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

Graphene-tuned EIT-like effect in photonic multilayers for actively controlled light absorption of topological insulators

Open Access Open Access

Abstract

As newly emerging nanomaterials, topological insulators with unique conducting surface states that are protected by time-reversal symmetry present excellent prospects in electronics and photonics. The active control of light absorption in topological insulators are essential for the achievement of novel optoelectronic devices. Herein, we investigate the controllable light absorption of topological insulators in Tamm plasmon multilayer systems composed of a Bi1.5Sb0.5Te1.8Se1.2 (BSTS) film and a dielectric Bragg mirror with a graphene-involved defect layer. The results show that an ultranarrow electromagnetically induced transparency (EIT)-like window can be generated in the broad absorption spectrum. Based on the EIT-like effect, the Tamm plasmon enhanced light absorption of topological insulators can be dynamically tuned by adjusting the gate voltage on graphene in the defect layer. These results will pave a new avenue for the realization of topological insulator-based active optoelectronic functionalities, for instance light modulation and switching.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Topological insulators, as novel quantum nanomaterials, have recently gained particular attention in electronics, optoelectronics and optics owing to unique electronic and optical characteristics [16]. Different from original insulators and metals, topological insulators exhibit a topologically-protected conducting surface (or edge) state and insulating bulk state simultaneously [13]. In 2006, Zhang et al. theoretically predicted the topological properties of two-dimensional (2D) HgTe quantum wells with the generation of the quantum spin Hall effect [5]. Next year, the prediction was verified experimentally by Molenkamp et al. [6]. In the 2D structured material, gapless Dirac fermions appear in the edge state because of strong spin-orbit coupling in the bulk state [5,6]. Subsequently, similar topological Dirac surface states were theoretically proposed and experimentally demonstrated in three-dimensional (3D) materials including Bi2Se3, Bi2Te3, Sb2Te3 and their composition Bi2-xSbxTe3-ySey [710]. The time-reversal symmetry of the surface states in topological insulators renders the travelling electrons avoiding from the back scattering of non-magnetic impurities [1,2,5,6]. Until now, some attractive physical phenomena have been reported in topological insulators, including magnetoelectric effects, Majorana fermions and superconductivity [13]. Topological insulators with fascinating electronic features offer new routes to the development of quantum computing, spintronics and new electronic devices [11]. Furthermore, exciting optical characteristics were reported in 3D topological insulator materials, for instance ultrahigh optical nonlinearities [12], saturable absorption [13], photonic Weyl points [14], optical resonance [15], photothermal conversion [16], surface plasmon polaritons [1720] and Tamm plasmon polaritons (TPPs) [21]. Taking advantage of these optical responses, a large number of high-performance functionalities were proposed based on topological insulators, such as all-optical switching [22], mode-locking fiber laser [23], nanometric hologram [15], optical modulation [19], angular-momentum nanometrology [24], light energy harvesting [25,26] and refractive index monitoring [27]. The light absorption of topological insulators is one of the most fundamental physical issues for optical nonlinearity and optoelectronic functionalities [23,25,26]. The dynamical tunability of light absorption is crucial for active optical and optoelectronic devices, especially modulators and switches [19,25,26,28]. The light absorption of topological insulators can be enhanced and tailored by TPPs [21]. But, how to realize active control of light absorption from topological insulators? The electrical control of Fermi level for topological insulators was used to modulate light transmission/reflection, which may offer a good method to tune light absorption of topological insulators [29]. The active tunability of light absorption without changing material properties is also significant.

We propose an effective method to actively control light absorption of topological insulators in a multilayer photonic system composed of a BSTS film and a dielectric Bragg mirror with a graphene-involved defect layer. TPPs generated between the BSTS film and Bragg mirror contribute to the strong light absorption of BSTS at near-infrared wavelengths. The introduction of a defect layer in the Bragg mirror enables the appearance of ultranarrow induced transparency in the absorption spectrum, derived from the formation of an electromagnetically induced transparency (EIT)-like effect. Particularly, we find that light absorption of a topological insulator can be effectively tuned by adjusting the gate voltage on graphene in the defect layer. The mechanism of light absorption control was reasonably analyzed using a two-oscillator model. The results will open a new route for topological insulator-based active electro-optic devices.

2. Structure and model

As depicted in Fig. 1, the multilayer photonic structure consists of a BSTS topological insulator film, a dielectric Bragg mirror with alternately stacked SiO2/Si3N4 layers, and a defect layer. The defect layer contains a doped Si layer and an Al2O3 layer with an inserted single-layer graphene. The gate voltage Vb is exerted between the graphene and Si layers. The incident light impinges on the BSTS film with an angle of θ. dt, ds, da, db, dg, and dc stand for the thicknesses of BSTS, Si, Si3N4, SiO2, graphene, and defect layers, respectively. di represents the thickness of the spacer between the graphene and Si layer. The period number of SiO2/Si3N4 Bragg mirror is set as N. At the wavelengths of interest, the refractive indices of SiO2, Si3N4, Al2O3, and Si can be set as 1.45, 2.2, 1.76, and 3.5, respectively. The refractive index of graphene can be derived from the Kubo formula [30,31]. According to the Kubo formula, the surface conductivity of single-layer graphene can be described as σg=σintra+σinter. Here, σintra and σinter correspond to the intraband and interband transitions of electrons in graphene, respectively. σintra can be described as

$${\sigma _{intra}} = \frac{{i{\textrm{e}^2}{k_B}T}}{{\pi {\hbar ^2}(\omega + i{\tau ^{ - 1}})}}[\frac{{{\mu _c}}}{{{k_B}T}} + 2\ln (\textrm{exp} ( - \frac{{{\mu _c}}}{{{k_B}T}}) + 1)],$$
where e, kB, T, and μc denote the electron charge, Boltzmann’s constant, temperature, and chemical potential of graphene, respectively. ω, ħ, and τ stand for the angular frequency of incident light, reduced Planck's constant, and momentum relaxation time of charge carrier scattering, respectively. When |μc|>>kBT and ħω>>kBT, σinter is governed by
$${\sigma _{inter}} = \frac{{i{\textrm{e}^2}}}{{4\mathrm{\pi}\hbar }}\ln [\frac{{2|{{\mu_c}} |- \hbar (\omega + i{\tau ^{ - 1}})}}{{2|{{\mu_c}} |+ \hbar (\omega + i{\tau ^{ - 1}})}}].$$
Here, the relaxation time depends on the carrier mobility μ and can be expressed as τ=μμc/(evf2). The Fermi velocity can be set as vf=106 m/s [31]. As previously reported, the carrier mobility of single-layer graphene suspended in air could exceed 2×105 cm2V−1s−1 [31]. The carrier mobility of graphene on a SiO2 substrate could approach 4×104 cm2V−1s−1 [32]. The carrier mobility can exceed 2×103 cm2V−1s−1 for CVD-grown graphene on Al2O3 [33]. To ensure the credibility of results, the carrier mobility of graphene is set as μ=1×104 cm2V−1s−1 in this work. The chemical potential of graphene can be described as μc=ħvfns)1/2 [34]. The doping level of graphene can be tailored by altering the gate voltage Vb on graphene through ns=εaε0|Vb|/(edi), where εa is the relative permittivity of Al2O3 layer. The single-layer graphene can be equivalent to an ultrathin film with a thickness dg=0.34 nm, thus the relative permittivity of graphene can be described as εg=1 + g/(ωε0dg) [30]. The topological insulator can be regarded as a bulk insulator coated with a surface conductor [18,35]. The BSTS topological insulator has attracted special attention due to the good bulk interior and surface electronic transport [35]. The surface conductor can be treated as an ultrathin layer with a thickness of t = 1.5 nm [18]. The relative permittivities of the surface and bulk states in the near-infrared region can be obtained by fitting the measured data with the Drude and Tauc-Lorentz models, respectively [18]. The relative permittivity of the bulk state is mainly attributed to the interband transition, whose band structure can be governed by Kramers-Kronig relations [18,36]. The imaginary and real parts (εb'’ and εb) of relative permittivity for the bulk state can be described as
$$\varepsilon _b^{\prime\prime}(E )= \left\{ \begin{array}{ll} \frac{{A{E_0}C{{({E - {E_g}} )}^2}}}{{{C^2}{E^2} + {{({{E^2} - E_0^2} )}^2}}} \cdot \frac{1}{E}, &E > {E_g} \\ 0 &E \le {E_g} \end{array} \right.$$
$$\varepsilon _b^{\prime}(E )= {\varepsilon _b}(\infty )+ \frac{2}{\pi}P\int_{E_g}^\infty {\frac{{\xi \varepsilon _b^{\prime\prime}}(\xi )}{{{\xi ^2} - {E^2}}}}d \xi ,$$
where A, E0, Eg, and C denote the absorption peak amplitude, peak in joint density of states, band gap energy, and broadening factor, respectively; E=ħω is the energy of incident photons; εb(∞) and P represent the relative permittivity at high frequency and Cauchy principal part of the integral, respectively [36]. The relative permittivity of the surface state can be described using the classical Drude model: εs(ω)=ε-ω2 p/[ω(ω+)], where ε, ωp, and γ are the relative permittivity at the infinite frequency, bulk plasma frequency, and electron collision frequency, respectively. The parameters for the BSTS material can be set as ε=1.3, ωp=7.5 eV, γ=0.05 eV, A = 65.9, Eg=0.25 eV, E0=1.94 eV, C = 1.94, and εb(∞) = 0 [18]. The results in [18] show that the relative permittivity of the BSTS bulk state presents the lossy insulating characteristic in the near-infrared region. The BSTS surface state has negative real permittivities and exhibits a metal-like property. The transfer matrix method (TMM) is an effective theoretical method to study the light propagations in the multilayer structures [37,38]. Derived from Maxwell’s equations and boundary conditions of electric and magnetic fields, the transfer matrix equations of the multilayer structure can be described as
$$\left[ \begin{array}{l} E_{up}^ - \\ E_{up}^ + \end{array} \right] = {M_1}{P_1}{M_2}{P_2} \cdots {M_{2N + 8}}\left[ \begin{array}{l} E_{down}^ - \\ E_{down}^ + \end{array} \right] = \left( {\begin{array}{cc} {{Q_{11}}}&{{Q_{12}}}\\ {{Q_{21}}}&{{Q_{22}}} \end{array}} \right)\left[ \begin{array}{l} E_{down}^ - \\ E_{down}^ + \end{array} \right],$$
where $E_{\rm{up}}^{+}$ and $E_{\rm{up}}^{-}$ are the electric fields of incident and reflection light on the top of the multilayer structure, respectively; $E_{\rm{down}}^{+}$ and $E_{\rm{down}}^{-}$ are the electric fields of output and input light at the bottom of the structure, respectively. Thus, the absorption spectra of the structure can be expressed as A = 1-|Q21/Q11|2-|1/Q11|2. For TM-polarized incident light, the transfer matrixes Mj (j = 1, 2, …, 2N + 8) and Pj (j = 1, 2, …, 2N + 7) can be described as
$${M_j} = \frac{1}{{2{n_{j - 1}}\cos {\theta _{j - 1}}}}\left( {\begin{array}{cc} {{n_{j - 1}}\cos {\theta_j} + {n_j}\cos {\theta_{j - 1}}}&{{n_{j - 1}}\cos {\theta_j} - {n_j}\cos {\theta_{j - 1}}}\\ {{n_{j - 1}}\cos {\theta_j} - {n_j}\cos {\theta_{j - 1}}}&{{n_{j - 1}}\cos {\theta_j} + {n_j}\cos {\theta_{j - 1}}} \end{array}} \right),$$
$${P_j} = \left( {\begin{array}{cc} {\textrm{exp} ( - i2\pi {d_j}{n_j}\cos {\theta_j}/\lambda )}&0\\ 0&{\textrm{exp} (i2\pi {d_j}{n_j}\cos {\theta_j}/\lambda )} \end{array}} \right),$$
where θj and nj respectively stand for the propagation angle of light and refractive index of the j-th layer, which can be governed by Snell’s law: njsinθj=nj−1sinθj−1 (θ0=θ and n0=1). dj represents the thickness of the j-th layer. λ=2πc/ω denotes the incident wavelength.

 figure: Fig. 1.

Fig. 1. 3D diagram of the multilayer photonic system consisting of a thin BSTS topological insulator film, a Bragg mirror with stacked Si3N4/SiO2 layers and a defect layer (containing Al2O3, graphene, and doped Si) in the middle of Bragg mirror.

Download Full Size | PDF

3. Results and analysis

The parameters of the BSTS film/Bragg mirror structure are originally set as dt=58 nm, da=150 nm, db=240 nm, dc=0 nm, N = 20, and θ=0° to explore the light propagation properties. The TMM theoretical calculations illustrate that the absorption spectrum of multilayer photonic structures presents a distinct peak at the wavelength of 1309.6 nm, as shown in Fig. 2(a). The strong light absorption is derived from the generation of TPP mode between the BSTS and Bragg mirror [21,37]. The topological insulators possess a metal-like surface state with negative permittivities, providing required condition for the formation of near-infrared TPPs [37]. To clarify the theoretical results, we employ the finite-difference time-domain (FDTD) method to simulate the light propagation [38,39]. In the simulations, the top and bottom of multilayer structures are set as the absorbing boundary conditions with perfectly matched layers. Periodic boundary conditions can be set on the other sides of computational space. The non-uniform mesh is used to simulate the response of all the layers. The mesh sizes of surface and bulk layers are set as 0.3 nm and 2 nm, respectively. The mesh size of graphene is set as 0.06 nm. The absorption spectra of multilayer structures can be obtained through A = 1-|Pt/Pi|-|Pr/Pi|, where Pt, Pr, and Pi stand for the transmitted, reflected, and incident light powers, respectively. The electric field amplitude of incident light is set as 1 V/m. As depicted in Fig. 2(a), the simulations are in good agreement with the TMM calculations. Figure 2(b) shows the electric field distribution and intensity profile in the multilayer structure at the wavelength of 1309.6 nm. An obvious TPP mode is observed between the BSTS film and Bragg mirror [21,38]. The field intensity of BSTS-based TPP mode is ∼20 times less than that of Au-based TPP mode [40]. Even so, the light absorption of a topological insulator can be significantly enhanced by the formation of TPPs [21].

 figure: Fig. 2.

Fig. 2. (a) Absorption spectrum of the multilayer structure with dc=0 nm. (b) Corresponding electric field distribution and intensity profile in the structure at the absorption peak. (c) Absorption spectrum of the multilayer structure with an Al2O3 defect layer (dc=154.5 nm and dg=ds=0 nm) in the Bragg mirror. The inset shows the prototype three-level model of the EIT effect. (d) Corresponding electric field distribution and intensity profile at the induced absorption wavelength (λ=1309.9 nm). Here, dt=58 nm, da=150 nm, db=240 nm, θ=0°, and N = 20. The circles and curves in (a) and (c) denote the FDTD and TMM results, respectively.

Download Full Size | PDF

Subsequently, we study the light propagation characteristics of the multilayer structure with an Al2O3 defect layer in the middle of Bragg mirror. It shows in Fig. 2(c) that an ultranarrow dip appears in the broad absorption spectrum at the wavelength of 1309.9 nm when dc=154.5 nm. This nonintuitive phenomenon can be reasonably attributed to the generation of optical effect in analogy to the EIT in atomic systems [41]. According to the prototype three-level model of the EIT effect, the TPP mode between the BSTS film and Bragg mirror can be analogous to the upper state 2>, and thus the excitation of the topological insulator-based TPP mode can be considered as the transition from the ground state 1 > to 2 > [41,42]. The incident light is regarded as the “probe light”. An optical mode formed in the defect layer can be in analogy to the upper state 3 > . The defect mode can be excited by the coupling of TPP mode, similar to the transition from the state 2 > to 3 > . Even though the distance between the BSTS and defect layers is ∼3.95 µm, the TPP field can be effectively coupled with the defect mode due to the relatively weak confinement of TI-based TPPs. The coupling from defect mode to TPP mode can be considered as “pump light”. The two transition pathways: 1>→2 > and 1>→2>→3>→2 > will destructively interfere with each other, resulting in the weakness of the TPP field, as shown in Fig. 2(d). The weakened TPP field gives rise to the reduction of light absorption from the topological insulator and the formation of ultranarrow induced transparency with a spectral width of 0.63 nm, as shown in Fig. 2(c). The numerical simulations are in excellent agreement with the theoretical calculations. Additionally, we investigate the dependence of the EIT-like response on the thickness of BSTS film dt. As depicted in Fig. 3(a), the broad absorption peak possesses a red shift with increasing dt, while the induced transparency wavelength keeps unchanged. The position of induced transparency is determined by the wavelength of defect mode [43]. The defect layer can be regarded as a Fabry-Pérot (FP) cavity, whose resonance wavelength can be described as λ=4πncdc/(2kπ-φ1-φ2) [44]. Here, φ1 and φ2 are the phase shifts of reflection on the two sides of defect cavity, respectively; nc is the refractive index of defect cavity; k is an integer, representing resonant order. The TPP wavelength exhibits a red shift with increasing dt [21], while the wavelength of defect mode is unrelated with dt. Therefore, we can observe a nearly unchanged induced transparency window in the red-shift absorption peak. As shown in Fig. 3(b), both the induced transparency window and the absorption peak present a red shift as the Bragg mirror layer thickness da increases. Moreover, the induced transparency has a linearly red shift in an invariant absorption peak with increasing dc, as depicted in Fig. 3(c). According to the resonance condition, we can see that the wavelength of defect mode is proportional to dc. However, the TPP mode is nearly independent on the defect layer. Thus, we can see the red-shift induced transparency in a fixed absorption peak with increasing dc. In Fig. 3(d), the induced transparency in the middle of absorption peak presents a blue shift with increasing θ. Similar to the previous reports, the TPP wavelength possesses a blue shift as θ increases [21,37]. The wavelength of the defect mode in the Bragg mirror also has a blue shift with increasing θ [45]. Thus, we observe the blue shift of induced transparency window with increasing θ. The simulations agree well with theoretical calculations. Figure 4 depicts the electric field amplitudes of EIT resonances in multilayer structures with different dt, da, dc, and θ. It shows that the EIT electric field amplitude nearly keeps unchanged with altering dt, da and dc, while decreases with the increase of θ.

 figure: Fig. 3.

Fig. 3. Evolution of absorption spectrum from the multilayer structure with (a) the thickness of the BSTS film dt when da=150 nm, dc=154.5 nm, and θ=0°, (b) the thickness of Si3N4 layer da when dt=58 nm, dc=154.5 nm, and θ=0°, (c) the thickness of Al2O3 defect layer dc when dt=58 nm, da=150 nm, and θ=0°, and (d) the light incident angle θ when dt=58 nm, da=150 nm, and dc=154.5 nm. Here, db=240 nm and N = 20. The circles denote the induced transparency wavelengths obtained by numerical simulations.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. (a)-(d) Electric field amplitudes of EIT resonances (defect modes) versus dt, da, dc, and θ in the multilayer structure with the same parameters as in Figs. 3(a)–3(d).

Download Full Size | PDF

The TPP excitation between the BSTS film and Bragg mirror can effectively boost light absorption of the topological insulator. The coupling between the TPP and defect modes in the multilayer structure enables the tailoring of light absorption from the topological insulator with the generation of the EIT-like effect. To realize the active control of light absorption, we integrate the graphene and Si layer into the defect layer, as shown in Fig. 1. The parameters are set as da=150 nm, db=240 nm, du=50 nm, di=5 nm, ds=60 nm, θ=0°, and N = 20. The in-plane electric fields dominate in the TPP and defect modes of multilayer structure. Graphene can be simply treated as an isotropic material [46]. Figure 5(a) depicts the spectral evolution of light absorption in multilayer structure with the gate voltage Vb between the graphene and Si layers. An ultranarrow transparency window obviously appears as Vb exceeds about 4.8 V, at which the interband transition of electrons in graphene will be hindered as the photon energy of defect mode (λ≈1310 nm) is less than 2μc (i.e., ħω<2μc). Moreover, the induced transparency wavelength originally presents a red shift and then has a blue shift with increasing Vb, which derives from the alternation of graphene refractive index. Figure 5(b) shows the electric field intensity profiles in the structure at the wavelength of 1310 nm when Vb=0 V and 6 V. We can see that the electric field of defect mode is relatively low when Vb=0 V, compared with the case when Vb=6 V. This is due to that the dissipative loss of graphene is greatly reduced when Vb changes from 0 V to 6 V. The electric field of TPP mode is suppressed when Vb=6 V, giving rise to the weakening of light absorption from the BSTS film, as depicted in the insets of Fig. 5(b). Thus, the gate-modulated BSTS absorption can be realized at the induced transparency wavelength, as shown in Figs. 5(c) and 5(d). The modulation depth of light absorption from the topological insulator is ∼3.67 dB when Vb alters from 0 V to 6 V. When the carrier mobility of graphene is decreased to 2000 cm2V−1s−1 with Vb=6 V, the real permittivity of graphene is almost unchanged, but the imaginary permittivity increases from 0.067 to 0.33. The BSTS absorption dip has only a slight rise of 0.02.

 figure: Fig. 5.

Fig. 5. (a) Evolution of absorption spectrum from the multilayer structure with the gate voltage Vb on graphene. (b) Electric field intensity profile in the structure when Vb=0 V and 6 V. The insets depict the intensity profiles in the BSTS film when Vb=0 V and 6 V. (c)-(d) Absorption spectra of the BSTS film and graphene when Vb=0 V and 6 V. The parameters are set as da=150 nm, db=240 nm, du=50 nm, di=5 nm, ds=60 nm, θ=0°, and N = 20.

Download Full Size | PDF

To clarify the physical mechanism of the tunability of light absorption in the multilayer structure, we theoretically investigate the dependence of the EIT-like response on the gate voltage on graphene. The EIT-like spectrum can be described by using a two-oscillator model. As mentioned above, the TPP structure composed of BSTS film and a Bragg mirror can be regarded as “oscillator 1”, where the TPP mode is directly excited by the incident light. The defect layer can be considered as “oscillator 2”, where the defect mode can be generated through coupling with the TPP mode. As depicted in the inset of Fig. 2(c), ω0 and δ stand for the resonance frequency of oscillator 1 and the resonance frequency detuning between the oscillators 1 and 2, respectively; γ1 and γ2 are the decaying rates derived from the intrinsic loss of the oscillators 1 and 2, respectively; κe are the coupling coefficient between the oscillators 1 and 2. When ω0>>γ1>>γ2, γ1>>|δ| and ω0>>|ω-ω0|, the light absorption of multilayer structure can be described as [47]

$$A\textrm{ = }{\mathop{\rm Im}\nolimits} \left\{ {\frac{{f(\omega - {\omega_0} + \delta + i{\gamma_2}/2)}}{{{{(\kappa {e^{i\varphi }})}^2} - (\omega - {\omega_0} + \delta + i{\gamma_2}/2)(\omega - {\omega_0} + i{\gamma_1}/2)}}} \right\},$$
where f is an amplitude coefficient. By using Eq. (8), we can achieve the light absorption spectra of a multilayer system with different gate voltages Vb by fitting the above calculated results. As shown in Fig. 6(a), the fitting results agree well with the TMM calculations. The fitted physical parameters f, ω0, γ1 and φ in Eq. (8) are about 0.892, 1.439×1015 rad/s, 1.55×1014 rad/s and π, respectively. The values of δ, γ2 and κ change with Vb, as depicted in Figs. 6(b)–6(d). We can see in Fig. 6(b) that the resonance frequency detuning δ gradually increases when Vb is adjusted from 0 V to 4.8 V, then decreases when Vb continuously increases. This contributes to the evolution of induced transparency wavelength in Fig. 5(a). The decay rate γ2 nearly remain constant when Vb<4.7 V, and then quickly alters when Vb is around 4.8 V. γ2 decreases from 1.72×1012 rad/s to 0.19×1012 rad/s when Vb changes from 4.7 V to 5V. It illustrates that the dissipative loss of defect layer dramatically alters when the gate voltage approaches ∼4.8 V, which is well consistent with the above analysis. The spectra in Fig. 3 can also be fitted using the above equation. The fitting process is still valid when δ is much larger than γ2. It is worth noting that the coupling strength κ is also related to the gate voltage Vb, as shown in Fig. 6(d). κ has unchanged value of 2.4×1012 rad/s when Vb<4.7 V, and then quickly ascends to 3.4×1014 rad/s when Vb=5 V. These theoretical calculations show that the active tuning of the EIT-like effect derives from the alternation of the decaying rate for the defect mode and the coupling strength between the TPP and defect modes, which strongly depend on the gate voltage on graphene. This offers a new avenue for active tunability of light absorption of topological insulators.

 figure: Fig. 6.

Fig. 6. (a) TMM calculated (dots) and fitting (curves) results of light absorption spectrum from the multilayer structures with different gate voltages Vb on graphene. (b) Fitting values of resonance frequency detuning δ between the oscillators 1 and 2 with different Vb. (c) Fitting values of decay rate γ2 in the oscillator 2 with different Vb. (d) Fitting values of coupling strength κ between the oscillators 1 and 2 with different Vb. Here, da=150 nm, db=240 nm, du=50 nm, di=5 nm, ds=60 nm, θ=0°, and N = 20.

Download Full Size | PDF

4. Conclusions

We have investigated light propagation characteristics in multilayer photonic structures composed of BSTS topological insulator film and a SiO2/Si3N4 Bragg mirror. The TPP mode can be generated between the BSTS film and Bragg mirror, inducing the strong light absorption of the topological insulator. When an Al2O3 defect layer is introduced in the Bragg mirror, an ultranarrow induced transparency can be realized in the absorption spectrum with the formation of the EIT-like effect in the multilayer structure because of the destructive interference between the TPP mode and defect mode. It is found that the EIT-like response is strongly dependent on the thicknesses of the BSTS film, Bragg mirror layer, defect layer and light incident angle. The theoretical results are in excellent agreement with numerical simulations. Especially, a single-layer graphene is integrated into the defect layer, and thus the dynamic tunability of light absorption in the topological insulator can be achieved by adjusting the gate voltage on graphene. According to two-oscillator model, we find that the tunability of light absorption is attributed to the graphene-controlled decaying rate of defect mode and coupling strength between the TPP and defect modes. Our results will pave a new pathway toward the active control of light absorption and novel topological insulator optoelectronic devices, especially modulators and switches.

Funding

National Key Research and Development Program of China (2017YFA0303800); National Natural Science Foundation of China (11974283, 11774290, 61705186, 11634010); Natural Science Basic Research Plan in Shaanxi Province of China (2020JM-130, 2018JM6013).

Disclosures

The authors declare no conflicts of interest.

References

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

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

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

4. J. Toudert and R. Serna, “Interband transitions in semi-metals, semiconductors, and topological insulators: a new driving force for plasmonics and nanophotonics [Invited],” Opt. Mater. Express 7(7), 2299–2325 (2017). [CrossRef]  

5. B. Bernevig, T. Hughes, and S. Zhang, “Quantum spin Hall effect and topological phase transition in HgTe quantum wells,” Science 314(5806), 1757–1761 (2006). [CrossRef]  

6. M. König, S. Wiedmann, C. Brüne, A. Roth, H. Buhmann, L. Molenkamp, X. Qi, and S. Zhang, “Quantum Spin Hall Insulator State in HgTe Quantum Wells,” Science 318(5851), 766–770 (2007). [CrossRef]  

7. H. Zhang, C. Liu, X. Qi, X. Dai, Z. Fang, and S. Zhang, “Topological insulators in Bi2Se3, Bi2Te3 and Sb2Te3 with a single Dirac cone on the surface,” Nat. Phys. 5(6), 438–442 (2009). [CrossRef]  

8. T. Zhang, P. Cheng, X. Chen, J. Jia, X. Ma, K. He, L. Wang, H. Zhang, X. Dai, Z. Fang, X. Xie, and Q. Xue, “Experimental demonstration of topological surface states protected by time-reversal symmetry,” Phys. Rev. Lett. 103(26), 266803 (2009). [CrossRef]  

9. Y. Xia, D. Qian, D. Hsieh, L. Wray, A. Pal, H. Lin, A. Bansil, D. Grauer, Y. Hor, R. Cava, and M. Hasan, “Observation of a large-gap topological-insulator class with a single Dirac cone on the surface,” Nat. Phys. 5(6), 398–402 (2009). [CrossRef]  

10. T. Arakane, T. Sato, S. Souma, K. Kosaka, K. Nakayama, M. Komatsu, T. Takahashi, Z. Ren, K. Segawa, and Y. Ando, “Tunable Dirac cone in the topological insulator Bi2-xSbxTe3-ySey,” Nat. Commun. 3(1), 636 (2012). [CrossRef]  

11. J. E. Moore, “The birth of topological insulators,” Nature 464(7286), 194–198 (2010). [CrossRef]  

12. S. Lu, C. Zhao, Y. Zou, S. Chen, Y. Chen, Y. Li, H. Zhang, S. Wen, and D. Tang, “Third order nonlinear optical property of Bi2Se3,” Opt. Express 21(2), 2072–2082 (2013). [CrossRef]  

13. Y. Tan, H. Zhang, C. Zhao, S. Akhmadaliev, S. Zhou, and F. Chen, “Bi2Se3 Q-switched Nd: YAG ceramic waveguide laser,” Opt. Lett. 40(4), 637–640 (2015). [CrossRef]  

14. E. Goi, Z. J. Yue, B. P. Cumming, and M. Gu, “Observation of type I photonic Weyl points in optical frequencies,” Laser Photonics Rev. 12(2), 1700271 (2018). [CrossRef]  

15. Z. Yue, G. Xue, J. Liu, Y. Wang, and M. Gu, “Nanometric holograms based on a topological insulator material,” Nat. Commun. 8(1), 15354 (2017). [CrossRef]  

16. H. Xie, Z. Li, Z. Sun, J. Shao, X. Yu, Z. Guo, J. Wang, Q. Xiao, H. Wang, Q. Wang, H. Zhang, and P. Chu, “Metabolizable ultrathin Bi2Se3 nanosheets in imaging-guided photothermal therapy,” Small 12(30), 4136–4145 (2016). [CrossRef]  

17. P. Pietro, M. Ortolani, O. Limaj, A. 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(8), 556–560 (2013). [CrossRef]  

18. J. Ou, J. So, G. Adamo, A. Sulaev, L. Wang, and N. Zheludev, “Ultraviolet and visible range plasmonics in the topological insulator Bi1.5Sb0.5Te1.8Se1.2,” Nat. Commun. 5(1), 5139 (2014). [CrossRef]  

19. S. Sim, H. Jang, N. Koirala, M. Brahlek, J. Moon, J. Sung, J. Park, S. Cha, S. Oh, M. Jo, J. Ahn, and H. Choi, “Ultra-high modulation depth exceeding 2,400% in optically controlled topological surface plasmons,” Nat. Commun. 6(1), 8814 (2015). [CrossRef]  

20. J. Yuan, W. Ma, L. Zhang, Y. Lu, M. Zhao, H. Guo, J. Zhao, W. Yu, Y. Zhang, K. Zhang, H. Hoh, X. Li, K. Loh, S. Li, C. Qiu, and Q. Bao, “Infrared nanoimaging reveals the surface metallic plasmons in topological insulator,” ACS Photonics 4(12), 3055–3062 (2017). [CrossRef]  

21. H. Lu, Y. Li, Z. Yue, D. Mao, and J. Zhao, “Topological insulator based Tamm plasmon polaritons,” APL Photonics 4(4), 040801 (2019). [CrossRef]  

22. S. Chen, L. Miao, X. Chen, Y. Chen, C. Zhao, S. Datta, Y. Li, Q. Bao, H. Zhang, Y. Liu, S. Wen, and D. Fan, “Few-layer topological insulator for all-optical signal processing using the nonlinear Kerr effect,” Adv. Opt. Mater. 3(12), 1769–1778 (2015). [CrossRef]  

23. C. Zhao, Y. Zou, Y. Chen, Z. Wang, S. Lu, H. Zhang, S. Wen, and D. Tang, “Wavelength-tunable picosecond soliton fiber laser with topological insulator: Bi2Se3 as a mode locker,” Opt. Express 20(25), 27888–27895 (2012). [CrossRef]  

24. Z. Yue, B. Cai, L. Wang, X. Wang, and M. Gu, “Intrinsically core-shell plasmonic dielectric nanostructures with ultrahigh refractive index,” Sci. Adv. 2(3), e1501536 (2016). [CrossRef]  

25. M. Zhao, J. Zhang, N. Gao, P. Song, M. Bosman, B. Peng, B. Sun, C. Qiu, Q. Xu, Q. Bao, and K. Loh, “Actively tunable visible surface plasmons in Bi2Te3 and their energy-harvesting applications,” Adv. Mater. 28(16), 3138–3144 (2016). [CrossRef]  

26. Z. Yue, H. Ren, S. Wei, J. Lin, and M. Gu, “Angular-momentum nanometrology in an ultrathin plasmonic topological insulator film,” Nat. Commun. 9(1), 4413 (2018). [CrossRef]  

27. H. Lu, S. Dai, Z. Yue, Y. Fan, H. Cheng, J. Di, D. Mao, E. Li, T. Mei, and J. Zhao, “Sb2Te3 topological insulator: surface plasmon resonance and application in refractive index monitoring,” Nanoscale 11(11), 4759–4766 (2019). [CrossRef]  

28. P. Chen and A. Alù, “Atomically thin surface cloak using graphene monolayers,” ACS Nano 5(7), 5855–5863 (2011). [CrossRef]  

29. W. Whitney, V. Brar, Y. Ou, Y. Shao, A. Davoyan, D. Basov, K. He, Q. Xue, and H. Atwater, “Gate-variable mid-infrared optical transitions in a (Bi1-xSbx)2Te3 topological insulator,” Nano Lett. 17(1), 255–260 (2017). [CrossRef]  

30. H. Lin, B. Sturmberg, K. Lin, Y. Yang, X. Zheng, T. Chong, C. Sterke, and B. Jia, “A 90-nm-thick graphene metamaterial for strong and extremely broadband absorption of unpolarized light,” Nat. Photonics 13(4), 270–276 (2019). [CrossRef]  

31. K. Bolotina, K. Sikes, Z. Jianga, M. Klima, G. Fudenberg, J. Hone, P. Kim, and H. Stormer, “Ultrahigh electron mobility in suspended graphene,” Solid State Commun. 146(9-10), 351–355 (2008). [CrossRef]  

32. J. Chen, C. Jang, S. Xiao, M. Ishigami, and M. Fuhrer, “Intrinsic and extrinsic performance limits of graphene devices on SiO2,” Nat. Nanotechnol. 3(4), 206–209 (2008). [CrossRef]  

33. N. Mishra, S. Forti, F. Fabbri, L. Martini, C. McAleese, B. Conran, P. Whelan, A. Shivayogimath, B. Jessen, L. Buß, J. Falta, I. Aliaj, S. Roddaro, J. I. Flege, P. Bøggild, K. Teo, and C. Coletti, “Wafer-scale synthesis of graphene on sapphire: toward fab-compatible graphene,” Small 15(50), 1904906 (2019). [CrossRef]  

34. H. Lu, X. Gan, D. Mao, and J. Zhao, “Graphene-supported manipulation of surface plasmon polaritons in metallic nanowaveguides,” Photonics Res. 5(3), 162–167 (2017). [CrossRef]  

35. A. Dubrovkin, G. Adamo, J. Yin, L. Wang, C. Soci, Q. Wang, and N. Zheludev, “Visible range plasmonic modes on topological insulator nanostructures,” Adv. Opt. Mater. 5(3), 1600768 (2017). [CrossRef]  

36. G. Jellison and F. Modine, “Parameterization of the optical functions of amorphous materials in the interband region,” Appl. Phys. Lett. 69(3), 371–373 (1996). [CrossRef]  

37. M. Kaliteevski, I. Iorsh, S. Brand, R. Abram, J. Chamberlain, A. Kavokin, and I. Shelykh, “Tamm plasmon-polaritons: Possible electromagnetic states at the interface of a metal and a dielectric Bragg mirror,” Phys. Rev. B 76(16), 165415 (2007). [CrossRef]  

38. H. Lu, X. Gan, D. Mao, Y. Fan, D. Yang, and J. Zhao, “Nearly perfect absorption of light in monolayer molybdenum disulfide supported by multilayer structures,” Opt. Express 25(18), 21630–21636 (2017). [CrossRef]  

39. A. Taflove and S. Hagness, Computational Electrodynamics: The Finite-Difference Time-Domain Method (Artech House, 2005).

40. O. Gazzano, S. Vasconcellos, K. Gauthron, C. Symonds, J. Bloch, P. Voisin, J. Bellessa, A. Lemaître, and P. Senellart, “Evidence for confined Tamm plasmon modes under metallic microdisks and application to the control of spontaneous optical emission,” Phys. Rev. Lett. 107(24), 247402 (2011). [CrossRef]  

41. N. Liu, L. Langguth, T. Weiss, J. Kästel, M. Fleischhauer, T. Pfau, and H. Giessen, “Plasmonic analogue of electromagnetically induced transparency at the Drude damping limit,” Nat. Mater. 8(9), 758–762 (2009). [CrossRef]  

42. B. Peng, Ş. K. Özdemir, W. Chen, F. Nori, and L. Yang, “What is and what is not electromagnetically induced transparency in whispering-gallery microcavities,” Nat. Commun. 5(1), 5082 (2014). [CrossRef]  

43. H. Cheng, S. Chen, P. Yu, X. Duan, B. Xie, and J. Tian, “Dynamically tunable plasmonically induced transparency in periodically patterned graphene nanostrips,” Appl. Phys. Lett. 103(20), 203112 (2013). [CrossRef]  

44. H. Lu, X. Liu, Y. Gong, D. Mao, and L. Wang, “Optical bistability in metal-insulator-metal plasmonic Bragg waveguides with Kerr nonlinear defects,” Appl. Opt. 50(10), 1307–1311 (2011). [CrossRef]  

45. C. Wu and Z. Wang, “Properties of defect modes in one-dimensional photonic crystals,” Prog. Electromagn. Res. 103, 169–184 (2010). [CrossRef]  

46. L. Falkovsky, “Optical properties of graphene,” J. Phys.: Conf. Ser. 129, 012004 (2008). [CrossRef]  

47. R. Taubert, M. Hentschel, J. Kastel, and H. Giessen, “Classical analog of electromagnetically induced absorption in plasmonics,” Nano Lett. 12(3), 1367–1371 (2012). [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 (6)

Fig. 1.
Fig. 1. 3D diagram of the multilayer photonic system consisting of a thin BSTS topological insulator film, a Bragg mirror with stacked Si3N4/SiO2 layers and a defect layer (containing Al2O3, graphene, and doped Si) in the middle of Bragg mirror.
Fig. 2.
Fig. 2. (a) Absorption spectrum of the multilayer structure with dc=0 nm. (b) Corresponding electric field distribution and intensity profile in the structure at the absorption peak. (c) Absorption spectrum of the multilayer structure with an Al2O3 defect layer (dc=154.5 nm and dg=ds=0 nm) in the Bragg mirror. The inset shows the prototype three-level model of the EIT effect. (d) Corresponding electric field distribution and intensity profile at the induced absorption wavelength (λ=1309.9 nm). Here, dt=58 nm, da=150 nm, db=240 nm, θ=0°, and N = 20. The circles and curves in (a) and (c) denote the FDTD and TMM results, respectively.
Fig. 3.
Fig. 3. Evolution of absorption spectrum from the multilayer structure with (a) the thickness of the BSTS film dt when da=150 nm, dc=154.5 nm, and θ=0°, (b) the thickness of Si3N4 layer da when dt=58 nm, dc=154.5 nm, and θ=0°, (c) the thickness of Al2O3 defect layer dc when dt=58 nm, da=150 nm, and θ=0°, and (d) the light incident angle θ when dt=58 nm, da=150 nm, and dc=154.5 nm. Here, db=240 nm and N = 20. The circles denote the induced transparency wavelengths obtained by numerical simulations.
Fig. 4.
Fig. 4. (a)-(d) Electric field amplitudes of EIT resonances (defect modes) versus dt, da, dc, and θ in the multilayer structure with the same parameters as in Figs. 3(a)–3(d).
Fig. 5.
Fig. 5. (a) Evolution of absorption spectrum from the multilayer structure with the gate voltage Vb on graphene. (b) Electric field intensity profile in the structure when Vb=0 V and 6 V. The insets depict the intensity profiles in the BSTS film when Vb=0 V and 6 V. (c)-(d) Absorption spectra of the BSTS film and graphene when Vb=0 V and 6 V. The parameters are set as da=150 nm, db=240 nm, du=50 nm, di=5 nm, ds=60 nm, θ=0°, and N = 20.
Fig. 6.
Fig. 6. (a) TMM calculated (dots) and fitting (curves) results of light absorption spectrum from the multilayer structures with different gate voltages Vb on graphene. (b) Fitting values of resonance frequency detuning δ between the oscillators 1 and 2 with different Vb. (c) Fitting values of decay rate γ2 in the oscillator 2 with different Vb. (d) Fitting values of coupling strength κ between the oscillators 1 and 2 with different Vb. Here, da=150 nm, db=240 nm, du=50 nm, di=5 nm, ds=60 nm, θ=0°, and N = 20.

Equations (8)

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

σ i n t r a = i e 2 k B T π 2 ( ω + i τ 1 ) [ μ c k B T + 2 ln ( exp ( μ c k B T ) + 1 ) ] ,
σ i n t e r = i e 2 4 π ln [ 2 | μ c | ( ω + i τ 1 ) 2 | μ c | + ( ω + i τ 1 ) ] .
ε b ( E ) = { A E 0 C ( E E g ) 2 C 2 E 2 + ( E 2 E 0 2 ) 2 1 E , E > E g 0 E E g
ε b ( E ) = ε b ( ) + 2 π P E g ξ ε b ( ξ ) ξ 2 E 2 d ξ ,
[ E u p E u p + ] = M 1 P 1 M 2 P 2 M 2 N + 8 [ E d o w n E d o w n + ] = ( Q 11 Q 12 Q 21 Q 22 ) [ E d o w n E d o w n + ] ,
M j = 1 2 n j 1 cos θ j 1 ( n j 1 cos θ j + n j cos θ j 1 n j 1 cos θ j n j cos θ j 1 n j 1 cos θ j n j cos θ j 1 n j 1 cos θ j + n j cos θ j 1 ) ,
P j = ( exp ( i 2 π d j n j cos θ j / λ ) 0 0 exp ( i 2 π d j n j cos θ j / λ ) ) ,
A  =  Im { f ( ω ω 0 + δ + i γ 2 / 2 ) ( κ e i φ ) 2 ( ω ω 0 + δ + i γ 2 / 2 ) ( ω ω 0 + i γ 1 / 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.