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

Active quasi-BIC metasurfaces assisted by epsilon-near-zero materials

Open Access Open Access

Abstract

Active devices play a critical role in modern electromagnetic and photonics systems. To date, the epsilon ($\varepsilon$)-near-zero (ENZ) is usually integrated with the low Q-factor resonant metasurface to achieve active devices, and enhance the light-matter interaction significantly at the nanoscale. However, the low Q-factor resonance may limit the optical modulation. Less work has been focused on the optical modulation in the low-loss and high Q-factor metasurfaces. Recently, the emerging optical bound states in the continuum (BICs) provides an effective way for achieving high Q-factor resonators. In this work, we numerically demonstrate a tunable quasi-BICs (QBICs) by integrating a silicon metasurface with ENZ ITO thin film. Such a metasurface is composed of five square holes in a unit cell, and hosts multiple BICs by engineering the position of centre hole. We also reveal the nature of these QBICs by performing multipole decomposition and calculating near field distribution. Thanks to the large tunability of ITO’s permittivity by external bias and high-Q factor enabled by QBICs, we demonstrate an active control on the resonant peak position and intensity of transmission spectrum by integrating ENZ ITO thin films with QBICs supported by silicon metasurfaces. We find that all QBICs show excellent performance on modulating the optical response of such a hybrid structure. The modulation depth can be up to 14.8 dB. We also investigate how the carrier density of ITO film influence the near-field trapping and far-field scattering, which in turn influence the performance of optical modulation based on this structure. Our results may find promising applications in developing active high-performance optical devices.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Active optical modulation plays an important role in developing high-performance reconfigurable nanophotonic-devices [13]. The epsilon ($\varepsilon$)-near-zero (ENZ) materials, whose real part of the dielectric constant is zero or near-zero at specific wavelengths, becomes an exciting platform to achieve the tunability of the device. Due to the fascinating capabilities, nanodevices integrating ENZ materials exhibiting interesting optical coupling effects has enabled many advanced optical applications [4]. Indium tin oxide (ITO) is one of the most widely used ENZ materials, many potential applications of ITO have been proposed or demonstrated by harnessing the ENZ behavior of ITO [4,5]. The ENZ wavelength of ITO is closely related to the electron density and can be obtained by applying an electrostatic field [68] or optical pumping [9,10] to tune the electron density, which can change its dielectric constant in the near-infrared state from positive (dielectric-like) to negative (metal-like) [11,12]. When the dielectric constant of the ITO material is close to zero, which is the ENZ state, the intrinsic absorption loss inside the ITO material can be effectively changed. This may open the door of realizing strong electro-optical effects [1317]. ITO is often combined with metal nanocavities for efficient optical modulation. For example: Alam et al. exploited an optical metasurface coupled with the dipole resonant mode of the gold antenna and the ENZ mode of the ITO thin film, thereby exhibiting a broadband and ultrafast nonlinear response [18]. Kuttruff et al. exploited linear photon absorption of ENZ-mode modulation systems of metal-insulator-metal (MIM) nanocavities, and achieved nondegenerate all-optical ultrafast modulation [19]. Tao et al. realized an electrically tunable ultra-compact plasmonic modulator with high modulation intensity using plasmonic-induced transparent structures [20]. Guo et al. proposed a hybrid metasurface composed of indium tin oxide (ITO) nanolayers and plasmonic metasurfaces, which can support two different coupling modes to realize double transparent windows in the communication frequency band [21]. Xie et al. proposed a tunable optical switch based on an ENZ metasurface, which can function as an electro-optical switch and an all-optical switch [22]. However, metal metasurfaces have large ohmic losses in the visible and near-infrared range, so metal devices based on localized surface plasmon resonance have lower Q-factor, and the optical modulation based on hybrid plasmonic metasurfaces and ENZ materials has been limited.

In recent years, dielectric metasurfaces have received extensive attention due to their low loss, CMOS-compatible fabrication process, and abundant resonant modes, providing a new platform for the realization of high Q-factor [2327]. The perturbation of optical bound states in the continuum (BICs) is a simple and effective method to obtain high Q-factor resonators [2841]. BIC has an infinite Q-factor, and is not coupled to free space [42,43]. It usually appears at highly symmetric dielectric metasurfaces such as lattice with $C_{\rm 4v}$ group symmetry. When perturbing the symmetry of the structure, the higher order group $C_{\rm 4v}$ evolves into lower order group $C_{\rm s}$, which converts non-radiative optical BICs modes into quasi-BICs (QBICs) [4448]. The QBICs in metasurface can be usually accompanied by the generation of high Q-factor Fano resonances, which can greatly enhance the light-matter interaction. For example: QBICs can reduce the threshold of the laser [49], improve the nonlinear conversion efficiency [5054], realize the high-efficiency electro-optic modulators [55,56], improve the sensitivity of optical sensors [57,58], etc. Moreover, the Q-factor of QBIC in the dielectric metasurface can be arbitrarily regulated by the asymmetric parameter, which can control the light-matter interaction strength. Among the BIC family, the symmetry-protected BIC are is the most common and are is easy to realize experimentally. For practical application, one should break the unit cell’s symmetry by removing or adding part of the geometric structure in the metasurface to transform such a BIC into a QBIC with a high Q-factor. Some typical examples of broken-symmetry in unit cell include split rings [59], asymmetric nanorods [52,60], notched cubes [61,62] and nanodisks [39,63]. It is worth noting that the smaller the asymmetric parameter, the larger the Q-factor. However, it is very challenging to fabricate such nanostructures with an ultra-small asymmetry parameter in the experiment due to the limitation of fabrication, thus limiting the actual Q-factor of QBIC. In addition, to date, there are few studies on active control on QBICs by either optical pumping, thermal-tuning or nano-electromechanical tuning [56,6468]. In real applications, it is also highly desired to achieve tunable QBICs by electrical gating. Given the salient property of ENZ ITO that can be controlled by external bias, a large spatial modulation and fast response should be expected by integrating ENZ ITO thin film with dielectric metasurface supporting QBICs.

In this work, we demonstrate an active control on the QBICs with ENZ materials. We consider a silicon metasurface comprising five square holes in a unit cell. Such a metasurface hosts multiple BICs realized by moving the position of center hole, where the disturbance parameter of QBICs can be accurately controlled in the experiment by this way. By integrating this metasurface with an ENZ ITO film, we demonstrate that multiple QBICs can be tuned by the carrier density of ITO film, evidenced by transmission spectra, far-field radiation, and near-field distributions. Thanks to the high-Q nature of QBICs, the transmission intensity can be modulated from 0 to >90% at certain wavelength by controlling electron density of ITO, corresponding to modulation depth of 14.8 dB. Our results may pave the way toward developing high-performance spatial optical modulators.

2. Results and discussion

2.1 Multiple QBICs excited by Si metasurfaces

The Si metasurface is composed of a periodic array of Si thin films embedded with five compound square air-holes, as shown in Fig. 1(a). The top view of unit cell is shown in Fig. 1(b), the five square air-holes are marked as H1, H2, H3, H4 and H5, respectively. If we set the centre of unit cell as origin, the coordinates of these four holes are H1=(-200 nm, 200 nm), H2=(200 nm, 200 nm), H3=(-200 nm, -200 nm) and H4=(200 nm, -200 nm). The period of the device is $P_x=P_y=600$ nm, the side length of the square hole is $l=120$ nm, the depth of the hole is 220 nm, and the substrate is ${\rm SiO_2}$. When H5 is located at the center of the device, namely H5(0, 0), the device satisfies the $C_{\rm 4v}$ rotational symmetry group, and there are multiple non-radiative modes, also known as BICs. When the air-hole H5 is moved up and down along the $y$-axis, the high-order group $C_{\rm 4v}$ degenerates into the low-order group $C_{\rm s}$, thus opening the radiation channel, so that these non-radiative BICs transform into leaky mode quasi BICs. In addition, when the device parameters are changed, interference cancellation occurs between the multi-dipoles, which drastically reduces the far-field radiation of light, and another BICs mode appears. The complex eigenfrequency ($\omega =\omega _0-i\gamma$) can be calculated by commercial software Comsol Multiphysics based on the finite-element method (FEM) (Study$\to$Eigenvalue or Eigenfrequency). The Q-factor is calculated by $Q=\omega _0/2\gamma$. The transmission spectra, field distributions, scattered power of different multipoles, and average field enhancement factor are calculated by the finite difference-time-domain method (Lumerical-Ansys FDTD Solution). In the simulation, the $x$- and $y$-directions are set as periodic boundary conditions, the $z$-direction is set as a perfectly matched layer boundary condition, the $x$-polarized incident light is perpendicular to the metasurface along the $z$-axis, and the optical constants of Si and ${\rm SiO_2}$ are taken from the Palik Handbook [69].

 figure: Fig. 1.

Fig. 1. (a) Schematic diagram of the nanostructure array. (b) Geometry of the unit cell, the off-set distance of the air-hole H5 from the center of unit cell is marked $d$. (c) Transmission spectra for different $d$ values. The three modes are marked with Mode I, Mode II, and Mode III, respectively, and A, B, C, and D corresponds to the four BICs, respectively. (d), (e), (f) Q-factors of the three modes as the function of off-set distance $d$, respectively.

Download Full Size | PDF

When H5 is moved along the $y$-axis, the two symmetry-protected quasi-BICs are excited because the in-plane mirror symmetry is broken with respect to $xoz$ plane. In Fig. 1(c), we calculate the transmission spectra of metasurface when H5 is located at different off-set distances $d$. When $d$=0, there is only one leaky mode (Mode II) at the wavelength of 1455.5 nm. As $d$ increases, the symmetry of the structure is broken, and two leaky modes appear on both sides of Mode II, which we mark as Mode I and Mode III, they are governed by symmetry-protected BIC. Figure 2 shows the electric field distribution of three modes at different off-set distances. Interestingly, the evolution of these two modes differs greatly when $d$ increases gradually. For Mode III, it always conforms to the rules of normal symmetry-protected BIC, the larger the perturbation parameter, the wider the spectra, thus the smaller the Q-factor. However, for Mode I, as $d$ increases, the resonance spectrum first gradually broadens, then narrows, and then broadens again. The resonance spectrum disappears at $d$=70 nm, implying that it is a non-radiative BIC mode but different from symmetry-protected BIC, which is formed by completely destructive interference between multiple dipoles. Therefore, Mode I can be evolved into two types of BICs as $d$ varies. For Mode II, it can be also evolved into BIC when center hole H5 moves along $y$ axis. When $d$=0, Mode II has a finite Q-factor and manifests itself as a Fano resonance in transmission spectrum. As $d$ increases, the resonance linewidth becomes narrower. At $d$=86.5 nm, the resonance disappears, then reappears and gradually broadens. The vanished resonance linewidth marks the appearance of BICs. The same BIC can be found if H5 is moved in the opposite direction. In addition, the same optical response can be observed when H5 is moved along the $x$-axis under $y$-polarized light.

 figure: Fig. 2.

Fig. 2. Electric field distribution and field vector distribution at the $x-y$ plane for the three modes, (a)-(c) $d=30$ nm, (d)-(f) $d=170$ nm. Here, the vector arrows are normalized.

Download Full Size | PDF

To prove that multiple BICs indeed exist at these critical positions, we calculate the Q-factors of each mode under different $d$, respectively. For free-standing metasurface, the Q-factors of the four BICs are all over $10^9$ (see Fig. 7). The Q-factor can reach infinity for the extremely enhanced numerical resolution, confirming the existence of these BICs. The Q-factors of accidental BICs (Modes I and II) for a photonic crystal slab embedded in a symmetric environment with respect to $z$=0 plane are calculated (see Supplement 1, Fig. S1), which is set as ${\rm SiO_2}$. Indeed, the Q-factor of accidental BICs (B and C) can be Q>$10^8$. If only the bottom ${\rm SiO_2}$ substrate is introduced, the Q-factor of accidental BICs will drop significantly due to the out-plane mirror symmetry-breaking, as shown in Figs. 1(d)-(f). The larger the refractive index of substrate, the smaller the Q-factor of accidental BICs. In a word, to achieve an ultrahigh Q-factor of accidental BICs, it is required to create a symmetry environment. It is noted that the other two BICs (A and D) are maintained because they are normal symmetry-protected BICs which are not influence by the substrate. For Mode I, there is an infinite Q-factor when $d$=0. As $d$ increases, the Q-factor first decreases gradually, and then increases to maximum $1.48\times 10^6$ at $|d|$=68.5 nm ($2.8\times 10^9$ for free-standing metasurface). Further increase $d$ results in the decreased Q-factor. The maximum of Q-factor is also confirmed by transmission spectra in Fig. 1(c). Similar trend is expected when H5 moves in an opposite direction. For Mode II, when $d$=0, it has a moderate Q-factor, and with the increase of $d$, the Q-factor gradually increases, and a high Q-factor appears at $|d|$=86.5 nm, and its Q-factor is up to $4.2\times 10^6$. Then, the Q-factor gradually decreases again, agreeing with transmission spectra evolution in Fig. 1(c). It is noted that BICs B and C are accidental BICs, which arise from the destructive interference of the multipoles response. For Mode III, as $d$ increases, the Q-factor decreases, it is an ordinary symmetry-protected BIC. Therefore, the Q-factor of three modes and transmission spectra at different $d$ confirm each other. It is noted that the Q-factor of the quasi symmetry-protected BIC satisfies the formula $Q\propto \alpha ^{-2}$, $\alpha$ is the asymmetry degree [30]. In this work, $\alpha =2d/P$. For Mode I, the Q-factor is proportional to $\alpha$ within the quasi symmetry-protected BIC range. In the accidental BIC region, the inverse quadratic law for the Q-factor is not satisfied. For Mode III, there exists only symmetry-protected BIC, and the inverse quadratic law for the Q-factor of the quasi-BIC is satisfied (Supplement 1, see Fig. S2).

Next, we discuss the local field properties of the above three resonance. The calculation results are exhibited in Fig. 2, showing the electric field intensity and field vector distribution at the resonance wavelength of the three QBICs for $d$=30 nm and 170 nm. Here, the color scale corresponds to the field intensity, the black dashed area represents the square air-hole. The arrows describe the field vector distribution in the $x-y$ plane at $z$ = 110 nm. For Mode I, as shown in Fig. 2(a), it can be observed that there is an obvious electric quadrupole distribution in the $x-y$ plane. The electric field is mainly distributed around the square air-holes H3 and H4. In Fig. 2(d), for $d$=170 nm, the electric field distribution and field vector distribution have changed significantly, and the electric field is mainly distributed inside of the square air-hole and its edge area. In the region below the air-holes H3 and H4, the electric field vector exhibits an obvious circular distribution, so the magnetic dipole of $z$-direction dominates the response. Here, the near-field distribution is also greatly affected by other dipoles, and each dipole moment interferes each other, so the total field distribution is the result of the interaction of multiple dipoles. The significant multipole interference induces the non-radiative BIC mode. For Mode II, when $d=30$ nm and $d=170$ nm, as shown in Figs. 2(b) and (e), the distinct circular electric field distribution is located the region of below H5, exciting the magnetic dipoles oscillation along the $z$-direction. The electric field is mainly distributed in the square air-hole and its edge area, showing slight differences, and they are almost mirror-symmetrical along the $y$-axis. It is suggested that, with the change of structural parameters, Mode II are relatively stable. For Mode III, as shown in Figs. 2(c) and (f), when $d=30$ nm, a very obvious circular electric field distribution appears in the H5 region of the $x-y$ plane, which excites the magnetic dipole along the $z$ direction, the electric field mainly distributed in the square air holes H1, H2 and H3 and their edge areas. When $d=170$ nm, two opposite circular electric field distributions appear at the $x-y$ plane, and excite a circular magnetic field distribution on the $x-z$ plane, which eventually leads to a electric toroidal dipole response in the $x$-direction. We perform an in-depth discussion on the near-field distribution characteristics of the three QBICs, and these results are mutually confirmed with the results of multipole decomposition. It is noted that the eigenfield fields of the BIC mode (A, D) at $d$=0 are highly symmetrically distributed, corresponding to non-degenerate modes, which do not couple to the normal incident light, as shown in Supplement 1, Fig. S3.

In order to reveal the nature of these QBICs, we perform the multipole decomposition on the structure [7072]. For Mode I, when $d=30$ nm, as shown in Fig. 3(a), the far-field scattering power of the nanostructure at the resonance wavelength is dominated by the electric quadrupole (EQ) at the $x-y$ plane. Also, a mild contribution comes from the magnetic dipole (MD). When $d=170$ nm, as shown in Fig. 3(d), the MD response along the $z$-direction dominates this mode, followed by EQ, the other dipole radiations are suppressed. With the increase of parameter $d$, the dominant multipole of Mode I evolves from EQ to MD. These results are consistent with the field distribution in Figs. 2(a) and (d). For Mode II, as shown in Figs. 3(b) and (d), when $d=30$ nm and 170 nm, the MD response in the $z$-direction is always dominant for this mode. Therefore, Mode II is relatively stable, the result is in good agreement with the field distribution in Figs. 2(b) and (e). For Mode III, as shown in Figs. 3(c) and (f), when $d=30$ nm, the dominant far-field scattering is MD along the $z$-direction, followed by toroidal dipole (TD). For $d=$170 nm, the strongest contribution to the resonances is provided by the TD along $x$-direction. As the parameter $d$ changes, the dominant dipole evolves from MD to TD. This result is consistent with the one in Figs. 2(c) and (f). Therefore, when the perturbation parameter is changed, the coupling between the multiple dipoles will change significantly, resulting in an obvious change in the resonance properties of Mode I and Mode III. Among them, the dominant contribution of Mode I evolves from EQ to MD, and Mode III evolves from MD to TD. Mode II is relatively stable and is always the MD response.

 figure: Fig. 3.

Fig. 3. Far-field scattered power and dominant multipole components of the multipole decomposition of Mode I, Mode II and Mode III; (a), (b), (c) are $d$=30 nm (d), (e ), (f) corresponds to $d$=170 nm. ED, MD, TD, EQ, MQ correspond to the response of electric dipole, magnetic dipole, toroidal dipole, electric quadrupole, and magnetic quadrupole, respectively.

Download Full Size | PDF

2.2 Active modulation of QBICs metasurface by ENZ materials

After establishing the knowledge of multiple BICs supported by such a metasurface, we move to investigate the optical modulation of the metasurface by introducing ENZ material ITO. The ITO film is sandwiched between the Si metasurface and ${\rm SiO_2}$ substrate, as shown in Fig. 4(a). The off-set distance of air-hole H5 is $d=$170 nm, the thickness of the ITO film is 10 nm, such the total hole depth is 230 nm.

In the visible and near-infrared regions, the dielectric constant of ITO can be described by the Drude model [9,73]:

$$\varepsilon=\varepsilon_{\infty}-\frac{\omega_p^2}{\omega^2+i\Gamma\omega},$$
$$\omega_p=\sqrt{\frac{Ne^2}{m_e^*\varepsilon_0}},$$
here, $\varepsilon$ is the permittivity, $\omega$ is the angular frequency of the incident light, $\omega _p$ is the plasma frequency, and $\Gamma$ is the Drude damping rate. It is worth mentioning that the plasma frequency is a function of the carrier density $N$, and $\Gamma =0.0468\omega _p$. $m_e^*$ is the effective mass of the electron, $e$ is the charge of the electron, and $\varepsilon _0$ is the vacuum permittivity, $\varepsilon _{\infty }=3.8055$ [9]. As illustrated in Figs. 4(b) and (c), the real and imaginary parts of the dielectric constant of ITO at different N values are calculated, the green dotted line shows the change of ENZ wavelength with $N_{\rm ITO}$, and the ENZ wavelength is blue-shifted as $N_{\rm ITO}$ increases. When $d$ = 170 nm, the resonance spectra and electric field distributions of the nanostructure without and with ITO film are calculated, as shown in Figs. 4(d)-(f). The resonance wavelengths of the three modes are 1373 nm, 1440 nm and 1528 nm without ITO case. Herein, we first consider that the ENZ wavelength is equal to the resonance wavelength of the three modes (see Fig. 8). It can be clearly seen that the transmission spectra of the three modes exhibit different degrees of red shift, and the electric field distribution does not change significantly, while the intensity becomes smaller, which can be ascribed to the energy leakages into the ITO layer. In addition, the extinction ratio shows different degree of reduction due to the introduction of loss.

 figure: Fig. 4.

Fig. 4. (a) Schematic diagram of the Si-ITO metasurface. The thickness of Si is 220 nm, the thickness of ITO is 10 nm. (b), (c) The changes of the real and imaginary parts of the dielectric constant of ITO with the current density $N_{\rm ITO}$, and the green dotted line is the corresponding ENZ position. (d), (e), and (f) are the transmission spectra of the metasurface without and with ITO film, respectively. The inset is the electric field distribution on the $x-y$ plane at the resonance wavelength, and $N_{\rm ITO}$ corresponds to the overlap between the resonance wavelength and the ENZ wavelength.

Download Full Size | PDF

Next, we show how such an ENZ material combining with QBIC helps to modulate the transmission spectra. The ENZ wavelength will change significantly when the carrier density $N_{\rm ITO}$ is changed, and the resonance mode will be affected. Here, in order to achieve an excellent optimal modulation, we manipulate the three modes respectively, as shown in Fig. 5. As $N_{\rm ITO}$ increases, the resonances are all blue-shifted due to the decrease of the effective refractive index of the nanostructures. As demonstrated above, the Q-factor and field distribution of the three modes are different under the same parameters, thus the influence of different $N_{\rm ITO}$ on the transmission amplitude and resonance linewidth is different, as illustrated in Figs. 5(a)-(c). We evaluate the modulation depth by the following formula [74] :

$$ER=10{\rm log}\frac{T_1}{T_0},$$
here, $T_0$ and $T_1$ are the original and modulated transmission spectra, respectively. The calculated extinction ratio (ER) is shown in Fig. 5. For Mode I, as $N_{\rm ITO}$ increases, the ER increases, and a larger modulation depth of 10.8 dB is obtained at the wavelength of 1362.6 nm. For mode II, the initial resonance spectrum is wider and the extinction ratio is low. With the decrease of $N_{\rm ITO}$, the spectral line is narrower and the extinction ratio becomes lager, and the modulation depth reaches 7.9 dB at 1444.3 nm. Mode III, holes a relatively wider the resonance spectrum, its modulation depth reaches 14.8 dB at 1514.3 nm. Therefore, by tuning the current density $N_{\rm ITO}$ of ITO, the transmission spectrum of spatial light can be significantly modulated in amplitude and resonance position, and the modulation characteristics are remarkably different for different resonance modes.

 figure: Fig. 5.

Fig. 5. Modulation of transmission spectrum with different $N_{\rm ITO}$ values. (a)-(c) The transmission spectra and corresponding modulation depths of Mode I, Mode II, and Mode III under different $N_{\rm ITO}$, respectively.

Download Full Size | PDF

Above we explored the modulation of the transmission for the three QBIC modes, and now we calculate the corresponding total far-field scattering and near-field light trapping modulation. Figures 6(a)-(c) show the normalized scattering intensity of three QBICs for different $N_{\rm ITO}$, it can be clearly seen that the three modes are significantly modulated in the scattering peak position and intensity. In addition, the corresponding contributions of multiple dipoles are shown in Fig. 9, and the different $N_{\rm ITO}$ value does not affect the dominant dipole response.

 figure: Fig. 6.

Fig. 6. (a)-(c) QBICs normalized far-field scattering modulation and (d)-(f) near-field capture modulation for different $N_{\rm ITO}$ values.

Download Full Size | PDF

The QBICs can generate huge local field enhancement within nanostructures. We finally investigate the modulation of light trapping under different $N_{\rm ITO}$ values. It is calculated by integrating the enhanced electric field intensity within the nanostructure and normalizing its incident intensity within the same volume. The calculated average field enhancement factor (EF) is shown in the following formula [75,76]:

$$EF=\frac{\iiint\left|E\right|^{2}dV}{\left|E_{0}\right|^{2}V},$$
$V$ is the volume of the Si-ITO metasurface nanostructure, $|E_0|$ and $|E|$ are the amplitude of the incident electric field and electric-field intensity of resonator, respectively. The calculated EF is shown in Figs. 6(d)-(f). Note that the maximum of the EF of these three modes are located at the resonance wavelengths. For Mode I and Mode III, with increasing the current density $N_{\rm ITO}$, the EF peaks are blue-shifted, but the EF maximum does not change significantly. While at specific wavelengths, EF intensity is modulated dramatically, thus we can obtain the strong local field at the desired wavelength by tuning $N_{\rm ITO}$. For Mode II, because it has a higher Q-factor, the modulation of $N_{\rm ITO}$ on EF is reflected in both wavelength position and spectrum intensity. The results show that when $N_{\rm ITO}=2.28\times 10^{21}$ ${\rm cm}^{-3}$ and $N_{\rm ITO}=1.04\times 10^{21}$ ${\rm cm}^{-3}$, the maximum EF are 39 and 101, respectively. The field distributions also confirm this modulation effect (see Supplement 1, Fig. S4). Therefore, the intensity modulation of high-Q resonances is more significant. When adjusting the asymmetric parameters of the metasurface, an arbitrary Q-factor can be obtained, therefore QBICs exhibit significant advantages in light modulation [77]. In addition, the light modulation is also discussed when ITO is not etched in Appendix D.

The trapping and manipulation of light at the nanoscale are the important aim of nanophotonics. Enhancing the dielectric constant contrast between different materials serves as an effective way of controlling the electromagnetic field at the nanoscale. The ITO is one of the most widely used ENZ materials. Its carrier density can be effectively modified with an external bias. ENZ can be easily realized by controlling its carrier density, inducing large variations of its optical properties, for which near-unity modulation, as well as short, subpicosecond, modulation times can be achieved. In contrast to the other active materials commonly used to achieve the active optical modulation, such as phase change materials [75,78,79] and two-dimensional materials graphene [80], the dielectric constant of ENZ material easily reaches zero, and it holds a relatively large dielectric constant contrast when adjust the carrier density, which can lead to a more significant optical modulation.

3. Conclusion

In summary, we investigated the multi-BIC modes in Si nanostructures, and discussed the modulation of the transmission, far-field scattering, and near-field trapping of QBICs based on ENZ (ITO) materials. We found that by changing the relative psitions of the hole in a highly symmetric composite air-hole nanostructure, silicon metasurfaces with unit cell consisting of five square holes in silicon slab host multiple BICs. The nature of these QBICs properties is fully discussed through the near field distribution and the far-field scattering contribution of multipole via multipole decomposition. By introducing ENZ material (ITO) thin film, we systematically investigated the modulation properties of the hybrid Si-ITO metasurface. The modulation of transmission, normalized far-field scattering intensity, and near-field enhancement factor of the three QBICs modes are calculated by changing current density $N_{\rm ITO}$. The different light modulation characteristics are shown for different resonance modes. It is noted that the disturbance parameter of these QBICs can be accurately controlled in experiment for this design, which provides a way for the realization of an ultrahigh Q-factor. Our results pave a solid avenue to achieve the high-performance light modulation devices.

Appendix

A. Q-factors of free-standing metasurfaces

 figure: Fig. 7.

Fig. 7. (a) Schematics of unit cell for free-standing metasurfaces. (b)-(d) The Q-factors of Modes I, II and III with respect to the offset distance, respectively.

Download Full Size | PDF

B. Permittivity of ITO film

 figure: Fig. 8.

Fig. 8. Real (solid) and imaginary (dashed) parts of the permittivity of ITO as a function of wavelength. Points A, B, and C represent zero-crossing at $\lambda _{\rm ENZ}=1373$ nm, $\lambda _{\rm ENZ}=1440$ nm, and $\lambda _{\rm ENZ}=1528$ nm, respectively.

Download Full Size | PDF

C. Far-field scattering power of multipole for ITO with different current densities

 figure: Fig. 9.

Fig. 9. Multipole contributions of Si-ITO metasurface from three QBICs at different current densities $N_{\rm ITO}$ at $d=$170 nm.

Download Full Size | PDF

D. Optical modulation of Si-ITO metasurfaces without etching air-holes in ITO films

In the main text, we calculate the modulation properties of transmission spectra, far-field scattering intensity, and near-field trapping of the Si-ITO metasurface with etched air-hole in ITO layer. Here, we discuss the optical modulation properties of Si-ITO metasurfaces without etched air-holes in ITO film. The calculated results are shown in Fig.  10. Similar modulation performance is shown, and the modulation effect is slightly worse than the case where ITO is etched with air-holes.

 figure: Fig. 10.

Fig. 10. (a)-(c) Transmission spectra and (a)-(c) corresponding modulation depths under different $N_{\rm ITO}$ values.

Download Full Size | PDF

Funding

National Natural Science Foundation of China (12004084, 12164008, 11564005); Guizhou Provincial Science and Technology Projects (ZK[2021]030, ZK(2022)198, ZK[2022]203); Science and Technology Talent Support Project of the Department of Education in the Guizhou Province (KY[2018]043); Natural Science Research Project of Guizhou Minzu University (GZMU[2019]YB29, GZMUZK[2021]YB06, GZMUZK[2021]YB08); Youth Science and Technology Talent Project of Guizhou Province (KY[2021]105, KY[2021]117); Key laboratory of Guizhou Minzu University (GZMUSYS[2021]03); Construction Project of Characteristic Key Laboratory in Guizhou Colleges and Universities (Y[2021]003); Central Guiding Local Science and Technology Development Foudation of China (QK ZYD[2019]4012); Shanghai Pujiang Program (22PJ1402900).

Disclosures

The authors declare that they have no conflict of interest.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. O. A. Abdelraouf, Z. Wang, H. Liu, Z. Dong, Q. Wang, M. Ye, X. R. Wang, Q. J. Wang, and H. Liu, “Recent advances in tunable metasurfaces: Materials, design, and applications,” ACS Nano 16(9), 13339–13369 (2022). [CrossRef]  

2. S. Xiao, T. Wang, T. Liu, C. Zhou, X. Jiang, and J. Zhang, “Active metamaterials and metadevices: a review,” J. Phys. D: Appl. Phys. 53(50), 503002 (2020). [CrossRef]  

3. C. Zou, J. Sautter, F. Setzpfandt, and I. Staude, “Resonant dielectric metasurfaces: active tuning and nonlinear effects,” J. Phys. D: Appl. Phys. 52(37), 373002 (2019). [CrossRef]  

4. J. Wu, Z. T. Xie, Y. Sha, H. Fu, and Q. Li, “Epsilon-near-zero photonics: infinite potentials,” Photonics Res. 9(8), 1616–1644 (2021). [CrossRef]  

5. B. Johns, N. M. Puthoor, H. Gopalakrishnan, A. Mishra, R. Pant, and J. Mitra, “Epsilon-near-zero response in indium tin oxide thin films: Octave span tuning and ir plasmonics,” J. Appl. Phys. 127(4), 043102 (2020). [CrossRef]  

6. S. J. Kim and M. L. Brongersma, “Active flat optics using a guided mode resonance,” Opt. Lett. 42(1), 5–8 (2017). [CrossRef]  

7. A. Forouzmand, M. M. Salary, S. Inampudi, and H. Mosallaei, “A tunable multigate indium-tin-oxide-assisted all-dielectric metasurface,” Adv. Opt. Mater. 6(7), 1701275 (2018). [CrossRef]  

8. A. Forouzmand, M. M. Salary, G. K. Shirmanesh, R. Sokhoyan, H. A. Atwater, and H. Mosallaei, “Tunable all-dielectric metasurface for phase modulation of the reflected and transmitted light via permittivity tuning of indium tin oxide,” Nanophotonics 8(3), 415–427 (2019). [CrossRef]  

9. M. Z. Alam, I. De Leon, and R. W. Boyd, “Large optical nonlinearity of indium tin oxide in its epsilon-near-zero region,” Science 352(6287), 795–797 (2016). [CrossRef]  

10. A. Forouzmand and H. Mosallaei, “Electro-optical amplitude and phase modulators based on tunable guided-mode resonance effect,” ACS Photonics 6(11), 2860–2869 (2019). [CrossRef]  

11. J. Park, J.-H. Kang, X. Liu, and M. L. Brongersma, “Electrically tunable epsilon-near-zero (enz) metafilm absorbers,” Sci. Rep. 5(1), 15754 (2015). [CrossRef]  

12. J. Park, J.-H. Kang, S. J. Kim, X. Liu, and M. L. Brongersma, “Dynamic reflection phase and polarization control in metasurfaces,” Nano Lett. 17(1), 407–413 (2017). [CrossRef]  

13. M. G. Silveirinha and N. Engheta, “Theory of supercoupling, squeezing wave energy, and field confinement in narrow channels and tight bends using $\varepsilon$ near-zero metamaterials,” Phys. Rev. B 76(24), 245109 (2007). [CrossRef]  

14. B. Edwards, A. Alù, M. E. Young, M. Silveirinha, and N. Engheta, “Experimental verification of epsilon-near-zero metamaterial coupling and energy squeezing using a microwave waveguide,” Phys. Rev. Lett. 100(3), 033903 (2008). [CrossRef]  

15. R. Maas, J. Parsons, N. Engheta, and A. Polman, “Experimental realization of an epsilon-near-zero metamaterial at visible wavelengths,” Nat. Photonics 7(11), 907–912 (2013). [CrossRef]  

16. A. Alu, M. G. Silveirinha, A. Salandrino, and N. Engheta, “Epsilon-near-zero metamaterials and electromagnetic sources: Tailoring the radiation phase pattern,” Phys. Rev. B 75(15), 155410 (2007). [CrossRef]  

17. G. Sinatkas, A. Pitilakis, D. C. Zografopoulos, R. Beccherelli, and E. E. Kriezis, “Transparent conducting oxide electro-optic modulators on silicon platforms: A comprehensive study based on the drift-diffusion semiconductor model,” J. Appl. Phys. 121(2), 023109 (2017). [CrossRef]  

18. M. Z. Alam, S. A. Schulz, J. Upham, I. De Leon, and R. W. Boyd, “Large optical nonlinearity of nanoantennas coupled to an epsilon-near-zero material,” Nat. Photonics 12(2), 79–83 (2018). [CrossRef]  

19. J. Kuttruff, D. Garoli, J. Allerbeck, R. Krahne, A. De Luca, D. Brida, V. Caligiuri, and N. Maccaferri, “Ultrafast all-optical switching enabled by epsilon-near-zero-tailored absorption in metal-insulator nanocavities,” Commun. Phys. 3(1), 114 (2020). [CrossRef]  

20. L. Tao, A. Anopchenko, S. Gurung, J. Zhang, and H. W. H. Lee, “Gate-tunable plasmon-induced transparency modulator based on stub-resonator waveguide with epsilon-near-zero materials,” Sci. Rep. 9(1), 2789 (2019). [CrossRef]  

21. M. Guo, L. Huang, W. Liu, and J. Ding, “Hybrid metasurface comprising epsilon-near-zero material for double transparent windows in optical communication band,” Opt. Mater. 112, 110802 (2021). [CrossRef]  

22. Z. T. Xie, J. Wu, H. Fu, and Q. Li, “Tunable electro-and all-optical switch based on epsilon-near-zero metasurface,” IEEE Photonics J. 12, 1–10 (2020). [CrossRef]  

23. A. I. Kuznetsov, A. E. Miroshnichenko, M. L. Brongersma, Y. S. Kivshar, and B. Luk’yanchuk, “Optically resonant dielectric nanostructures,” Science 354(6314), aag2472 (2016). [CrossRef]  

24. Y. Kivshar, “The rise of mie-tronics,” (2022).

25. P. Tonkaev, I. S. Sinev, M. V. Rybin, S. V. Makarov, and Y. Kivshar, “Multifunctional and transformative metaphotonics with emerging materials,” Chem. Rev. 122(19), 15414–15449 (2022). [CrossRef]  

26. J. Tian, Q. Li, P. A. Belov, R. K. Sinha, W. Qian, and M. Qiu, “High-q all-dielectric metasurface: super and suppressed optical absorption,” ACS Photonics 7(6), 1436–1443 (2020). [CrossRef]  

27. I. Staude and J. Schilling, “Metamaterial-inspired silicon nanophotonics,” Nat. Photonics 11(5), 274–284 (2017). [CrossRef]  

28. C. Zhou, L. Huang, R. Jin, L. Xu, G. Li, M. Rahmani, X. Chen, W. Lu, and A. E. Miroshnichenko, “Bound states in the continuum in asymmetric dielectric metasurfaces,” Laser Photonics Rev. 17(3), 2200564 (2023). [CrossRef]  

29. L. Huang, L. Xu, M. Rahmani, D. N. Neshev, and A. E. Miroshnichenko, “Pushing the limit of high-q mode of a single dielectric nanocavity,” Adv. Photonics 3(01), 016004 (2021). [CrossRef]  

30. 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]  

31. M. V. Gorkunov, A. A. Antonov, and Y. S. Kivshar, “Metasurfaces with maximum chirality empowered by bound states in the continuum,” Phys. Rev. Lett. 125(9), 093903 (2020). [CrossRef]  

32. S. Li, C. Zhou, T. Liu, and S. Xiao, “Symmetry-protected bound states in the continuum supported by all-dielectric metasurfaces,” Phys. Rev. A 100(6), 063803 (2019). [CrossRef]  

33. A. S. Kupriianov, Y. Xu, A. Sayanskiy, V. Dmitriev, Y. S. Kivshar, and V. R. Tuz, “Metasurface engineering through bound states in the continuum,” Phys. Rev. Appl. 12(1), 014024 (2019). [CrossRef]  

34. B. Meng, J. Wang, C. Zhou, and L. Huang, “Bound states in the continuum supported by silicon oligomer metasurfaces,” Opt. Lett. 47(6), 1549–1552 (2022). [CrossRef]  

35. P. Vaity, H. Gupta, A. Kala, S. Dutta Gupta, Y. S. Kivshar, V. R. Tuz, and V. G. Achanta, “Polarization-independent quasibound states in the continuum,” Adv. Photonics Res. 3(2), 2100144 (2022). [CrossRef]  

36. J. Yu, B. Ma, A. Ouyang, P. Ghosh, H. Luo, A. Pattanayak, S. Kaur, M. Qiu, P. Belov, and Q. Li, “Dielectric super-absorbing metasurfaces via pt symmetry breaking,” Optica 8(10), 1290–1295 (2021). [CrossRef]  

37. W. Liu, B. Wang, Y. Zhang, J. Wang, M. Zhao, F. Guan, X. Liu, L. Shi, and J. Zi, “Circularly polarized states spawning from bound states in the continuum,” Phys. Rev. Lett. 123(11), 116104 (2019). [CrossRef]  

38. M. Qin, J. Duan, S. Xiao, W. Liu, T. Yu, T. Wang, and Q. Liao, “Manipulating strong coupling between exciton and quasibound states in the continuum resonance,” Phys. Rev. B 105(19), 195425 (2022). [CrossRef]  

39. C. Zhou, X. Qu, S. Xiao, and M. Fan, “Imaging through a fano-resonant dielectric metasurface governed by quasi–bound states in the continuum,” Phys. Rev. Appl. 14(4), 044009 (2020). [CrossRef]  

40. I. A. Al-Ani, K. As’ Ham, L. Huang, A. E. Miroshnichenko, and H. T. Hattori, “Enhanced strong coupling of tmdc monolayers by bound state in the continuum,” Laser Photonics Rev. 15(12), 2100240 (2021). [CrossRef]  

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

42. C. W. Hsu, B. Zhen, A. D. Stone, J. D. Joannopoulos, and M. Soljačić, “Bound states in the continuum,” Nat. Rev. Mater. 1(9), 16048 (2016). [CrossRef]  

43. E. N. Bulgakov and A. F. Sadreev, “Bound states in the continuum in photonic waveguides inspired by defects,” Phys. Rev. B 78(7), 075105 (2008). [CrossRef]  

44. P. Yu, A. S. Kupriianov, V. Dmitriev, and V. R. Tuz, “All-dielectric metasurfaces with trapped modes: Group-theoretical description,” J. Appl. Phys. 125(14), 143101 (2019). [CrossRef]  

45. A. C. Overvig, S. C. Malek, M. J. Carter, S. Shrestha, and N. Yu, “Selection rules for quasibound states in the continuum,” Phys. Rev. B 102(3), 035434 (2020). [CrossRef]  

46. F. Wu, D. Liu, and S. Xiao, “Bandwidth-tunable near-infrared perfect absorption of graphene in a compound grating waveguide structure supporting quasi-bound states in the continuum,” Opt. Express 29(25), 41975–41989 (2021). [CrossRef]  

47. J. Algorri, F. Dell’Olio, P. Roldán-Varona, L. Rodríguez-Cobo, J. López-Higuera, J. Sánchez-Pena, V. Dmitriev, and D. Zografopoulos, “Analogue of electromagnetically induced transparency in square slotted silicon metasurfaces supporting bound states in the continuum,” Opt. Express 30(3), 4615–4630 (2022). [CrossRef]  

48. V. Dmitriev, D. C. Zografopoulos, S. D. S. Santos, and G. F. da Silva Barros, “Flat metasurfaces with square supercells of dielectric disk quadrumers: tailoring the fine structure of toroidal mode local field,” J. Phys. D: Appl. Phys. 55(20), 205104 (2022). [CrossRef]  

49. M.-S. Hwang, H.-C. Lee, K.-H. Kim, K.-Y. Jeong, S.-H. Kwon, K. Koshelev, Y. Kivshar, and H.-G. Park, “Ultralow-threshold laser using super-bound states in the continuum,” Nat. Commun. 12(1), 4135 (2021). [CrossRef]  

50. 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]  

51. G. Zograf, K. Koshelev, A. Zalogina, V. Korolev, R. Hollinger, D.-Y. Choi, M. Zuerch, C. Spielmann, B. Luther-Davies, D. Kartashov, S. V. Makarov, S. S. Kruk, and Y. Kivshar, “High-harmonic generation from resonant dielectric metasurfaces empowered by bound states in the continuum,” ACS Photonics 9(2), 567–574 (2022). [CrossRef]  

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

53. L. Xu, K. Zangeneh Kamali, L. Huang, M. Rahmani, A. Smirnov, R. Camacho-Morales, Y. Ma, G. Zhang, M. Woolley, D. Neshev, and A. E. Miroshnichenko, “Dynamic nonlinear image tuning through magnetic dipole quasi-bic ultrathin resonators,” Adv. Sci. 6(15), 1802119 (2019). [CrossRef]  

54. K. Sun, H. Jiang, D. A. Bykov, V. Van, U. Levy, Y. Cai, and Z. Han, “1d quasi-bound states in the continuum with large operation bandwidth in the ω k space for nonlinear optical applications,” Photonics Res. 10(7), 1575–1581 (2022). [CrossRef]  

55. I.-C. Benea-Chelmus, S. Mason, M. L. Meretska, D. L. Elder, D. Kazakov, A. Shams-Ansari, L. R. Dalton, and F. Capasso, “Gigahertz free-space electro-optic modulators based on mie resonances,” Nat. Commun. 13(1), 3170 (2022). [CrossRef]  

56. S. Han, L. Cong, Y. K. Srivastava, B. Qiang, M. V. Rybin, A. Kumar, R. Jain, W. X. Lim, V. G. Achanta, S. S. Prabhu, Q. J. Wang, Y. S. Kivshar, and R. Singh, “All-dielectric active terahertz photonics driven by bound states in the continuum,” Adv. Mater. 31(37), 1901921 (2019). [CrossRef]  

57. F. Yesilkoy, E. R. Arvelo, Y. Jahani, M. Liu, A. Tittl, V. Cevher, Y. Kivshar, and H. Altug, “Ultrasensitive hyperspectral imaging and biodetection enabled by dielectric metasurfaces,” Nat. Photonics 13(6), 390–396 (2019). [CrossRef]  

58. M. L. Tseng, Y. Jahani, A. Leitis, and H. Altug, “Dielectric metasurfaces enabling advanced optical biosensors,” ACS Photonics 8(1), 47–60 (2020). [CrossRef]  

59. L. Cong and R. Singh, “Symmetry-protected dual bound states in the continuum in metamaterials,” Adv. Opt. Mater. 7, 1900383 (2019). [CrossRef]  

60. L. Xu, M. Rahmani, Y. Ma, D. A. Smirnova, K. Z. Kamali, F. Deng, Y. K. Chiang, L. Huang, H. Zhang, S. Gould, D. N. Neshev, and A. E. Miroshnichenko, “Enhanced light–matter interactions in dielectric nanostructures via machine-learning approach,” Adv. Photonics 2(02), 1 (2020). [CrossRef]  

61. S. Campione, S. Liu, L. I. Basilio, L. K. Warne, W. L. Langston, T. S. Luk, J. R. Wendt, J. L. Reno, G. A. Keeler, I. Brener, and M. B. Sinclair, “Broken symmetry dielectric resonators for high quality factor fano metasurfaces,” ACS Photonics 3(12), 2362–2367 (2016). [CrossRef]  

62. Z. Liu, Y. Xu, Y. Lin, J. Xiang, T. Feng, Q. Cao, J. Li, S. Lan, and J. Liu, “High-q quasibound states in the continuum for nonlinear metasurfaces,” Phys. Rev. Lett. 123(25), 253901 (2019). [CrossRef]  

63. J. Wang, J. Kühne, T. Karamanos, C. Rockstuhl, S. A. Maier, and A. Tittl, “All-dielectric crescent metasurface sensor driven by bound states in the continuum,” Adv. Funct. Mater. 31(46), 2104652 (2021). [CrossRef]  

64. N. Karl, P. P. Vabishchevich, S. Liu, M. B. Sinclair, G. A. Keeler, G. M. Peake, and I. Brener, “All-optical tuning of symmetry protected quasi bound states in the continuum,” Appl. Phys. Lett. 115(14), 141103 (2019). [CrossRef]  

65. D. R. Abujetas, N. de Sousa, A. García-Martín, J. M. Llorens, and J. A. Sánchez-Gil, “Active angular tuning and switching of brewster quasi bound states in the continuum in magneto-optic metasurfaces,” Nanophotonics 10(17), 4223–4232 (2021). [CrossRef]  

66. K. Fan, I. V. Shadrivov, and W. J. Padilla, “Dynamic bound states in the continuum,” Optica 6(2), 169–173 (2019). [CrossRef]  

67. S. C. Malek, A. C. Overvig, S. Shrestha, and N. Yu, “Active nonlocal metasurfaces,” Nanophotonics 10(1), 655–665 (2021). [CrossRef]  

68. H. Kwon, T. Zheng, and A. Faraon, “Nano-electromechanical tuning of dual-mode resonant dielectric metasurfaces for dynamic amplitude and phase modulation,” Nano Lett. 21(7), 2817–2823 (2021). [CrossRef]  

69. E. D. Palik, Handbook of optical constants of solids, vol. 3 (Academic press, 1998).

70. P. C. Wu, C. Y. Liao, V. Savinov, T. L. Chung, W. T. Chen, Y.-W. Huang, P. R. Wu, Y.-H. Chen, A.-Q. Liu, N. I. Zheludev, and D. P. Tsai, “Optical anapole metamaterial,” ACS Nano 12(2), 1920–1927 (2018). [CrossRef]  

71. V. Savinov, V. Fedotov, and N. I. Zheludev, “Toroidal dipolar excitation and macroscopic electromagnetic properties of metamaterials,” Phys. Rev. B 89(20), 205112 (2014). [CrossRef]  

72. C. Zhou, S. Li, Y. Wang, and M. Zhan, “Multiple toroidal dipole fano resonances of asymmetric dielectric nanohole arrays,” Phys. Rev. B 100(19), 195306 (2019). [CrossRef]  

73. S. A. Maier, “Surface plasmon polaritons at metal/insulator interfaces,” in Plasmonics: Fundamentals and Applications, (Springer, 2007), pp. 21–37.

74. S. Tavana, S. Bahadori-Haghighi, and M. H. Sheikhi, “High-performance electro-optical switch using an anisotropic graphene-based one-dimensional photonic crystal,” Opt. Express 30(6), 9269–9283 (2022). [CrossRef]  

75. J. Huang, B. Meng, L. Chen, X. Wang, X. Qu, M. Fan, and C. Zhou, “Light trapping and manipulation of quasibound states in continuum Ge2Sb2Se4Te metasurfaces,” Phys. Rev. B 106(4), 045416 (2022). [CrossRef]  

76. T. Shibanuma, G. Grinblat, P. Albella, and S. A. Maier, “Efficient third harmonic generation from metal–dielectric hybrid nanoantennas,” Nano Lett. 17(4), 2647–2651 (2017). [CrossRef]  

77. M. M. Salary and H. Mosallaei, “Tunable all-dielectric metasurfaces for phase-only modulation of transmitted light based on quasi-bound states in the continuum,” ACS Photonics 7(7), 1813–1829 (2020). [CrossRef]  

78. A. Barreda, C. Zou, A. Sinelnik, E. Menshikov, I. Sinev, T. Pertsch, and I. Staude, “Tuning and switching effects of quasi-bic states combining phase change materials with all-dielectric metasurfaces,” Opt. Mater. Express 12(8), 3132–3142 (2022). [CrossRef]  

79. D. V. Bochek, N. S. Solodovchenko, D. A. Yavsin, A. B. Pevtsov, K. B. Samusev, and M. F. Limonov, “Bound states in the continuum versus material losses: Ge2Sb2Se4Te as an example,” Phys. Rev. B 105(16), 165425 (2022). [CrossRef]  

80. H. Wang, J. Zhou, D. Song, and Y. Wang, “Active tuning of near-infrared electromagnetic responses in the graphene/silicon hybrid nanohole arrays by electrical control,” Phys. Rev. B 105(3), 035407 (2022). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       This document provides supplementary information to “Active quasi-BIC metasurfaces assisted by Epsilon-Near-Zero materials”.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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. (a) Schematic diagram of the nanostructure array. (b) Geometry of the unit cell, the off-set distance of the air-hole H5 from the center of unit cell is marked $d$. (c) Transmission spectra for different $d$ values. The three modes are marked with Mode I, Mode II, and Mode III, respectively, and A, B, C, and D corresponds to the four BICs, respectively. (d), (e), (f) Q-factors of the three modes as the function of off-set distance $d$, respectively.
Fig. 2.
Fig. 2. Electric field distribution and field vector distribution at the $x-y$ plane for the three modes, (a)-(c) $d=30$ nm, (d)-(f) $d=170$ nm. Here, the vector arrows are normalized.
Fig. 3.
Fig. 3. Far-field scattered power and dominant multipole components of the multipole decomposition of Mode I, Mode II and Mode III; (a), (b), (c) are $d$=30 nm (d), (e ), (f) corresponds to $d$=170 nm. ED, MD, TD, EQ, MQ correspond to the response of electric dipole, magnetic dipole, toroidal dipole, electric quadrupole, and magnetic quadrupole, respectively.
Fig. 4.
Fig. 4. (a) Schematic diagram of the Si-ITO metasurface. The thickness of Si is 220 nm, the thickness of ITO is 10 nm. (b), (c) The changes of the real and imaginary parts of the dielectric constant of ITO with the current density $N_{\rm ITO}$, and the green dotted line is the corresponding ENZ position. (d), (e), and (f) are the transmission spectra of the metasurface without and with ITO film, respectively. The inset is the electric field distribution on the $x-y$ plane at the resonance wavelength, and $N_{\rm ITO}$ corresponds to the overlap between the resonance wavelength and the ENZ wavelength.
Fig. 5.
Fig. 5. Modulation of transmission spectrum with different $N_{\rm ITO}$ values. (a)-(c) The transmission spectra and corresponding modulation depths of Mode I, Mode II, and Mode III under different $N_{\rm ITO}$, respectively.
Fig. 6.
Fig. 6. (a)-(c) QBICs normalized far-field scattering modulation and (d)-(f) near-field capture modulation for different $N_{\rm ITO}$ values.
Fig. 7.
Fig. 7. (a) Schematics of unit cell for free-standing metasurfaces. (b)-(d) The Q-factors of Modes I, II and III with respect to the offset distance, respectively.
Fig. 8.
Fig. 8. Real (solid) and imaginary (dashed) parts of the permittivity of ITO as a function of wavelength. Points A, B, and C represent zero-crossing at $\lambda _{\rm ENZ}=1373$ nm, $\lambda _{\rm ENZ}=1440$ nm, and $\lambda _{\rm ENZ}=1528$ nm, respectively.
Fig. 9.
Fig. 9. Multipole contributions of Si-ITO metasurface from three QBICs at different current densities $N_{\rm ITO}$ at $d=$170 nm.
Fig. 10.
Fig. 10. (a)-(c) Transmission spectra and (a)-(c) corresponding modulation depths under different $N_{\rm ITO}$ values.

Equations (4)

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

ε = ε ω p 2 ω 2 + i Γ ω ,
ω p = N e 2 m e ε 0 ,
E R = 10 l o g T 1 T 0 ,
E F = | E | 2 d V | E 0 | 2 V ,
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.