Abstract
Bound states in the continuum (BICs) provide exceptional light confinement due to their inherent decoupling from radiative channels. Small symmetry breaking transforms BIC into quasi-BIC (qBIC) that couples to free-space radiation enabling ultra-high-quality-factor (Q-factor) resonances desirable for refractive index (RI) sensing. In practical implementations, geometric asymmetry is typically employed. However, since the radiative loss remains fixed once fabricated, such metasurfaces exhibit only a horizontal shift of the resonance spectrum in RI sensing, without modification of its overall shape. Here, we demonstrate a permittivity-asymmetric qBIC (ε-qBIC) metasurface, which encodes environmental refractive index variations directly into the asymmetry factor, resulting in an index response involving both resonance wavelength shift and modulation variation. In addition to exhibiting a competitive transmittance sensitivity of ∼5,300 %/RIU under single-wavelength conditions, the ε-qBIC design provides a substantially improved linear response. Specifically, the linear window area of its sensing data distribution, calculated as the integrated wavelength region where the linearity parameter remains above the preset threshold, is 104 times larger than that of the geometry-asymmetric qBIC (g-qBIC), enabling more robust and reliable single-wavelength signal readout. Additionally, numerical results reveal that environmental permittivity asymmetry can optically restore the g-qBIC to a state with ultra-high-Q (over 107), approaching the BIC condition. Unlike traditional BICs, which are typically inaccessible once perturbed, the permittivity-restored BIC becomes accessible through environmental perturbations. These findings suggest an alternative design strategy for developing high-performance photonic devices for practical sensing applications.
1 Introduction
Among the various optical resonances in nanophotonics, bound states in the continuum (BICs) stand out for their ability to confine light without radiative loss, enabled by symmetry or interference that decouples them from free-space modes [1], [2]. In particular, if the coupling vanishes due to symmetry, such states are referred to as symmetry-protected BICs [3], [4], [5]. To make these idealized, non-radiative states accessible in practical photonic devices, a slight symmetry-breaking is typically introduced, transforming them into quasi-BICs (qBICs) that retain a high quality factor (Q factor), which is suitable for compact devices [6], [7], [8], [9]. Conventional implementations usually achieve this by introducing minor geometrical asymmetries, such as tilted elliptical rods [10], nanodisks with off-centered holes [11], or asymmetric split-rings [12], among other structures [13], [14], [15].
Among the many emerging applications, refractive index (RI) sensing based on symmetry-protected BIC metasurfaces has recently attracted growing interest. Compared to other types of metasurfaces [16], [17], [18], [19], qBIC resonances offer the advantage that their Q factor can be tuned simply by adjusting the geometrical asymmetry [20], [21], [22], [23], [24]. However, like most resonance-based sensing platform, the readout method for the environmental RI changes relies on tracking the spectral shift of an optical resonance due to changes in the surrounding medium [25], [26]. It typically requires high-resolution spectrometers and stable broadband light sources, which add cost and complexity [27], [28].
To address these practical constraints, recent efforts have explored single-wavelength intensity variation, where the sensing signal is extracted from the intensity modulation at a fixed probe wavelength near the resonance [27], [29], [30]. Such an approach simplifies the hardware and enables integration into compact, low-cost platforms.
However, for conventional geometry-asymmetric qBIC (g-qBIC) metasurfaces, the radiative loss is essentially fixed during sensing. In low-loss environments, changes in the surrounding refractive index do not alter the radiative loss, and thus mainly induce a lateral spectral shift with minimal impact on the vertical modulation depth of the resonance. [31], [32], [33], [34], [35].
In contrast, permittivity asymmetric qBIC (ε-qBIC) metasurfaces can offer a fundamentally different mechanism for interacting with RI environment. It is well know that radiative coupling of qBICs with the far field is governed by the asymmetry factor, which conventionally defined by structure’s geometrical asymmetry as mentioned before [21], [24]. However, according to our previous research, based on the permittivity asymmetric qBIC metasurfaces, we can encode the environmental RI into asymmetry factor of the system [36]. As a result, the qBIC becomes highly responsive to RI variations, resulting not only a shift of the resonance position, but also a pronounced modulation of its amplitude in the low loss system. This response improves the linearity of the sensing signal in single-wavelength analysis, which, while not strictly required for all sensing approaches, facilitates simpler and more robust signal interpretation compared to nonlinear sensing data distributions with environmental RI, thereby enhancing sensing stability.
In this work, the sensing performance of ε-qBIC metasurfaces is experimentally demonstrated, and the concept of an environment-accessible restored BIC enabled by permittivity asymmetry is proposed through simulations. We first confirm that the ε-qBIC metasurface exhibits not only resonance wavelength shifts but also additional resonance modulation with changes in the RI, in contrast to the purely wavelength-shift response of g-qBIC’s spectrum, using commercially available refractive index oils (Δn = 0.02). To evaluate the relative quality of signals between wavelength shift and single wavelength intensity modulation, sensing experiments were further conducted with smaller RI intervals (Δn = 0.004). The intensity variation signal (ΔT s) at a probe wavelength achieved a higher signal-to-noise ratio (SNR ≈ 17 dB) compared to the wavelength-shift signal (Δλ s) (SNR ≈ 5 dB).
Consequently, the intensity variation signal (ΔT s) was further analyzed, and the experimental responses of ε-qBIC and g-qBIC metasurfaces were compared. Comparable sensitivities were obtained for both metasurfaces (∼5,000 %/RIU). However, the ε-qBIC metasurface exhibited a markedly improved linearity of the sensing response, with a linear window, defined as the wavelength range where the extracted linearity parameter exceeds the preset threshold, whose integrated area is approximately 104 times larger than that of the g-qBIC. This wider window indicates lower noise across a broader wavelength range, thereby improving robustness and stability for practical sensing applications.
Finally, from a fundamental perspective, numerical results reveal that geometric asymmetry, which typically converts BICs into quasi-BICs, can be compensated by precisely tailoring the environmental permittivity profile within the unit cell. This enables the recovery of BICs with radiative losses approaching zero. Notably, the restored symmetry protected BICs (RSP-BIC [37]) become optically accessible to changes in the surrounding environment, an interaction that is otherwise forbidden for conventional BICs. These findings suggest a broader design strategy: radiative channels in BIC systems can be tuned not only through geometry but also via environmental permittivity engineering, enabling new possibilities in high performance photonics platform for practical sensing application.
2 Results and discussion
2.1 Environmental RI-controlled asymmetry factor in ε-qBIC metasurfaces
To construct the ε-qBIC in our metasurface, the design follows two steps. First, the unit cell is composed of two identical TiO2 nanorods, which support a symmetry-protected BIC due to their C2 rotational symmetry. Second, a dielectric cover layer is selectively applied to one of the rods, introducing a difference in permittivity between the surroundings of the nanorods. This breaks the C2 symmetry and transforms the original BIC into a quasi-BIC, accompanied by a radiative resonance and a distinct electric field profile (Figure 1a and b). A more detailed discussion of the ε-qBIC origin is presented in the supplementary information (Fig. S6).

The concept of environmental permittivity-asymmetric quasi-BIC metasurfaces for refractive index sensing. (a) Illustrating the unit cell of the ε-qBIC metasurfaces, consisting of two identical dielectric nanorods made of TiO2 with periodicities P x = P y = 410 nm. The length l, width d, and height h of the nanobar are 320 nm, 105 nm, and 110 nm respectively. One of the resonators is encapsulated in a dielectric medium with refractive index of ncov and the height of H. (b) Transition from BIC to ε-qBIC by adding a cladding layer on one of the nanobars, and corresponding electric field distribution of the ε-qBIC. (c) Asymmetry factor defined by α = n env – n cov, which is related to radiative loss γ rad, where n cov = 1.49 and n env represent the refractive index of the dielectric medium covering on the one of nanorods, and the global environment of the whole ε-qBIC metasurfaces, respectively. (d) The Q factor exhibits an inverse-square dependence on the asymmetry factor, indicating that the asymmetry factor can modulate the radiative loss. (e) Spectral response under varying environmental RI, where changes in radiative loss additionally induce variations in transmittance intensity.
The asymmetry in permittivity leads to unequal dipole strengths between the two out-of-phase electric dipoles, resulting in a nonzero net dipole moment. This enables the resonance to couple to the far field and become observable in the transmission spectrum.
Specifically, the two TiO2 nanorods have dimensions of l = 320 nm in length, d = 105 nm in width, and h = 110 nm in height, with a unit cell periodicity of P x = P y = 410 nm. One of the rods is fully encapsulated in a dielectric medium with refractive index n cov, while the other remains exposed to an external medium with refractive index n env. For example, this can be implemented using a PMMA layer (n cov = 1.49) with height H of 300 nm as the cladding material.
The prepared metasurface is then placed in the sensing environment (Figure 1c). Unlike conventional g-qBIC designs, where the structural asymmetry, and thus the radiative loss, is fixed after fabrication, the ε-qBIC structure introduces asymmetry through the refractive index contrast between the cladding and the environment. This contrast is quantified by the asymmetry factor α = n env − n cov, allowing the radiative loss γ rad to vary dynamically with the surrounding medium. This work first demonstrates in RI sensing where environmental changes are directly translated into both the asymmetry factor and the associated radiative loss. As shown in Figure 1d, the Q-factor exhibits an inverse-square dependence on the asymmetry factor (related with environment RI), consistent with the characteristic behavior of qBIC modes. This dependence is also reflected in the optical spectra: as the refractive index of the environment varies (from 1.30 to 1.39), the system exhibits not only a shift in resonance wavelength but also a change in resonance modulation depth (Figure 1e). This feature improves the linearity of the sensing signal readout, resulting in better stability over a broader wavelength range. We will return to this point later. Before that, we experimentally verify the simulation results, as discussed in the next section. In addition, it is worth noting that PMMA, while convenient for proof-of-concept demonstrations, suffers from poor durability in practical applications such as biosensing environments [38]. In particular, it is soluble in many common organic solvents such as acetone and chloroform, and is easily removed during plasma cleaning processes [39]. Moreover, although PMMA can remain stable in aqueous buffer solutions, it shows limited resistance to strong acidic or basic regeneration steps that are often required in biosensing protocols [40]. These limitations indicate that more durable polymers (e.g., SU-8 [41]) or inorganic coatings (e.g., SiO2) should be considered for future applications.
2.2 Experimental validation of both resonance wavelength and intensity modulation in ε-qBIC metasurface
To experimentally validate the refractive index sensing capabilities of both g-qBIC and ε-qBIC metasurfaces, we performed transmittance measurements (Figure 2a) in a controlled aqueous environment using commercial index-matching oils with well-defined refractive indices (details in Method). Both metasurfaces share the similar Q factor measured in the air to ensure the fair comparison (Figure S3). Their fabrication has been confirmed through the SEM images (Figure 2b). The fabrication process of the ε-qBIC metasurfaces can be found in Figure S1.

Experimental characterization of the qBIC metasurfaces for refractive index sensing. (a) Schematic of the experimental setup for transmittance measurement for g-qBIC and ε-qBIC metasurfaces in an aqueous environment. (b) The SEM images for the g-qBIC metasurfaces, ε-qBIC metasurfaces, and BIC metasurfaces. (c) Measured transmittance spectra of g-qBIC metasurfaces immersed in commercial index-matching oils with refractive indices n i = 1.30, 1.32, 1.34, 1.36, 1.38. The PMMA (n cov = 1.49) serves as the cladding material. The corresponding resonance shifts are analyzed using a centroid method, shown on the right, where λ 0 to λ 4 denote the centroid wavelengths of the transmission dips, highlighting a clear monotonic wavelength shift as the RI increases. (d) Measured transmittance spectra of ε-qBIC metasurfaces under the same set of environmental refractive indices. In addition to the spectral shift, the modulation depth of the resonance also varies with n i. The right panel presents changes in both centroid wavelength and normalized modulation depth.
The transmittance spectra of g-qBIC metasurfaces are shown in Figure 2c, measured across a series of index-matching oils (n i = 1.30, 1.32, 1.34, 1.36, 1.38) with a refractive index step size of Δn = 0.02. As expected, the resonance wavelength exhibits a monotonic redshift with increasing environmental index, consistent with the behavior of conventional single-parameter high-Q RI sensing platforms, with a sensitivity of 118 nm/RIU. The centroid analysis provides the wavelength centroid of a resonance within a half-wavelength range, enabling a clearer tracking of its wavelength shift under different refractive indices (details of the centroid analysis are given in Figure S2). The centroid analysis on the right of Figure 2c confirms the linear trend, with a spectral shift of around 2 nm per RI step. Notably, the resonance modulation depth remains nearly constant, even at n i = 1.49, indicating that the optical response is primarily governed by phase effects rather than radiative coupling changes.
In contrast, Figure 2d demonstrates the spectral response of ε-qBIC metasurfaces under the same set of environmental refractive indices. The resonance wavelength undergoes a redshift of approximately 2 nm with increasing ni, similar to the g-qBIC case. Specifically, the sensitivity in terms of wavelength shift is 85 nm/RIU. It is slightly lower than that of the g-qBIC because the initial Q-factors of the two samples are not exactly the same, with the Q-factor of the ε-qBIC (Q = 124) being slightly smaller than that of the g-qBIC (Q = 131) (see supplementary information Figure S3). This results in a slightly lower sensitivity for the ε-qBIC. Nevertheless, it can be seen that the sensing capability in terms of wavelength shift remains on the same order of magnitude for both. In addition, the ε-qBIC provides an extra modulation of the resonance intensity, which can be further exploited in single-wavelength sensing to achieve improved sensing linearity. In Figure 2d, the modulation depth gradually decreases and eventually disappears at n i = 1.49. This critical point corresponds to a fully permittivity symmetric configuration, where the system reverts to a BIC. The complete disappearance of the resonance provides a distinct optical signature, highlighting the strong potential sensing capability enabled by permittivity-induced asymmetry.
This variation in modulation depth reflects the dependence of the Q factor on the asymmetry factor α, as previously discussed in Figure 1c–e. The right panel of Figure 2d provides a quantitative analysis of both the resonance centroid and the modulation depth, clearly showing that ε-qBIC metasurfaces exhibit not only a resonance wavelength shift but also a change in resonance intensity.
After validation of the simulation results, the next step is to assess two common sensing readouts: the resonance wavelength shift and the transmittance change at a fixed wavelength. The analysis focuses on determining which signal provides a stronger and more robust response under background noise for the ε-qBIC metasurface.
2.3 Single-wavelength intensity modulation outperforms wavelength shift in signal robustness
To evaluate which of the two available signal channels in the ε-qBIC metasurface, resonance wavelength shift or intensity variation at a fixed wavelength, offers better performance in RI sensing, we conducted a set of high-resolution sensing experiments, implemented by narrowing the refractive index step, to extract the signal-to-noise ratio (SNR) under system noise conditions.
In this experiment, glycerol–water mixtures with refractive indices from 1.386 to 1.402 were prepared in increments of Δn = 0.004. These five RI values (n 1 to n 5) correspond to the spectra presented in Figure 3a. To highlight the relative changes between spectra, the transmission spectrum from the first measurement (n 1) was taken as a reference. Its maximum and minimum values were normalized to 1 and 0, respectively, so that all subsequent spectra could be compared on the same scale. This normalization preserves the overall spectral shape while enabling the relative variations to be clearly observed. For each RI, the transmission spectrum was recorded, and two sensing parameters were extracted. The intensity variation signal ΔT s is defined as the change in transmittance at a fixed probe wavelength, ΔT s = T i − T 1, where T i is the normalized transmittance for the i-th RI value. The resonance shift signal Δλ s is defined as the change in resonance wavelength, where λ i represent the resonance wavelength extracted from the i-th measurement.

Comparative analysis of signal quality between wavelength shift and intensity variation in RI fine sensing. (a) Schematic representation of two types of signals used in RI fine sensing, along with normalized experimental spectra. The intensity variation ΔT s refers to the transmittance difference at a fixed wavelength, while Δλ s denotes the shift in resonance wavelength. Experiments were conducted in five refractive index environments (n 1 to n 5), ranging from 1.386 to 1.402, stepsize Δn = 0.004. (b) Illustration of system noise, showing small variations in transmittance (ΔT n) and resonance position (Δλ n) based on repeated measurements (m 0 to m i) under identical air conditions. Quantitative evaluation of noise levels was performed across five repeated measurements on the same ε-qBIC metasurface. (c–d) Signal-to-noise ratio (SNR) analysis for ΔT s and Δλ s based on the extracted signals in (a). Error bars represent the experimentally determined system noise. The signal ΔT s exhibits a significantly higher SNR, indicating better robustness for sensing applications.
To evaluate the system noise, five repeated measurements in air were performed after each sensing step. Prior to each repetition, the sample was rinsed, dried and measured under identical conditions, simulating practical scenarios involving metasurface reuse and capturing typical instrumental and operational fluctuations. The corresponding variations in transmittance (ΔT n) and resonance wavelength (Δλ n) are presented in the right panel of Figure 3b, with a schematic illustration provided to the left. These repeated measurements were used to quantify the noise levels, which were then incorporated as error bars in the analysis of ΔT s and Δλ s. As shown in Figure 3c, the intensity variation signal exhibits consistently higher SNR values across the tested RI range, exceeding 15 dB. By contrast, the resonance shift signal Δλ s, although above the noise threshold, displays greater fluctuations and a lower average SNR (Figure 3d).
These results indicate that, when accounting for measurement repetition and associated operational fluctuations, the intensity variation channel provides a more robust signal. In the following section, the advantages of using ΔT s for RI sensing with ε-qBIC metasurfaces are examined in comparison with g-qBIC.
2.4 Higher linearity of ε-qBIC sensing data distribution under single wavelength detection
To further evaluate the sensing performance of ε-qBIC metasurfaces based on intensity variation signals, we conducted a direct comparison with g-qBIC under identical refractive index conditions (n 1 to n 5, ranging from 1.386 to 1.402 with a step size of Δn = 0.004). Figure 4a and b show the transmittance spectra and extracted intensity variation signals for ε-qBIC and g-qBIC, respectively.

Comparison of sensing performance between ε-qBIC and g-qBIC metasurfaces under identical RI conditions. Refractive indices n 1 to n 5 range from 1.386 to 1.402, with a step size of Δn = 0.004. (a) Experimental transmittance spectra of the ε-qBIC metasurface under different RI environments, along with the extracted intensity variation signals. (b) Transmittance spectra and corresponding intensity variation signals for the g-qBIC metasurface under the same RI conditions. (c) Linear fitting of intensity variation signals at a single representative wavelength, comparing the responses of ε-qBIC and g-qBIC. (d) R 2 values obtained from linear fitting across multiple wavelengths, used to evaluate the linearity of the intensity variation response. Higher R 2 values indicate better linearity and lower fitting residuals. The ε-qBIC metasurface shows consistently higher R 2 values across a broader wavelength range, indicating enhanced linearity.
Linear fitting of the intensity variation ΔT s across the five RI values at each wavelength was performed. Figure 4c compares the fitted results for both metasurfaces at a selected representative wavelength. The slope of each fit corresponds to the sensitivity (ΔT s / Δn), while the coefficient of determination R 2 is used to evaluate the fitting quality. The R 2 value is defined as R 2 = 1 − S res / S tot, where S res = Σi(y i − ŷ i)2 is the residual sum of squares between the measured values yi and the fitted values ŷ i, and S tot = Σi(y i − ȳ)2 is the total sum of squares with respect to the mean ȳ of all yi. yi is the corresponding sensing data ΔT s. A higher R 2 value for a fitting indicates better linearity and lower noise in the signal.
As shown in Figure 4c, both ε-qBIC and g-qBIC exhibit high sensitivity in the range of ∼5,000 % / RIU. While some recent studies using metallic sensing structures have demonstrated extremely high spectral resolution [42], the sensitivity values reported in recent works based on dielectric metasurfaces, particularly those employing single-wavelength transmittance variation readout, are typically around 2000 % / RIU [29], [43], [44].
For the g-qBIC, the maximum R 2 observed across the wavelength range is 0.9617 at 727.5 nm. Since none of the wavelengths for g-qBIC reach R 2 = 0.98 which is the commonly referenced benchmark [45], [46], [47], we selected the highest R 2 point for comparison. In contrast, ε-qBIC achieves a competitive sensitivity (∼5,382 % / RIU) at a wavelength where R 2 = 0.9813, exceeding the 0.98 threshold. This indicates that ε-qBIC not only offers excellent sensitivity but also delivers improved signal quality and linearity.
Moreover, simulations in Figure S4 suggest that the sensitivity of ε-qBIC can be further enhanced by adjusting the cladding index ncov to closely match the refractive index of the sensing environment. In such high-Q regimes, even a small RI change can significantly alter the radiative loss channel to zero and restore the ε-qBIC resonance to a BIC, as the metasurface effectively exhibits symmetry both in geometry (identical rods) and in the surrounding dielectric environment. This leads to a sharp transition from a resonant to a non-resonant state and results in a unity intensity modulation at a fixed wavelength, representing an ideal condition for signal readout in RI sensing applications.
Beyond individual points, we further compared the R 2 distribution across the wavelength range (Figure 4d). While g-qBIC fails to exceed threshold within the range, ε-qBIC consistently maintains R 2 values above it. To quantify the effective linearity window, we applied a relaxed threshold of R 2 = 0.96 and calculated the integrated area between the R 2 curve and the threshold line. The ε-qBIC shows an area approximately 104 times larger than that of the g-qBIC, clearly demonstrating its much broader and more stable sensing window with reduced sensitivity to noise.
2.5 Environmentally permittivity restored symmetry-protected BIC
We demonstrate how a g-qBIC can be optically restored to a state that approaches a BIC with zero radiative losses by applying a compensating permittivity asymmetry. Specifically, the simulations start from a structure with the same periodicity, resonator height, and width as used in Figure 1. The initial configuration is a symmetric geometry with equal rod lengths (L 1 = L 2 = 320 nm), as shown on the right side of Figure 5a. In this case, the effective dipole moments (p 1 and p 2) cancel each other, resulting in a decoupled state with no resonance observable in the transmittance spectrum.

Permittivity-controlled restoration of symmetry-protected BICs. (a) Conceptual illustration of transformation of g-qBIC into restored symmetry-protected BIC (RSP-BIC) via environmental permittivity control. Starting from a symmetric structure supporting a BIC (left), introducing a geometric perturbation (shorter L 1) breaks the in-plane symmetry and produce a g-qBIC. By selectively covering the shorter resonator with a dielectric layer of refractive index n cov, the permittivity asymmetry compensates the geometric imbalance, restoring the BIC (right) with radiative losses approaching to zero. The middle panel illustrates the corresponding electric dipole distributions, while the bottom panel shows the associated transmittance spectra during this process. Simulated transmittance spectra showing the evolution of a BIC under geometric perturbation (L 1) and its recovery via covering the shorter rod (L 1 = 209 nm) with growing height H of a dielectric layer (n cov = 1.49). (b) Transforming of RSP-BIC into radiative qBIC due to immersing the metasurface into environment with n env. In contrast, conventional BICs are optically inaccessible and remain completely isolated from the environment with nenv. (c) Calculated Q factor map as a function of rod length L 1 and dielectric cover layer height H, demonstrating the condition for RSP-BIC characterized by a Q factor >107. This indicates the compensation point where the permittivity asymmetry cancels the geometric perturbation.
Next, by decreasing the length of L 1, we break the in-plane symmetry, resulting in a nonzero net dipole moment. This leads to a radiative qBIC mode that becomes observable as a resonance peak in the spectrum.
We then fix the short rod length (L 1 = 209 nm) and gradually add a PMMA cladding layer (n = 1.49) over that rod, increasing the coverage height H from 0 to 110 nm. This dielectric coating selectively enhances the local permittivity around the short rod, increasing its effective dipole moment p 1. As p 1 approaches p 2 in magnitude in the given antiparallel configuration, the effective total dipole moment of the system approaches zero again, restoring the destructive interference condition. As a result, the system transitions to a new restored symmetry-protected BIC (RSP-BIC), and the resonance disappears in the spectrum, as shown in the bottom panel of Figure 5a.
To quantify this transition, we extracted the Q-factors from the spectra across a range of L 1 and H values. Figure 5c shows the Q-factor map, clearly indicating the region where a RSP-BIC emerges. At this RSP-BIC condition, the Q factor exceeds 107, substantially higher than surrounding regions, confirming that a permittivity-induced compensation can effectively reestablish the BIC condition from a geometrically asymmetric structure.
Crucially, unlike conventional BICs, which are isolated from the external environment and remain stable under permittivity perturbations, the RSP-BIC is environmentally accessible. Small changes in the surrounding refractive index can significantly perturb this balance, driving a transition from the RSP-BIC back to a qBIC. This approach, which leverages the environmental RI to reversibly tune the system between RSP-BIC and qBIC, offers a novel mechanism that has not been addressed in prior studies and may open new directions in metasurface-based sensing. Possible applications are discussed in the supplementary information, Note 1.
3 Conclusions
We have presented a permittivity-driven quasi-bound state in the continuum metasurface design, where the environmental refractive index is directly encoded into the system’s asymmetry factor, enabling not only the resonance wavelength but also the resonance intensity to vary systematically with the surrounding medium. Under a single-wavelength condition, the intensity variation response with fine RI steps was analyzed. Although both the ε-qBIC and the g-qBIC exhibited high sensitivity, the ε-qBIC demonstrated a considerably broader and more stable linear window across the spectrum, suggesting enhanced robustness and reliability. Importantly, our numerical results reveal that permittivity asymmetry can be used not only to generate a quasi-BIC from a symmetric BIC, but also to optically restore a geometrically symmetry broken qBIC into a new state approaching a BIC with Q > 107 that remains accessible to environmental changes. Unlike conventional BICs, this restored BIC is responsive to RI variations and represents an optical state between nonradiative and radiative regimes, broadening the design space for BIC-based photonics, and offering a versatile framework for high-Q devices that merge fundamental control over light–matter interaction with practical sensing capabilities.
4 Methods
4.1 Optical characterizations
Transmittance spectra were measured using a white-light transmission microscope (Witec Alpha Series 300). The samples were illuminated with linearly polarized white light, and the transmitted signal was collected using a 20× immersion objective with a numerical aperture of 0.5. The collected light was coupled into a multimode fiber and directed to a grating-based spectrometer equipped with a silicon CCD detector. All measurements were normalized to its corresponding signal from a bare fused silica substrate. For the initial sensing experiments in Figure 2, commercial refractive index liquids from Cargille Labs (Series: AAA) were used, with refractive indices ranging from n = 1.30 to 1.39 in steps of 0.01. To further evaluate the sensing performance in Figures 3 and 4, additional refractive index liquids were prepared by mixing water and glycerol to obtain a finer range of indices between n = 1.386 and 1.414, with an increment of Δn = 0.004.
4.2 Numerical simulations
Simulations were performed using CST Studio Suite, a commercial finite-element-based solver. The model was configured in the frequency domain with periodic boundary conditions and employed adaptive mesh refinement to ensure convergence. The dielectric functions of TiO2 and PMMA used in the simulations were obtained from spectroscopic ellipsometry measurements and subsequently imported into CST. For sensing simulations, the background refractive index was varied to simulate changes in the surrounding environment.
4.3 Nano fabrications
Fused silica substrates were first cleaned in an ultrasonic acetone bath, then rinsed with isopropanol (IPA) and treated with oxygen plasma to remove any remaining contaminants. A 110 nm TiO2 layer was then deposited by sputtering a titanium target in an oxygen-containing plasma (Angstrom Engineering). A layer of PMMA 950k A4 resist was spin-coated and baked at 180 °C for 3 min. To avoid charging during electron beam lithography (EBL), a conductive polymer (E-Spacer 300Z) was spin-coated on top of the PMMA. In the initial patterning step, a 30 nm gold markers system was defined on the TiO2 film for alignment in the following processes. Using these markers, the second patterning step positioned two-rod nanostructure metasurfaces. After exposure, the sample was developed for 135 s in a 3:1 IPA–MIBK solution. A 50 nm chromium layer was then deposited by e-beam evaporation to serve as a hard mask, and lift-off took place in Microposit Remover 1,165 at 80 °C overnight. The structures were transferred into the TiO2 layer by reactive ion etching (RIE) in a PlasmaPro 100 ICP-RIE (Oxford Instruments), and the chromium mask was then removed in a wet Cr etchant. In the final patterning step, another PMMA layer was spin-coated and patterned onto the fabricated BIC metasurfaces using the same marker system. The patterned regions were cleaned using the same development protocol.
Funding source: European Union
Award Identifier / Grant number: EIC, OMICSENS, 101129734
Award Identifier / Grant number: ERC, METANEXT, 101078018
Funding source: Deutsche Forschungsgemeinschaft
Award Identifier / Grant number: EXC 2089/1-390776260
Award Identifier / Grant number: TI 1063/1
Funding source: Chinese scholarship council
Award Identifier / Grant number: Chinese scholarship
Funding source: Center for NanoScience Solar Energies Go Hybrid (SolTech)Bavarian programme
Award Identifier / Grant number: Bavarian programme
Acknowledgements
The authors thank Jonas Biechteler and Maxim Gorkunov for valuable discussions.
-
Research funding: This project was funded by the Deutsche Forschungsgemeinschaft (German Research Foundation) under grant numbers EXC 2089/1–390776260 (Germany’s Excellence Strategy) and TI 1063/1 (Emmy Noether Program), the Bavarian programme Solar Energies Go Hybrid (SolTech), and the Center for NanoScience at LMU. It was also funded by the European Union (ERC, METANEXT, 101078018, and EIC, OMICSENS, 101129734) and Chinese scholarship council. Views and opinions expressed are however those of the author(s) only and do not necessarily reflect those of the European Union, the European Research Council Executive Agency, or the European Innovation Council and SMEs Executive Agency (EISMEA). Neither the European Union nor the granting authority can be held responsible for them.
-
Author contributions: HH, AT, XY, and AA contributed to the conceptual design of the study. XY carried out the simulations, experiments, and data analysis, with support from HH and AA. XY and HH prepared the initial draft of the manuscript. All authors contributed to the methodology, critically revised the manuscript, and approved the final version. AT supervised the project. All authors have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results and approved the final version of the manuscript.
-
Conflict of interest: Authors state no conflicts of interests.
-
Data availability: The data that support the findings of this study are available from Zenodo (https://doi.org/10.5281/zenodo.17600249).
References
[1] 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., vol. 121, no. 19, p. 193903, 2018, https://doi.org/10.1103/PhysRevLett.121.193903.Suche in Google Scholar PubMed
[2] C. W. Hsu, B. Zhen, A. D. Stone, J. D. Joannopoulos, and M. Soljačić, “Bound states in the continuum,” Nat. Rev. Mater., vol. 1, no. 9, pp. 1–13, 2016, https://doi.org/10.1038/natrevmats.2016.48.Suche in Google Scholar
[3] M. Kang, T. Liu, C. T. Chan, and M. Xiao, “Applications of bound states in the continuum in photonics,” Nat. Rev. Phys., vol. 5, pp. 659–678, 2023. https://doi.org/10.1038/s42254-023-00642-8.Suche in Google Scholar
[4] M. S. Hwang et al.., “Ultralow-threshold laser using super-bound states in the continuum,” Nat. Commun., vol. 12, no. 1, p. 4135, 2021, https://doi.org/10.1038/s41467-021-24502-0.Suche in Google Scholar PubMed PubMed Central
[5] H. Hu et al.., “Surface-Enhanced raman scattering in BIC-Driven semiconductor metasurfaces,” Adv. Opt. Mater., vol. 12, p. 2302812, 2024. https://doi.org/10.1002/adom.202302812.Suche in Google Scholar
[6] Z. Sadrieva, K. Frizyuk, M. Petrov, Y. Kivshar, and A. Bogdanov, “Multipolar origin of bound states in the continuum,” Phys. Rev. B, vol. 100, no. 11, pp. 1–12, 2019, https://doi.org/10.1103/PhysRevB.100.115303.Suche in Google Scholar
[7] S. Li, C. Zhou, T. Liu, and S. Xiao, “Symmetry-protected bound states in the continuum supported by all-dielectric metasurfaces,” Phys. Rev. A (Coll Park), vol. 100, no. 6, p. 063803, 2019, https://doi.org/10.1103/PhysRevA.100.063803.Suche in Google Scholar
[8] C. Zhou et al.., “Ultrahigh-Q Quasi-BICs via precision-controlled asymmetry in dielectric metasurfaces,” Nano Lett., vol. 25, no. 14, pp. 5916–5924, 2025, https://doi.org/10.1021/acs.nanolett.5c00967.Suche in Google Scholar PubMed
[9] Z. Liu et al.., “High- Q quasibound States in the continuum for nonlinear metasurfaces,” Phys. Rev. Lett., vol. 123, no. 25, pp. 1–6, 2019, https://doi.org/10.1103/PhysRevLett.123.253901.Suche in Google Scholar PubMed
[10] A. Tittl et al.., “Imaging-based molecular barcoding with pixelated dielectric metasurfaces,” Science (1979), vol. 1109, no. June, pp. 1105–1109, 2018.10.1126/science.aas9768Suche in Google Scholar PubMed
[11] V. R. Tuz et al.., “High-quality trapped modes in all-dielectric metamaterials,” Opt. Express, vol. 26, no. 3, pp. 2905–2916, 2018, https://doi.org/10.1364/oe.26.002905.Suche in Google Scholar PubMed
[12] A. Forouzmand and H. Mosallaei, “All-Dielectric C-Shaped nanoantennas for light manipulation: tailoring both magnetic and electric resonances to the desire,” Adv. Opt. Mater., vol. 5, no. 14, p. 1700147, 2017, https://doi.org/10.1002/adom.201700147.Suche in Google Scholar
[13] R. Singh et al.., “Observing metamaterial induced transparency in individual Fano resonators with broken symmetry,” Appl. Phys. Lett., vol. 99, no. 20, p. 201107, 2011, https://doi.org/10.1063/1.3659494.Suche in Google Scholar
[14] S. Campione et al.., “Broken symmetry dielectric resonators for high quality factor fano metasurfaces,” ACS Photonics, vol. 3, no. 12, pp. 2362–2367, 2016, https://doi.org/10.1021/acsphotonics.6b00556.Suche in Google Scholar
[15] F. Zhang et al.., “Fano resonance of an asymmetric dielectric wire pair,” Appl. Phys. Lett., vol. 105, no. 17, p. 172901, 2014, https://doi.org/10.1063/1.4900757.Suche in Google Scholar
[16] S. A. Khan et al.., “Optical sensing by metamaterials and metasurfaces: from physics to biomolecule detection,” Adv. Opt. Mater., vol. 10, p. 2200500, 2022. https://doi.org/10.1002/adom.202200500.Suche in Google Scholar
[17] H. Li, J. T. Kim, J. S. Kim, D. Y. Choi, and S. S. Lee, “Metasurface-incorporated optofluidic refractive index sensing for identification of liquid chemicals through vision intelligence,” ACS Photonics, vol. 10, no. 3, pp. 780–789, 2023, https://doi.org/10.1021/acsphotonics.3c00057.Suche in Google Scholar
[18] L. La Spada, “Metasurfaces for advanced sensing and diagnostics,” Sensors, vol. 19, no. 2, p. 355, 2019, https://doi.org/10.3390/s19020355.Suche in Google Scholar PubMed PubMed Central
[19] T. Chung, H. Wang, and H. Cai, “Dielectric metasurfaces for next-generation optical biosensing: a comparison with plasmonic sensing,” Nanotechnology, vol. 34, no. 40, p. 402001, 2023, https://doi.org/10.1088/1361-6528/ace117.Suche in Google Scholar PubMed PubMed Central
[20] D. N. Maksimov, V. S. Gerasimov, A. A. Bogdanov, and S. P. Polyutov, “Enhanced sensitivity of an all-dielectric refractive index sensor with an optical bound state in the continuum,” Phys. Rev. A, vol. 105, no. 3, p. 033518, 2022, https://doi.org/10.1103/PhysRevA.105.033518.Suche in Google Scholar
[21] W. Chen, M. Li, W. Zhang, and Y. Chen, “Dual-resonance sensing for environmental refractive index based on quasi-BIC states in all-dielectric metasurface,” Nanophotonics, vol. 12, no. 6, pp. 1147–1157, 2023, https://doi.org/10.1515/nanoph-2022-0776.Suche in Google Scholar PubMed PubMed Central
[22] H. H. Hsiao, Y. C. Hsu, A. Y. Liu, J. C. Hsieh, and Y. H. Lin, “Ultrasensitive refractive index sensing based on the quasi-bound states in the continuum of all-dielectric metasurfaces,” Adv. Opt. Mater., vol. 10, no. 19, p. 2270075, 2022, https://doi.org/10.1002/adom.202200812.Suche in Google Scholar
[23] S. Romano et al.., “Ultrasensitive surface refractive index imaging based on quasi-bound states in the continuum,” ACS Nano, vol. 14, no. 11, pp. 15417–15427, 2020, https://doi.org/10.1021/acsnano.0c06050.Suche in Google Scholar PubMed
[24] K. Watanabe, T. Nagao, and M. Iwanaga, “Low-contrast BIC metasurfaces with quality factors exceeding 100,000,” Nano Lett., vol. 25, no. 7, pp. 2777–2784, 2025. https://doi.org/10.1021/acs.nanolett.4c05880.Suche in Google Scholar PubMed PubMed Central
[25] X. Wang and J. L. Wang, “Refractive index sensing characteristics of electromagnetic metamaterial absorber in terahertz band,” Wuli Xuebao/Acta Physica Sinica, vol. 70, no. 3, p. 038102, 2021, https://doi.org/10.7498/aps.70.20201054.Suche in Google Scholar
[26] Y. Xu et al.., “Optical refractive index sensors with plasmonic and photonic structures: promising and inconvenient truth,” Adv. Opt. Mater., vol. 7, no. 9, p. 1801433, 2019. https://doi.org/10.1002/adom.201801433.Suche in Google Scholar
[27] A. E. Cetin et al.., “Handheld high-throughput plasmonic biosensor using computational on-chip imaging,” Light Sci. Appl., vol. 3, no. Jan, p. e122, 2014, https://doi.org/10.1038/lsa.2014.3.Suche in Google Scholar
[28] G. A. Lopez, M. C. Estevez, M. Soler, and L. M. Lechuga, “Recent advances in nanoplasmonic biosensors: applications and lab-on-a-chip integration,” Nanophotonics, vol. 6, no. 1, pp. 123–136, 2017. https://doi.org/10.1515/nanoph-2016-0101.Suche in Google Scholar
[29] A. J. Ollanik, I. O. Oguntoye, G. Z. Hartfield, and M. D. Escarra, “Highly sensitive, affordable, and adaptable refractive index sensing with silicon-based dielectric metasurfaces,” Adv. Mater. Technol., vol. 4, no. 2, p. 1800567, 2019, https://doi.org/10.1002/admt.201800567.Suche in Google Scholar
[30] A. Kapoor and E. K. Sharma, “Long period grating refractive-index sensor: optimal design for single wavelength interrogation,” Appl. Opt., vol. 48, no. 31, pp. G88–G94, 2009, https://doi.org/10.1364/AO.48.000G88.Suche in Google Scholar PubMed
[31] E. Zanganeh, Z. Sadrieva, P. Kapitanova, and A. Bogdanov, “High- Q Mie resonators for refractive-index sensing,” Phys. Rev. Appl., vol. 21, no. 2, p. 024028, 2024, https://doi.org/10.1103/PhysRevApplied.21.024028.Suche in Google Scholar
[32] J. Zhou et al.., “Visualizing Mie resonances in low-index dielectric nanoparticles,” Phys. Rev. Lett., vol. 120, no. 25, p. 253902, 2018, https://doi.org/10.1103/PhysRevLett.120.253902.Suche in Google Scholar PubMed
[33] B. Gallinet and O. J. F. Martin, “Refractive index sensing with subradiant modes: a framework to reduce losses in plasmonic nanostructures,” ACS Nano, vol. 7, no. 8, pp. 6978–6987, 2013, https://doi.org/10.1021/nn4021967.Suche in Google Scholar PubMed
[34] Y. Shen et al.., “Plasmonic gold mushroom arrays with refractive index sensing figures of merit approaching the theoretical limit,” Nat. Commun., vol. 4, p. 2381, 2013. https://doi.org/10.1038/ncomms3381.Suche in Google Scholar PubMed
[35] K. Lodewijks, W. Van Roy, G. Borghs, L. Lagae, and P. Van Dorpe, “Boosting the figure-of-merit of LSPR-based refractive index sensing by phase-sensitive measurements,” Nano Lett., vol. 12, no. 3, pp. 1655–1659, 2012, https://doi.org/10.1021/nl300044a.Suche in Google Scholar PubMed
[36] H. Hu, W. Lu, A. Antonov, R. Berté, S. A. Maier, and A. Tittl, “Environmental permittivity-asymmetric BIC metasurfaces with electrical reconfigurability,” Nat. Commun., vol. 15, no. 1, p. 7050, 2024, https://doi.org/10.1038/s41467-024-51340-7.Suche in Google Scholar PubMed PubMed Central
[37] A. Aigner et al.., “Optical control of resonances in temporally symmetry-broken metasurfaces,” Nature, vol. 644, pp. 896–902, 2025. https://doi.org/10.1038/s41586-025-09363-7.Suche in Google Scholar PubMed PubMed Central
[38] L. E. Ocola and A. Stein, “Effect of cold development on improvement in electron-beam nanopatterning resolution and line roughness,” J. Vac. Sci. Technol. B: Microelectron. Nanometer Struct. Process., Meas., Phenom., vol. 24, no. 6, pp. 3061–3065, 2006, https://doi.org/10.1116/1.2366698.Suche in Google Scholar
[39] F. Rahman, D. Carbaugh, J. Wright, P. Rajan, S. Pandya, and S. Kaya, “A review of polymethyl methacrylate (PMMA) as a versatile lithographic resist – with emphasis on UV exposure,” Microelectron. Eng., vol. 224, p. 111238, 2020, https://doi.org/10.1016/j.mee.2020.111238.Suche in Google Scholar
[40] C. Oehr, “Plasma surface modification of polymers for biomedical use,” Nucl. Instrum. Methods Phys. Res. B: Beam Interact. Mat. Atom, vol. 208, pp. 40–47, 2003. https://doi.org/10.1016/S0168-583X(03)00650-5.Suche in Google Scholar
[41] S. Anbumani et al.., “Physiochemically distinct surface properties of SU-8 polymer modulate bacterial cell-surface holdfast and colonization,” ACS Appl. Bio Mater., vol. 5, no. 10, pp. 4903–4912, 2022, https://doi.org/10.1021/acsabm.2c00632.Suche in Google Scholar PubMed PubMed Central
[42] K. V. Sreekanth et al.., “Extreme sensitivity biosensing platform based on hyperbolic metamaterials,” Nat. Mater., vol. 15, no. 6, pp. 621–627, 2016, https://doi.org/10.1038/nmat4609.Suche in Google Scholar PubMed PubMed Central
[43] Y. Li, H. Fei, X. Liu, and H. Lin, “Design of refractive index sensors based on Valley photonic crystal Mach–Zehnder interferometer,” Sensors, vol. 25, no. 11, p. 3289, 2025, https://doi.org/10.3390/s25113289.Suche in Google Scholar PubMed PubMed Central
[44] H. Li, Z. Shi, H. Zhang, S. Qiu, and Z.-K. Zhou, “Hybrid metasurface based on Si3N4 nanopillar for optical sensing with dual channels,” ACS Appl. Nano Mater., vol. 8, no. 6, pp. 2965–2973, 2025, https://doi.org/10.1021/acsanm.4c06648.Suche in Google Scholar
[45] F. Usman et al.., “Acetone vapor-sensing properties of chitosan-polyethylene glycol using surface plasmon resonance technique,” Polymers, vol. 12, no. 11, p. 2586, 2020, https://doi.org/10.3390/polym12112586.Suche in Google Scholar PubMed PubMed Central
[46] L. Suo, H. Zhou, Y.-P. Peng, F. Yang, H.-C. Chui, and N.-K. Chen, “High sensitivity fiber refractive index sensors based on asymmetric supermodes interference in tapered four core fiber,” Photonics, vol. 9, no. 1, p. 45, 2022, https://doi.org/10.3390/photonics9010045.Suche in Google Scholar
[47] Z. Li, L. Hou, L. Ran, J. Kang, and J. Yang, “Ultra-sensitive fiber refractive index sensor with intensity modulation and self-temperature compensation,” Sensors, vol. 19, no. 18, p. 3820, 2019, https://doi.org/10.3390/s19183820.Suche in Google Scholar PubMed PubMed Central
Supplementary Material
This article contains supplementary material (https://doi.org/10.1515/nanoph-2025-0415).
© 2025 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Artikel in diesem Heft
- Frontmatter
- Reviews
- Light-driven micro/nanobots
- Tunable BIC metamaterials with Dirac semimetals
- Large-scale silicon photonics switches for AI/ML interconnections based on a 300-mm CMOS pilot line
- Perspective
- Density-functional tight binding meets Maxwell: unraveling the mysteries of (strong) light–matter coupling efficiently
- Letters
- Broadband on-chip spectral sensing via directly integrated narrowband plasmonic filters for computational multispectral imaging
- Sub-100 nm manipulation of blue light over a large field of view using Si nanolens array
- Tunable bound states in the continuum through hybridization of 1D and 2D metasurfaces
- Integrated array of coupled exciton–polariton condensates
- Disentangling the absorption lineshape of methylene blue for nanocavity strong coupling
- Research Articles
- Demonstration of multiple-wavelength-band photonic integrated circuits using a silicon and silicon nitride 2.5D integration method
- Inverse-designed gyrotropic scatterers for non-reciprocal analog computing
- Highly sensitive broadband photodetector based on PtSe2 photothermal effect and fiber harmonic Vernier effect
- Online training and pruning of multi-wavelength photonic neural networks
- Robust transport of high-speed data in a topological valley Hall insulator
- Engineering super- and sub-radiant hybrid plasmons in a tunable graphene frame-heptamer metasurface
- Near-unity fueling light into a single plasmonic nanocavity
- Polarization-dependent gain characterization in x-cut LNOI erbium-doped waveguide amplifiers
- Intramodal stimulated Brillouin scattering in suspended AlN waveguides
- Single-shot Stokes polarimetry of plasmon-coupled single-molecule fluorescence
- Metastructure-enabled scalable multiple mode-order converters: conceptual design and demonstration in direct-access add/drop multiplexing systems
- High-sensitivity U-shaped biosensor for rabbit IgG detection based on PDA/AuNPs/PDA sandwich structure
- Deep-learning-based polarization-dependent switching metasurface in dual-band for optical communication
- A nonlocal metasurface for optical edge detection in the far-field
- Coexistence of weak and strong coupling in a photonic molecule through dissipative coupling to a quantum dot
- Mitigate the variation of energy band gap with electric field induced by quantum confinement Stark effect via a gradient quantum system for frequency-stable laser diodes
- Orthogonal canalized polaritons via coupling graphene plasmon and phonon polaritons of hBN metasurface
- Dual-polarization electromagnetic window simultaneously with extreme in-band angle-stability and out-of-band RCS reduction empowered by flip-coding metasurface
- Record-level, exceptionally broadband borophene-based absorber with near-perfect absorption: design and comparison with a graphene-based counterpart
- Generalized non-Hermitian Hamiltonian for guided resonances in photonic crystal slabs
- A 10× continuously zoomable metalens system with super-wide field of view and near-diffraction–limited resolution
- Continuously tunable broadband adiabatic coupler for programmable photonic processors
- Diffraction order-engineered polarization-dependent silicon nano-antennas metagrating for compact subtissue Mueller microscopy
- Lithography-free subwavelength metacoatings for high thermal radiation background camouflage empowered by deep neural network
- Multicolor nanoring arrays with uniform and decoupled scattering for augmented reality displays
- Permittivity-asymmetric qBIC metasurfaces for refractive index sensing
- Theory of dynamical superradiance in organic materials
- Second-harmonic generation in NbOI2-integrated silicon nitride microdisk resonators
- A comprehensive study of plasmonic mode hybridization in gold nanoparticle-over-mirror (NPoM) arrays
- Foundry-enabled wafer-scale characterization and modeling of silicon photonic DWDM links
- Rough Fabry–Perot cavity: a vastly multi-scale numerical problem
- Classification of quantum-spin-hall topological phase in 2D photonic continuous media using electromagnetic parameters
- Light-guided spectral sculpting in chiral azobenzene-doped cholesteric liquid crystals for reconfigurable narrowband unpolarized light sources
- Modelling Purcell enhancement of metasurfaces supporting quasi-bound states in the continuum
- Ultranarrow polaritonic cavities formed by one-dimensional junctions of two-dimensional in-plane heterostructures
- Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2
- Ultrafast optical modulation of vibrational strong coupling in ReCl(CO)3(2,2-bipyridine)
- Chirality-driven all-optical image differentiation
- Wafer-scale CMOS foundry silicon-on-insulator devices for integrated temporal pulse compression
- Monolithic temperature-insensitive high-Q Ta2O5 microdisk resonator
- Nanogap-enhanced terahertz suppression of superconductivity
- Large-gap cascaded Moiré metasurfaces enabling switchable bright-field and phase-contrast imaging compatible with coherent and incoherent light
- Synergistic enhancement of magneto-optical response in cobalt-based metasurfaces via plasmonic, lattice, and cavity modes
- Scalable unitary computing using time-parallelized photonic lattices
- Diffusion model-based inverse design of photonic crystals for customized refraction
- Wafer-scale integration of photonic integrated circuits and atomic vapor cells
- Optical see-through augmented reality via inverse-designed waveguide couplers
- One-dimensional dielectric grating structure for plasmonic coupling and routing
- MCP-enabled LLM for meta-optics inverse design: leveraging differentiable solver without LLM expertise
- Broadband variable beamsplitter made of a subwavelength-thick metamaterial
- Scaling-dependent tunability of spin-driven photocurrents in magnetic metamaterials
- AI-based analysis algorithm incorporating nanoscale structural variations and measurement-angle misalignment in spectroscopic ellipsometry
Artikel in diesem Heft
- Frontmatter
- Reviews
- Light-driven micro/nanobots
- Tunable BIC metamaterials with Dirac semimetals
- Large-scale silicon photonics switches for AI/ML interconnections based on a 300-mm CMOS pilot line
- Perspective
- Density-functional tight binding meets Maxwell: unraveling the mysteries of (strong) light–matter coupling efficiently
- Letters
- Broadband on-chip spectral sensing via directly integrated narrowband plasmonic filters for computational multispectral imaging
- Sub-100 nm manipulation of blue light over a large field of view using Si nanolens array
- Tunable bound states in the continuum through hybridization of 1D and 2D metasurfaces
- Integrated array of coupled exciton–polariton condensates
- Disentangling the absorption lineshape of methylene blue for nanocavity strong coupling
- Research Articles
- Demonstration of multiple-wavelength-band photonic integrated circuits using a silicon and silicon nitride 2.5D integration method
- Inverse-designed gyrotropic scatterers for non-reciprocal analog computing
- Highly sensitive broadband photodetector based on PtSe2 photothermal effect and fiber harmonic Vernier effect
- Online training and pruning of multi-wavelength photonic neural networks
- Robust transport of high-speed data in a topological valley Hall insulator
- Engineering super- and sub-radiant hybrid plasmons in a tunable graphene frame-heptamer metasurface
- Near-unity fueling light into a single plasmonic nanocavity
- Polarization-dependent gain characterization in x-cut LNOI erbium-doped waveguide amplifiers
- Intramodal stimulated Brillouin scattering in suspended AlN waveguides
- Single-shot Stokes polarimetry of plasmon-coupled single-molecule fluorescence
- Metastructure-enabled scalable multiple mode-order converters: conceptual design and demonstration in direct-access add/drop multiplexing systems
- High-sensitivity U-shaped biosensor for rabbit IgG detection based on PDA/AuNPs/PDA sandwich structure
- Deep-learning-based polarization-dependent switching metasurface in dual-band for optical communication
- A nonlocal metasurface for optical edge detection in the far-field
- Coexistence of weak and strong coupling in a photonic molecule through dissipative coupling to a quantum dot
- Mitigate the variation of energy band gap with electric field induced by quantum confinement Stark effect via a gradient quantum system for frequency-stable laser diodes
- Orthogonal canalized polaritons via coupling graphene plasmon and phonon polaritons of hBN metasurface
- Dual-polarization electromagnetic window simultaneously with extreme in-band angle-stability and out-of-band RCS reduction empowered by flip-coding metasurface
- Record-level, exceptionally broadband borophene-based absorber with near-perfect absorption: design and comparison with a graphene-based counterpart
- Generalized non-Hermitian Hamiltonian for guided resonances in photonic crystal slabs
- A 10× continuously zoomable metalens system with super-wide field of view and near-diffraction–limited resolution
- Continuously tunable broadband adiabatic coupler for programmable photonic processors
- Diffraction order-engineered polarization-dependent silicon nano-antennas metagrating for compact subtissue Mueller microscopy
- Lithography-free subwavelength metacoatings for high thermal radiation background camouflage empowered by deep neural network
- Multicolor nanoring arrays with uniform and decoupled scattering for augmented reality displays
- Permittivity-asymmetric qBIC metasurfaces for refractive index sensing
- Theory of dynamical superradiance in organic materials
- Second-harmonic generation in NbOI2-integrated silicon nitride microdisk resonators
- A comprehensive study of plasmonic mode hybridization in gold nanoparticle-over-mirror (NPoM) arrays
- Foundry-enabled wafer-scale characterization and modeling of silicon photonic DWDM links
- Rough Fabry–Perot cavity: a vastly multi-scale numerical problem
- Classification of quantum-spin-hall topological phase in 2D photonic continuous media using electromagnetic parameters
- Light-guided spectral sculpting in chiral azobenzene-doped cholesteric liquid crystals for reconfigurable narrowband unpolarized light sources
- Modelling Purcell enhancement of metasurfaces supporting quasi-bound states in the continuum
- Ultranarrow polaritonic cavities formed by one-dimensional junctions of two-dimensional in-plane heterostructures
- Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2
- Ultrafast optical modulation of vibrational strong coupling in ReCl(CO)3(2,2-bipyridine)
- Chirality-driven all-optical image differentiation
- Wafer-scale CMOS foundry silicon-on-insulator devices for integrated temporal pulse compression
- Monolithic temperature-insensitive high-Q Ta2O5 microdisk resonator
- Nanogap-enhanced terahertz suppression of superconductivity
- Large-gap cascaded Moiré metasurfaces enabling switchable bright-field and phase-contrast imaging compatible with coherent and incoherent light
- Synergistic enhancement of magneto-optical response in cobalt-based metasurfaces via plasmonic, lattice, and cavity modes
- Scalable unitary computing using time-parallelized photonic lattices
- Diffusion model-based inverse design of photonic crystals for customized refraction
- Wafer-scale integration of photonic integrated circuits and atomic vapor cells
- Optical see-through augmented reality via inverse-designed waveguide couplers
- One-dimensional dielectric grating structure for plasmonic coupling and routing
- MCP-enabled LLM for meta-optics inverse design: leveraging differentiable solver without LLM expertise
- Broadband variable beamsplitter made of a subwavelength-thick metamaterial
- Scaling-dependent tunability of spin-driven photocurrents in magnetic metamaterials
- AI-based analysis algorithm incorporating nanoscale structural variations and measurement-angle misalignment in spectroscopic ellipsometry