respectively. Due to monotonic behavior of the dispersion curves and propagation lengths we omit the intermediate values of the Fermi levels. The range of frequencies 30 cm1 to 200 cm1 is chosen in order to relate our numerical results to the available experimental data on SPP dispersion relation in Bi2Se3 [22].

4. Results and discussion

4.1. Bi2Se3

In Fig. 4(a, b) dispersion curves ω(q) are shown for various combinations of TI film thickness d and the Fermi level EF, while Fig. 4(c, d) present the propagation lengths L for the same parameters. Cirlce, square and triangle marks correspond to the experimental values taken from [22]. While there are visible changes in the dispersion curves, the most notable feature is the drastic enhancement of the propagation length as either the film thickness or the Fermi level is decreased. The non-monotonous behavior of dispersion curves, evident around ωα ≈ 63 cm1 and ωβ 133 cm1, comes from the interaction between the SPPs (ωq) and IR-active phonon modes represented by the terms Lorentz-α and Lorentz-β in Drude-Lorentz model. This interaction leads to the “avoided-crossing” type of behavior of the dispersion curves near ωα and ωβ.

 figure: Fig. 4

Fig. 4 (a, b) Dispersion relations of SPPs in Bi2Se3 films of several thicknesses for (a) kF = 0.15 × 109 m1 (EF = 50 meV) and (b) kF = 1.52 × 109 m1 (EF = 500 meV). Cirlce, square and triangle marks represent the experimental values from [22]. (c, d) Propagation lengths of SPPs in Bi2Se3 films for corresponding parameters: (c) for kF = 0.15 × 109 m1, (d) for kF = 1.52 × 109 m1. The lines marked α ≈ 63 cm1 and β ≈ 133 cm1 indicate positions of IR-active modes interacting with SPPs and causing non-monotinic behavior of dispersion and propagation length.

Download Full Size | PDF

We note that if the contribution of the Drude term in the model (3) is naively set to zero then similar degrees of sensitivity of dispersion relations ω(q) and the propagation lengths are exhibited: About fivefold increase of the propagation length happens when there are no free carriers in the bulk, while ω(q) does not show such a significant change. These results indicate that the propagation length of SPP in Bi2Se3 is very sensitive to the concentration of both surface and bulk carriers. Since the Fermi level can be controlled by gating, this sensitivity suggests a way of tuning the propagation length by almost two orders of magnitude without significantly affecting the SPP dispersion relation. Tuning Fermi-level from 50 meV to 500 meV would change the surface carrier density from n = 2.55 × 1011 cm2 to n = 2.55 × 1013 cm2. In [32] a tunability of n from zero (Dirac point) up to 1.0× 1013 cm2 has been reported, using voltages up to 80 V in a bottom-gate configuration with Si/SiO2. Slightly larger range (up to n ≈ 2.2 × 1013 cm2) was reported in [33], where SrTiO3 was used as a dielectric substrate.

We next perform calculations of the dispersion curves and propagation lengths for the experimental parameters given in [22]: vF = 6 × 105 m/s, kF = 1.37 × 109 m1, d = 60 and 120 nm. Relevant parameters for the Drude-Lorentz model are given in Table 2 and agree well with the numbers given in [27] and reproduced in Table 1. The high-frequency Lorentz oscillator with Ωgap 2030 cm1 = 252 meV can not be neglected because its contribution to εTI for the frequencies 30 cm1–200 cm1 is significant (ε ≈ 30) and does not vary appreciably. We thus take the values for the oscillators α and β from [22] (Table 2), while borrowing the value for Ωgap from [27] (Table 1). The dispersion curves and propagation lengths are given in Fig. 5 and Fig. 6. To achieve a reasonable fit to the experimental data for both films thicknesses the magnitude of the Drude term was reduced to 60% of its reported value. Since the contribution of this term is not precisely known (see the discussion above) this adjustment is justified. This calculation demonstrates that a reasonable agreement of the theory and experiment may be achieved without additional two-dimensional electron gas proposed in [23].

 figure: Fig. 5

Fig. 5 Dispersion curves of SPPs in Bi2Se3 thin films, calculated using the Drude-Lorentz model and the parameters reported in [22] (kF = 1.37 × 109 m1, EF = 541 meV). Cirlce, square and triangle marks correspond to the experimental results [22].

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Propagation lengths of SPPs in Bi2Se3 thin films, calculated using the Drude-Lorentz model and the parameters reported in [22] (kF = 1.37 × 109 m1, EF = 541 meV).

Download Full Size | PDF

4.2. Bi2Te3 and Sb2Te3

Optical characterization of Bi2Te3 has been performed in a number of works [27, 28, 34], with substantial variations of the reported values for optical parameters. Such variations indicate that for realistic calculations the bulk dielectric function must be estimated from reflectance or ellipsometry measurements on a particular sample. In our calculations the hybrid Drude-Lorentz model was used, with parameters from both Richter [28] and Wolf [27] (see Tables 3 and 4, respectively). This is done to capture contributions from in-plane IR-active vibrational modes present in Bi2Te3 around 50 cm1 and 90 cm1, as well as describe the plasma edge clearly observed near 500 cm1.

Tables Icon

Table 3. Parameters of the Drude-Lorentz model with ε = 85, extracted from the reflectance measurements on Bi2Te3 at room temperatures in far-IR [28].

Tables Icon

Table 4. Parameters for the Drude-Lorentz model with ε = 1 extracted from the reflectance measurements on Bi2Te3 at room temperatures in far- and mid-IR [27].

The results of calculations are presented in Figs. 7. Compared to Bi2Se3, Bi2Te3 demonstrates 1) greater sensitivity of dispersion curves ω(q) to the film thickness; and 2) significantly smaller propagation lengths for almost all values of the Fermi level. These differences become less pronounced if the Drude term is omitted, suggesting that overall higher level of free electrons in the bulk of Bi2Te3 strongly affects the propagation of the SPP in this material.

 figure: Fig. 7

Fig. 7 (a, b) Dispersion relations of SPPs in Bi2Te3 films of several thicknesses for (a) kF = 0.15 × 109 m1 (EF = 50 meV) and (b) kF = 1.52 × 109 m1 (EF = 500 meV). (c, d) Propagation lengths of SPPs in Bi2Te3 films for corresponding parameters: (c) for kF = 0.15 × 109 m1, (d) for kF = 1.52 × 109 m1.

Download Full Size | PDF

Similar to Bi2Te3 the data on Sb2Te3 [27, 28] was combined to make the hybrid Drude-Lorentz model with the parameters presented in Table 5. The behavior of the dispersion curves and propagation lengths is expectedly similar to Bi2Te3, given the similarity between the optical parameters for these materials. We therefore do not include the results of calculations here.

Tables Icon

Table 5. Parameters for the Drude-Lorentz model with ε = 51 extracted from the reflectance measurements on Sb2Te3 at room temperatures in far and mid-IR [27, 28].

4.3. SPP in superlattice

The study of SPPs in multi-layered graphene waveguides [35] shows that in the limit of long wavelengths (qd ≪ 1) the effective optical conductance increases linearly with the number of layers in the stack. Therefore it seems plausible to use superlattices made of alternating layers of thin films of TI and dielectric in order to tune the dispersion curve of a fundamental mode of SPP by growing the required number of layers. The growth of such superlattice using Bi2Se3 films has been reported by Chen et al. [36]. Since the SPP frequency grows with conductance, increasing the number of unit cells in the superlattice allows to shift-up the SPP frequency for a given wave-vector q. The illustration of such a shift is given in Fig. 8. The energy dispersion curves for SPP in layers of Bi2Se3/ZnSe (9 nm and 10 nm thick, respectively) were calculated using transfer matrix method. Optical conductance of TI surface was modeled using the Drude-like expression G(ω) = 0/[1 + (4ħωσ0)2] with σ0 = 138G0, G0 = 2e2/h, and µ = 0.500 eV. Bulk dielectric function of ZnSe in far-IR is taken from [37]. The results of calculations support the idea of changing the SPP dispersion relations by increasing number of “unit cells” in the superlattice.

 figure: Fig. 8

Fig. 8 Tuning SPP dispersion by varying the number of units cells in superlattice of TI and dielectric (ZnSe in this case). Transfer matrix calculations for the structure reported in [36].

Download Full Size | PDF

5. Conclusions

The results of this work can be summarized as follows:

  1. Dispersion relations and propagation lengths of SPPs in thin films of Bi2Se3, Bi2Te3, and Sb2Te3 were determined using realistic material parameters in far-IR. Bi2Se3 is identified as the material of choice if larger propagations lengths are desired.
  2. Key parameters influencing the propagation length are found to be 1) the Fermi level, EF, and 2) film thickness. Lowering EF by gating is a feasible way to control the propagation lengths of SPPs. Additional enhancement of propagation lengths can be achieved when working with thin films (d ≤ 20 nm). As an example, SPP in 15 nm thin film of Bi2Se3 with EF = 50 meV will propagate about 100 times farther than in 120 nm film with EF = 500 meV.
  3. The disagreement between the theory of SPP in TIs [23] and the first experimental measurements on Bi2Se3 [22] is removed by simply considering realistic optical properties of Bi2Se3. We stress again the importance of allowing the bulk dielectric function to vary when analyzing experiments or discussing possible applications.
  4. It was demonstrated that stacking of TI films and dielectrics into superlattice is a promising way to modify dispersion relations of SPPs. This raises an intriguing question: Can stacking (possibly combined with other controlling factors) shift the SPP frequency up into the spectral range of e.g. quantum cascade lasers. Such a scenario, if experimentally realized, may open a new venue for manipulating SPPs in terahertz range.

Acknowledgments

We would like to thank Dr. Pouyan Ghaemi and Dr. Alexander Punnoose for helpful discussions. We wish to acknowledge funding from CUNY ASRC Joint Seed Program. Research was carried out in part at the Center for Functional Nanomaterials, Brookhaven National Laboratory, supported by the U.S. Department of Energy, Office of Basic Energy Sciences, under Contract No. DE-SC0012704. This work was also supported in part by NSF DMR-1420634 and DOD-W911NF-13-1-0159. Vinod Menon would like to acknowledge funding from NSF through the EFRI-2DARE program (EFMA-1542863).

References and links

1. Y. Ando, “Topological insulator materials,” J. Phys. Soc. Jap. 82(10), 102001 (2013). [CrossRef]  

2. M. Z. Hasan and C. L. Kane, “Colloquium : topological insulators,” Rev. Mod. Phys. 82(4), 3045–3067 (2010). [CrossRef]  

3. X. L. Qi and S. C. Zhang, “Topological insulators and superconductors,” Rev. Mod. Phys. 83(4), 1057–1110 (2011). [CrossRef]  

4. L. Fu, C. L. Kane, and E. J. Mele, “Topological insulators in three dimensions,” Phys. Rev. Lett. 98(10), 106803 (2007). [CrossRef]   [PubMed]  

5. H. Zhang, C. X. Liu, X. L. Qi, X. Dai, Z. Fang, and S. C. Zhang, “Topological insulators in Bi2Se3, Bi2Te3 and Sb2Te3 with a single Dirac cone on the surface,” Nat. Phys. 5, 438–442 (2009). [CrossRef]  

6. O. V. Yazyev, J. E. Moore, and S. G. Louie, “Spin polarization and transport of surface states in the topological insulators Bi2Se3 and Bi2Te3 from first principles,” Phys. Rev. Lett. 105(26), 266806 (2010). [CrossRef]  

7. Y. L. Chen, J. G. Analytis, J.-H. Chu, Z. K. Liu, S.-K. Mo, X. L. Qi, H. J. Zhang, D. H. Lu, X. Dai, Z. Fang, S. C. Zhang, I. R. Fisher, Z. Hussain, and Z.-X. Shen, “Experimental realization of a three-dimensional topological insulator, Bi2Te3,” Science 325(5937), 178–181 (2009). [PubMed]  

8. D. Hsieh, Y. Xia, D. Qian, L. Wray, F. Meier, J. H. Dil, J. Osterwalder, L. Patthey, A. V. Fedorov, H. Lin, A. Bansil, D. Grauer, Y. S. Hor, R. J. Cava, and M. Z. Hasan, “Observation of time-reversal-protected single-Diraccone topological-insulator states in Bi2Te3 and Sb2Te3,” Phys. Rev. Lett. 103(14), 146401 (2009). [CrossRef]   [PubMed]  

9. Y. H. Wang, D. Hsieh, E. J. Sie, H. Steinberg, D. R. Gardner, Y. S. Lee, P. Jarillo-Herrero, and N. Gedik, “Measurement of intrinsic Dirac fermion cooling on the surface of the topological insulator Bi2Se3 using time-resolved and angle-resolved photoemission spectroscopy,” Phys. Rev. Lett. 109(12), 127401 (2012). [CrossRef]   [PubMed]  

10. S. Raghu, S. B. Chung, X. L. Qi, and S. C. Zhang, “Collective modes of a helical liquid,” Phys. Rev. Lett. 104(11), 116401 (2010) [CrossRef]   [PubMed]  

11. I. Appelbaum, H. D. Drew, and M. S. Fuhrer, “Proposal for a topological plasmon spin rectifier,” http://arxiv.org/abs/1011.2697

12. A. Yu. Kitaev, “Unpaired Majorana fermions in quantum wires,” Phys. Usp. 44(10S), 131 (2001). [CrossRef]  

13. Y. P. Lai, I.T. Lin, K. Wu, and J. M. Liu, “Plasmonics in topological insulators,” Nanomater. and Nanotecho. 4(13), 1–8 (2014). [CrossRef]  

14. S. Sim, M. Brahlek, N. Koirala, S. Cha, S. Oh, and H. Choi, “Ultrafast terahertz dynamics of hot Dirac-electron surface scattering in the topological insulator Bi2Se3,” Phys. Rev. B 89, 165137 (2014). [CrossRef]  

15. S. Sim, N. Koirala, M. Brahlek, J. H. Sung, J. Park, S. Cha, M.-H. Jo, S. Oh, and H. Choi, “Tunable Fano quantum-interference dynamics using a topological phase transition in (Bi1x Inx)2Se3,” Phys. Rev. B 91, 235438 (2015). [CrossRef]  

16. M. Autore, F. D’Apuzzo, A. Di Gaspare, V. Giliberti, O. Limaj, P. Roy, M. Brahlek, N. Koirala, S. Oh, F. J. G. de Abajo, and S. Lupi, “Plasmon-phonon interactions in topological insulator microrings,” Adv. Opt. Mat. 3, 1257–1263 (2015). [CrossRef]  

17. M. Autore, H. Engelkamp, F. D’Apuzzo, A. Di Gaspare, P. Di Pietro, I. L. Vecchio, M. Brahlek, N. Koirala, S. Oh, and S. Lupi, “Observation of magnetoplasmons in Bi2Se3 topological insulator,” ASC Photonics 2, 1231–1235 (2015). [CrossRef]  

18. P. Di Pietro, F. M. Vitucci, D. Nicoletti, L. Baldassarre, P. Calvani, R. Cava, Y. S. Hor, U. Schade, and S. Lupi, “Optical conductivity of bismuth-based topological insulators,” Phys. Rev. B 86, 045439 (2012). [CrossRef]  

19. M. Zhao, M. Bosman, M. Danesh, M. Zeng, P. Song, Y. Darma, A. Rusydi, H. Lin, C.-W. Qiu, and K. P. Loh, “Visible surface plasmon modes in single Bi2Te3 nanoplate,” Nano Lett. 15(12), 8331–8335 (2015). [CrossRef]   [PubMed]  

20. J.-Y. Ou, J.-K. So, G. Adamo, A. Sulaev, L. Wang, and N. I. Zheludev, “Ultraviolet and visible range plasmonics in the topological insulator Bi1.5Sb0.5Te1.8Se1.2,” Nature Commun. 55139 (2014). [CrossRef]  

21. S. A. Maier, Plasmonics: Fundamentals and Applications(Springer, 2007).

22. P. Di Pietro, M. Ortolani, O. Limaj, A. Di Gaspare, V. Giliberti, F. Giorgianni, M. Brahlek, N. Bansal, N. Koirala, S. Oh, P. Calvani, and S. Lupi, “Observation of Dirac plasmons in a topological insulator,” Nat. Nanotechnol. 8, 556–560 (2013). [CrossRef]   [PubMed]  

23. T. Stauber, G. Gomez-Santós, and L. Brey, “Spin-charge separation of plasmonic excitations in thin topological insulators,” Phys. Rev. B 88(20), 205427 (2013). [CrossRef]  

24. S. Das Sarma and E. H. Hwang, “Collective modes of the massless Dirac plasma,” Phys. Rev. Lett. 102(20), 206412 (2009). [CrossRef]   [PubMed]  

25. S. Roberts and D.D. Coon, “Far-infrared properties of quartz and sapphire,” J. Opt. Soc. Am. 52(9), 1023–1029 (1962). [CrossRef]  

26. M. Schubert, T. E. Tiwald, and C. M. Herzinger, “Infrared dielectric anisotropy and phonon modes of sapphire,” Phys. Rev. B 61(12), 8187–8201 (2000). [CrossRef]  

27. M.S. Wolf, Infrared and Optical Studies of Topological Insulators Bi2Se3, Bi2Te3 and Sb2Te3, (Thesis, The University of Akron, 2011).

28. W. Richter, H. Kohler, and C. R. Becker, “A Raman and far-infrared investigation of phonons in the rhombohedral V2-VI3 compounds,” Phys. Stat. Sol. (b) 84, 619–628 (1977). [CrossRef]  

29. V. R. Aguilar, A. V. Stier, W. Liu, L. S. Bilbro, D. K. George, N. Bansal, L. Wu, J. Cerne, A. G. Markelz, S. Oh, and N. P. Armitage, “Terahertz response and colossal Kerr rotation from the surface states of the topological insulator Bi2Se3,” Phys. Rev. Lett. 108(8), 087403 (2012). [CrossRef]  

30. R. E. V. Profumo, R. Asgari, M. Polini, and A. H. MacDonald, “Double-layer graphene and topological insulator thin-film plasmons,” Phys. Rev. B 85(8), 085443 (2012). [CrossRef]  

31. Z. Xie, S. He, C. Chen, Y. Feng, H. Yi, A. Liang, Z. Lin, D. Mou, J. He, Y. Peng, X. Liu, Y. Liu, G. Liu, X. Dong, L. Yu, J. Zhang, S. Zhang, Z. Wang, F. Zhang, F. Yang, Q. Peng, X. Wang, C. Chen, Z. Xu, and X. J. Zhou, “Orbital-selective spin texture and its manipulation in a topological insulator,” Nat. Commun. 5, 1–9 (2014). [CrossRef]  

32. D. Kim, S. Paul, N. P. Butch, J. Paglione, and M. S. Fuhrer, “Ambipolar surface state thermoelectric power of topological insulator Bi2Se3,” Nano Lett. 141701–1706 (2014). [CrossRef]  

33. A. L. Yeats, Y. Pan, A. Richardella, P. J. Mintun, N. Samarth, and D. D. Awschalom, “Persistent optical gating of a topological insulator,” Sci. Adv.1 (2014).

34. El. H. Kaddouri, T. Maurice, X. Gratens, S. Charar, S. Benet, A. Mefleh, J. C. Tedenac, and B. Liautard, “Optical properties of bismuth telluride thin films, Bi2Te3/Si(100) and Bi2Te)3/SiO2/Si(100),” Phys. Stat. Sol.(a) 176(2), 1071–1076 (1999). [CrossRef]  

35. D. A. Smirnova, I. V. Iorsh, I. V. Shadrivov, and Y. S. Kivshar, “Multilayer graphene waveguides,” JETP Lett. 99(8), 456–460 (2014). [CrossRef]  

36. Z. Chen, L. Zhao, K. Park, T. A. Garcia, M. C. Tamargo, and L. Krusin-Elbaum, “Robust topological interfaces and charge transfer in epitaxial Bi2Se3/II–VI semiconductor superlattices,” Nano Lett. 15(10), 6365–6370 (2015). [CrossRef]   [PubMed]  

37. A. Deneuville, D. Tanner, and P. H. Holloway, “Optical constants of ZnSe in the far infrared,” Phys. Rev. B 43(8), 6544–6550 (1991). [CrossRef]  

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 (8)

Fig. 1
Fig. 1 Schematics of the experiment for observing SPP resonances in Bi2Se3 (see [22]). Thin film, MBE grown on sapphire, is etched into a periodic array of ribbons of the width W, separated by a gap of the same width. Incident linearly polarized far-IR light excites standing waves of collective oscillations of surface charges. The resonant excitation for given ribbon width W is observed in the drop of the transmitted light intensity T(ν). The surface current density j, plotted on the vertical axis, describes standing waves with the period 2W.
Fig. 2
Fig. 2 Contributions of various terms of the Drude-Lorentz model into the real part of the far-IR bulk dielectric function of Bi2Se3. Two IR-active phonon modes (Lorentz-α with ω ≈ 61 cm1 and Lorentz-β with ω ≈ 133 cm1) correspond to the first two terms in the Drude-Lorentz model (3). The oscillator Lorentz-Ω accounts for higher frequency absorption, in this case at ω ≈ 2029.5 cm1 = 252 meV which is close to the band-gap energy of Bi2Se3 bulk [28].
Fig. 3
Fig. 3 Real ε′ and imaginary ε″ parts of the bulk dielectric function of Bi2Se3, according to the Drude-Lorentz model. Solid lines show the bulk dielectric function without the Drude term, e.g. in a perfect crystal at low temperature. Dashed lines reproduce the bulk dielectric function with the Drude term present. The effect of Drude term is very strong in this spectral range.
Fig. 4
Fig. 4 (a, b) Dispersion relations of SPPs in Bi2Se3 films of several thicknesses for (a) kF = 0.15 × 109 m1 (EF = 50 meV) and (b) kF = 1.52 × 109 m1 (EF = 500 meV). Cirlce, square and triangle marks represent the experimental values from [22]. (c, d) Propagation lengths of SPPs in Bi2Se3 films for corresponding parameters: (c) for kF = 0.15 × 109 m1, (d) for kF = 1.52 × 109 m1. The lines marked α ≈ 63 cm1 and β ≈ 133 cm1 indicate positions of IR-active modes interacting with SPPs and causing non-monotinic behavior of dispersion and propagation length.
Fig. 5
Fig. 5 Dispersion curves of SPPs in Bi2Se3 thin films, calculated using the Drude-Lorentz model and the parameters reported in [22] (kF = 1.37 × 109 m1, EF = 541 meV). Cirlce, square and triangle marks correspond to the experimental results [22].
Fig. 6
Fig. 6 Propagation lengths of SPPs in Bi2Se3 thin films, calculated using the Drude-Lorentz model and the parameters reported in [22] (kF = 1.37 × 109 m1, EF = 541 meV).
Fig. 7
Fig. 7 (a, b) Dispersion relations of SPPs in Bi2Te3 films of several thicknesses for (a) kF = 0.15 × 109 m1 (EF = 50 meV) and (b) kF = 1.52 × 109 m1 (EF = 500 meV). (c, d) Propagation lengths of SPPs in Bi2Te3 films for corresponding parameters: (c) for kF = 0.15 × 109 m1, (d) for kF = 1.52 × 109 m1.
Fig. 8
Fig. 8 Tuning SPP dispersion by varying the number of units cells in superlattice of TI and dielectric (ZnSe in this case). Transfer matrix calculations for the structure reported in [36].

Tables (5)

Tables Icon

Table 1 Parameters for the Drude-Lorentz model with ε = 1, extracted from the reflectance measurements on Bi2Se3 at room temperatures in far and mid-IR [27].

Tables Icon

Table 2 Parameters for the Drude-Lorentz model for the unpatterned Bi2Se3 thin films at 300 K, see Supplementary Material to [22].

Tables Icon

Table 3 Parameters of the Drude-Lorentz model with ε = 85, extracted from the reflectance measurements on Bi2Te3 at room temperatures in far-IR [28].

Tables Icon

Table 4 Parameters for the Drude-Lorentz model with ε = 1 extracted from the reflectance measurements on Bi2Te3 at room temperatures in far- and mid-IR [27].

Tables Icon

Table 5 Parameters for the Drude-Lorentz model with ε = 51 extracted from the reflectance measurements on Sb2Te3 at room temperatures in far and mid-IR [27, 28].

Equations (4)

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

ω 2 = v F k F e 2 ε 0 h q ε T + ε B + q d ε T I
ε ( ω , cm 1 ) = n 0 2 + ( n 0 2 1 ) ( λ ω ) 2 + i γ ( n 0 2 1 ) ( λ ω ) ,
ε ( ω , cm 1 ) = ε ω D 2 ω 2 + i ω γ D + j = 1 j = 3 , 4 ω p j 2 ω 0 j 2 ω 2 i ω γ j .
q ( ω ) = A ω 2 [ ε T ( ω ) + ε B ( ω ) ] 1 A ω 2 d ε T I ( ω ) , A ε 0 h ν F k F e 2 .
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.