Open Access
3 April 2024 λ/20-Thick cavity for mimicking electromagnetically induced transparency at telecommunication wavelengths
Hua Lu, Shouhao Shi, Dikun Li, Shuwen Bo, Jianxu Zhao, Dong Mao, Jianlin Zhao
Author Affiliations +
Abstract

Optical cavities play crucial roles in enhanced light–matter interaction, light control, and optical communications, but their dimensions are limited by the material property and operating wavelength. Ultrathin planar cavities are urgently in demand for large-area and integrated optical devices. However, extremely reducing the planar cavity dimension is a critical challenge, especially at telecommunication wavelengths. Herein, we demonstrate a type of ultrathin cavities based on large-area grown Bi2Te3 topological insulator (TI) nanofilms, which present distinct optical resonance in the near-infrared region. The result shows that the Bi2Te3 TI material presents ultrahigh refractive indices of >6 at telecommunication wavelengths. The cavity thickness can approach 1/20 of the resonance wavelength, superior to those of planar cavities based on conventional Si and Ge high refractive index materials. Moreover, we observed an analog of the electromagnetically induced transparency (EIT) effect at telecommunication wavelengths by depositing the cavity on a photonic crystal. The EIT-like behavior is derived from the destructive interference coupling between the nanocavity resonance and Tamm plasmons. The spectral response depends on the nanocavity thickness, whose adjustment enables the generation of obvious Fano resonance. The experiments agree well with the simulations. This work will open a new door for ultrathin cavities and applications of TI materials in light control and devices.

1.

Introduction

Optical cavities with excellent resonance and light confinement activities play crucial roles in light generation, modulation, detection, sensing, and storage for the realization of optical functional devices, integrations, and communications.15 The dimensions of optical cavities are generally limited by the material characteristics and operating wavelength. Metallic cavities enable light control and functionalities at visible and near-infrared ranges with the excitation of localized surface plasmon resonance on hundred-nanometer structures.6,7 Silicon (Si) dielectric particles with several hundred nanometers can work as nanoresonators for functional devices at optical wavelengths.8,9 The complex geometries of metallic and dielectric cavities strongly depend on the precision fabrication process, limiting the fabrication of large-area devices. Ultrathin planar cavities play important roles in the large-area and integrated functional devices.10,11 High refractive index (HRI) materials attract considerable attention for achieving ultrathin nanocavities with strong light confinement and optical interference at the nanoscale.12,13 As conventional HRI materials, Si and germanium (Ge) semiconductors with refractive indices of 3.47 and 4.27 at 1550-nm wavelength have advanced the rapid developments of optical signal processing and communications.14,15 Ge material enables the realization of optical cavities with several hundred nanometers for photodetection at the telecommunication band.16 The materials with higher refractive indices are in great demand for acquiring extremely thin cavities operating at telecommunication wavelengths. Different from traditional HRI semiconductors, topological insulators (TIs) as newly emerging quantum materials exhibit topologically protected conducting surface (or edge) and semiconducting bulk states.1720 The state of matter with topological property was first reported in two-dimensional HgTe quantum wells.17 Subsequently, many three-dimensional (3D) materials with topological surface states were reported, such as Bi2Te3, Sb2Te3, Bi2Se3, and their compounds.19,20 Recently, topological materials have attracted broad attention in the fields of electronics, optics, and condensed matter physics because of their fantastic electronic and optical properties.1824 Especially, 3D TI materials generally present the characteristic of HRI exceeding the refractive indices of Si and Ge at infrared wavelengths.2123 TI materials will offer a promising prospect for achieving ultrathin optical cavities and their applications in novel optical effects and functional devices.

Electromagnetically induced transparency (EIT) is a counterintuitive quantum interference effect occurring in atomic systems.25 The EIT effect acts as an essential role in modern optical physics due to its significant applications in optical data storage, optical switching, and enhanced light nonlinearities.26 Even so, the practical applications of atomic EIT are limited by the rigorous requirements of low temperature or stable gas lasing conditions.27 Fortunately, the analogies to atomic EIT were observed in conventional optical cavities, such as photonic crystal cavities,28 whispering-gallery-mode cavities,29,30 and metallic cavities.3133 The EIT-like effects in coupled optical cavities can be used to realize advanced ultracompact devices, for instance, sensitive sensors,32 3D rulers,33 and optical switching.34 Mimicking EIT in TI material-based cavities will not only enrich the formation mechanism of EIT-like effects but also advance the novel optical applications of TI materials, which, however, have not been demonstrated until now.

Herein, we reported a type of ultrathin planar cavities based on the large-area grown Bi2Te3 TI nanofilms, which possess obvious optical resonance in the near-infrared region. The experimental results demonstrate that the Bi2Te3 TI material presents ultrahigh refractive index at near-infrared wavelengths, contributing to the realization of optical cavities that are tens of nanometers in thickness. The TI material-based nanocavity thickness is 1/20 of resonance wavelength (λ/20), much thinner than those of cavities based on the conventional HRI Si and Ge materials. We also experimentally observed an analogy to atomic EIT at telecommunication wavelengths by depositing the nanocavity onto a one-dimensional (1D) photonic crystal. A transparency window distinctly appears in the original cavity-induced resonance absorption spectrum owing to the coupling between the nanocavity resonance and Tamm plasmons. The spectral profile relies on the cavity thickness and can evolve to an asymmetric Fano resonance line shape. The experimental results agree well with the simulations. This work will pave a new pathway for ultrathin optical cavities as well as the applications of TI materials in the mimicking of EIT and optoelectronic devices.

2.

Results

Figure 1(a) shows the 3D diagram of the cavity with a TI material nanofilm coating on a metallic layer. To explore the optical properties, we fabricated the Bi2Te3 TI nanocavity by using a magnetron sputtering (MS) technique (see Appendix). Figure 1(b) depicts an angle-tilted cross-sectional scanning electron microscopic (SEM) image of a nanocavity with the Bi2Te3 and silver nanofilms deposited on a SiO2 substrate. d and t denote the thicknesses of Bi2Te3 and silver films, respectively. A 2-nm SiO2 layer was deposited to isolate the silver and Bi2Te3 films for avoiding the formation of the Schottky barrier and charge transport.35 The optical and top-view SEM images reveal the large area and good flatness properties of MS-grown Bi2Te3 nanofilms, respectively (see Fig. S1 in the Supplementary Material). Compared with the pulsed laser deposition, the MS technique enables the large-area growth of relatively flat Bi2Te3 films.36 The MS method also presents the better flatness and higher growth rate for fabricating Bi2Te3 films when compared with the atomic layer deposition.37 Figure 1(c) depicts the absorption spectrum of a 42-nm Bi2Te3 TI film on a 30-nm silver layer (i.e., d=42  nm and t=30  nm). The Bi2Te3 TI thickness can be clarified using the atomic force microscope (AFM) measurement (see Fig. S2 in the Supplementary Material). We measured the reflection, transmission, and absorption spectra of a TI nanocavity using a spectrophotometer (see Appendix). In Fig. S3 in the Supplementary Material, a distinct reflection dip appears at 1465-nm wavelength (i.e., λ=1465  nm) for the cavity with a thickness of λ/20, which is derived from the Fabry–Perot resonance in the asymmetric nanocavity. Here, the flat silver film with a large area can reflect a part of the incident light. The silver-reflected light will generate destructive interference with directly reflected light from the Bi2Te3 TI layer around the resonance wavelength.13 Thus, the reflection light of the cavity can be effectively suppressed. The cavity thickness of λ/20 is ultrathin, according to the reported cavities.10,38,39 The optical loss of Bi2Te3 TI gives rise to strong resonance absorption at a broad wavelength range covering 1100 to 1800 nm, as depicted in Fig. 1(c). Moreover, we proposed to integrate the TI nanocavity onto a 1D photonic crystal, as can be seen in Fig. 1(d). The Bi2Te3 TI nanocavity was deposited on a photonic crystal with alternately stacked Ta2O5 and SiO2 layers. The Ta2O5 and SiO2 layers were fabricated using electron beam deposition. Figure 1(e) shows the angle-tilted cross-sectional SEM image of a fabricated TI nanocavity/photonic crystal structure. We can see that the photonic crystal has good uniformity with the periodic number N=16. The thicknesses of Ta2O5 and SiO2 layers are 182 and 260 nm, respectively (i.e., dt=182  nm and ds=260  nm). Figure 1(f) depicts the absorption spectrum of TI nanocavity/photonic crystal structure with d=42  nm and t=30  nm. A distinct narrow dip appears around 1550 nm wavelength in the absorption spectrum. Thus, the original broad absorption of TI nanocavity exhibits a transparency window at the telecommunication wavelength, which is analogous to the EIT effect in atomic systems.28,29 Meanwhile, we characterized the grown Bi2Te3 nanofilm using a transmission electron microscope (TEM). Figure 2(a) shows the TEM image of a grown 42-nm Bi2Te3 TI film transferred on the supporter of a copper microgrid (see Appendix and Note 1 in the Supplementary Material). The chemical composition of Bi2Te3 nanofilm can be confirmed using the energy-dispersive X-ray (EDX) measurement, which reveals the atomic ratio of Bi:Te2:3 (see Fig. S4 in the Supplementary Material). The good quality of Bi2Te3 can also be verified by the Raman shift spectrum (see Fig. S5 in the Supplementary Material).40 Figure 2(b) depicts the selected area electron diffraction (SAED) pattern of Bi2Te3 nanofilm, which corresponds to the rhombohedral Bi2Te3 phase.41 It also denotes that the MS-grown Bi2Te3 nanofilm presents the preferred growth orientation of (015). The high-resolution TEM (HRTEM) image in Fig. 2(c) illustrates multi-orientation atomic lattice arrangements in the Bi2Te3 nanofilm. The HRTEM image exhibits distinct fringes with some lattice spacing containing 0.321 and 0.220 nm, corresponding to the (015) and (110) planes of Bi2Te3, respectively. The continuous rings of sharp diffraction pattern and atomic lattice arrangements reveal the polycrystalline nature of the Bi2Te3 nanofilm. We measured the optical constant of Bi2Te3 TI using a spectroscopic ellipsometer (see Appendix and Note 2 in the Supplementary Material). To differentiate the surface and bulk states, the ellipsometer-measured results can be fitted with the layer-on-bulk model.42,43 The surface and bulk were fitted with the Drude and Tauc–Lorentz dispersion relations, respectively.4244 From Fig. 2(d), we can see that the fitted complex refractive indices agree well with the measurement. Figures 2(e) and 2(f) depict the complex refractive indices of surface and bulk for the Bi2Te3 TI. For the surface, the extinction coefficient k is larger than the refractive index n at the wavelengths from 230 to 1930 nm, exhibiting metal-like characteristic. For the bulk, the refractive index can exceed 6.0 at the wavelengths of >1385  nm, which is much higher than the refractive indices of Si and Ge. The surface thickness was fitted as 2.29  nm, approaching the estimated 2.5 nm for 3D TI.45 The bandgap energy is 0.21  eV, which is close to the reported bandgap (see Table S1 in the Supplementary Material).46 The HRI property of our Bi2Te3 accords with the reported result of single-crystal Bi2Te3 film, even though there is a certain difference in the optical constants.47 The ultrahigh refractive indices contribute to the ultrathin TI nanocavities at near-infrared range, similar to the Ge nanocavities at visible range.13 To our knowledge, this is the first report of a planar TI cavity with tens of nanometers in thickness at telecommunication wavelengths.

Fig. 1

TI structures and absorption spectra. (a) 3D diagram of the ultrathin nanocavity with a TI nanofilm coated on a metallic layer. (b) Angle-tilted cross-sectional SEM image of a Bi2Te3 TI nanocavity deposited on a SiO2 substrate. The scale bar is 200 nm. Here, d and t are the thicknesses of Bi2Te3 and silver nanofilms, respectively. (c) Experimental measurement of the absorption spectrum of the nanocavity with a 42-nm Bi2Te3 film on a 30-nm silver layer (i.e., d=42  nm and t=30  nm). (d) 3D diagram of a TI nanocavity on a 1D photonic crystal. (e) Cross-sectional SEM image of a Bi2Te3 nanocavity on a photonic crystal with alternately stacked Ta2O5 and SiO2 layers. Here, dt and ds are the thicknesses of Ta2O5 and SiO2 layers, respectively. N is the periodic number. The scale bar is 2  μm. The inset shows the high-resolution SEM image of Bi2Te3 and silver films on the photonic crystal. The scale bar is 200 nm. (f) Measured absorption spectrum of the TI nanocavity/photonic crystal structure with dt=182  nm, ds=260  nm, d=42  nm, t=30  nm, and N=16. The inset depicts a three-level system of the EIT-like effect. ωc stands for the nanocavity resonant frequency. δ is the frequency detuning between the nanocavity resonance and Tamm plasmons in silver/photonic crystal structure. γ1 and γ2 are the decaying rates of nanocavity resonance and Tamm plasmons, respectively. κeiφ is the complex coupling coefficient between the nanocavity resonance and Tamm plasmons with a phase retardation φ.

AP_6_3_036001_f001.png

Fig. 2

Material characterization and optical constants of Bi2Te3 TI. (a) TEM image of a grown 42-nm Bi2Te3 TI film transferred on the supporter of a copper microgrid. (b) SAED pattern of the Bi2Te3 nanofilm. (c) HRTEM image of the Bi2Te3 nanofilm. (d) Ellipsometer-measured (circles) and fitted (curves) refractive indices and extinction coefficients of Bi2Te3 TI at the wavelengths from 230 to 1930 nm. (e), (f) Fitted refractive indices and extinction coefficients of Bi2Te3 TI surface and bulk states with the layer-on-bulk model [the inset in panel (d)].

AP_6_3_036001_f002.png

Subsequently, we measured the absorption spectra of TI nanocavities with different Bi2Te3 thicknesses. As shown in Fig. 3(a), the resonant absorption wavelength presents a redshift in the near-infrared region as d increases from 30 to 52 nm. Simultaneously, the absorption peak possesses an obvious rise. To verify the experimental results, the absorption spectra were theoretically calculated using the transfer matrix method (see Fig. S6 and Note 3 in the Supplementary Material).48 In the calculations, we employed the measured complex refractive indices of the Bi2Te3 TI surface and bulk in Figs. 2(e) and 2(f) as well as the measured optical constants of silver film in Fig. S7 in the Supplementary Material. As depicted in Fig. 3(b), there exists a nearly linear relation between the absorption peak wavelength and Bi2Te3 TI thickness. As mentioned above, the TI nanocavity thickness is 1/20 of the resonance wavelength. Moreover, we simulated the spectral response and field distributions of the TI nanocavities using the finite-difference time-domain (FDTD) method (see Appendix). We can see that the experimental measurements agree well with the theoretical and simulation results. Figure 3(c) depicts the intensity distribution of the magnetic field in the TI nanocavity at the absorption peak wavelength. It shows that the magnetic field is confined and distinctly enhanced in the nanocavity. As shown in Fig. 2(d), there are large extinction coefficients for the Bi2Te3 TI at near-infrared wavelengths, which means that the TI nanocavity possesses high optical loss. Thus, the Fabry–Perot resonance in the asymmetric cavity can give rise to relatively strong light absorption in the near-infrared region.13 The absorption ratio between the Bi2Te3 TI and silver nanofilms is 13:1, revealing that the light absorption mainly stems from the Bi2Te3 TI nanofilm [see Fig. S8(a) in the Supplementary Material]. From Fig. S8(b) in the Supplementary Material, we can see that the TI bulk dominates in the light absorption of Bi2Te3 nanofilm, and the surface layers induce a slight redshift of the resonance absorption peak. Moreover, we deposited a 30-nm silver film on the photonic crystal and measured the spectral response (see Fig. S9 in the Supplementary Material). The result demonstrates that there exists a narrow reflection dip and an absorption peak at the wavelength of 1560  nm, which results from the formation of Tamm plasmons in the silver/photonic crystal multilayer. Both the Bi2Te3/silver and silver/photonic crystal structures present relatively strong light absorption at the wavelengths from 1545 to 1580 nm. Intuitively, the TI nanocavity/photonic crystal structure would present stronger absorption at these wavelengths, but a distinct transparency window appears around 1550 nm in the absorption spectrum, as shown in Fig. 1(f). The generation of reflection dips in the silver/photonic crystal structure is similar to the Tamm plasmon response in previous structures.4951 But, the appearance of absorption dips in the Bi2Te3/silver/photonic crystal structure is fundamentally different from the spectral response (reflection dips) in the Tamm plasmon structures, even with a poly(vinyl alcohol) layer on the metal.50,51 As mentioned above, this counterintuitive phenomenon can be regarded as an analogy to the EIT effect in atomic systems.31,32,52 It can be analyzed with a prototype three-level model, as depicted in the inset of Fig. 1(f). In the model, |1> and |2> stand for the upper states, and |0> denotes the ground state. The light impinges on the Bi2Te3 nanofilm and excites the resonance in the TI nanocavity, which is analogous to the transition from |0> to |1>.31,53 Tamm plasmons can be generated by the evanescent field of nanocavity resonance, which will couple with each other. The coupling between the nanocavity resonance and Tamm plasmons is analogous to the transition between the states |1> and |2>. The two possible transition processes |0>|1>|2>|1> and |0>|1> will form the destructive interference. The interference coupling contributes to the appearance of a narrow transparency window in the broad absorption peak. Therefore, the EIT-like effect is generated due to the coupling between the Fabry–Perot resonance in the planar TI cavity and Tamm plasmons in the silver/photonic crystal structure. This is different from the generation mechanism of EIT-like transmission in previous plasmonic structures, e.g., the dielectric waveguide with silver nanoparticles.54 In Fig. 1(f), we can see some small grooves in the absorption spectrum around the transparency window. The small grooves correspond to the transmission peaks of the multilayer structure, which originate from the constructive interference at the output interface, not the EIT-like effect. We measured the spectral response of TI nanocavity/photonic crystal structures with different Bi2Te3 thicknesses d when t=30  nm. From Fig. 3(d), we can see that the EIT-like spectral profile is dependent on d. The EIT-like spectrum distinctly appears as d ranges from 42 to 52 nm, which denotes that the EIT-like behavior is robust to the small fabrication deviation of Bi2Te3 thickness. However, an obviously asymmetric Fano-type line shape can be generated when d=30  nm. The formation of Fano resonance can be derived from the coherent interference between a quasi-continuum state (broadly spectral nanocavity off-resonance) and a discrete excited state (narrowly spectral Tamm plasmons). The numerical simulations are in excellent agreement with the experimental results, as depicted in Fig. 3(e). Figure 3(f) shows the magnetic field distribution of the TI nanocavity/photonic crystal structure with a 42-nm Bi2Te3 film at the absorption dip wavelength (i.e., 1550 nm). It is found that Tamm plasmons between the silver film and photonic crystal can be excited with a distinct field enhancement. It verifies the nanocavity coupling-induced excitation of Tamm plasmons. Interestingly, the coupling-induced Tamm plasmons can effectively reduce the strong light absorption in traditional Tamm plasmon systems, which will be beneficial for reinforcing light–matter interactions. Finally, we studied the dependence of the EIT-like spectrum on the silver film thickness t in the TI nanocavity/photonic crystal structure, as shown in Fig. 4(a). The theoretical calculations illustrate that the induced transparency window becomes narrower and shallower with the increase of t. To verify it, we fabricated the TI nanocavity/photonic crystal structures with t=15, 30, 40, and 50 nm when d=44  nm and measured their absorption spectra.

Fig. 3

TI thickness-dependent nanocavity resonance and induced transparency. (a) Experimentally measured absorption spectra of the TI nanocavities on the SiO2 substrate with different Bi2Te3 film thicknesses d when t=30  nm. (b) Corresponding resonance wavelengths of the TI nanocavity with different d. The curve, red circles, and blue circles denote the theoretical, experimental, and simulation results, respectively. (c) Corresponding magnetic field distribution of the TI nanocavity with a 42-nm Bi2Te3 film (i.e., d=42  nm) at the absorption peak wavelength. (d), (e) Measured and simulated absorption spectra of the TI nanocavity/photonic crystal structures with different d when t=30  nm. (f) Corresponding magnetic field distribution of the TI nanocavity/photonic crystal structure with a 42-nm Bi2Te3 film at the absorption dip wavelength.

AP_6_3_036001_f003.png

Fig. 4

EIT-like response with different silver film thicknesses. (a) Theoretical evolution of absorption spectrum from the TI nanocavity/photonic crystal structures with different silver film thicknesses t when d=44  nm. (b) Measured (solid curves) and fitted (dashed curves) absorption spectra of the TI nanocavity/photonic crystal structures with different t when d=44  nm. Here, the experimentally measured spectra were fitted with the TCO model. (c) Fitted decaying rates γ2 and coupling strengths κ in the TCO model with different t. The inset shows the corresponding decaying rates γ1.

AP_6_3_036001_f004.png

As depicted in Fig. 4(b), the experimental results verify the theoretical prediction of spectral evolution. To clarify the mechanism of the EIT-like effect, we theoretically fitted the measured absorption spectra with the two-coupled-oscillator (TCO) model (see Appendix). The results show that the decaying rate γ1 of oscillator 1 (TI nanocavity) slightly changes as the Bi2Te3 TI thickness varies from 15 to 50 nm. γ1 is 1.02×1015  rad/s when t=30  nm. The decaying rate γ2 of oscillator 2 (silver/photonic crystal structure) slowly increases with t and can approach 0.85×1013  rad/s when t=30  nm. The coupling strength κ between oscillators 1 and 2 rapidly decreases as t increases. κ reaches 5.5×1013  rad/s when t=30  nm, which is much less than the threshold coupling strength κT [κT=(γ1γ2)/4=5.06×1014  rad/s]. Therefore, the coupling between the nanocavity resonance and Tamm plasmons is located in the weak-driving regime when t=30  nm.29 The coupling phase retardation φ approaches π in the TI nanocavity/photonic crystal structures with different t (see Fig. S10 in the Supplementary Material), verifying the destructive interference between the nanocavity resonance and Tamm plasmons. Because the coupling strength satisfies the criterion κ<κT, the observed transparency windows in our systems are attributed to the EIT-like effect with the existence of large γ1 and small γ2 even when t declines to 15 nm, as depicted in Fig. 4(c). Thus, the EIT-like response can be achieved in the TI nanocavity/photonic crystal structure with a Bi2Te3 film thickness of λ/25. The absorption dip can approach a low value of 6% at the 1550-nm wavelength when t=15  nm (see Fig. S11 in the Supplementary Material). It means that the EIT-like effect can effectively bypass the optical loss in the structure with a small silver thickness.

3.

Discussion

The Bi2Te3 TI possesses an ultrahigh refractive index at near-infrared wavelengths, much exceeding the conventional HRI Si and Ge semiconductors [see Fig. 2(d)]. The ultrahigh refractive index of Bi2Te3 TI enables the realization of planar cavities with tens of nanometers in thickness, which can find promising applications in light–matter interaction, for instance, infrared optoelectronic conversion. From Fig. 1(c) and Fig. S3 in the Supplementary Material, we can see that the TI cavity with a 42-nm Bi2Te3 film presents an ultrabroad resonance absorption of light with high efficiency of >50% at the wavelengths from 970 to 1850 nm. The nanocavities can effectively break the absorption limit of 50% for ultrathin free-standing film at near-infrared range.55 As shown in Fig. S8 in the Supplementary Material, the light absorption mainly stems from the Bi2Te3 TI nanofilm with a narrow-bandgap semiconducting property. The absorption ratio between the TI and silver nanofilms is 13:1 around the resonance wavelength. The Bi2Te3 TI cavities will contribute to the high-efficiency, broad-spectrum, ultrathin, and large-area near-infrared optoelectronic devices, e.g., telecommunication band photodetectors.

From Fig. 3(b), we can see that the resonance wavelength presents a linear redshift with the increasing Bi2Te3 TI thickness. The resonance wavelength is 20 times as large as the TI nanocavity thickness (0.05λ), which is more remarkable than the subwavelength-thin (0.13λ) layer of Ge nanostructures.55 This property also enables the precision monitoring of nanofilm thickness in the deposition process.13 To achieve this kind of optical cavities, the thicknesses of Bi2Te3 TI are much thinner than those of HRI Si and Ge semiconductors (see Fig. S12 in the Supplementary Material). For example, the Bi2Te3 TI thickness is 53% (64%) of Si (Ge) thickness for optical resonance at the 1550-nm wavelength. Thus, the thickness of the Bi2Te3 TI cavity can be improved by 34% (24.5%) compared with that of the Si (Ge) cavity. Moreover, the nanocavities may find applications in ultrathin structures for wavefront shaping, similar to Si metasurfaces.8

The TI nanocavities will offer new building blocks for realizing novel optical effects. Herein, we have innovatively integrated the TI nanocavity onto a photonic crystal with alternately stacked Ta2O5 and SiO2 layers. Interestingly, an analogy to the atomic EIT effect has been observed at telecommunication wavelengths. This EIT-like behavior is attributed to the destructive interference coupling between the resonance in TI nanocavity and Tamm plasmons in silver/photonic crystal structure. Compared with the conventional materials, the Bi2Te3 TI with relatively high refractive index and extinction coefficient has the specific advantage for the generation of the EIT-like effect in this system (see Fig. S13 in the Supplementary Material). Besides the nanocavity thicknesses, the EIT-like response is also dependent on the incident light angle. From Fig. S14 in the Supplementary Material, we can see that the induced transparency wavelength exhibits a blueshift with the increase of incident light angle. This stems from the identical shift of the Tamm plasmon wavelength by altering the incident light angle. The shifted wavelength of s-polarized light is different from that of p-polarized light, which stems from the energy splitting between the s- and p-polarized Tamm states.49 The adjustment of the incident light angle offers an effective way for dynamic tunability of the induced transparency window. The induced transparency spectrum can also be tuned by controlling the polarization of light at oblique incidence, as shown in Fig. S14(f) in the Supplementary Material. This kind of cavity structures based on planar multilayers can effectively avoid the fabrication process and area limitations of the structures based on dielectric or metallic nanoparticles for the generation of EIT-like effects. The EIT-like effect presents good robustness to the alteration of the environmental refractive index (see Fig. S15 in the Supplementary Material). The TI nanocavity also enables the formation of EIT-like behavior by integrating with the thin Ge and silver layers (see Fig. S16 in the Supplementary Material). The thickness can be effectively improved when compared with the Ge-based cavity. This work will open a new door for ultrathin optical cavities and promote the applications of TIs in light control and optical functional devices.

4.

Appendix: Methods

4.1.

Fabrication

The Bi2Te3 TI and silver nanofilms were prepared by the MS deposition technique. The Bi2Te3 film was deposited using radio-frequency (RF) sputtering with a pressure of 0.9 Pa and power of 70 W. The silver film was deposited using the direct current sputtering with a pressure of 0.6 Pa and power of 50 W. The feasible grown area can be set as 4  cm×4  cm for the flat nanofilms in our MS deposition. The 2-nm SiO2 layer was deposited by employing the RF sputtering with a pressure of 1 Pa and power of 100 W. The flowing Ar (100 sccm) was employed as the sputtering gas. The 1D photonic crystal with alternately stacked Ta2O5 and SiO2 layers was prepared on the polished SiO2 substrate (Beijing Dream Material Technology Co., Ltd.) through electron beam deposition.

4.2.

Characterizations and Measurements

The surface flatness of grown Bi2Te3 TI nanofilm was characterized using the commercial SEM equipment (FEI Verios G4) and AFM system (Bruker Dimension FastScan). The AFM system was also used to test the thicknesses of deposited films. The material morphology of Bi2Te3 TI nanofilm was characterized through the SAED pattern and HRTEM image, which were measured using the TEM equipment (FEI Talos F200X) with a voltage of 200 kV. The Raman spectrum of Bi2Te3 TI was measured using a confocal microspectrometer (WITec Alpha300R) with a 532-nm laser. The complex refractive index of Bi2Te3 TI in the optical region was tested by employing a spectroscopic ellipsometer (HORIBA UVISEL Plus) with a 70-deg incident light angle. The absorption, reflection, and transmission spectra of fabricated samples were measured using a spectrophotometer (Hitachi UH5700).

4.3.

Simulations

To verify the experimental results, we employed the FDTD method to numerically simulate the optical spectra and field distributions of TI nanocavity and TI nanocavity/photonic crystal structures.56 In FDTD simulations, the bottom and top of structures were set as the perfectly matched layer absorbing boundary conditions. The other sides of computational space were set as the periodic boundary conditions. The non-uniform mesh was utilized in the y-axis direction to fully consider the influence of each layer. The mesh size of the Bi2Te3 TI surface layer was set as 0.3 nm. The maximum mesh size of other layers was 5 nm. The absorption spectra of the entire structures were achieved with A=1|Pr/Pi||Pt/Pi|, where Pr, Pt, and Pi denote the powers of reflected, transmitted, and incident light, respectively. The complex refractive indices of Bi2Te3 TI surface and bulk were set as the measured values in Figs. 2(e) and 2(f). The measured complex refractive index of MS-deposited silver film was used in the simulations (see Fig. S7 in the Supplementary Material). The refractive indices of SiO2 and Ta2O5 were set as 1.44 and 2.058 at the wavelengths of interest, respectively.57,58 The refractive indices of Si and Ge in the simulations come from the experimental data.59,60

4.4.

Theory

To clarify the mechanism of the EIT-like effect in the TI nanocavity/photonic crystal structures, we theoretically studied the EIT-like spectra with the TCO model.31,53 As shown in Fig. 1(c), the resonant mode in the TI nanocavity (i.e., oscillator 1) can be generated when the light is incident on the Bi2Te3 TI nanofilm. Tamm plasmons in silver/photonic crystal structure (i.e., oscillator 2) can be excited by the resonant coupling of TI nanocavity. As depicted in the inset of Fig. 1(f), ωc stands for the resonant frequency of oscillator 1. δ is the frequency detuning between oscillators 1 and 2. γ1 and γ2 are the decaying rates of oscillators 1 and 2, respectively. κeiφ is the complex coupling coefficient between the two oscillators with a coupling phase retardation φ. When |ωωc|ωc, γ2γ1ωc, and |δ|γ1, the light absorption in the entire structure can be described as

Eq. (1)

A(ω)=Im{f(ωωc+δ+iγ2/2)(κeiφ)2(ωωc+iγ1/2)(ωωc+δ+iγ2/2)},
where A(ω) is the imaginary part of the solution for the coupled differential equations of the two oscillators.31,53 f is an amplitude coefficient. According to Eq. (1), we can fit the experimentally measured absorption spectra to extract the related physical parameters. As shown in Fig. 4(b), the fitting curves are well consistent with the experimental results, verifying the feasibility of the theoretical model.

Disclosures

The authors declare no competing interests.

Code and Data Availability

All the data that support the findings of this study are available from the corresponding authors upon reasonable request.

Author Contributions

H.L. conceived the idea; carried out the theoretical calculations, numerical simulations, sample fabrications, and characterizations; analyzed results; drew the figures; and wrote the manuscript text. S.H.S. participated in the fabrications of Bi2Te3 and silver films. D.K.L. took part in the sample preparation for TEM measurements. S.W.B. and D.K.L. conducted the experimental measurement of optical spectra. J.X.Z. participated in the fitting of complex refractive index for Bi2Te3 TI. D.M. and J.L.Z. discussed the results and helped to promote the manuscript presentation. All authors substantially contributed to the manuscript.

Acknowledgments

This work was supported by the National Key R&D Program of China (Grant No. 2022YFA1404800), the National Natural Science Foundation of China (Grant Nos. 11974283, 61705186, and 11774290), the Natural Science Basic Research Plan in Shaanxi Province of China (Grant No. 2020JM-13), the “Double First-Class” Construction Fund Project (Grant No. 0206022GH0202), and the Fundamental Research Funds for the Central Universities (Grant No. D5000220175). The authors thank the Analytical & Testing Center of Northwestern Polytechnical University (NPU) for the TEM, AFM, SEM, Raman, and EDX measurements.

References

1. 

L. Yang et al., “Topological-cavity surface-emitting laser,” Nat. Photonics, 16 (4), 279 –283 https://doi.org/10.1038/s41566-022-00972-6 NPAHBY 1749-4885 (2022). Google Scholar

2. 

H. Wang et al., “Towards optimal single-photon sources from polarized microcavities,” Nat. Photonics, 13 (11), 770 –775 https://doi.org/10.1038/s41566-019-0494-3 NPAHBY 1749-4885 (2019). Google Scholar

3. 

J. Ni et al., “Multidimensional phase singularities in nanophotonics,” Science, 374 (6566), eabj0039 https://doi.org/10.1126/science.abj0039 SCIEAS 0036-8075 (2021). Google Scholar

4. 

X. Zhang et al., “Symmetry-breaking-induced nonlinear optics at a microcavity surface,” Nat. Photonics, 13 (1), 21 –24 https://doi.org/10.1038/s41566-018-0297-y NPAHBY 1749-4885 (2019). Google Scholar

5. 

T. Shi et al., “Planar chiral metasurfaces with maximal and tunable chiroptical response driven by bound states in the continuum,” Nat. Commun., 13 4111 https://doi.org/10.1038/s41467-022-31877-1 NCAOBW 2041-1723 (2022). Google Scholar

6. 

H. Altug et al., “Advances and applications of nanophotonic biosensors,” Nat. Nanotechnol., 17 (1), 5 –16 https://doi.org/10.1038/s41565-021-01045-5 NNAABX 1748-3387 (2022). Google Scholar

7. 

C. Chi et al., “Selectively steering photon spin angular momentum via electron-induced optical spin Hall effect,” Sci. Adv., 7 (18), eabf8011 https://doi.org/10.1126/sciadv.abf8011 STAMCV 1468-6996 (2021). Google Scholar

8. 

I. Staude and J. Schilling, “Metamaterial-inspired silicon nanophotonics,” Nat. Photonics, 11 (5), 274 –284 https://doi.org/10.1038/nphoton.2017.39 NPAHBY 1749-4885 (2017). Google Scholar

9. 

T. Liu et al., “High-efficiency optical frequency mixing in an all-dielectric metasurface enabled by multiple bound states in the continuum,” Phys. Rev. B, 107 075441 https://doi.org/10.1103/PhysRevB.107.075441 (2023). Google Scholar

10. 

A. Shaltout et al., “Ultrathin and multicolour optical cavities with embedded metasurfaces,” Nat. Commun., 9 2673 https://doi.org/10.1038/s41467-018-05034-6 NCAOBW 2041-1723 (2018). Google Scholar

11. 

J. Wu et al., “Graphene oxide for integrated photonics and flat optics,” Adv. Mater., 33 (3), 2006415 https://doi.org/10.1002/adma.202006415 ADVMEW 0935-9648 (2021). Google Scholar

12. 

M. L. Brongersma et al., “Light management for photovoltaics using high-index nanostructures,” Nat. Mater., 13 (5), 451 –460 https://doi.org/10.1038/nmat3921 NMAACR 1476-1122 (2014). Google Scholar

13. 

M. A. Kats et al., “Nanometre optical coatings based on strong interference effects in highly absorbing media,” Nat. Mater., 12 (1), 20 –24 https://doi.org/10.1038/nmat3443 NMAACR 1476-1122 (2013). Google Scholar

14. 

X. Gan et al., “Chip-integrated ultrafast graphene photodetector with high responsivity,” Nat. Photonics, 7 (11), 883 –887 https://doi.org/10.1038/nphoton.2013.253 NPAHBY 1749-4885 (2013). Google Scholar

15. 

S. Lischke et al., “Ultra-fast germanium photodiode with 3-dB bandwidth of 265 GHz,” Nat. Photonics, 15 (12), 925 –931 https://doi.org/10.1038/s41566-021-00893-w NPAHBY 1749-4885 (2021). Google Scholar

16. 

W. Hu et al., “Germanium/perovskite heterostructure for high-performance and broadband photodetector from visible to infrared telecommunication band,” Light Sci. Appl., 8 106 https://doi.org/10.1038/s41377-019-0218-y (2019). Google Scholar

17. 

M. König et al., “Quantum spin Hall insulator state in HgTe quantum wells,” Science, 318 (5851), 766 –770 https://doi.org/10.1126/science.1148047 SCIEAS 0036-8075 (2007). Google Scholar

18. 

H. Krishnamoorthy et al., “Topological insulator metamaterials,” Chem. Rev., 123 (8), 4416 –4442 https://doi.org/10.1021/acs.chemrev.2c00594 CHREAY 0009-2665 (2023). Google Scholar

19. 

H. Zhang et al., “Topological insulators in Bi2Se3, Bi2Te3 and Sb2Te3 with a single Dirac cone on the surface,” Nat. Phys., 5 (6), 438 –442 https://doi.org/10.1038/nphys1270 NPAHAX 1745-2473 (2009). Google Scholar

20. 

D. Kong et al., “Ambipolar field effect in the ternary topological insulator (BixSb1x)2Te3 by composition tuning,” Nat. Nanotechnol., 6 (11), 705 –709 https://doi.org/10.1038/nnano.2011.172 NNAABX 1748-3387 (2011). Google Scholar

21. 

H. Lu et al., “Topological insulator plasmonics and enhanced light-matter interactions,” Plasmon-Enhanced Light-Matter Interactions, 89 –116 Springer International Publishing, Cham (2022). Google Scholar

22. 

H. Lu et al., “Magnetic plasmon resonances in nanostructured topological insulators for strongly enhanced light-MoS2 interactions,” Light Sci. Appl., 9 191 https://doi.org/10.1038/s41377-020-00429-x (2020). Google Scholar

23. 

J. Yin et al., “Plasmonics of topological insulators at optical frequencies,” NPG Asia Mater., 9 (8), e425 https://doi.org/10.1038/am.2017.149 1884-4057 (2017). Google Scholar

24. 

X. Sun et al., “Topological insulator metamaterial with giant circular photogalvanic effect,” Sci. Adv., 7 eabe5748 https://doi.org/10.1126/sciadv.abe5748 STAMCV 1468-6996 (2021). Google Scholar

25. 

K. J. Boller et al., “Observation of electromagnetically induced transparency,” Phys. Rev. Lett., 66 (20), 2593 –2596 https://doi.org/10.1103/PhysRevLett.66.2593 PRLTAO 0031-9007 (1991). Google Scholar

26. 

T. F. Krauss, “Why do we need slow light?,” Nat. Photonics, 2 (8), 448 –450 https://doi.org/10.1038/nphoton.2008.139 NPAHBY 1749-4885 (2008). Google Scholar

27. 

L. V. Hau et al., “Light speed reduction to 17 metres per second in an ultracold atomic gas,” Nature, 397 (6720), 594 –598 https://doi.org/10.1038/17561 (1999). Google Scholar

28. 

N. Caselli et al., “Generalized Fano lineshapes reveal exceptional points in photonic molecules,” Nat. Commun., 9 396 https://doi.org/10.1038/s41467-018-02855-3 NCAOBW 2041-1723 (2018). Google Scholar

29. 

B. Peng et al., “What is and what is not electromagnetically induced transparency in whispering-gallery microcavities,” Nat. Commun., 5 5082 https://doi.org/10.1038/ncomms6082 NCAOBW 2041-1723 (2014). Google Scholar

30. 

C. Wang et al., “Electromagnetically induced transparency at a chiral exceptional point,” Nat. Phys., 16 (3), 334 –340 https://doi.org/10.1038/s41567-019-0746-7 NPAHAX 1745-2473 (2020). Google Scholar

31. 

N. Liu et al., “Plasmonic analogue of electromagnetically induced transparency at the Drude damping limit,” Nat. Mater., 8 758 –762 https://doi.org/10.1038/nmat2495 NMAACR 1476-1122 (2009). Google Scholar

32. 

N. Liu et al., “Planar metamaterial analogue of electromagnetically induced transparency for plasmonic sensing,” Nano Lett., 10 (4), 1103 –1107 https://doi.org/10.1021/nl902621d NALEFD 1530-6984 (2010). Google Scholar

33. 

N. Liu et al., “Three-dimensional plasmon rulers,” Science, 332 (6036), 1407 –1410 https://doi.org/10.1126/science.1199958 SCIEAS 0036-8075 (2011). Google Scholar

34. 

T. Kim et al., “Electrically tunable slow light using graphene metamaterials,” ACS Photonics, 5 (5), 1800 –1807 https://doi.org/10.1021/acsphotonics.7b01551 (2018). Google Scholar

35. 

M. Sabarinathan et al., “n-type to p-type transition of electrical conduction in silver (Ag)-modified Bi2Te3 nanosheets,” J. Electron. Mater., 51 (12), 7275 –7282 https://doi.org/10.1007/s11664-022-09974-0 JECMA5 0361-5235 (2022). Google Scholar

36. 

A. Bailini et al., “Pulsed laser deposition of Bi2Te3 thermoelectric films,” Appl. Surf. Sci., 254 (4), 1249 –1254 https://doi.org/10.1016/j.apsusc.2007.09.039 ASUSEE 0169-4332 (2007). Google Scholar

37. 

M. Rusek et al., “Bismuth amides as promising ALD precursors for Bi2Te3 films,” J. Crys. Growth, 470 128 –134 https://doi.org/10.1016/j.jcrysgro.2017.04.019 GROWAH 0017-4793 (2017). Google Scholar

38. 

K. Sreekanth et al., “Biosensing with the singular phase of an ultrathin metal-dielectric nanophotonic cavity,” Nat. Commun., 9 369 https://doi.org/10.1038/s41467-018-02860-6 NCAOBW 2041-1723 (2018). Google Scholar

39. 

S. N. Chowdhury et al., “Wide-range angle-sensitive plasmonic color printing on lossy-resonator substrates,” Adv. Opt. Mater., 12 (4), 2301678 https://doi.org/10.1002/adom.202301678 2195-1071 (2024). Google Scholar

40. 

L. Goncalves et al., “Optimization of thermoelectric properties on Bi2Te3 thin films deposited by thermal co-evaporation,” Thin Solid Films, 518 (10), 2816 –2821 https://doi.org/10.1016/j.tsf.2009.08.038 THSFAP 0040-6090 (2010). Google Scholar

41. 

V. Y. Kolosov and A. A. Yushkov, “Microstructures in thin Bi2Te3 films according to transmission electron microscopy,” AIP Conf. Proc., 2313 030019 https://doi.org/10.1063/5.0033544 APCPCS 0094-243X (2020). Google Scholar

42. 

J. Y. Ou et al., “Ultraviolet and visible range plasmonics in the topological insulator Bi1.5Sb0.5Te1.8Se1.2,” Nat. Commun., 5 5139 https://doi.org/10.1038/ncomms6139 NCAOBW 2041-1723 (2014). Google Scholar

43. 

H. Lu et al., “Sb2Te3 topological insulator: surface plasmon resonance and application in refractive index monitoring,” Nanoscale, 11 (11), 4759 –4766 https://doi.org/10.1039/C8NR09227C NANOHL 2040-3364 (2019). Google Scholar

44. 

G. Jellison and F. Modine, “Parameterization of the optical functions of amorphous materials in the interband region,” Appl. Phys. Lett., 69 (3), 371 –373 https://doi.org/10.1063/1.118064 APPLAB 0003-6951 (1996). Google Scholar

45. 

B. Xia et al., “Indications of surface-dominated transport in single crystalline nanoflake devices of topological insulator Bi1.5Sb0.5Te1.8Se1.2,” Phys. Rev. B, 87 (8), 085442 https://doi.org/10.1103/PhysRevB.87.085442 (2013). Google Scholar

46. 

Y. Lin et al., “Thermocatalytic hydrogen peroxide generation and environmental disinfection by Bi2Te3 nanoplates,” Nat. Commun., 12 180 https://doi.org/10.1038/s41467-020-20445-0 NCAOBW 2041-1723 (2021). Google Scholar

47. 

N. Krishnamoorthy et al., “Infrared dielectric metamaterials from high refractive index chalcogenides,” Nat. Commun., 11 1692 https://doi.org/10.1038/s41467-020-15444-0 NCAOBW 2041-1723 (2020). Google Scholar

48. 

H. Lu et al., “Nearly perfect absorption of light in monolayer molybdenum disulfide supported by multilayer structures,” Opt. Express, 25 (18), 21630 –21636 https://doi.org/10.1364/OE.25.021630 OPEXFF 1094-4087 (2017). Google Scholar

49. 

B. I. Afinogenov et al., “Ultrafast all-optical light control with Tamm plasmons in photonic nanostructures,” ACS Photonics, 6 (4), 844 –850 https://doi.org/10.1021/acsphotonics.8b01792 (2019). Google Scholar

50. 

R. Badugu and J. R. Lakowicz, “Tamm state-coupled emission: effect of probe location and emission wavelength,” J. Phys. Chem. C, 118 (37), 21558 –21571 https://doi.org/10.1021/jp506190h JPCCCK 1932-7447 (2014). Google Scholar

51. 

R. Badugu et al., “Radiative decay engineering 7: Tamm state-coupled emission using a hybrid plasmonic-photonic structure,” Anal. Biochem., 445 1 –13 https://doi.org/10.1016/j.ab.2013.10.009 ANBCA2 0003-2697 (2014). Google Scholar

52. 

S. Xiao et al., “Active modulation of electromagnetically induced transparency analogue in terahertz hybrid metal-graphene metamaterials,” Carbon, 126 271 https://doi.org/10.1016/j.carbon.2017.10.035 CRBNAH 0008-6223 (2018). Google Scholar

53. 

L. Mao et al., “Reversible switching of electromagnetically induced transparency in phase change metasurfaces,” Adv. Photonics, 2 (5), 056004 https://doi.org/10.1117/1.AP.2.5.056004 (2020). Google Scholar

54. 

Y. He et al., “Plasmon induced transparency in a dielectric waveguide,” Appl. Phys. Lett., 99 (4), 043113 https://doi.org/10.1063/1.3621860 APPLAB 0003-6951 (2011). Google Scholar

55. 

J. Tian et al., “Near-infrared super-absorbing all-dielectric metasurface based on single-layer germanium nanostructures,” Laser Photonics Rev., 12 (9), 1800076 https://doi.org/10.1002/lpor.201800076 (2018). Google Scholar

56. 

A. Taflove and S. Hagness, Computational Electrodynamics: The Finite-Difference Time-Domain Method, 2nd ed.Artech House, Boston, MA (2000). Google Scholar

57. 

T. J. Bright et al., “Infrared optical properties of amorphous and nanocrystalline Ta2O5 thin films,” J. Appl. Phys., 114 (8), 083515 https://doi.org/10.1063/1.4819325 JAPIAU 0021-8979 (2013). Google Scholar

58. 

J. Kischkat et al., “Mid-infrared optical properties of thin films of aluminum oxide, titanium dioxide, silicon dioxide, aluminum nitride, and silicon nitride,” Appl. Opt., 51 (28), 6789 –6798 https://doi.org/10.1364/AO.51.006789 APOPAI 0003-6935 (2012). Google Scholar

59. 

E. Palik, Handbook of Optical Constants of Solids, Academic Press( (1998). Google Scholar

60. 

M. J. Weber, Handbook of Optical materials, CRC Press, Boca Raton (2003). Google Scholar

Biography

Hua Lu is as a professor at Northwestern Polytechnical University in China. He received his PhD from the Chinese Academy of Sciences. He was a postdoc fellow at Swinburne University of Technology in Australia, and then joined Northwestern Polytechnical University in 2015. His current research interest is focused on nanophotonics, plasmonics, and nanomaterials. To date, he has published more than 100 articles in professional journals, including Light Sci. Appl., Adv. Photon., Laser Photon. Rev., and Small, with citation counts of >6400. He received the 13th Shaanxi Youth Science and Technology Award (2020) and was continuously selected as Elsevier’s Highly Cited Chinese Researcher (2021–2023).

Jianlin Zhao is a chief scientist at Northwestern Polytechnical University and the director of MOE Key Laboratory of Material Physics and Chemistry under Extraordinary Conditions. He has been selected as a fellow of the Chinese Optical Society and the executive editor-in-chief of Acta Optica Sinica. His research interests include micro-/nanophotonics, light-field manipulation, digital holography, and optical fiber sensing. To date, he has published three textbooks (sole author) and more than 500 articles in professional journals, including Science, Nat. Mater., Adv. Mater., Light Sci. Appl., and Nat. Commun., with citation counts of >13,000. He was continuously selected as Elsevier’s Highly Cited Chinese Researcher (2021–2023).

Biographies of the other authors are not available.

CC BY: © The Authors. Published by SPIE and CLP under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Hua Lu, Shouhao Shi, Dikun Li, Shuwen Bo, Jianxu Zhao, Dong Mao, and Jianlin Zhao "λ/20-Thick cavity for mimicking electromagnetically induced transparency at telecommunication wavelengths," Advanced Photonics 6(3), 036001 (3 April 2024). https://doi.org/10.1117/1.AP.6.3.036001
Received: 27 October 2023; Accepted: 14 March 2024; Published: 3 April 2024
Advertisement
Advertisement
KEYWORDS
Bismuth

Tellurium

Refractive index

Nanofilms

Photonic crystals

Transparency

Silver

Back to Top