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

Efficient polarization-insensitive quasi-BIC modulation by VO2 thin films

Open Access Open Access

Abstract

Bound states in the continuum (BIC) offer great design freedom for realizing high-quality factor metasurfaces. By deliberately disrupting the inherent symmetries, BIC can degenerate into quasi-BIC exhibiting sharp spectra with strong light confinement. This transformation has been exploited to develop cutting-edge sensors and modulators. However, most proposed quasi-BICs in metasurfaces are composed of unit cells with Cs symmetry that may experience performance degradation due to polarization deviation, posing challenges in practical applications. Addressing this critical issue, our research introduces an innovative approach by incorporating metasurfaces with C4v unit cell symmetry to eliminate polarization response sensitivity. Vanadium Dioxide (VO2) is a phase-change material with a relatively low transition temperature and reversibility. Here, we theoretically investigate the polarization-insensitive quasi-BIC modulation in Si-VO2 hybrid metasurfaces. By introducing defects into metasurfaces with Cs, C4, and C4v symmetries, we enable the emergence of quasi-BICs characterized by strong Fano resonance in their transmission spectra. Via numerically calculating the multipole decomposition, distinct dominant multipoles for different quasi-BICs are identified. A comprehensive investigation into the polarization responses of these structures under varying directions of linearly polarized light reveals the superior polarization-independent characteristics of metasurfaces with C4 and C4v symmetries, a feature that ensures the maintenance of maximum resonance peaks irrespective of polarization direction. Utilizing the polarization-insensitive quasi-BIC, we thus designed two different Si-VO2 hybrid metasurfaces with C4v symmetry. Each configuration presents complementary benefits, leveraging the VO2 phase transition's loss change to facilitate efficient modulation. Our quantitative calculation indicates notable achievements in modulation depth, with a maximum relative modulation depth reaching up to 342%. For the first time, our research demonstrates efficient modulation using polarization-insensitive quasi-BICs in designed Si-VO2 hybrid metasurfaces, achieving identical polarization responses for quasi-BIC-based applications. Our work paves the way for designing polarization-independent quasi-BICs in metasurfaces and marks a notable advancement in the field of tunable integrated devices.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Bound states in the continuum (BIC) are a captivating physical phenomenon that has garnered widespread attention in the field of nanostructures [14]. Initially conceptualized in quantum mechanics [5], BIC refers to a counterintuitive situation where a bound state (a state with discrete energy levels) exists in a continuum band, allowing it to remain localized and not decay despite being in a spectrum of lossy states [6]. The incorporation of BIC in nanostructures offers notable advantages like a novel light confinement mechanism, high-quality factors, intense field localization, and enhanced light-matter interactions [7]. Harnessing these characteristics has led to significant advances in areas such as resonator design [8], low-loss optical transmission [9], and efficient nonlinear generation [10,11], improving functionalities in diverse nanostructures including metasurfaces [1215]. Known for their ability to achieve precise control over light properties such as phase, amplitude, and polarization, metasurfaces are composed of subwavelength nanostructures with extensive design flexibility [16,17]. To date, a variety of BIC-based applications in metasurfaces have leveraged the well-known symmetry-protected BIC in metasurfaces with C4v group symmetry for their design simplicity and fabrication robustness [1825]. By introducing specific defects in C4v metasurfaces, non-radiative BIC can degenerate into quasi-BIC that couple with continuum modes, exhibiting Fano resonances with high-quality factors immensely beneficial for developing high-performance optical sensors [2628].

Besides, the quasi-BIC can be employed to achieve efficient optical modulation by exploiting its extremely sharp peak in the transmission (reflection) spectrum. To realize such modulation effects, a variety of functional materials have been integrated into dielectric metasurfaces for their unique tunability, including 2D materials [2932], electro-optic materials [33,34], and phase-change materials [3537]. Among diverse phase-change materials, Vanadium dioxide (VO2) has been proven a promising candidate to achieve efficient modulation for its relatively low transition temperature (about 68°) and reversibility nature. These properties enable VO2-based modulation devices to operate at lower temperatures and offer enhanced durability [4,3841]. By incorporating VO2 materials into BIC design, the efficient modulations of quasi-BIC peaks have been achieved in previous works [37,42]. However, while various BIC-based modulation devices have exhibited exceptional tunability, the majority is composed of unit cells with Cs symmetry. Consequently, they may experience significant performance degradation due to polarization deviations [43,44]. This issue could present challenges in practical applications that utilize the resonance peaks of quasi-BICs, such as sensing [45,46], bio-detection [28,47], and dynamic imaging [31,48]. Therefore, it is crucial to design quasi-BIC modulation devices beyond Cs unit cell symmetry with better robustness against polarization deviations. Despite previous demonstrations of polarization-independent quasi-BICs with C4 and C4v unit cell symmetries [19,4954], the integration of existing tuning methods with these modes still awaits further exploration. To the best of our knowledge, our work presents the first attempt to utilize these modes to achieve efficient modulation with enhanced polarization responses.

In this paper, we propose the polarization-insensitive quasi-BIC modulation in Si-VO2 hybrid metasurfaces. For comparing the polarization response properties, three types of quasi-BICs are designed by introducing specific defects in different metasurfaces with the unit cell symmetries of Cs, C4, C4v. These quasi-BICs exhibit strong resonance in the transmission spectrum at different wavelengths, which can be described by the Fano formula. To gain further insights into their physical origins, the far-field scattering of multipoles is calculated, revealing their distinct components with strongly localized fields observed near the resonant wavelengths. Next, we move to investigate the responses under linearly polarized waves with varying directions for three metasurfaces. The resonance peak in the Cs metasurface exhibits significant degradation, while C4 and C4v systems maintain their maximum peaks regardless of the polarization direction. To facilitate transmission modulation, a VO2 layer is integrated into C4v metasurface aiming to harness its loss change during phase transition. As the temperature rises from 20° to 80°, considerable transmission modulations are observed with a maximum relative modulation depth approaching 350%. Compared to existing quasi-BIC-enabled tunable optical devices [29,31,37,55], our design shows an excellent modulation performance with more fabrication-friendly structures and exhibits identical polarization responses, safeguarding it against performance degradation due to polarization deviation. These features ensure consistent and reliable operation across a range of polarization states, making it a significant advancement in the field of tunable optical devices.

2. Results

2.1 Quasi-BIC design in different structures

By introducing defects to disrupt the symmetry of metasurfaces, quasi-BICs in different metasurfaces are engineered. We examine the dielectric metasurfaces composed of silicon cylinder arrays on a SiO2 substrate with three symmetry unit cells: Cs, C4, and C4v, as depicted in Fig. 1.(a1-a3). Through the use of electron-beam lithography (EBL) techniques, several experimentally-oriented studies have reported high-precision fabrications of designed nanostructures with similar patterns to realize quasi-BICs [5658]. By adopting a parallel fabrication approach, the etched nanodisk structures in Fig. 1.(a1-a3) are also achievable. The Cs symmetry metasurface as depicted in Fig. 1.(a1), consists of two ellipses, each having a major axis Ly of 360 nm, a minor axis Lx of 180 nm, and a height of 200 nm, spaced 500 nm apart at a specific angle. The lattice geometry is Px =1.26 µm, Py =900 nm. Using the Finite-Difference Time-Domain (FDTD) method with periodic boundary conditions in the x- and y-directions and perfectly matched layers in the z-direction, the transmission spectrum under x-polarized incident light is calculated for varying θ values. For θ ≠ 0°, distinct resonance peaks around 2 µm emerge, broadening and shifting to shorter wavelengths as θ increases. At θ = 0°, the resonance peak vanishes, indicating a symmetry-protected BIC in Fig. 1.(b1). To explore BIC in more complex structures, we construct metasurfaces with C4 and C4v symmetries using four identical cylinders with specific defects arranged in a square grid pattern, as illustrated in Fig. 1.(a2-a3). These cylinders have a radius r = 320 nm and a height of 200 nm, with a length of d = 160 nm and varying width w defect (marked in Fig. 2.(b2-b3)). The lattice geometry is ax = ay = a = 860 nm. In Fig. 1.(a2), each cylinder is oriented such that the defect appears to incrementally rotate in a clockwise direction by 90° from one to the next. Such configuration suggests a sequential quarter-turn rotational symmetry to ensure the C4 symmetry of the unit cell. While maintaining the rotational symmetry, the defect in Fig. 1.(a3) is oriented diagonally at a 45° angle relative to the vertical axis to provide a C4v symmetry, rather than vertically as in the former case. As shown in Fig. 1.(b2-b3), adjusting the defect widths w in C4 and C4v metasurfaces also results in sharp resonance peaks caused by quasi-BICs. A comparison of the transmission spectrum between C4 and C4v metasurfaces that share identical parameters but different symmetries reveals close similarity, indicating the resonance peaks are mainly dominated by geometries rather than unit cell symmetries.

 figure: Fig. 1.

Fig. 1. Quasi-BIC design in different metasurfaces. (a1-a3) Schematics of metasurfaces with different symmetries. Relevant geometrical parameters are marked with Italics. In (a1), Px =1.26 µm, Py =900 nm, Lx = 180 nm, Ly = 360 nm. In (a2-a3), ax = ay = a = 860 nm, r = 320 nm, d = 160 nm. (b1-b3) Transmission spectra at varying defect sizes. The absence of a resonance peak when no defects are introduced indicates the existence of BIC when no defect is introduced. (c1-c3) Curve fitting of the scatters with the Fano formula. (d1-d3) Quality factors versus asymmetry parameters. The data are mainly extracted from (b1-b3), with the quality factors decreasing as the asymmetry parameters increase, also indicating the tuning of quasi-BICs.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. BIC properties in different metasurfaces. (a1-a3) Electric field distributions of quasi-BICs at resonance wavelength. Significant field localization effects are observed. (b1-b3) Multipole decompositions near the resonant wavelength. (c1-c3) Transmission spectra at increasing scaling factors. With the proportionally increasing geometries, the resonance peak exhibits a considerable shift.

Download Full Size | PDF

Relevant studies suggest that the resonant peaks in the three cases arise from the coupling between discrete quasi-BICs and modes in the continuous spectrum, which can be described using the typical Fano formula [59,60]:

$$T(\lambda )= {T_0} + {A_0}\frac{{{{[{q + 2({\lambda - {\lambda_0}} )/\Gamma } ]}^2}}}{{1 + {{[{2({\lambda - {\lambda_0}} )/\Gamma } ]}^2}}}.$$

Here, T0 is the background transmission, A0 is the coupling strength between the continuous and discrete modes, λ0 is the center wavelength of the resonance, Γ is the resonance linewidth, and q represents the Fano parameter, describing the asymmetry of the spectral line. Employing the Matlab curve fitting toolbox, the calculated transmission spectrum is fitted with the Fano formula as shown in Fig. 1.(c1-c3). The obtained regular R2 values are 0.9983, 0.9999, and 0.9998, indicating a strong correlation to Fano resonances. Besides, the quality factor Q defined by $\frac{{2\lambda }}{\Gamma }$ yields 4.1 × 103, 2.85 × 103, and 2.38 × 103 for the three resonance peaks respectively. Other values extracted from the lineshapes are summarized in Fig. 1.(d1-d3). As the asymmetry parameters increase, the quality factors drop rapidly, providing another evidence for the existence of quasi-BICs.

To gain deeper insights into the physical origins of BIC, the multipole decompositions near quasi-BIC resonance peaks for the three types of BIC are calculated using Comsol Multiphysics to analyze the dominant multipole components. Additional details about the multipole decomposition method can be found in Ref. [6165]. Firstly, the electric field distributions at resonance wavelengths are calculated as shown in Fig. 2.(a1-a3). The field intensities exhibit localized enhancement at specific locations, especially the defects in Fig. 2.(a2-a3). Building on this, the computed multipole component contributions for the far-field scattering power are shown in Fig. 2.(b1-b3), revealing significant differences for distinct quasi-BICs. Scattering powers of different orders of multipoles exhibit sharp peaks near the resonance centers, while other wavelengths only have negligible contributions. The quasi-BIC with Cs symmetry predominantly consists of magnetic dipole (MD) and electric quadrupole (EQ). Conversely, the rest are both primarily comprising toroidal dipole (TD) and magnetic quadrupole (MQ) elements.

Moreover, leveraging the robustness of the symmetry-protected BIC [47,66], we highlight that the center resonance wavelengths can be manipulated by uniformly scaling the geometric structures as shown in Fig. 2.(c1-c3).

2.2 Polarization response property of quasi-BICs

The obtained quasi-BIC Fano resonance peaks have considerable high-quality factors with sharp transmission peaks, paving the way to achieve efficient modulations. However, owing to the distinct unit cell symmetries, designed quasi-BICs would have divergent responses to linearly polarized light in different directions.

To explore their polarization responses, the transmission spectra are calculated for the three structures under linearly polarized incident light with four different polarization directions (φ = 0°, 30°, 60°, 90°), as shown with solid lines in Fig. 3.(a1-a3). Considering the anisotropy of Cs symmetric structure, the corresponding background transmission spectrum (θ=0°, no defects are introduced) is also depicted with a dot line for better comparing the peak values of quasi-BICs. From Fig. 3.(a1), the resonance peak progressively weakens as the φ angle increases, and becomes totally undetectable at φ= 90° (corresponding to y-direction). Thus, the decreased resonance peak in the Cs symmetry metasurface will lead to performance degradation in applications depending on the resonance peaks, such as sensing [45,46], bio-detection [28,47], and dynamic imaging [31,48]. In contrast, metasurfaces with C4 and C4v symmetries exhibit consistent responses to linearly polarized light from any direction as shown in Fig. 3.(a2-a3). This uniformity enables these structures to achieve maximum quasi-BIC resonance peaks without polarization alignment, thus minimizing the flux losses and contributing to a more compact, robust system.

 figure: Fig. 3.

Fig. 3. The polarization response properties of different symmetry systems. (a1-a3) Polarization responses of different metasurfaces. The resonance peak of Cs systems is strongly affected by the polarization directions, while the spectra of C4 and C4v systems remain the same regardless of polarization directions. (b1-b3) Electric field distributions at resonance wavelength for polarization direction φ=90°.

Download Full Size | PDF

To investigate the physical origins of these varying polarization responses, the electric field distributions at the resonance wavelength for φ=90° (y-polarized incident light) are calculated, as illustrated in Fig. 3.(b1-b3). Analysis of these visualizations shows that y-polarized incident light excites the other two defects in the C4 structure. Besides, apart from the two vertical nodal lines, the electric field distribution in C4v structure is almost identical to the case in the x-polarized incident light. Indeed, the electric field distributions shown in Fig. 2.(a2-a3) and Fig. 3.(b2-b3) differ only by a 90° planar rotation. The relative orientation of the electric field does not impact the transmission spectrum. Consequently, this similarity in electric field distributions ensures the identical spectrum of C4 and C4v structures in φ=0°, 90°. In contrast, the absence of a direct correlation between the electric field distributions in Fig. 2.(a1) and Fig. 3.(b1) results in distinct responses of the Cs structure.

For a more in-depth analysis of the polarization response property, define a 2 × 2 matrix S that relates the transmitted electric fields Ex', Ey' to the incident electric fields Ex, Ey:

$$\left( {\begin{array}{{c}} {{E_{x^{\prime}}}}\\ {{E_{y^{\prime}}}} \end{array}} \right) = \left( {\begin{array}{{cc}} {{S_{11}}}&{{S_{12}}}\\ {{S_{21}}}&{{S_{22}}} \end{array}} \right)\left( {\begin{array}{{c}} {{E_x}}\\ {{E_y}} \end{array}} \right).$$

The scattering matrix related to different metasurfaces is derived using group representation theory, $S = \left( {\begin{array}{{cc}} {{S_{11}}}&0\\ 0&{{S_{22}}} \end{array}} \right)$, $\left( {\begin{array}{{cc}} {{S_{11}}}&{ - {S_{12}}}\\ {{S_{12}}}&{{S_{11}}} \end{array}} \right)$, $\left( {\begin{array}{{cc}} {{S_{11}}}&0\\ 0&{{S_{11}}} \end{array}} \right)$ for Cs, C4, C4v respectively, with each configuration being dominated by the unit cell symmetries. These varying configurations of scattering matrix cause the distinct polarization responses of different metasurfaces. For the Cs symmetry, the non-zero elements in its scattering matrix are confined to the diagonal with different values (S11≠ S22). The varying diagonal elements lead to the polarization response sensitivity in Fig. 3.(a1). Conversely, the scattering matrices for C4 and C4v symmetries have equal diagonal elements, rendering them polarization-insensitive properties in Fig. 3.(a2-a3). A notable distinction of scattering matrix for C4 lies in its opposite off-diagonal elements (S12=-S21), potentially exhibiting circular dichroism beyond polarization insensitivity [43].

2.3 Polarization-insensitive quasi-BIC modulation by VO2 thin films

With the simplest scattering matrix, the constructed C4v metasurface is a promising choice to achieve polarization-independent quasi-BIC modulation. By employing the pulsed laser deposition (PLD) technique, thin-film VO2 with a thickness of several tens of nanometers have been fabricated with excellent performance characteristics [38,67]. These studies provide solid foundation for designing VO2 hybrid systems. In consideration of the potential impact of VO2 layer thickness on modulation effects, our study initially focuses on examining these effects by integrating a 10 nm thick VO2 layer onto each cylinder within a silicon array, as shown in Fig. 4.(a). The modulations of resonance peaks are achieved by leveraging the phase-change property of VO2, which switches from a low-loss dielectric state to a high-loss metallic state at a specific temperature. This transition significantly affects the loss of VO2 layers, thus facilitating effective tuning of quasi-BIC resonance peaks. Our simulations adopt the temperature-dependent refractive index and extinction coefficient data of VO2 thin films from Ref. [68]. Considering the VO2 layer's thin thickness, a refined mesh at the VO2-Si interface is implemented to ensure computational precision. The calculated transmission spectra across rising temperatures are shown in Fig. 4.(b). At 20°C, the VO2 remains in its dielectric state with minimal loss. The lower loss enables quasi-BIC to maintain a strong Fano resonance with a significant transmission peak. As the temperature increases, the extinction coefficient of VO2 gradually rises, diminishing the resonance peak and causing spectral broadening due to the higher loss. Upon the temperature surpassing the phase transition threshold (about 68°C), VO2 enters its metallic phase with dramatically increased loss, causing a significant modulation with negligible resonance effect observed in the transmission spectrum. For a more quantitative measurement, the relative modulation depths are calculated using the following formula [37,69] (Fig. 4.c):

$$\Delta T = \frac{{{T_1} - {T_0}}}{{{T_0}}}.$$

From the results, ΔT shows sharp peaks near the resonance wavelength, corresponding to the modulation effect of quasi-BIC induced by temperature changes. While at other wavelengths, only very low changes are observed. The modulation depths of the two temperatures hold 202%, and 342% respectively, delivering a notable progress compared to existing researches [37,42].

 figure: Fig. 4.

Fig. 4. Resonance peak modulation induced by VO2 layers. (a) Schematic of the Si-VO2 hybrid metasurface. The VO2 layer is on each Si cylinder. (b) Transmission spectra at different temperatures. (c) Calculated relative modulation depths ΔT for 55°C and 80°C, with a maximum depth of 202% and 342%, respectively.

Download Full Size | PDF

Figure 4.(c) demonstrates the relative modulation depth depends on the loss of VO2 layers. To further examine the effects of different VO2 thickness on light modulation, the transmission spectra for VO2 layers with thicknesses of 3 nm and 50 nm are calculated at three representative temperatures (Fig. 5.a1-a2), and their corresponding modulation depths are determined (Fig. 5.b1-b2). By comparing the results, it is observed that thinner VO2 layers restrict the loss caused by the VO2 phase transition, leading to less degradation as the temperature rises. As a result, the transmission spectra can still maintain distinguishable resonance peaks despite VO2 entering its metal phase. The relative modulation depth of 20°C to 55°C holds only very low values near the resonance wavelength due to an inadequate peak change.

 figure: Fig. 5.

Fig. 5. (a1-a2) Transmission spectra at three representative temperatures for 3 nm (a1) and 50 nm (a2) thick VO2 layers. (b1-b2) Corresponding relative modulation depths.

Download Full Size | PDF

Conversely, excessively thick VO2 layers incur considerable loss even at 20°C, adversely reducing the peak and quality factor of the transmission spectrum. The significant spectrum degradation at 20°C leads to a much smaller ΔT than the previous two cases. Therefore, controlling the thickness of VO2 layers is crucial for achieving efficient modulation. The 10 nm thick VO2 is a better choice compared to the other two for providing sufficient loss change to strike a balance between maintaining a strong resonance in the dielectric state and achieving efficient temperature-switching modulation capabilities.

By adding the VO2 layer to each Si cylinder, significant transmission modulation is observed. However, this approach requires modifying numerous VO2 thin layers on Si cylinders, potentially increasing fabrication difficulty and influencing the thickness uniformity. Hence, we propose another more fabrication-friendly configuration that yields comparable modulation effects. As depicted in Fig. 6.(a), this approach involves growing the VO2 layer directly on a SiO2 substrate. Without additional VO2 shaping processes, this method streamlines the fabrication procedure and enhances the uniformity of VO2 thin films. Following a similar manner as above, the transmission spectrum (Fig. 6.b1-b2), and their relative modulation depths (Fig. 6.c1-c2) are calculated with VO2 thicknesses of 3 nm and 10 nm. Compared to the previous method, using VO2 as a direct-contact substrate results in a more pronounced impact on the resonance peak at the same VO2 thicknesses, which can be attributed to the increased loss due to a larger VO2 area. Therefore, to achieve comparable modulation depth, a thinner VO2 layer is required. Despite vibrant advances in VO2 growth technology, producing such thin VO2 layers on SiO2 substrates still awaits more opportunities.

 figure: Fig. 6.

Fig. 6. (a) Schematic of Si-VO2 hybrid metasurface. The VO2 layer is only on the SiO2 substrate. (b1-b2) Transmission spectrum at three representative temperatures for 3 nm (b1) and 10 nm (b2) thick VO2 layers. (c1-c2) Corresponding relative modulation depths.

Download Full Size | PDF

Overall, each proposed approach has distinct advantages and limitations. The choice of the most appropriate structure should be guided by practical fabrication conditions.

3. Conclusion

In summary, we have achieved the modulation of polarization-insensitive quasi-BICs via the Si-VO2 hybrid metasurfaces with C4v symmetry. By exploiting the symmetry-broken induced quasi-BICs, we design the high-Q Fano resonance peaks in metasurface structures with Cs, C4, and C4v symmetries, and investigate their various properties including the localized enhanced fields, dominant multipole components as well as the tunning effects induced by proportional geometry scaling. The comparison of their responses to differently polarized incident light reveals that the metasurfaces with C4 and C4v symmetries respond uniformly, maintaining maximum resonance peaks regardless of the polarization orientation. Building on this, the modulations in two specially designed Si-VO2 hybrid metasurfaces with C4v symmetry are achieved by leveraging the phase transition loss of VO2. Quantitative calculations show that the relative modulation depth approaches 350%. Beyond the loss modulation induced by VO2 materials, our method can be easily extended to other modulation schemes like refractive index tuning devices by integrating the polarization-insensitive quasi-BICs with C4v unit cell symmetry into such structures [33], which can enhance their device performance with better polarization response robustness. We believe our work will offer valuable insights for the development of designing polarization-insensitive quasi-BICs in metasurfaces.

Funding

National Natural Science Foundation (62075113, 62122040).

Disclosures

The authors declare no conflicts 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.

References

1. S. I. Azzam and A. V. Kildishev, “Photonic Bound States in the Continuum: From Basics to Applications,” Adv. Opt. Mater. 9(1), 2001469 (2021). [CrossRef]  

2. S. Joseph, Saurabh Pandey, Swagato Sarkar, et al., “Bound states in the continuum in resonant nanostructures: an overview of engineered materials for tailored applications,” Nanophotonics 10(17), 4175–4207 (2021). [CrossRef]  

3. H. Zhong, Tiantian He, Yuan Meng, et al., “Photonic Bound States in the Continuum in Nanostructures,” Materials 16(22), 7112 (2023). [CrossRef]  

4. K. Koshelev, Andrey Bogdanov, Yuri Kivshar, et al., “Meta-optics and bound states in the continuum,” Sci. Bull. 64(12), 836–842 (2019). [CrossRef]  

5. F. H. Stillinger and D. R. Herrick, “Bound states in the continuum,” Phys. Rev. A 11(2), 446–454 (1975). [CrossRef]  

6. C. W. Hsu, Bo Zhen, A. Douglas Stone, et al., “Bound states in the continuum,” Nat. Rev. Mater. 1(9), 16048 (2016). [CrossRef]  

7. E. Maggiolini, Laura Polimeno, Francesco Todisco, et al., “Strongly enhanced light–matter coupling of monolayer WS2 from a bound state in the continuum,” Nat. Mater. 22(8), 964–969 (2023). [CrossRef]  

8. M. Kang, Li Mao, Shunping Zhang, et al., “Merging bound states in the continuum by harnessing higher-order topological charges,” Light: Sci. Appl. 11(1), 228 (2022). [CrossRef]  

9. C. L. Zou, Jin-Ming Cui, Fang-Wen Sun, et al., “Guiding light through optical bound states in the continuum for ultrahigh-Q microresonators,” Laser Photonics Rev. 9(1), 114–119 (2015). [CrossRef]  

10. T. C. Tan, Yogesh Kumar Srivastava, Rajour Tanyi Ako, et al., “Active control of nanodielectric-induced THz quasi-BIC in flexible metasurfaces: a platform for modulation and sensing,” Adv. Mater. 33(27), 2100836 (2021). [CrossRef]  

11. Y. Meng, Yizhen Chen, Longhui Lu, et al., “Optical meta-waveguides for integrated photonics and beyond,” Light: Sci. Appl. 10(1), 235 (2021). [CrossRef]  

12. Y. Meng, Futai Hu, Zhoutian Liu, et al., “Chip-integrated metasurface for versatile and multi-wavelength control of light couplings with independent phase and arbitrary polarization,” Opt. Express 27(12), 16425–16439 (2019). [CrossRef]  

13. T. Ning, Xin Li, Yan Zhao, et al., “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]  

14. F. Ye, Yue Yu, Xiang Xi, et al., “Second-Harmonic Generation in Etchless Lithium Niobate Nanophotonic Waveguides with Bound States in the Continuum,” Laser Photonics Rev. 16(3), 2100429 (2022). [CrossRef]  

15. X. Zong, Lixia Li, Yufang Liu, et al., “Bound states in the continuum in all-van der Waals photonic crystals: a route enabling electromagnetically induced transparency,” Opt. Express 30(11), 17897–17908 (2022). [CrossRef]  

16. Y. Meng, Zhoutian Liu, Zhenwei Xie, et al., “Versatile on-chip light coupling and (de)multiplexing from arbitrary polarizations to controlled waveguide modes using an integrated dielectric metasurface,” Photonics Res. 8(4), 564–576 (2020). [CrossRef]  

17. Z. Liu, Yuan Meng, Futai Hu, et al., “Largely Tunable Terahertz Circular Polarization Splitters Based on Patterned Graphene Nanoantenna Arrays,” IEEE Photonics J. 11(5), 1–11 (2019). [CrossRef]  

18. Ibrahim A. M. Al-Ani, As’Ham, Lujun Huang, et al., “Strong coupling of exciton and high-Q mode in all-perovskite metasurfaces,” Adv. Opt. Mater. 10, 2101120 (2022). [CrossRef]  

19. Y. Cai, Yong Huang, Keyong Zhu, et al., “Symmetric metasurface with dual band polarization-independent high-Q resonances governed by symmetry-protected BIC,” Opt. Lett. 46(16), 4049–4052 (2021). [CrossRef]  

20. X. Chen and W. Fan, “Tunable bound states in the continuum in all-dielectric terahertz metasurfaces,” Nanomaterials 10(4), 623 (2020). [CrossRef]  

21. C. F. Doiron, Igal Brener, Alexander Cerjan, et al., “Realizing symmetry-guaranteed pairs of bound states in the continuum in metasurfaces,” Nat. Commun. 13(1), 7534 (2022). [CrossRef]  

22. K. Koshelev, Sergey Lepeshov, Mingkai Liu, et al., “Asymmetric metasurfaces with high-Q resonances governed by bound states in the continuum,” Phys. Rev. Lett. 121(19), 193903 (2018). [CrossRef]  

23. T. Shi, Zi-Lan Deng, Guangzhou Geng, et al., “Planar chiral metasurfaces with maximal and tunable chiroptical response driven by bound states in the continuum,” Nat. Commun. 13(1), 4111 (2022). [CrossRef]  

24. Y. Tang, Yao Liang, Jin Yao, et al., “Chiral bound states in the continuum in plasmonic metasurfaces,” Laser Photonics Rev. 17(4), 2200597 (2023). [CrossRef]  

25. X. Zhang, Yilin Liu, Jiecai Han, et al., “Chiral emission from resonant metasurfaces,” Science 377(6611), 1215–1218 (2022). [CrossRef]  

26. W. Cen, Tingting Lang, Zhi Hong, et al., “Ultrasensitive flexible terahertz plasmonic metasurface sensor based on bound states in the continuum,” IEEE Sens. J. 22(13), 12838–12845 (2022). [CrossRef]  

27. S. Isaacs, Ansar Hajoj, Mohammad Abutoama, et al., “Resonant grating without a planar waveguide layer as a refractive index sensor,” Sensors 19(13), 3003 (2019). [CrossRef]  

28. R. Wang, Lei Xu, Lujun Huang, et al., “Ultrasensitive Terahertz Biodetection Enabled by Quasi-BIC-Based Metasensors,” Small 19(35), 2301165 (2023). [CrossRef]  

29. Y. Chen, Zexu Liu, Yuke Li, et al., “Adjustable converter of bound state in the continuum basing on metal-graphene hybrid metasurfaces,” Opt. Express 30(13), 23828–23839 (2022). [CrossRef]  

30. Y. Meng, Jiangang Feng, Sangmoon Han, et al., “Photonic van der Waals integration from 2D materials to 3D nanomembranes,” Nat. Rev. Mater. 8(8), 498–517 (2023). [CrossRef]  

31. J. Li, Jie Li, Chenglong Zheng, et al., “Free switch between bound states in the continuum (BIC) and quasi-BIC supported by graphene-metal terahertz metasurfaces,” Carbon 182, 506–515 (2021). [CrossRef]  

32. Y. Meng, Hongkun Zhong, Zhihao Xu, et al., “Functionalizing nanophotonic structures with 2D van der Waals materials,” Nanoscale Horiz. 8(10), 1345–1365 (2023). [CrossRef]  

33. I.-C. Benea-Chelmus, Sydney Mason, Maryna L. Meretska, et al., “Gigahertz free-space electro-optic modulators based on Mie resonances,” Nat. Commun. 13(1), 3170 (2022). [CrossRef]  

34. X. Li, Di Liu, Yanyan Huo, et al., “Low-voltage electro-optic switching and modulation of second harmonic generation in lithium niobate resonant waveguide gratings assisted by quasi-BICs,” Opt. Laser Technol. 160, 109083 (2023). [CrossRef]  

35. A. Barreda, Chengjun Zou, Artem Sinelnik, et al., “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]  

36. D. Wang, JiangKun Tian, Tian Ma, et al., “Terahertz modulators based on VO2–metal hybridized metamaterials for free switching between BIC and quasi-BIC states,” Opt. Commun. 551, 130040 (2024). [CrossRef]  

37. Y. Zhang, Deliang Chen, Wenbin Ma, et al., “Active optical modulation of quasi-BICs in Si-VO2 hybrid metasurfaces,” Opt. Lett. 47(21), 5517–5520 (2022). [CrossRef]  

38. T. Kang, Zongwei Ma, Jun Qin, et al., “Large-scale, power-efficient Au/VO2 active metasurfaces for ultrafast optical modulation,” Nanophotonics 10(2), 909–918 (2020). [CrossRef]  

39. L. Long, Sydney Taylor, Liping Wang, et al., “Enhanced Infrared Emission by Thermally Switching the Excitation of Magnetic Polariton with Scalable Microstructured VO2 Metasurfaces,” ACS Photonics 7(8), 2219–2227 (2020). [CrossRef]  

40. S. Abdollahramezani, Omid Hemmatyar, Hossein Taghinejad, et al., “Tunable nanophotonics enabled by chalcogenide phase-change materials,” Nanophotonics 9(5), 1189–1241 (2020). [CrossRef]  

41. L. Huang, Lei Xu, David A. Powell, et al., “Resonant leaky modes in all-dielectric metasystems: Fundamentals and applications,” Phys. Rep. 1008, 1–66 (2023). [CrossRef]  

42. H. Dai, Jukun Liu, Jiaqi Ju, et al., “Absorption modulation of quasi-BIC in Si–VO2 composite metasurface at near-infrared wavelength,” New J. Phys. 25(10), 103010 (2023). [CrossRef]  

43. P. Yu, Anton S. Kupriianov, Victor Dmitriev, et al., “All-dielectric metasurfaces with trapped modes: Group-theoretical description,” J. Appl. Phys. 125(14), 1 (2019). [CrossRef]  

44. Y. Chen, Meijie Li, Jiankun Wang, et al., “Polarization-selective excitation of multiband quasibound states in the continuum supported by symmetry-perturbed silicon metasurface,” Opt. Commun. 527, 128904 (2023). [CrossRef]  

45. H.-H. Hsiao, Yi-Chien Hsu, Ai-Yin Liu, et al., “Ultrasensitive Refractive Index Sensing Based on the Quasi-Bound States in the Continuum of All-Dielectric Metasurfaces,” Adv. Opt. Mater. 10(19), 2200812 (2022). [CrossRef]  

46. Y. Wang, Zhanghua Han, Yong Du, et al., “Ultrasensitive terahertz sensing with high-Q toroidal dipole resonance governed by bound states in the continuum in all-dielectric metasurface,” Nanophotonics 10(4), 1295–1307 (2021). [CrossRef]  

47. A. Tittl, Aleksandrs Leitis, Mingkai Liu, et al., “Imaging-based molecular barcoding with pixelated dielectric metasurfaces,” Science 360(6393), 1105–1109 (2018). [CrossRef]  

48. S. Romano, Maria Mangini, Erika Penzo, et al., “Ultrasensitive Surface Refractive Index Imaging Based on Quasi-Bound States in the Continuum,” ACS Nano 14(11), 15417–15427 (2020). [CrossRef]  

49. J. Feng, Xianghui Wang, Jierong Cheng, et al., “Polarization-independent bound state in the continuum without the help of rotational symmetry,” Opt. Lett. 48(18), 4829–4832 (2023). [CrossRef]  

50. X. Wang, Shiyu Li, Chaobiao Zhou, et al., “Polarization-independent toroidal dipole resonances driven by symmetry-protected BIC in ultraviolet region,” Opt. Express 28(8), 11983–11989 (2020). [CrossRef]  

51. S. Yu, Yusen Wang, Ziang Gao, et al., “Dual-band polarization-insensitive toroidal dipole quasi-bound states in the continuum in a permittivity-asymmetric all-dielectric meta-surface,” Opt. Express 30(3), 4084–4095 (2022). [CrossRef]  

52. A. Sayanskiy, Anton S. Kupriianov, Su Xu, et al., “Controlling high-Q trapped modes in polarization-insensitive all-dielectric metasurfaces,” Phys. Rev. B 99(8), 085306 (2019). [CrossRef]  

53. K. Wu, Haosen Zhang, Guo Ping Wang, et al., “Polarization-independent high-Q monolayer Ge-assisted near-perfect absorber through quasibound states in the continuum,” Phys. Rev. B 108(8), 085412 (2023). [CrossRef]  

54. R. Masoudian Saadabad, Lujun Huang, Andrey E. Miroshnichenko, et al., “Polarization-independent perfect absorber enabled by quasibound states in the continuum,” Phys. Rev. B 104(23), 235405 (2021). [CrossRef]  

55. X. Wang, Junyi Duan, Wenya Chen, et al., “Controlling light absorption of graphene at critical coupling through magnetic dipole quasi-bound states in the continuum resonance,” Phys. Rev. B 102(15), 155432 (2020). [CrossRef]  

56. D. R. Abujetas, Jorge Olmos-Trigo, José A. Sánchez-Gil, et al., “Tailoring Accidental Double Bound States in the Continuum in All-Dielectric Metasurfaces,” Adv. Opt. Mater. 10(15), 2200301 (2022). [CrossRef]  

57. L. Huang, Shuangli Li, Chaobiao Zhou, et al., “Realizing Ultrahigh-Q Resonances Through Harnessing Symmetry-Protected Bound States in the Continuum,” Advanced Functional Materials n/a, 2309982, 1 (2023). [CrossRef]  

58. C. Zhou, Rong Jin, Lei Xu, et al., “Bound States in the Continuum in Asymmetric Dielectric Metasurfaces,” Laser Photonics Rev. 17(3), 2200564 (2023). [CrossRef]  

59. M. F. Limonov, Mikhail V. Rybin, Alexander N. Poddubny, et al., “Fano resonances in photonics,” Nat. Photonics 11(9), 543–554 (2017). [CrossRef]  

60. E. Melik-Gaykazyan, Kirill Koshelev, Jae-Hyuck Choi, et al., “From Fano to Quasi-BIC Resonances in Individual Dielectric Nanoantennas,” Nano Lett. 21(4), 1765–1771 (2021). [CrossRef]  

61. C. Zhou, Shiyu Li, Yi Wang, et al., “Multiple toroidal dipole Fano resonances of asymmetric dielectric nanohole arrays,” Phys. Rev. B 100(19), 195306 (2019). [CrossRef]  

62. S. Li, Chaobiao Zhou, Tingting Liu, et al., “Symmetry-protected bound states in the continuum supported by all-dielectric metasurfaces,” Phys. Rev. A 100(6), 063803 (2019). [CrossRef]  

63. E. A. Gurvitz, Konstantin S. Ladutenko, Pavel A. Dergachev, et al., “The High-Order Toroidal Moments and Anapole States in All-Dielectric Photonics,” Laser Photonics Rev. 13(5), 1800266 (2019). [CrossRef]  

64. X. Wang, Deliang Chen, Wenbin Ma, et al., “Tuning the magnetic toroidal dipole response in dielectric metasurfaces,” J. Opt. Soc. Am. B 40(3), 560–566 (2023). [CrossRef]  

65. M. Zhou, Shaojun You, Lei Xu, et al., “Bound states in the continuum in all-dielectric metasurfaces with scaled lattice constants,” Sci. China Phys. Mech. Astron. 66(12), 124212 (2023). [CrossRef]  

66. A. Kodigala, Thomas Lepetit, Qing Gu, et al., “Lasing action from photonic bound states in continuum,” Nature 541(7636), 196–199 (2017). [CrossRef]  

67. R. Yuan, Qingxi Duan, Pek Jun Tiw, et al., “A calibratable sensory neuron based on epitaxial VO2 for spike-based neuromorphic multisensory system,” Nat. Commun. 13(1), 3973 (2022). [CrossRef]  

68. I. O. Oguntoye, Siddharth Padmanabha, Max Hinkle, et al., “Continuously Tunable Optical Modulation Using Vanadium Dioxide Huygens Metasurfaces,” ACS Appl. Mater. Interfaces 15(34), 41141–41150 (2023). [CrossRef]  

69. P. Hosseini, C. David Wright, Harish Bhaskaran, et al., “An optoelectronic framework enabled by low-dimensional phase-change films,” Nature 511(7508), 206–211 (2014). [CrossRef]  

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 (6)

Fig. 1.
Fig. 1. Quasi-BIC design in different metasurfaces. (a1-a3) Schematics of metasurfaces with different symmetries. Relevant geometrical parameters are marked with Italics. In (a1), Px =1.26 µm, Py =900 nm, Lx = 180 nm, Ly = 360 nm. In (a2-a3), ax = ay = a = 860 nm, r = 320 nm, d = 160 nm. (b1-b3) Transmission spectra at varying defect sizes. The absence of a resonance peak when no defects are introduced indicates the existence of BIC when no defect is introduced. (c1-c3) Curve fitting of the scatters with the Fano formula. (d1-d3) Quality factors versus asymmetry parameters. The data are mainly extracted from (b1-b3), with the quality factors decreasing as the asymmetry parameters increase, also indicating the tuning of quasi-BICs.
Fig. 2.
Fig. 2. BIC properties in different metasurfaces. (a1-a3) Electric field distributions of quasi-BICs at resonance wavelength. Significant field localization effects are observed. (b1-b3) Multipole decompositions near the resonant wavelength. (c1-c3) Transmission spectra at increasing scaling factors. With the proportionally increasing geometries, the resonance peak exhibits a considerable shift.
Fig. 3.
Fig. 3. The polarization response properties of different symmetry systems. (a1-a3) Polarization responses of different metasurfaces. The resonance peak of Cs systems is strongly affected by the polarization directions, while the spectra of C4 and C4v systems remain the same regardless of polarization directions. (b1-b3) Electric field distributions at resonance wavelength for polarization direction φ=90°.
Fig. 4.
Fig. 4. Resonance peak modulation induced by VO2 layers. (a) Schematic of the Si-VO2 hybrid metasurface. The VO2 layer is on each Si cylinder. (b) Transmission spectra at different temperatures. (c) Calculated relative modulation depths ΔT for 55°C and 80°C, with a maximum depth of 202% and 342%, respectively.
Fig. 5.
Fig. 5. (a1-a2) Transmission spectra at three representative temperatures for 3 nm (a1) and 50 nm (a2) thick VO2 layers. (b1-b2) Corresponding relative modulation depths.
Fig. 6.
Fig. 6. (a) Schematic of Si-VO2 hybrid metasurface. The VO2 layer is only on the SiO2 substrate. (b1-b2) Transmission spectrum at three representative temperatures for 3 nm (b1) and 10 nm (b2) thick VO2 layers. (c1-c2) Corresponding relative modulation depths.

Equations (3)

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

T ( λ ) = T 0 + A 0 [ q + 2 ( λ λ 0 ) / Γ ] 2 1 + [ 2 ( λ λ 0 ) / Γ ] 2 .
( E x E y ) = ( S 11 S 12 S 21 S 22 ) ( E x E y ) .
Δ T = T 1 T 0 T 0 .
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.