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

Guiding-mode-assisted double-BICs in an all-dielectric metasurface

Open Access Open Access

Abstract

The electromagnetically induced transparency (EIT) effect realized in a metasurface is potential for slow light applications for its extreme dispersion variation in the transparency window. Herein, we propose an all-dielectric metasurface to generate a double resonance-trapped quasi bound states in the continuum (BICs) in the form of EIT or Fano resonance through selectively exciting the guiding modes with the grating. The group delay of the EIT is effectively improved up to 2113 ps attributing to the ultrahigh Q-factor resonance carried by the resonance-trapped quasi-BIC. The coupled harmonic oscillator model and a full multipole decomposition are utilized to analyze the physical mechanism of EIT-based quasi-BIC. In addition, the BIC based on Fano and EIT resonance can simultaneously exist at different wavelengths. These findings provide a new feasible platform for slow light devices in the near-infrared region.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Electromagnetically induced transparency (EIT) is caused by quantum destructive interference between two different excitation paths in a three-level atomic system. When the energy levels of the interaction resonance coincide, the classical analogue of EIT can evolve from the Fano resonance [1]. The EIT effect with a transparent window in the transmission spectra exhibits the remarkable feature of strong dispersion, which enhances the nonlinear effect [25] and leads to the slow light phenomenon [69]. However, the scathing experimental requirements of stable lasers and a cryogenic temperature environment hinder its practical applications [10]. To overcome the barriers, the analogue of the EIT effect is implemented in the classical systems such as waveguide systems [11,12], photonic crystals [13], and metamaterials [1416]. Particularly, due to the easy design of resonances and subwavelength effective mode volumes in metamaterials, much attention has turned to EIT in various oscillator systems [17,18]. With the characteristics of the EIT-like effect realized in metamaterials, it has a great ability to drastically slow down the group delay of light, promising to be used in slow light devices [19]. However, the Q-factor of the EIT-like effect is relatively restricted, thus the group delay in the transparent window is limited.

Bound state in the continuum (BIC) represents perfectly localized states with no leakage even though they coexist with a continuous spectra of radiating waves, which effectively manipulates the light-matter interaction and generates an ultrahigh Q-factor resonance [20,21]. Originally, this concept was explored in quantum mechanics [22], and in recent years, has received much attention in photonics [2325]. As a dark mode, the optical BIC is invisible in the optical spectra with zero linewidth and infinite Q-factor. When it collapses to the quasi-BIC in the form of Fano resonance or EIT resonance with perturbations, it can be observed experimentally and widely investigated in the linear optical regime [26,27], the nonlinear enhancement [2830], coherent light generation [31,32], and integrated circuits [33,34]. Using BICs instead of the traditional low-quality optical modes will greatly improve the performance of the optical devices. For example, in the compound grating waveguide structure, the Goos-Hänchen shift [35,36] and harmonic generation can be greatly enhanced [37]. BIC can be divided into two major categories according to the formation of coupling between the resonant modes and all radiation channels in the surrounding space. Symmetry-protected BIC [38,39] is defined if the coupling between light in the free space and eigenmode of the resonator vanishes by symmetry breaking while the resonance-trapped BIC [4042] like Friedrich-Wingten BIC [24,4346]) is formed due to the coupling coefficient disappears by tuning the system parameters [4749]. Particularly, the resonance-trapped BIC can be realized at an arbitrary point of reciprocal space rather than just restricted to the edge of the Brillouin region, realizing a propagating BIC in the continuous range of the free space [41]. So far, their unique properties have attracted much attention and are explored in various material systems, such as nanoresonators (change radius) [27], resonator array [50], and metallic metasurface [47,48]. For the single subwavelength nanoresonator, the BIC can be accessed both numerically and experimentally at normal incidence by using vector beams [27,45,51]. However, most resonance-trapped quasi-BICs in the metasurface are implemented by the oblique incident light in the metasurface, which complicates the measurement of the resonance-trapped quasi-BICs. In addition, metal-based metasurfaces decrease the Q-factor of the quasi-BICs for their inevitable intrinsic high damping losses. Hence, the metasurface used to generate resonance-trapped quasi-BIC with a high Q-factor by normal incident light and improving the group delay of the EIT effect by resonance-trapped quasi-BIC still needs to be further investigated [52].

In this paper, we design an all-dielectric metasurface to achieve resonance-trapped BICs assisted by guiding modes at normal incidence. The double resonance-trapped quasi-BICs in the form of Fano resonance or EIT resonance are realized through selectively exciting different guiding modes in the proposed metasurface for its useful frequency tunability. In addition, the EIT-based quasi-BIC can actively switch to Fano resonance-based quasi-BIC by modulating the polarization of the incident light. Particularly, coupled harmonic oscillator model is employed to analyze the generation mechanism of the quasi-BICs in the form of EIT resonance. Attributing to the high Q-factor of the quasi-BIC, the dispersion in the transparency window of the EIT extremely changes and the group delay is up to 2113 ps, which may offer possible applications for slow light in the near-infrared region.

2. Structure design

The all-dielectric metasurface utilized for exciting the quasi-BICs in the form of EIT or Fano resonance and the specific geometrical parameters are schematically illustrated in Figs. 1(a) and 1(b). The periods of the unit cell in the x and y directions are Px = 1300 nm and Py = 1050 nm, respectively. The upper layer of a unit cell is the grating layer with three bars, which is utilized to increase the wave vector of the incident light and excite the guiding modes in the middle layer, while the guiding mode is supported by the micro-structure composed of the Si waveguide layer and the Si3N4 substrate. The material of the upper and middle layers is Si with the refractive index of n1 = 3.5 while the substrate layer is Si3N4 with the refractive index of n2 = 2 [53]. The thickness of the Si grating, guiding layer, and substrate are H = 150 nm, D = 150 nm, and T = 150 nm, respectively. The length of the Si-bars is L = 500 nm. The width of the middle Si-bar W2 is variable while the side Si-bars are fixed as W1 = 150 nm in the following simulations. All the calculations of the all-dielectric metasurface are numerically simulated by the finite element method (FEM) with the commercial software COMSOL Multiphysics 5.4. The Floquet-periodic boundary conditions are employed in the x and y directions, and the perfectly matching layers are adopted in the z-direction.

 figure: Fig. 1.

Fig. 1. Schematic of the unit cell of the all-dielectric metasurface structure under TE-polarized illumination. (a) The unit cell is arranged with a period Px = 1300 nm, Py = 1050 nm. The Si grating height H = 150 nm, the Si guiding layer thickness D = 150 nm, and the Si3N4 substrate thickness T = 150 nm. (b) Top view of the unit cell. The length of the Si-bars L = 500 nm. Meanwhile, W1 = 150 nm and W2 represent the width of the Si-bars on both sides and in the middle respectively.

Download Full Size | PDF

3. Results and discussions

To demonstrate the mechanism of the BICs in the metasurface, the in-plane band structure of eigenmodes is calculated by a three-dimensional finite element method eigenfrequency solver as shown in Fig. 2(a). Herein, the width of the triple Si bar is fixed as 150 nm. In the investigating range from 1600 nm to 2100 nm, there are three groups of doubly degenerate mode at the $\mathrm{\Gamma }$ point, which completely decouple with the external environment, leading to the formation of three resonance-trapped BICs [41,54]. Figure 2(b) illustrates the corresponding electromagnetic field distributions at the Γ point of the six bands, which are well confined in the proposed metasurface and do not decay. The quasi-BICs will be established away from the $\mathrm{\Gamma }$ point where the modes are no longer degenerate.

 figure: Fig. 2.

Fig. 2. (a) Band structure of the metasurface and (b) the corresponding electromagnetic field distribution of eigenmodes at the $\mathrm{\Gamma }$ point. The structural parameters are: W1= W2 = 150 nm, T = D = H = 150 nm, L = 500 nm, Px = 1300 nm, Py = 1050 nm.

Download Full Size | PDF

In order to better understand the optical response of the proposed metasurface, the transmission spectra are investigated with a TE-polarized plane wave propagating along the z-direction. The calculated transmission spectra with varying W2 are shown in Fig. 3(a), which exhibits two lineshapes corresponding to EIT resonance around 1685 nm and Fano resonance around 1980 nm. When the width of the triple Si bar is equal to 150 nm, the linewidth of the resonance turns to zero and the Q-factor is infinite, which means the establishment of the BICs where no leaky energy from the resonances to the free space attributing to the destructive inference of the guiding modes and grating modes. When the width of the middle Si bars is different from the side Si bars, one can find the Q-factor and linewidth of the EIT resonance and Fano resonance is finite, which means the BICs collapse into quasi-BICs. The linewidth of resonance peaks of the quasi-BICs widen when the W2 is away from 150 nm. To further confirm the existence of the BICs, we calculate both the transmission spectra and the eigenmodes spectra with W2 as depicted in Figs. 3(b) and 3(c). For clarity, we define the two eigenmodes as quasi-BIC1 and quasi-BIC2, which radiate as EIT resonance at short wavelength and Fano resonance at long wavelength, respectively.

 figure: Fig. 3.

Fig. 3. The transmission spectra and eigenmode analysis of a metasurface under TE-polarization incident light. (a) Evolution of the transmission spectra with different W2. Transmission spectra and eigenmodes spectra for (b) BIC1 radiating as EIT resonance and (c) BIC2 radiating as Fano resonance with error bars represent the magnitude of the mode inverse radiation lifetime. All the resonance energy is completely restricted to a bound state in the radiation continuum at W2 = 150 nm, which are defined as the BICs. When the linewidths appear in the spectra (W2≠150 nm), they collapse into quasi-BICs.

Download Full Size | PDF

According to the eigenmode analysis shown in Figs. 3(b) and 3(c), the metasurface supports two resonance-trapped BICs that are unstable against perturbations. These leaky modes can be represented by a complex eigenfrequency $\omega = {\omega _0} - i\gamma$, where ${\omega _0}$ and $\gamma$ are resonant frequency and leakage rates of modes, respectively. A slight perturbation induces energy leakage of an ideal bound state, resulting in a quasi-BIC with the observable spectral feature of finite linewidth. The error bars indicate the resonance linewidth/leakage rates, and both eigenmodes disappear at W2 = 150 nm. It is worth mentioning that BIC1 establishes in a cross-section of two modes. Herein, the mode inverse radiation lifetime is calculated for the hybrid mode in the transmission peak, where the dispersion rapidly changes and is helpful for the slow light devices. We can observe from the transmission spectra that BIC1 and BIC2 with infinite Q-factors at W2 = 150 nm transform into high Q-factors quasi-BICs, which verify the eigenmode analysis of BICs. It should be pointed out that the double-BICs are resonance-trapped BICs since the varying W2 does not break the in-plane inversion symmetry of the metasurface.

The Q-factors of the quasi-BICs can be extracted and further defined as $Q = {{{\omega _0}} / {2\gamma }}$. Since the resonance linewidths of BIC disappear, the general method to verify the existence of BIC is to track the diverging trajectory of Q-factor from the quasi-BICs. As shown in Figs. 4(a) and 4(b), it reveals that the Q-factors of two quasi-BICs of both the EIT and Fano modes tend to be infinite at W2 = 150 nm and the Q-factors diverge when W2 changes away from 150 nm, which validates the existence of BICs

 figure: Fig. 4.

Fig. 4. Extracted Q-factors of two quasi-BICs for (a) EIT and (b) Fano mode.

Download Full Size | PDF

The electric field distributions with different W2 are calculated to investigate the generated mechanism of the quasi-BICs and BICs in the metasurface. One should note that the incident light is above the wavevector of the guiding mode in the bare guiding layer, which means that the guiding mode cannot be excited and leak into the free space [55]. When the grating layer is mounted above the guiding layer, the incident light diffracts in the grating layer and the wave vector of the incident light is modulated to match the guiding modes, leading to the radiation of the guiding modes. For the quasi-BIC1 in the form of EIT resonance, one can find that one of the guiding modes named transverse-electric (TE4,1) mode is excited and radiating to free space. The middle Si-bar grating provides a leaky channel to the TE4,1 mode in the shorter wavelength of the dip of EIT resonance while the longer wavelength of TE4,1 mode is excited by side Si-bar grating as shown in Figs. 5(a) and 5(b).

 figure: Fig. 5.

Fig. 5. Electric field distributions in the x-z plane at the resonant wavelength. (a), (b) for two resonant dips of quasi-BIC1 at W2 = 100 nm, (c) for BIC1 in the dip of the transmission spectra at W2 = 150 nm, and (d) for the resonant dip of quasi-BIC2 at W2 = 100 nm. The white arrows indicate the magnetic field directions.

Download Full Size | PDF

When the W2 = W1 = 150 nm, the electric field localizes in the grating layer, and TE4,1 mode cannot be excited at the 1689 nm as shown in Fig. 5(c), which leads to the establishment of BIC. The phenomenon can be explained as follow: the period of grating in the x-direction decreases rapidly from the triple Si bars to a single Si bar as the effect of them are the same in the unit cell when W2 = W1 = 150 nm, leading the wavevector of grating increases sharply. The matching wavevector frequency point between the TE4,1 mode and the refraction light will blue shift out of the investigating frequency region, which means that the TE4,1 mode cannot be excited from 1600 nm to 2100 nm when W2 = W1 = 150 nm. Thus, only the grating mode and single transmission dip remain as shown in Fig. 5(c) and Fig. 3(a). In addition, the electric field distribution of quasi-BIC2 is depicted in Fig. 5(d) where the TE2,1 mode is excited and radiating to the free space through all the triple Si-bars. Herein, the central wavelength of quasi-BIC2 is larger than quasi-BIC1. For the TE2,1 mode, the wavevector is larger than the TE4,1 mode at the same frequency in the guiding layer, Thus, the TE4,1 mode is excited by grating at a short wavelength, while the TE2,1 at a long wavelength with the same grating [56].

Using the coupled harmonic oscillator model, the strong coupling regime in the all-dielectric metasurface can be described to explore the formation of quasi-BIC1. The respective contributions of the grating mode and guiding mode can be calculated as [57]:

$$\left( {\begin{array}{*{20}{c}} {{E_{gra}} - {{i\hbar {\gamma_{gra}}} / 2}}&g\\ g&{{E_{gui}} - {{i\hbar {\gamma_{gui}}} / 2}} \end{array}} \right)\left( {\begin{array}{*{20}{c}} \alpha \\ \beta \end{array}} \right) ={\pm} \textrm{E}\left( {\begin{array}{*{20}{c}} \alpha \\ \beta \end{array}} \right),$$
where ${E_{gra}}$ and ${E_{gui}}$ are the energies of grating mode and guiding mode while ${\gamma _{gra}}$ and ${\gamma _{gui}}$ are the dissipation rates of the two oscillators, respectively. g is the coupling strength, and $\mathrm{\ \pm E}$ are the eigen-energies of the hybrid states. The parameters $\alpha $ and $\beta $ are the eigenvector components and satisfy ${|\alpha |^2} + {|\beta |^2} = 1$. By considering the detuning, the eigenvalues can be described as:
$${\textrm{E}_ \pm } = ({E_{gra}} + {E_{gui}})/2 \pm \sqrt {4{g^2} + {{\left( {\delta - \frac{i}{2}({{\gamma_{gra}} - {\gamma_{gui}}} )} \right)}^2}} /2,$$
with the energy difference $\delta = {E_{gra}} - {E_{gui}}$ between grating mode and guiding mode. The criterion for the strong coupling regime is described as $\hbar \varOmega > ({{\gamma_{gra}} - {\gamma_{gui}}} )/2$ or $\hbar \varOmega > ({{\gamma_{gra}} + {\gamma_{gui}}} )/2$, where $\hbar \Omega = \sqrt {4{g^2} - {{({{\gamma_{gra}} + {\gamma_{gui}}} )}^2}/4}$ is the Rabi splitting energy at ${E_{gra}} = {E_{gui}}$. In this case, ${E_{gra}}$ and ${E_{gui}}$ can be extracted from the transmission spectra. The top figure of Fig. 3(b) shows the transmission spectra of BIC1. Since the guiding mode is not excited at W2 = 100 nm, we observed the line width of the resonant peak disappearance at the wavelength of 1686 nm. ${E_{gui}}$ is derived to be ∼ 735.9 meV by the half-width at the half-maximum (HWHM). We can fit the results (yellow dotted lines) to indicate the appearance of anti-crossing behavior using the coupled harmonic oscillator model shown in Fig. 3(b). From the fitted curve, we obtain the coupling constant g to be ∼ 6.6 meV, whereas the corresponding dissipation rates ${\gamma _{gra}}$ of ∼ 17.9 meV and ${\gamma _{gui}}$ of ∼ 0.2 meV. Thus, the criterion $\hbar \varOmega > ({{\gamma_{gra}} + {\gamma_{gui}}} )/2$ is satisfied with the Rabi splitting of around 9.8 meV, implying that the coupling between the grating mode and guiding mode enters the strong coupling regime.

To further explain the EIT resonance represented as quasi-BIC1, the coupled Lorentz oscillator model is utilized to analyze the interaction between the grating modes and guiding modes. The grating mode in our structure is defined by oscillator 1. The TE4,1 mode is defined by oscillator 2, which is a dark mode and cannot be directly excited by the incident light. The coupled two-oscillator system to analyze the interaction between two modes is as follows [15]:

$${\ddot{x}_1} + {\gamma _1}{\dot{x}_1} + \omega _0^2{x_1} + \kappa {x_2} = E,$$
$${\ddot{x}_2} + {\gamma _2}{\dot{x}_2} + {({\omega + \delta } )^2}{x_2} + \kappa {x_1} = 0.$$
where ${x_i}$ and ${\gamma _i}$ are the resonance amplitude and damping rates of the resonance modes ($i = 1$ represents the grating mode and $i = 2$ refers to the TE4,1 mode), respectively. ${\omega _0}$ is the resonance frequency of oscillator 1 when it is isolated from oscillator 2. $\delta $ is the detuning frequency between the two oscillators. $\kappa $ is the coupling coefficient between the two oscillators and g is a geometric parameter related to the coupling strength between the resonant mode and the incident electric field $E$. The susceptibility $\chi $ of the system is proportional to the amplitude ${x_i}$ and can be described as:
$$\chi (\omega )= {\chi _{_r}} + i{\chi _i} \propto \frac{{({\omega - {\omega_0} - \delta } )+ i\frac{{{\gamma _2}}}{2}}}{{\left( {\omega - {\omega_0} + i\frac{{{\gamma_1}}}{2}} \right)\left( {\omega - {\omega_0} - \delta + i\frac{{{\gamma_{_2}}}}{2}} \right) - \frac{{{\kappa ^2}}}{4}}}.$$

As the energy dissipation is proportional to the imaginary part of susceptibility ${\chi _i}$, the transmission T can be described as:

$$T(\omega )= 1 - g{\chi _i}(\omega ),$$

In Figs. 6(a)–6(e), we define the difference between the simulation and the fitting as the error bars. Focus on the transparency window, the error bars are small. The theoretical fitting of transmission spectra marked by the red curve reveals good agreements with simulation results of different W2 values represented by black points. In addition, the fitting parameters of different W2 are provided in Fig. 6(f). It is apparent that ${\gamma _2}$ show very slight variations with W2, while the detuning frequency $\delta $, the damping rate ${\gamma _1}$, and the coupling coefficient $\kappa $ of the grating mode display significant variations. In this case, ${\gamma _2}$ gets several times smaller than ${\gamma _1}$, which means that the excitation of the guiding mode can reach a higher Q-factor. As the W2 moves away from 150 nm, $\delta $ and $\kappa $ increase and ${\gamma _1}$ decreases due to the stronger coupling of TE4,1 mode and grating mode, which refers that the switchable establishment between BIC1 and quasi-BIC1 and the active modulation of EIT resonance can be attributed to the change of the damping rate of the grating mode. When W2 = 150 nm, ${\gamma _1}$ becomes large enough to completely suppress the TE4,1 mode excitation and hampers the destructive interference between the TE4,1 mode and grating modes, which eventually leads to the disappearance of the EIT resonance peak. Herein, the only grating mode can remain as the ${\gamma _2}$ turns to zero when W2 = 150 nm, which validates the above analysis in Fig. 5(c).

 figure: Fig. 6.

Fig. 6. (a)-(e) The numerical transmission spectra (black dots) and corresponding analytically fitted (red curves) transmission spectra with the error bars of the proposed metasurface with different W2. (f) The fitting parameters $\delta $, $\kappa $, ${\gamma _1}$ and ${\gamma _2}$ of coupled two-oscillator model with the variation of W2.

Download Full Size | PDF

We use cartesian electromagnetic multipole decomposition to analyze the contribution of different multipole components of quasi-BICs in far-field radiation, which can be defined as:

$$\vec{P} = \frac{1}{{i\omega }}\smallint \vec{j}{d^3}r,$$
$$\vec{M} = \frac{1}{{2c}}\smallint ({\vec{r} \times \vec{j}} ){d^3}r,$$
$$\vec{T} = \frac{1}{{10c}}\smallint [{({\vec{r} \cdot \vec{j}} )\vec{r} - 2{r^2}\vec{j}} ]{d^3}r,$$
$$Q_{\alpha \beta }^{(e )} = \frac{1}{{2i\omega }}\smallint \left[ {{r_\alpha }{j_\beta } + {r_\beta }{j_\alpha } - \frac{2}{3}({\vec{r} \cdot \vec{j}} ){\delta_{\alpha ,\beta }}} \right]{d^3}r,$$
$$Q_{\alpha \beta }^{(m )} = \frac{1}{{3c}}\smallint [{{{({\vec{r} \times \vec{j}} )}_\alpha }{r_\beta } + {{({\vec{r} \times \vec{j}} )}_\beta }{r_\alpha }} ]{d^3}r,$$
where c and $\omega $ are the speed and angular frequency for light, respectively. $\alpha ,\beta = x,y,z$ and $\vec{j} ={-} i\omega {\varepsilon _0}({{n^2} - 1} )\vec{E}$ is the current density distribution in a unit cell. The $\vec{P}$, $\vec{M}$, $\vec{T}$, ${Q^{(e )}}$, ${Q^{(m )}}$ are the electric dipole (ED) moment, magnetic dipole (MD) moment, toroidal dipole (TD) moment, electric quadrupole (EQ) moment, and magnetic quadrupole (MD) moment, respectively. The far-field scattered power of them can be expressed by the following formulas [58]:
$$\begin{array}{l} {I_P} = \frac{{2{\omega ^4}}}{{3{c^3}}}{|{\vec{P}} |^2},{I_M} = \frac{{2{\omega ^4}}}{{3{c^3}}}{|{\vec{M}} |^2},{I_T} = \frac{{2{\omega ^6}}}{{3{c^5}}}{|{\vec{T}} |^2},\\ {I_{{Q^{(e )}}}} = \frac{{{\omega ^6}}}{{5{c^5}}}\sum {|{Q_{\alpha \beta }^{(e )}} |^2},{I_{{Q^{(m )}}}} = \frac{{{\omega ^6}}}{{40{c^5}}}\sum {|{Q_{\alpha \beta }^{(e )}} |^2}. \end{array}$$

As shown in Figs. 7(a) and 7(b), one can find that the scattering power of the ED moment and the EQ moment are dominant for quasi-BIC1 while the MQ moment and the TD moment for quasi-BIC2 play a dominant role at W2 = 100 nm. For quasi-BIC1, the ED moment was attributed to the guiding mode waves with energy fluxes. The generation of transmission dips around 1670 nm is mainly due to the destructive interference of the ED moment and the EQ moment owing to their similar far-field scattering patterns. Due to the large scattering power difference between the ED moment and the EQ moment, the destructive interference effect weakens creating a peak between the two dips. For quasi-BIC2, the scattering of the MQ moment is significant, while the scattering of the ED moment is suppressed dramatically. Owing to the scattered power difference between the MQ moment and the TD moment, the radiation channel is provided, resulting in a Fano resonance around 1976 nm.

 figure: Fig. 7.

Fig. 7. Multipole decomposition analysis of the metasurface of (a) quasi-BIC1 and (b) quasi-BIC2 when W2 = 100 nm, respectively.

Download Full Size | PDF

Then the transmission spectra with a TM-polarized plane wave propagating along the z-axis are calculated to investigate the responses of the metasurface to polarization. As shown in Fig. 8(a), the transmission spectra with varying W2 infer that two Fano resonances excite around 1800 nm and 1970 nm. Compared with Fig. 3(a), it can be noted that the switch between Fano and EIT mode BIC is realized. It is obvious that the switch between EIT and Fano modes can be achieved not only by altering different wavelengths of the same polarization but also by converting polarization. As the geometric parameter W2 changes from 50 nm to 250 nm, two Fano resonances with vanished linewidth demonstrate the existence of the BIC states with infinite Q-factors at W2 = 150 nm in Figs. 8(b) and 8(c). For clarity, we define the two states as BIC3 and BIC4. According to the eigenmode analysis, the linewidths of the two resonances become disappeared at W2 = 150 nm, and the corresponding Q-factors also diverge as the W2 gradually changes, validating two resonance-trapped BICs. The Q-factor of the metasurface as a function of the W2 is presented in Fig. 8(d), it becomes an ideal BIC at W2 = 150 nm.

 figure: Fig. 8.

Fig. 8. The transmission spectra and eigenmode analysis of the metasurface under TM-polarization incident light. (a) Evolution of the transmission spectra with different W2. Transmission spectra and eigenmodes spectra for (b) BIC3 and (c) BIC4 radiating as Fano resonances with error bars represent the magnitude of the mode inverse radiation lifetime. All the resonance energy is completely restricted to a bound state in the radiation continuum at W2 = 150 nm, which are defined as the BICs. (d) Extracted Q-factors of two quasi-BICs for Fano modes respectively.

Download Full Size | PDF

To gain insight into the generated mechanism of two resonances, we analyze the contributions of different multipole components in the far-field radiation. The magnetic field distribution in the grating layer is excited around 1790 nm as shown in Fig. 9(a), which leads to the formation of quasi-BIC3. Different from the TE4,1 mode in TE-polarization, the guiding mode named transverse-magnetic (TM2,1) mode is excited inside the guiding layer in TM-polarization, resulting in the switching from EIT to Fano resonance at the longer wavelengths [59,60]. According to the Ref. [60], the relationship of the propagation constant of the eigenmode is ${\beta _{\textrm{T}{\textrm{E}_{2,1}}}}$>${\beta _{\textrm{T}{\textrm{M}_{2,1}}1}}$>${\beta _{\textrm{T}{\textrm{E}_{4,1}}}}$ above the light line under the same wavelength in the grating, which means that the central wavelength of the resonance dominated by guided modes is ${\lambda _{\textrm{TE}2,1}}$>${\lambda _{\textrm{TM}2,1}}$>${\lambda _{\textrm{TE}4,1}}$ under the same excitation condition (the same grating). Therefore, the TM2,1 mode is excited at a longer wavelength in contrast to the TE4,1 mode. Similar quasi-BIC features are observed in the Fano resonance at the long wavelengths in different polarization directions, which means they originate from the same mechanisms. The electric field distributions clearly display the mode pattern of the quasi-BICs in Fig. 9(b). The quasi-BIC4 is excited around 1968 nm while quasi-BIC2 is excited around 1976 nm. As the order is three, the effective refractive index of TM mode is smaller than the TE mode, so the resonance wavelength of TE2,1 mode is greater than it of TM2,1 mode under the excitation of the same propagation constant. The dominant role of scattering power distribution changes from the ED moment to the MD moment, and the MD moment and EQ moment play a dominant role in forming quasi-BIC3 at W2 = 100 nm, as shown in Fig. 9(b). Meanwhile, it can be seen in Fig. 9(d) that the coupling affects the formation of quasi-BIC4 is still dominated by the TD moment and MQ moment at W2 = 100 nm. Owing to the scattered power difference between the two dominant moments, respectively, the radiation channel is provided, resulting in Fano resonances around 1790 nm and 1967.5 nm.

 figure: Fig. 9.

Fig. 9. (a) Magnetic and (c) electric field distributions in the x-z plane at the resonant wavelength for both quasi-BIC1 and quasi-BIC2 at W2 = 100 nm. (b), (d) Multipole decomposition analysis of the metasurface, which shows the scattered power contribution of different multipoles at W2 = 100 nm.

Download Full Size | PDF

Characterized by the group time delay, the slow light phenomenon accompanied by the spectral response is a significant feature of the EIT resonance response for the metasurfaces. Based on the simulated phase shift $\phi $ spectra, the group delay is described as $G\textrm{ = } - \partial \phi (\omega )/\partial \omega$, where $\phi (\omega )$ and $\omega = 2\pi f$ represent the transmission phase and angular frequency of the light, respectively [61]. Figure 10(a) plots the transmission spectra of the light passing through the metasurface with different structure parameters W2. It can be obviously shown that the large group time delay occurs near the transparent window. When the width of the middle bar is close to W2 = 150 nm gradually, the tunable group delay is greatly enhanced correspondingly in Fig. 10(b). Figure 10(c) shows the wavelength of transmission peaks and corresponding group delays with W2 varying from 145 nm to 150 nm. As W2 is set to 149.5 nm, the maximum group delay can run up to around 2113 ps near the transparency window. In this way, we can control the group delay accompanied by the EIT effect, which can be further investigated for slow light application.

 figure: Fig. 10.

Fig. 10. (a) The transmission spectra and (b) group delays of EIT-like effect under the excitation of TE-polarized incident wave, where the group delay data is plotted by logarithm. (c) The wavelength of transmission peaks and corresponding group delays with W2 varying from 145 nm to 150 nm.

Download Full Size | PDF

For the fabrication of the proposed metasurface, we discuss the possibility of utilizing the standard nanofabrication techniques. The material Si and Si3N4 are used to epitaxially grow the 300 nm Si layer and 150 nm Si3N4 layer on the InP substrate. Then, using electron-beam lithography and reactive ion etching to define the 150 nm grating layer with the help of a mask. The InP substrate can be ultrasonically removed with the help of photolithography and chemical solution. Our structure can be interconnected by a network of supporting bridges without the glass substrate like the Ref. [41].

4. Conclusions

In summary, we present a numerical and theoretical analysis of double resonance-trapped BIC in the form of EIT or Fano resonance in an all-dielectric metasurface. The eigenmodes analysis reveals the existence of resonance-trapped BICs with infinite leakage and the multipole analysis exhibits the dominant pole of the resonance-trapped BICs upon the proper choice of structure parameters. To further insight into the physical mechanism of the two BICs, the coupled harmonic oscillator model and coupled Lorentz oscillator model are employed and the electric distribution is calculated, which reveals the grating can selectively excite different guiding modes in the guiding layer for the realization of the two BICs. In addition, the BIC based on Fano and EIT resonance can simultaneously exist at different wavelengths. Particularly, the group delay in the transparency window of the EIT can effectively reach 2113 ps attributing to the ultrahigh Q-factor resonance caused by the resonance-trapped BICs. Therefore, these findings provide a new feasible platform for slow light applications in the near-infrared region.

Funding

Natural Science Foundation of Guangdong Province (2022A1515010817); National Natural Science Foundation of China (11874019).

Disclosures

The authors declare no conflicts of interest.

Data availability

No data were generated or analyzed in the presented research.

References

1. H. S. Khakestar, Subradiant, superradiant and Fano resonances in dimer and ring/disk nanocavities (Rice University, 2010).

2. H. Schmidt and A. Imamoǧlu, “High-speed properties of a phase-modulation scheme based on electromagnetically induced transparency,” Opt. Lett. 23(13), 1007–1009 (1998). [CrossRef]  

3. Y. Wu, J. Saldana, and Y. Zhu, “Large enhancement of four-wave mixing by suppression of photon absorption from electromagnetically induced transparency,” Phys. Rev. A 67(1), 013811 (2003). [CrossRef]  

4. Y. Zhang, A. W. Brown, and M. Xiao, “Opening Four-Wave Mixing and Six-Wave Mixing Channels via Dual Electromagnetically Induced Transparency Windows,” Phys. Rev. Lett. 99(12), 123603 (2007). [CrossRef]  

5. Y. Mu, L. Qin, Z. Shi, and G. Huang, “Giant Kerr nonlinearities and magneto-optical rotations in a Rydberg-atom gas via double electromagnetically induced transparency,” Phys. Rev. A 103(4), 043709 (2021). [CrossRef]  

6. L. V. Hau, S. E. Harris, Z. Dutton, and C. H. Behroozi, “Light speed reduction to 17 metres per second in an ultracold atomic gas,” Nature 397(6720), 594–598 (1999). [CrossRef]  

7. M. D. Lukin and A. Imamoğlu, “Controlling photons using electromagnetically induced transparency,” Nature 413(6853), 273–276 (2001). [CrossRef]  

8. J. J. Longdell, E. Fraval, M. J. Sellars, and N. B. Manson, “Stopped Light with Storage Times Greater than One Second Using Electromagnetically Induced Transparency in a Solid,” Phys. Rev. Lett. 95(6), 063601 (2005). [CrossRef]  

9. I. Novikova, R. L. Walsworth, and Y. Xiao, “Electromagnetically induced transparency-based slow and stored light in warm atoms,” Laser Photonics Rev. 6(3), 333–353 (2012). [CrossRef]  

10. J. P. Marangos, “Electromagnetically induced transparency,” J. Mod. Opt. 45(3), 471–503 (1998). [CrossRef]  

11. H. Tang, L. Zhou, J. Xie, L. Lu, and J. Chen, “Electromagnetically Induced Transparency in a Silicon Self-Coupled Optical Waveguide,” J. Lightwave Technol. 36(11), 2188–2195 (2018). [CrossRef]  

12. H. Lu, X. Liu, and D. Mao, “Plasmonic analog of electromagnetically induced transparency in multi-nanoresonator-coupled waveguide systems,” Phys. Rev. A 85(5), 053803 (2012). [CrossRef]  

13. X. Yang, M. Yu, D.-L. Kwong, and C. W. Wong, “All-Optical Analog to Electromagnetically Induced Transparency in Multiple Coupled Photonic Crystal Cavities,” Phys. Rev. Lett. 102(17), 173902 (2009). [CrossRef]  

14. 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,” Nature Mater. 8(9), 758–762 (2009). [CrossRef]  

15. J. Gu, R. Singh, X. Liu, X. Zhang, Y. Ma, S. Zhang, S. A. Maier, Z. Tian, A. K. Azad, H.-T. Chen, A. J. Taylor, J. Han, and W. Zhang, “Active control of electromagnetically induced transparency analogue in terahertz metamaterials,” Nat. Commun. 3(1), 1151 (2012). [CrossRef]  

16. S. Xiao, T. Wang, T. Liu, X. Yan, Z. Li, and C. Xu, “Active modulation of electromagnetically induced transparency analogue in terahertz hybrid metal-graphene metamaterials,” Carbon 126, 271–278 (2018). [CrossRef]  

17. S. G. Lee, S. Y. Jung, H. S. Kim, S. Lee, and J. M. Park, “Electromagnetically induced transparency based on guided-mode resonances,” Opt. Lett. 40(18), 4241–4244 (2015). [CrossRef]  

18. G. Li, J. Yang, Z. Zhang, K. Wen, Y. Tao, Y. Han, and Z. Zhang, “Double Spectral Electromagnetically Induced Transparency Based on Double-Bar Dielectric Grating and Its Sensor Application,” Appl. Sci. 10(9), 3033 (2020). [CrossRef]  

19. G. Li, J. Yang, Z. Zhang, Y. Tao, L. Zhou, H. Huang, Z. Zhang, and Y. Han, “Double Electromagnetically Induced Transparency and Its Slow Light Application Based On a Guided-Mode Resonance Grating Cascade Structure,” Materials 13(17), 3710 (2020). [CrossRef]  

20. C. W. Hsu, B. Zhen, A. D. Stone, J. D. Joannopoulos, and M. Soljačić, “Bound states in the continuum,” Nat. Rev. Mater. 1(9), 1–13 (2016). [CrossRef]  

21. R. F. Ndangali and S. V. Shabanov, “Electromagnetic bound states in the radiation continuum for periodic double arrays of subwavelength dielectric cylinders,” J. Math. Phys. 51(10), 102901 (2010). [CrossRef]  

22. J. von Neumann and E. P. Wigner, “Über das Verhalten von Eigenwerten bei adiabatischen Prozessen,” in The Collected Works of Eugene Paul Wigner: Part A: The Scientific Papers, A. S. Wightman, eds. (Springer Berlin Heidelberg, 1993), pp. 294–297.

23. K. Koshelev, G. Favraud, A. Bogdanov, Y. Kivshar, and A. Fratalocchi, “Nonradiating Photonics with Resonant Dielectric Nanostructures,” Nanophotonics 8(5), 725–745 (2019). [CrossRef]  

24. K. Koshelev, A. Bogdanov, and Y. Kivshar, “Meta-optics and bound states in the continuum,” Sci. Bull. 64(12), 836–842 (2019). [CrossRef]  

25. S. Weimann, Y. Xu, R. Keil, A. E. Miroshnichenko, A. Tünnermann, S. Nolte, A. A. Sukhorukov, A. Szameit, and Y. S. Kivshar, “Compact Surface Fano States Embedded in the Continuum of Waveguide Arrays,” Phys. Rev. Lett. 111(24), 240403 (2013). [CrossRef]  

26. M. Odit, K. Koshelev, S. Gladyshev, K. Ladutenko, Y. Kivshar, and A. Bogdanov, “Observation of Supercavity Modes in Subwavelength Dielectric Resonators,” Adv. Mater. 33(1), 2003804 (2021). [CrossRef]  

27. E. Melik-Gaykazyan, K. Koshelev, J.-H. Choi, S. S. Kruk, A. Bogdanov, H.-G. Park, and Y. Kivshar, “From Fano to Quasi-BIC Resonances in Individual Dielectric Nanoantennas,” Nano Lett. 21(4), 1765–1771 (2021). [CrossRef]  

28. K. Koshelev, Y. Tang, K. Li, D. Choi, G. Li, and Y. Kivshar, “Nonlinear metasurfaces governed by bound states in the continuum,” ACS Photonics 6(7), 1639–1644 (2019). [CrossRef]  

29. K. Koshelev, S. Kruk, E. Melik-Gaykazyan, J.-H. Choi, A. Bogdanov, H.-G. Park, and Y. Kivshar, “Subwavelength dielectric resonators for nonlinear nanophotonics,” Science 367(6475), 288–292 (2020). [CrossRef]  

30. H. Can, Z. Chen, X. Shumin, W. Yuhan, F. Yubin, L. Yilin, Z. Nan, Q. Geyang, J. Hongjun, H. Jiecai, G. Li, K. Yuri, and S. Qinghai, “Ultrafast control of vortex microlasers,” Science 367(6481), 1018–1021 (2020). [CrossRef]  

31. M. Wu, S. T. Ha, S. Shendre, E. G. Durmusoglu, W.-K. Koh, D. R. Abujetas, J. A. Sánchez-Gil, R. Paniagua-Domínguez, H. V. Demir, and A. I. Kuznetsov, “Room-Temperature Lasing in Colloidal Nanoplatelets via Mie-Resonant Bound States in the Continuum,” Nano Lett. 20(8), 6005–6011 (2020). [CrossRef]  

32. V. Mylnikov, S. T. Ha, Z. Pan, V. Valuckas, R. Paniagua-Domínguez, H. V. Demir, and A. I. Kuznetsov, “Lasing Action in Single Subwavelength Particles Supporting Supercavity Modes,” ACS Nano 14(6), 7338–7346 (2020). [CrossRef]  

33. Z. Yu, X. Xi, J. Ma, H. Tsang, C.-L. Zou, and X. Sun, “Photonic integrated circuits with bound states in the continuum,” Optica 6(10), 1342–1348 (2019). [CrossRef]  

34. Z. Yu, Y. Tong, H. K. Tsang, and X. Sun, “High-dimensional communication on etchless lithium niobate platform with photonic bound states in the continuum,” Nat. Commun. 11(1), 1–9 (2020). [CrossRef]  

35. F. Wu, J. Wu, Z. Guo, H. Jiang, Y. Sun, Y. Li, J. Ren, and H. Chen, “Giant Enhancement of the Goos-Hänchen Shift Assisted by Quasibound States in the Continuum,” Phys. Rev. Appl. 12(1), 014028 (2019). [CrossRef]  

36. F. Wu, M. Luo, J. Wu, C. Fan, X. Qi, Y. Jian, D. Liu, S. Xiao, G. Chen, H. Jiang, Y. Sun, and H. Chen, “Dual quasibound states in the continuum in compound grating waveguide structures for large positive and negative Goos-Hänchen shifts with perfect reflection,” Phys. Rev. A 104(2), 023518 (2021). [CrossRef]  

37. T. Ning, X. Li, Y. Zhao, L. Yin, Y. Huo, L. Zhao, and Q. Yue, “Giant enhancement of harmonic generation in all-dielectric resonant waveguide gratings of quasi-bound states in the continuum,” Opt. Express 28(23), 34024–34034 (2020). [CrossRef]  

38. K. Koshelev, S. Lepeshov, M. Liu, A. Bogdanov, and Y. Kivshar, “Asymmetric metasurfaces with high-Q resonances governed by bound states in the continuum,” Phys. Rev. Lett. 121(19), 193903 (2018). [CrossRef]  

39. D. R. Abujetas, N. Van Hoof, S. ter Huurne, J. G. Rivas, and J. A. J. O. Sánchez-Gil, “Spectral and temporal evidence of robust photonic bound states in the continuum on terahertz metasurfaces,” Optica 6(8), 996–1001 (2019). [CrossRef]  

40. D. C. Marinica, A. G. Borisov, and S. V. Shabanov, “Bound States in the Continuum in Photonics,” Phys. Rev. Lett. 100(18), 183902 (2008). [CrossRef]  

41. A. Kodigala, T. Lepetit, Q. Gu, B. Bahari, Y. Fainman, and B. J. N. Kanté, “Lasing action from photonic bound states in continuum,” Nature 541(7636), 196–199 (2017). [CrossRef]  

42. C. W. Hsu, B. Zhen, J. Lee, S.-L. Chua, S. G. Johnson, J. D. Joannopoulos, and M. Soljačić, “Observation of trapped light within the radiation continuum,” Nature 499(7457), 188–191 (2013). [CrossRef]  

43. H. Friedrich and D. Wintgen, “Interfering resonances and bound states in the continuum,” Phys. Rev. A 32(6), 3231–3242 (1985). [CrossRef]  

44. J. Niu, Y. Zhai, Q. Han, J. Liu, and B. Yang, “Resonance-trapped bound states in the continuum in metallic THz metasurfaces,” Opt. Lett. 46(2), 162–165 (2021). [CrossRef]  

45. A. Bogdanov, K. Koshelev, P. Kapitanova, M. Rybin, S. Gladyshev, Z. Sadrieva, K. Samusev, Y. Kivshar, and M. Limonov, “Bound states in the continuum and Fano resonances in the strong mode coupling regime,” Adv. Photonics 1(01), 1 (2019). [CrossRef]  

46. M. V. Rybin, K. L. Koshelev, Z. F. Sadrieva, K. B. Samusev, A. A. Bogdanov, M. F. Limonov, and Y. S. Kivshar, “High-Q Supercavity Modes in Subwavelength Dielectric Resonators,” Phys. Rev. Lett. 119(24), 243901 (2017). [CrossRef]  

47. X. Zhao, C. Chen, K. Kaj, I. Hammock, Y. Huang, R. Averitt, and X. Zhang, “Terahertz investigation of bound states in the continuum of metallic metasurfaces,” Optica 7(11), 1548–1554 (2020). [CrossRef]  

48. J. Li, J. Li, C. Zheng, Z. Yue, S. Wang, M. Li, H. Zhao, Y. Zhang, and J. Yao, “Free switch between bound states in the continuum (BIC) and quasi-BIC supported by graphene-metal terahertz metasurfaces,” Carbon 182, 506–515 (2021). [CrossRef]  

49. D. R. Abujetas, J. A. Sánchez-Gil, and J. J. A. P. A. Olmos-Trigo, “Tailoring accidental double bound states in the continuum in all-dielectric metasurfaces,” Adv. Opt. Mater. 2200301(2022).

50. L. Cong and R. Singh, “Symmetry-Protected Dual Bound States in the Continuum in Metamaterials,” Adv. Opt. Mater. 7(13), 1900383 (2019). [CrossRef]  

51. I. Volkovskaya, L. Xu, L. Huang, A. I. Smirnov, A. E. Miroshnichenko, and D. Smirnova, “Multipolar second-harmonic generation from high-Q quasi-BIC states in subwavelength resonators,” Nanophotonics 9(12), 3953–3963 (2020). [CrossRef]  

52. D. Abujetas, A. Barreda, F. Moreno, A. Litman, G. Jean-Michel, and J. Sánchez-Gil, “High-Q Transparency Band in All-Dielectric Metasurfaces Induced by a Quasi Bound State in the Continuum,” Laser Photonics Rev. 15(1), 2000263 (2021). [CrossRef]  

53. E. J. Prucha and E. D. Palik, “Handbook of Optical Constants of Solids,” (Academic, 1985)

54. T. Shi, Z.-L. Deng, Q.-A. Tu, Y. Cao, and X. J. P. Li, “Displacement-mediated bound states in the continuum in all-dielectric superlattice metasurfaces,” PhotoniX 2(1), 1–10 (2021). [CrossRef]  

55. S. A. Maier, Plasmonics: fundamentals and applications (Springer, 2007).

56. M. Kim, E. S. Hwang, O. Prudnikov, and B.-H. Cheong, “Guided-mode waves structure of electric and magnetic dipole resonances in a metamaterial slab,” Opt. Express 29(19), 30610–30624 (2021). [CrossRef]  

57. F. Deng, H. Liu, L. Xu, S. Lan, and A. Miroshnichenko, “Strong Exciton–Plasmon Coupling in a WS2 Monolayer on Au Film Hybrid Structures Mediated by Liquid Ga Nanoparticles,” Laser Photonics Rev. 14(4), 1900420 (2020). [CrossRef]  

58. Z. Chaobiao, S. Li, M. Fan, X. Wang, X. Yanli, W. Xu, S. Xiao, M. Hu, and J.-T. Liu, “Optical radiation manipulation of Si-Ge2Sb2Te5 hybrid metasurfaces,” Opt. Express 28(7), 9690–9701 (2020). [CrossRef]  

59. D. Shin, S. Tibuleac, T. Maldonado, and R. Magnusson, “Thin-film optical filters with diffractive elements and waveguides,” Opt. Eng. 37(9), 2634–2646 (1998). [CrossRef]  

60. C. J. Chang-Hasnain and W. Yang, “High-contrast gratings for integrated optoelectronics,” Adv. Opt. Photon. 4(3), 379–440 (2012). [CrossRef]  

61. T. Yang, X. Liu, and J. Zhou, “Terahertz Polarization Conversion in an Electromagnetically Induced Transparency (EIT)-Like Metamaterial,” Ann. Phys-Berlin. 533(3), 2000528 (2021). [CrossRef]  

Data availability

No data were generated or analyzed in the presented research.

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 (10)

Fig. 1.
Fig. 1. Schematic of the unit cell of the all-dielectric metasurface structure under TE-polarized illumination. (a) The unit cell is arranged with a period Px = 1300 nm, Py = 1050 nm. The Si grating height H = 150 nm, the Si guiding layer thickness D = 150 nm, and the Si3N4 substrate thickness T = 150 nm. (b) Top view of the unit cell. The length of the Si-bars L = 500 nm. Meanwhile, W1 = 150 nm and W2 represent the width of the Si-bars on both sides and in the middle respectively.
Fig. 2.
Fig. 2. (a) Band structure of the metasurface and (b) the corresponding electromagnetic field distribution of eigenmodes at the $\mathrm{\Gamma }$ point. The structural parameters are: W1= W2 = 150 nm, T = D = H = 150 nm, L = 500 nm, Px = 1300 nm, Py = 1050 nm.
Fig. 3.
Fig. 3. The transmission spectra and eigenmode analysis of a metasurface under TE-polarization incident light. (a) Evolution of the transmission spectra with different W2. Transmission spectra and eigenmodes spectra for (b) BIC1 radiating as EIT resonance and (c) BIC2 radiating as Fano resonance with error bars represent the magnitude of the mode inverse radiation lifetime. All the resonance energy is completely restricted to a bound state in the radiation continuum at W2 = 150 nm, which are defined as the BICs. When the linewidths appear in the spectra (W2≠150 nm), they collapse into quasi-BICs.
Fig. 4.
Fig. 4. Extracted Q-factors of two quasi-BICs for (a) EIT and (b) Fano mode.
Fig. 5.
Fig. 5. Electric field distributions in the x-z plane at the resonant wavelength. (a), (b) for two resonant dips of quasi-BIC1 at W2 = 100 nm, (c) for BIC1 in the dip of the transmission spectra at W2 = 150 nm, and (d) for the resonant dip of quasi-BIC2 at W2 = 100 nm. The white arrows indicate the magnetic field directions.
Fig. 6.
Fig. 6. (a)-(e) The numerical transmission spectra (black dots) and corresponding analytically fitted (red curves) transmission spectra with the error bars of the proposed metasurface with different W2. (f) The fitting parameters $\delta $, $\kappa $, ${\gamma _1}$ and ${\gamma _2}$ of coupled two-oscillator model with the variation of W2.
Fig. 7.
Fig. 7. Multipole decomposition analysis of the metasurface of (a) quasi-BIC1 and (b) quasi-BIC2 when W2 = 100 nm, respectively.
Fig. 8.
Fig. 8. The transmission spectra and eigenmode analysis of the metasurface under TM-polarization incident light. (a) Evolution of the transmission spectra with different W2. Transmission spectra and eigenmodes spectra for (b) BIC3 and (c) BIC4 radiating as Fano resonances with error bars represent the magnitude of the mode inverse radiation lifetime. All the resonance energy is completely restricted to a bound state in the radiation continuum at W2 = 150 nm, which are defined as the BICs. (d) Extracted Q-factors of two quasi-BICs for Fano modes respectively.
Fig. 9.
Fig. 9. (a) Magnetic and (c) electric field distributions in the x-z plane at the resonant wavelength for both quasi-BIC1 and quasi-BIC2 at W2 = 100 nm. (b), (d) Multipole decomposition analysis of the metasurface, which shows the scattered power contribution of different multipoles at W2 = 100 nm.
Fig. 10.
Fig. 10. (a) The transmission spectra and (b) group delays of EIT-like effect under the excitation of TE-polarized incident wave, where the group delay data is plotted by logarithm. (c) The wavelength of transmission peaks and corresponding group delays with W2 varying from 145 nm to 150 nm.

Equations (12)

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

( E g r a i γ g r a / 2 g g E g u i i γ g u i / 2 ) ( α β ) = ± E ( α β ) ,
E ± = ( E g r a + E g u i ) / 2 ± 4 g 2 + ( δ i 2 ( γ g r a γ g u i ) ) 2 / 2 ,
x ¨ 1 + γ 1 x ˙ 1 + ω 0 2 x 1 + κ x 2 = E ,
x ¨ 2 + γ 2 x ˙ 2 + ( ω + δ ) 2 x 2 + κ x 1 = 0.
χ ( ω ) = χ r + i χ i ( ω ω 0 δ ) + i γ 2 2 ( ω ω 0 + i γ 1 2 ) ( ω ω 0 δ + i γ 2 2 ) κ 2 4 .
T ( ω ) = 1 g χ i ( ω ) ,
P = 1 i ω j d 3 r ,
M = 1 2 c ( r × j ) d 3 r ,
T = 1 10 c [ ( r j ) r 2 r 2 j ] d 3 r ,
Q α β ( e ) = 1 2 i ω [ r α j β + r β j α 2 3 ( r j ) δ α , β ] d 3 r ,
Q α β ( m ) = 1 3 c [ ( r × j ) α r β + ( r × j ) β r α ] d 3 r ,
I P = 2 ω 4 3 c 3 | P | 2 , I M = 2 ω 4 3 c 3 | M | 2 , I T = 2 ω 6 3 c 5 | T | 2 , I Q ( e ) = ω 6 5 c 5 | Q α β ( e ) | 2 , I Q ( m ) = ω 6 40 c 5 | Q α β ( 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.