Abstract
Structural colors generated by optical micro-/nanostructures offer a notable advantage over traditional chemical pigments, including higher purity, greater brightness, resistance to fading, and enhanced environmental friendliness. However, achieving dynamically switchable color displays with high performances and without resorting to complex nanofabrication methods remain a challenge. Here, we present a simple method using grayscale lithography and conformal coating to create Salisbury screen (SS) cavities with variable resonant wavelengths, enabling the formation of tunable colorful patterns. The dynamic color display is achieved through the phase change of vanadium dioxide (VO2) nanostructures under electrothermal effects. At a low actuation voltage of 1.4 V, high performances of color switching such as high sensitivity, fast speed, high repeatability, and wide-view angle are achieved. The tunable structural colors, featuring a simple preparation process and high-speed switching, represent a promising alternative for applications such as thermal sensors, security information encryption, and dynamic full-color displays.
1 Introduction
Structural colors are a fascinating and vibrant phenomenon that fundamentally differs from chemical colors [1], [2]. While chemical colors are produced by the absorption of specific wavelengths of light by materials, structural colors arise from the physical interaction of light with microscopic and nanoscopic structures on the surface of the material. In recent years, significant advances have been achieved in metasurface-based structural colors [3], [4], [5]. These metasurface interactions include thin-film interference effects [6], [7], [8], [9], diffraction from periodic structures [4], [5], and wavelength-selective scattering produced by particles, all of which contribute to the generation of brighter and more vivid colors [10]. For instance, the color gamut area of silicon metasurface structural colors reaches a relatively high proportion, increasing the spatial resolution to the diffraction limit [4]. These colors, formed through the manipulation of nanoscale structures, offer enhanced stability and resistance to fading, making them ideal for advanced display technologies. However, many methods for creating structural colors involve complex fabrication processes, limiting their practical application and integration on a device scale. One notable approach for producing structural colors is the metal–insulator–metal Fabry–Pérot (MIMFP) cavity, a fundamental and widely used design [11], [12]. An asymmetric variant of this cavity is the Salisbury screen (SS) structure, which features a reflective metal layer at the bottom, a dielectric layer in the middle, and a semi-transparent thin metal layer on top [13]. While static structural colors have achieved high-resolution, high-saturation, and wide-gamut color display applications [14], [15], [16], the potential advancements for dynamic color displays are highly appealing. The challenge lies in developing adjustable structural colors that can meet the high frame rate requirements necessary for dynamic color displays. Due to constraints in modulation mechanisms and methods, existing adjustable structural colors fall short in this regard. For instance, stimuli-responsive liquid crystals (LCs) can provide dynamic and reversible tunability with rapid color switching upon external stimuli [8]. However, the slow response time of LCs and their birefringence characteristics restrict their application in high-speed displays.
In the current studies, external stimuli for changing structural colors include electrical signals [8], [17], mechanical deformation [18], [19], temperature changes [6], [20], [21], and chemical interactions [22], [23]. Electrical tuning is a dynamic modulation method that is relatively easy to integrate and apply in practice. However, adjustable structural colors with fast response, low power dissipation, and long lifetime are still challenging. On the other hand, electrically driven phase change materials have sparked numerous innovations and breakthroughs in the field of tunable nanophotonics [8], [24], [25], [26]. Vanadium dioxide (VO2) is a phase change material that undergoes the reversible phase transition from an insulating to a metallic state near 68 °C, which significantly alters its optical properties [27], [28]. This transition can be triggered rapidly by an external stimuli such as heat [29], [30], [31], electric [25], [32], [33], [34], [35], [36], [37], and light [38], [39], allowing for high-speed and reversible changes in colors. These features make VO2 a promising material for advanced display technologies [20], [32], [40], smart windows [41], [42], [43], [44], switchable holography [30], and other applications requiring controllable and responsive color changes. Actively tunable structural colors hold great potential for applications in adaptive displays [45], smart textiles [46], [47], and sensors [21], [48], [49]. However, the huge thermal capacity often limits the high-speed modulation and power dissipation reduction of the device, and the ultrathin pixel structure faces significant challenges compared to passive structured color devices.
In this work, we demonstrate a strategy of tunable structural colors based on conformal VO2-coated Salisbury screen cavities with different resonant wavelengths. The hybrid structures are fabricated by employing an electron beam lithography (EBL)-based grayscale exposure and a conformal coating method to combine vanadium dioxide (VO2) and polymethyl methacrylate (PMMA) on a gold-coated ultra-thin silicon nitride (Si3N4) windows. Versatile structural colors formed by the freestanding hybrid nanostructures exhibit high performance of dynamic tunability when actuated by a low voltage of 1.4 V. Prototype applications of electrothermal color display and thermal-induced “converging” colors are demonstrated. The combination of a straightforward fabrication process, fast modulation speed, and low triggering power underscores the broad application prospects of this technology in dynamic color displays, which may be useful to achieve smooth display effects with high frame rates and provide an attractive solution for advanced display technologies.
2 Results
2.1 Structural design and nanofabrication
The schematic diagram of the structural color display and switching under external voltage stimulation is shown in Figure 1(a). The device consists of a VO2 layer on top, a PMMA layer in the middle, and a thick Au film on the bottom. Due to the ultralow thermal capacity of the freestanding multilayer film, the trigger voltage of the structure is only 1.4 V. During the transition from insulator to metal, the lattice structure of VO2 changes from monoclinic phase (M1-VO2) to rutile phase (R-VO2), as shown in the inset of Figure 1(a). To investigate the color-changing ability of VO2, the reflection spectra in both insulating and metallic states are simulated (see Supplementary Figure S1(b) and (c)). This simulation provides insights into the changes in the optical properties of VO2 during its phase transition, which is crucial for designing devices that leverage its dynamic tunable color capabilities. For example, by analyzing the reflection spectra, one can predict the colors produced during the phase change and optimize the VO2 layer for various applications in dynamic color display, which also enables the fine-tuning of the optical responses of VO2 to enhance the overall functionality and versatility of dynamic display technologies. Based on the simulated reflection spectra, Figure 1(b) and (c) plot the versatile colors and their transitions by varying the structural parameters. For the structure with 40-nm-thick VO2 and without the PMMA layer (marked by the stars), Figure 1(b) and (c) show that the change in refractive index during phase transition promotes the significant color change from blue to orange. Furthermore, color mixing is a common problem in structural colors, mainly due to nonresonant components and broad resonance peaks. Therefore, color purity is an important characterization parameter that represents the purity of color without mixing white light or other colors. This purity can be calculated by using the following equation [50]:
where (x, y), (x i , y i ), and (x d , y d ) represent the corresponding CIE coordinates of the samples, standard white light (0.310, 0.316), and dominant wavelength, respectively. The obtained orange purity of the nanostructure is 92 %. The color gamut of standard red-green-blue (sRGB) color space is marked by the black triangle. By adjusting the thickness of the VO2 and PMMA layers, the insulating and metallic VO2 phases can achieve a comprehensive color gamut, covering 41 % and 67 % of the sRGB color space, respectively, as shown in Figure 1(d). This tunability highlights the substantial potential of VO2 for use in applications requiring dynamic and vivid color displays.

Schematic of Salisbury screen (SS) cavities for color display. (a) Schematic diagram of the VO2/PMMA/Au structures, illustrating the structural design and operational principles based on SS cavities. The inset displays the transformation of lattice structures during the phase transition of VO2, with large vanadium atoms in purple and small oxygen atoms in yellow. (b) and (c) Simulated color plates of SS cavity with VO2 and PMMA layers of different thickness as noted, in the insulator phase and metal phase of VO2, respectively. Each color square is calculated based on the simulated reflection spectra of the structure. (d) Plot of the simulated colors of (b) and (c) in CIE 1931 space. The positions of the stars correspond to specific structures depicted in (b) and (c), indicating the structural parameters with the largest color changes.
To realize such structural designs, the fabrication process includes three main steps: electron beam (E-beam) deposition of gold, EBL grayscale exposure of nanopatterns, and VO2 film transfer [27]. The high excitation threshold is mainly due to the large thermal capacity of the hard substrate material. In this study, a suspended thin-film configuration with ultralow thermal capacity and small thickness is employed to reduce the excitation threshold. As shown in the schematic diagram of Figure 2(a), a 100-nm-thick gold electrode is lithographically patterned on the Si3N4 window to suppress light transmission (the microscope image of the electrode pattern is shown in Figure S2(a). Then, the lossless PMMA films of different thickness are prepared under different exposure doses on the top of the Au layer. PMMA is chosen since it is a high-resolution electron beam resist with high sensitivity to electron beam dose and is thus desirable for applications with different thicknesses. Finally, the VO2 film was conformally transferred onto the PMMA layer to form the SS cavity with different spacer thicknesses (d PMMA). The freestanding VO2 film before transfer is shown in Figure S2(b) and (c). The transfer method is primarily employed because the performance of VO2 film is significantly influenced by the substrate. Additionally, the high temperatures necessary for the crystallization of the VO2 phase, typically exceeding 500 °C, can damage PMMA. This makes it unsuitable for direct deposition on PMMA polymer materials. To precisely control the height of the nanostructure, the relationship between the electron beam dose and the remaining PMMA thickness is measured by the stylus profiler. The resolved data in Figure 2(b) indicate that the thickness of the PMMA decreases gradually when the electron beam dose is more than 40 μC/cm2 and decreases significantly to 0 nm when the dose reaches ∼90 μC/cm2. To enhance the hydrophilicity of the PMMA surface prior to VO2 film transfer, we treated it with oxygen plasma. As demonstrated in Figure 2(b), after PMMA is treated with low-power oxygen plasma for 10 s, the PMMA thickness remains nearly unchanged (decreases by an average of 1.1 nm), resulting in negligible effect on the resonance wavelength of the FP cavity. Such a method allows us to fine-tune the structural parameters, ensuring precise control over the optical properties of the device while maintaining the integrity of the PMMA layer.

Fabrication method of gray scale exposure. (a) Flowchart of the grayscale exposure with EBL and conformal coating with VO2. (b) Thickness of PMMA as a function of the exposure dose. (c) Topographic image measured by AFM, showing square structures with different height, which clearly shows the conformal coating. (d) Captured images of the square structures with different height when temperature varies from 25 to 80 °C.
The local height of VO2 film after transferred onto the PMMA is further precisely characterized by atomic force microscopy (AFM), as the images shown in Figure 2(c). It can be seen that from the first row to the fifth row, the local contrast increases gradually, indicating a decreased PMMA thickness in the square and confirming the conformal coating of VO2 with PMMA after the transfer process, which is critical for the high-resolution color displays. As a typical stimulus-responsive phase change material, VO2 shows the obvious response to external electric, thermal, and light stimulation, making it one of the most promising materials for tunable structural colors. To test this capability, the hybrid structures are heated by a hot and cold stage and the changes in structural colors with temperature are plotted in Figure 2(d), showing dynamically tunable structural colors by adjusting the applied temperature. In addition, we observed that the VO2/PMMA/Au multilayer structure exhibits continuous color tunability from room temperature to 80 °C.
2.2 Electrothermally tunable structural colors
The asymmetric SS MIMFP cavity is very desirable to enhance the structural color performance by leveraging the principles of interference. In this design, multiple reflections between parallel surfaces create standing wave patterns that selectively reflect specific wavelengths, thereby producing bright colors. The thickness of the thin metal layer on top of the SS cavity is less than the skin depth of the incident light, allowing part of the light to penetrate the cavity. Figure 3(a) presents the schematic design of our reflective device, where the color change in the SS structure is attributed to the VO2 phase transition. Based on the SS cavity, by controlling the EBL exposure process, we can adjust the thickness of the PMMA in the middle layer to adjust the length of the SS cavity and thus regulate the displayed colors. The coherent accumulation of the reflection in the FP cavity satisfies the resonant wavelength condition [51], which can be expressed as:
where n is the effective refractive index of dielectric layers consisting of VO2, d PMMA is the thickness of PMMA layers, θ is the incident angle, and φ 1 and φ 2 represent the phase shift of the reflection coefficients at the two interfaces of the FP cavity.

Dynamically tunable structural colors achieved through electric stimulus. (a) Schematic of the working principle of the SS cavity structure before and after the phase change of VO2. (b) Color images of the squares with PMMA of different the thickness as noted at the applied voltage of 0 and 1.4 V. The thickness of the VO2 thin film is 120 nm. The blue solid frame and red dotted line frame represent the insulating and metal phases of VO2, respectively. ΔE in the right column means the color difference. (c) Measured and simulated reflection spectra of SS cavity that consist of a 120-nm-thick VO2 film under different d PMMA. Both spectra exhibit a red shift as d PMMA increases, as indicated by the gray dashed arrows. (d) Measured structural colors in CIE 1931 space when the thickness of VO2 is 40, 80, and 120 nm under d PMMA = 250 nm.
In this model, when VO2 undergoes a phase transition by external stimuli, its refractive index will experience a significant change, making VO2 highly promising for applications in tunable color displays. To validate this point, a voltage is applied to the structure, which induces thermal effects to control the phase state. In such a case, the structural colors could be dynamically tunable by adjusting the applied voltage. To achieve different color palettes in the visible wavelength range, we systematically varied the thickness of PMMA and VO2. The initial colors correspond to the structures with PMMA film of different thickness, as shown in the left column of Figure 3(b), which are significantly changed under a low voltage of 1.4 V (the middle column of Figure 3(b)). To better illustrate the color differences before and after the phase transition, the color difference ΔE is calculated by using the following formula:
where ΔL *, Δa *, and Δb * represent the difference between two colors on the three axes of L * (lightness of the color), a * (from green to red), and b * (from blue to yellow), respectively. As the results are shown in Figure 3(b), for the cavity with a thickness of approximately 90 nm, the ΔE value in the CIE Lab color space can reach 81. For structures with PMMA thicknesses of 70, 150, and 250 nm, the color difference exceeds 60, making the changes easily distinguishable by the human eye. Meanwhile, Figure 3(c) shows the reflection spectra for the PMMA films with thicknesses from 5 to 250 nm when VO2 is in the insulating and metal phases, respectively, where the experimental and simulated results are well consistent. The small deviation observed between the simulation and experimental results can be largely attributed to the inaccurate refractive index of VO2 thin film in simulations compared to its realistic value, as well as the experimental imperfections during measurements. It should be mentioned that by controlling the thickness of the top VO2 layer, one can adjust the loss within the SS cavity, thereby regulating the resultant color changes. Such a capability is well demonstrated by our experiments with VO2 of different thickness (40 and 80 nm in Supplementary Figures S3 and S4, respectively). As illustrated in Figure 3(d), the three data sets presented in the CIE 1931 color space demonstrate that under the same situation, the thicker the VO2 film, the larger the color change.
2.3 Performances of tunable colors
The sensitivity, speed, lifespan, and wide-view angle of electrically tunable structural colors are critical for device applications. To evaluate the sensitivity, Figure 4(a) plots the measured voltage-dependent reflection intensity of the transferred VO2 film at wavelength 729 nm. It can be seen that at a voltage of 0 V, the reflected light intensity is around 16,000. When the applied voltage increases to 1.4 V, the light intensity drops significantly to 1,200. At the critical voltage (V c) of 1 V, VO2 undergoes a fast transition from the insulator phase to the metal phase. Such a significant change results from the changes in the refractive index of VO2 during the phase transition, which causes changes in both spectral peaks and intensities, as shown by the voltage-dependent spectra in Figure 4(b) (also see simulation results in Supplementary Figure S1(c)). Specifically, with the increase of applied voltage, the reflection peak is gradually blue-shifted. Further evidence is provided in Figure 4(c), which illustrates the correlation between the peak wavelength and voltage. Within the voltage range of 1.1–1.4 V, the wavelength changes by 89.8 nm. Here, the sensitivity can be defined by:
where Δλ represents the change of resonance wavelength and ΔV represents the change of relative voltage. In such a case, the average sensitivity of the sample is calculated to be 299.3 nm/V.

The tunable performance and the dynamic response of the SS cavities. (a) Reflection intensity of incident light at wavelength 729 nm versus external voltage. (b) Reflection spectra of VO2/PMMA/Au cavity with the applied voltage increasing from 0 to 1.4 V. (c) Resonance wavelength of the reflection spectra in (b) versus the applied voltage with d PMMA = 250 nm. (d) Measured signal intensity versus time when the voltage is switched off-on-off at 1.4 V, showing a fast rising and fall time of ∼0.2 ms. (e) Measured reflectance at λ = 700 nm for ∼1,000,000 cycles of switching between 0 and 1 V. (f) Statistics and comparison of the response time and power consumption of electrically driven VO2 nanostructure, film, and GST devices in literatures and this work.
In addition, to measure the switching speed of the VO2-based tunable structural colors, the voltage at which each device is fully actuated is applied. As displayed in Figure 4(d), the preset structure initially displays a blue color, which then turns pink within 0.2 ms when a voltage of 1.4 V (with current 0.3 A) is applied. Once the applied voltage is removed, the color reverts to blue. Such a color change is controlled through changing the bias current, which consumes power of no more than 0.42 W. These results indicate that the ultralow thermal capacity of the freestanding VO2/PMMA/Au film effectively reduces the modulated voltage and response time. The rapid switching speed and low power consumption of this structure underscore its potential for advanced applications in dynamic displays and other optical technologies where fast and efficient color modulation is essential.
Furthermore, the durability of the structure is tested by applying an alternating voltage with a frequency of 50 Hz in the ambient environment. As shown in Figure 4(e), the reflection appears to be very stable under two states throughout one million switching cycles. Figure S5 shows microscope images of the pinwheel pattern on the first day (Figure S5(a)) and after 180-day storage (Figure S5(b)) in air. There is no obvious difference between the colors of the metallic phase and the insulating phase of the VO2/PMMA/Au structure, demonstrating excellent stable performance in air. The performance of the developed electrically controlled VO2 structural color device is further compared with several recently developed electrically controlled structures, as shown in Figure 4(f). It can be seen that compared with other VO2 nanostructures, VO2 thin films, and Ge2Sb2Te5 (GST) [52], the suspended VO2 nanostructure here exhibits advantages regarding both response time and power consumption.
It should be mentioned that various visual experiences and scenarios often require consistent color displays from wide angles. To investigate the angle-dependent performance of the color displays, the VO2/PMMA/Au nanostructures with different thickness are characterized. As shown in Figure 5(a), the microscopic images of the nanostructures show nearly the same colors when at observation angles θ = 0° and θ = 35° at each thickness. This is also verified by the measured angular reflection spectra (0–35°) in Figure 5(b) (d PMMA = 250 nm, d VO2 = 120 nm), demonstrating that the displayed colors are consistent with only slight changes at angles from 0° to 35°. When VO2 is in the insulating phase, the resonance valley of the reflection spectrum blue-shifts from 602 nm (0°) to 584 nm (35°), and the relative variation of resonant wavelength (Δλ R/λ 0) is only 2.99 %. As for the metallic phase VO2, the resonance valley blue-shifts from 557 nm (0°) to 548 nm (35°), and the relative variation is reduced to 1.62 %. This small shift in resonant wavelength does not cause a significant change in color perception. Therefore, unless the observation angle is large enough (for example >40°), the angle-dependent shift in resonant wavelength and the resulting change in color remain small, which is verified by the simulation results in Figure S6 (see more details in Supporting Information). Such characteristic makes the nanostructures suitable for practical applications where light sources may impinge at various angles, thereby enhancing their applicability in real-world dynamic color display technologies.

Angle-dependent performance of the nanostructure and its applications in color displays. (a) and (b) Measured images and reflection spectra of hybrid structures with PMMA of different thickness under different observation angles (θ) in the insulating and metal phase of VO2, respectively. (c) SEM image of a “BIT” pattern prepared after the transfer of VO2 film. (d) Optical images of “BIT” patterns at the insulating (right) and metal (left) phase of VO2 when turned on and off at 1.4 V. Structural parameters: d PMMA = 150 nm, d VO2 = 120 nm. (e) Optical images of a pinwheel pattern obtained by grayscale exposure before and after the phase change actuate by external voltage of 1.4 V. (f) Optical images of two squares with PMMA thickness of 250 and 240 nm in the background (BG) area and pattern (PAT) area, respectively (d VO2 = 120 nm). The initially different colors “converge” to the same color after the phase change. (g) Temperature-dependent evolution of the color of the pinwheel pattern when temperature increases from 25 to 70 °C, where the contrast of the image decreases and finally the image is “erased.”
2.4 Applications in color displays
The ability of the grayscale exposure to independently and precisely adjust pixel heights allows for the production of structured color graphics. In this case, Figure 5(c) shows the scanning electron microscope (SEM) image of a sample with the pattern “BIT,” where the thickness of PMMA and VO2 are 150 and 120 nm, respectively, and the background area remains unexposed. Such a pattern shows stable color changes between 0 and 1.4 V when the phase transition of VO2 occurs, as shown in Figure 5(d). In addition, by varying the electron beam exposure dose, patterns with PMMA of different thickness can be designed. As shown in Figure 5(e), a colorful pinwheel is formed, with four blades showing their own structural colors. As the phase change occurs, the color of each blade in the pinwheel pattern dynamically changes, showing the versatility of structured color graphics achieved through this approach.
Another interesting phenomenon is that for some structures with different initial colors (with PMMA of 10-nm difference in thickness) at the insulator state, the same colors are displayed after the phase change at the metal state, as shown in Figure 5(f). Since VO2 is sensitive to temperature, such a “convergence” effect in the color display can be employed for information erasing purposes or to monitor the temperature of the environment. For example, in Figure 5(g), the contrast of the image decreases with the increase of the temperature and the image is finally invisible when the temperature is higher than 70 °C. Therefore, by carefully controlling the PMMA thickness and exposure dose, the method not only achieves dynamic color tuning for aesthetic and decorative purposes but also enables functionalities such as anticounterfeiting, information erasing, and information encryption. This precise control over structural colors underscores the versatility of our approach, making it suitable for a wide range of applications beyond conventional display technologies. It should be mentioned that with the high dielectric constant of the spacer layer, the lossless dielectric materials could produce stronger effects in optical interferences, resulting in more vivid and saturated colors. In the future, by applying etching-based fabrication methods on the spacing materials, the high-dielectric-constant materials can be integrated into these systems. Therefore, this approach not only offers greater flexibility in material selection but also enables the combination of new materials. Such advancements may enable novel optical properties and functionalities for various applications.
3 Conclusions
In conclusion, a simple nanofabrication method with EBL-based grayscale exposure and VO2-based conformal coating was demonstrated. The VO2 layers were coated onto the PMMA spacer layers of variable thickness, with gold films in the bottom to create SS cavities with different resonant wavelengths. Versatile VO2-based structural colors were constructed, exhibiting high performance of dynamically tunable properties, such as a short switching time of 0.2 ms when actuated by a low voltage of 1.4 V. The applications of color displays and information erasing were briefly demonstrated. These findings highlight the great potential of our approach for dynamic tuning of structural colors and information encryption applications, offering a wide range of possibilities for development in these fields.
4 Methods
4.1 Numerical modeling
The optical responses of the structures were simulated by using the finite element method. Optical reflection spectra were obtained under light incidence propagating along the z-axis direction. Periodic boundary conditions were applied on a unit cell in the x and y directions. The optical properties of Au and Si3N4 were obtained using data from Johnson and Christy and Vogt, respectively. The refractive index of PMMA was set to 1.49. Additionally, Figure S1(a) presents the wavelength-dependent refractive index (n) and absorption coefficient (k) of VO2 thin films in both insulating and metallic phases, as measured by ellipsometry.
4.2 Sample fabrications
4.2.1 Synthesis of the VO2 film on SiO2 substrate
To prepare the freestanding VO2 films, magnetron sputtering was firstly employed to deposit high-purity vanadium to form the VO x layer on the SiO2 substrate. Sputtering was performed at 0.6 Pa with a flowing gas mixture (30 sccm Ar and 20 sccm Ar/O2 mixture (O2:4 %), DC power of 55 W) at 25 °C. After VO x deposition, the VO x film was annealed in a low-pressure oxygen atmosphere (4.5 Pa) at 450 °C for 10 min.
4.2.2 Device fabrications
To prepare the SS cavity based on VO2/PMMA/Au film, a layer of 100-nm-thick Au electrodes was first lithographically patterned and deposited on the Si3N4 window (Beijing ZXBR Technology Co. Ltd, 100 nm) with a window area of 400 × 400 µm by using E-beam evaporation. After the residual metal was lifted off in NMP at 60 °C, the PMMA (Allresist, 950 K) resist was then spin-coated (6,000 rpm, 30 s) with a thickness of 250 nm. Subsequently, the patterns with different heights were exposed under varied doses by EBL (20 kV, 0.3 nA, 20 μm aperture), while the irradiation dose can be selected from 0 to 120 μC/cm2. The samples were then immersed in MIBK solution at 10 °C for 40 s to remove the exposed resist. After that, the PMMA film was treated by oxygen plasma (100 sccm O2 for 10 s, power of 50 W) to enhance its hydrophilicity. As shown in Figure S2, a buffered oxide etch (BOE, 5:1 vol ratio of 40 % NH4F to 49 % HF) solution was used to wet etch the VO2/SiO2 interface to obtain a freestanding VO2 film. After etching the VO2/SiO2 interface layer, the VO2 layer was separated from the SiO2 substrate and transferred to deionized water. Finally, the released VO2 film was conformally transferred onto the PMMA surface, as verified by the AFM characterizations.
4.3 Material characterizations
A home-built optical system was utilized to measure the reflection spectra. The stabilized tungsten-halogen light source (360–2,600 nm, HL-2000, Ocean Optics) was launched into the microscopic system through a fiber with a core diameter of 100 μm. An objective lens (×10 with NA = 0.3, Olympus) was used to focus the incident light onto the sample. The reflected light in the normal direction was collected by the same objective and delivered to a spectrometer (400–1,000 nm, PG2000-Pro, Ideaoptics). Angle-resolved reflectance was characterized by the micro-angle-resolved spectroscopy system (−60°–60°, ARMS, Ideaoptics). An optical microscope (BX FM, Olympus) was used to take the optical images where the phase change was controlled by a hot plate and a source meter. Sample temperatures could be tuned precisely by a hot and cold stage (20–600 °C, HCS621GXY, INSTEC). For the electric control test, the DC voltage was supplied by a source meter (2,450, Keithley). The dynamic time response of the reflection intensity was obtained from the output of the biased Si detectors (200–1,100 nm, DET10A2, Thorlabs) and recorded by an oscilloscope (TBS2000 Series, Tektronix). AFM imaging was performed by an atomic force microscope (WITec, alpha 300).
Supporting Information
See the supplementary material for additional information.
Funding source: National Key Research and Development Program of China
Award Identifier / Grant number: 2023YFB3208700
Funding source: Beijing Natural Science Foundation
Award Identifier / Grant number: Z2006E01202301
Funding source: National Natural Science Foundation of China
Award Identifier / Grant number: 51802008
Award Identifier / Grant number: 52272286
Award Identifier / Grant number: 62375016
Award Identifier / Grant number: T2325005
Funding source: Science and Technology Project of Guangdong
Award Identifier / Grant number: 2020B010190001
Acknowledgments
The authors thank the Analysis and Testing Center of Beijing Institute of Technology and the Information Optoelectronics Research Institute of Beijing University of Technology for their assistance in facility support.
-
Research funding: This work was supported by the National Natural Science Foundation of China (under Grant Nos. T2325005, 62375016, 51802008, 52272286), Beijing Natural Science Foundation (Grant No. Z2006E01202301), National Key R&D Program of China (Grant No. 2023YFB3208700), and Science and Technology Project of Guangdong (2020B010190001).
-
Author contributions: JL, HM, and XZ conceived the idea. HM and XZ fabricated and characterized the SS cavity. HS, YL, QL, and YZ assisted in the fabrication process. XZ, JH, and HM designed and performed the spectral and dynamic response experiments. XZ and HM analyzed and interpreted the data. XZ performed the numerical simulations. XZ wrote the manuscript. JL and HM revised the manuscript and supervised the whole process. All authors have accepted responsibility for the entire content of this manuscript and approved its submission.
-
Conflict of interest: Authors state no conflicts of interest.
-
Data availability: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.
References
[1] S. Kinoshita, S. Yoshioka, and J. Miyazaki, “Physics of structural colors,” Rep. Prog. Phys., vol. 71, no. 7, 2008, Art. no. 076401. https://10.1088/0034-4885/71/7/076401 10.1088/0034-4885/71/7/076401Suche in Google Scholar
[2] Y. Fu, C. A. Tippets, E. U. Donev, and R. Lopez, “Structural colors: from natural to artificial systems,” Wiley Interdiscip. Rev.:Nanomed. Nanobiotechnol., vol. 8, no. 5, pp. 758–775, 2016. https://10.1002/wnan.1396 10.1002/wnan.1396Suche in Google Scholar PubMed
[3] X. Liu, Z. Huang, and J. Zang, “All-dielectric silicon nanoring metasurface for full-color printing,” Nano Lett., vol. 20, no. 12, pp. 8739–8744, 2020. https://10.1021/acs.nanolett.0c03596 10.1021/acs.nanolett.0c03596Suche in Google Scholar PubMed
[4] Y. Wenhong, et al.., “All-dielectric metasurface for high-performance structural color,” Nat. Commun., vol. 11, no. 1, p. 1864, 2020. https://10.1038/s41467-020-15773-0 10.1038/s41467-020-15773-0Suche in Google Scholar PubMed PubMed Central
[5] G. Geng, et al.., “Height‐gradiently‐tunable nanostructure arrays by grayscale assembly nanofabrication for ultra‐realistic imaging,” Laser Photonics Rev., vol. 17, no. 9, 2023. https://10.1002/lpor.202300073.10.1002/lpor.202300073Suche in Google Scholar
[6] D. T. Yimam, M. Liang, J. Ye, and B. J. Kooi, “3D nanostructuring of phase‐change materials using focused ion beam towards versatile optoelectronics applications,” Adv. Mater., vol. 36, no. 23, 2023, Art. no. 2303502. https://10.1002/adma.202303502 10.1002/adma.202303502Suche in Google Scholar PubMed
[7] S. Jana, et al.., “Aperiodic bragg reflectors for tunable high-purity structural color based on phase change material,” Nano Lett., vol. 24, no. 13, pp. 3922–3929, 2024. https://10.1021/acs.nanolett.4c00052 10.1021/acs.nanolett.4c00052Suche in Google Scholar PubMed
[8] M. Huang, et al.., “Dynamically tunable structural colors enabled by pixelated programming of soft materials on thickness,” Adv. Opt. Mater., vol. 11, no. 19, 2023, Art. no. 2300573. https://10.1002/adom.202300573 10.1002/adom.202300573Suche in Google Scholar
[9] H. Wei, et al.., “Tunable VO2 cavity enables multispectral manipulation from visible to microwave frequencies,” Light: Sci. Appl., vol. 13, no. 1, p. 54, 2024. https://10.1038/s41377-024-01400-w 10.1038/s41377-024-01400-wSuche in Google Scholar PubMed PubMed Central
[10] C. U. Hail, G. Schnoering, M. Damak, D. Poulikakos, and H. Eghlidi, “A plasmonic painter’s method of color mixing for a continuous red-green-blue palette,” ACS Nano, vol. 14, no. 2, pp. 1783–1791, 2020. https://10.1021/acsnano.9b07523 10.1021/acsnano.9b07523Suche in Google Scholar PubMed
[11] S. D. Rezaei, et al.., “Tunable, cost‐effective, and scalable structural colors for sensing and consumer products,” Adv. Opt. Mater., vol. 7, no. 20, 2019, Art. no. 1900735. https://10.1002/adom.201900735 10.1002/adom.201900735Suche in Google Scholar
[12] T. Guo, J. Evans, N. Wang, and S. He, “Monolithic chip-scale structural color filters fabricated with simple UV lithography,” Opt. Express, vol. 27, no. 15, pp. 21646–21651, 2019. https://10.1364/OE.27.021646 10.1364/OE.27.021646Suche in Google Scholar PubMed
[13] Y.-L. Meng, et al.., “Salisbury screen optical color filter with ultra-thin titanium nitride film,” Appl. Opt., vol. 58, no. 24, pp. 6700–6705, 2019. https://10.1364/ao.58.006700 10.1364/AO.58.006700Suche in Google Scholar PubMed
[14] C. Tan, et al.., “Deep-subwavelength multilayered meta-coatings for visible-infrared compatible camouflage,” Nanophotonics, vol. 13, no. 13, pp. 2391–2400, 2024. https://10.1515/nanoph-2024-0029 10.1515/nanoph-2024-0029Suche in Google Scholar PubMed PubMed Central
[15] E. Heydari, J. R. Sperling, S. L. Neale, and A. W. Clark, “Plasmonic color filters as dual‐state nanopixels for high‐density microimage encoding,” Adv. Funct. Mater., vol. 27, no. 35, 2017, Art. no. 1701866. https://10.1002/adfm.201701866 10.1002/adfm.201701866Suche in Google Scholar
[16] H. Pan, et al.., “Wide gamut, angle-insensitive structural colors based on deep-subwavelength bilayer media,” Nanophotonics, vol. 9, no. 10, pp. 3385–3392, 2020. https://10.1515/nanoph-2020-0106 10.1515/nanoph-2020-0106Suche in Google Scholar
[17] T. Guo, et al.., “Durable and programmable ultrafast nanophotonic matrix of spectral pixels,” Nat. Nanotechnol., vol. 19, no. 11, pp. 1635–1643, 2024.https://10.1038/s41565-024-01756-5 10.1038/s41565-024-01756-5Suche in Google Scholar PubMed PubMed Central
[18] G. Chen and W. Hong, “Mechanochromism of structural‐colored materials,” Adv. Opt. Mater., vol. 8, no. 19, 2020, Art. no. 2000984. https://10.1002/adom.202000984 10.1002/adom.202000984Suche in Google Scholar
[19] M. A. Haque, T. Kurokawa, G. Kamita, Y. Yue, and J. P. Gong, “Rapid and reversible tuning of structural color of a hydrogel over the entire visible spectrum by mechanical stimulation,” Chem. Mater., vol. 23, no. 23, pp. 5200–5207, 2011. https://10.1021/cm2021142 10.1021/cm2021142Suche in Google Scholar
[20] X. Duan, et al.., “Reconfigurable multistate optical systems enabled by VO2 phase transitions,” ACS Photonics, vol. 7, no. 11, pp. 2958–2965, 2020. https://10.1021/acsphotonics.0c01241 10.1021/acsphotonics.0c01241Suche in Google Scholar PubMed PubMed Central
[21] J. Han, K. Lin, H. Lin, K.-T. Lau, and B. Jia, “Tunable thermochromic graphene metamaterials with iridescen color,” Nano Lett., vol. 22, no. 11, pp. 6026–6033, 2022. https://10.1021/acs.nanolett.1c04768 10.1021/acs.nanolett.1c04768Suche in Google Scholar PubMed
[22] I. Kim, et al.., “Structural color switching with a doped indium-gallium-zinc-oxide semiconductor,” Photonics Res., vol. 8, no. 9, pp. 1409–1415, 2020. https://10.1364/prj.395749 10.1364/PRJ.395749Suche in Google Scholar
[23] X. Duan, S. Kamin, and N. Liu, “Dynamic plasmonic colour display,” Nat. Commun., vol. 8, no. 1, 2016, Art. no. 14606. https://10.48550/arXiv.2105.02647 10.1038/ncomms14606Suche in Google Scholar PubMed PubMed Central
[24] W. Pang, et al.., “Electro-mechanically controlled assembly of reconfigurable 3D mesostructures and electronic devices based on dielectric elastomer platforms,” Natl. Sci. Rev., vol. 7, no. 2, pp. 342–354, 2020. https://10.1093/nsr/nwz164 10.26226/morressier.5f5f8e69aa777f8ba5bd6085Suche in Google Scholar
[25] Y. G. Jeong, et al.., “Electrical control of terahertz nano antennas on VO2 thin film,” Opt. Express, vol. 19, no. 22, pp. 21211–21215, 2011. https://10.1364/OE.19.021211 10.1364/OE.19.021211Suche in Google Scholar PubMed
[26] S. Chen, et al.., “Electromechanically reconfigurable optical nano-kirigami,” Nat. Commun., vol. 12, no. 1, p. 1299, 2021. https://10.1038/s41467-021-21565-x 10.1038/s41467-021-21565-xSuche in Google Scholar PubMed PubMed Central
[27] H. Ma, et al.., “Wafer-Scale freestanding vanadium dioxide film,” Sci. Adv., vol. 7, no. 50, p. eabk3438, 2021. https://10.1126/sciadv.abk3438 10.1126/sciadv.abk3438Suche in Google Scholar PubMed PubMed Central
[28] C. Xu, et al.., “Transient dynamics of the phase transition in VO2 revealed by mega-electron-volt ultrafast electron diffraction,” Nat. Commun., vol. 14, no. 1, p. 1265, 2023. https://10.1038/s41467-023-37000-2 10.1038/s41467-023-37000-2Suche in Google Scholar PubMed PubMed Central
[29] A. M. Boyce, et al.., “Actively tunable metasurfaces via plasmonic nanogap cavities with sub-10-nm VO2 films,” Nano Lett., vol. 22, no. 9, pp. 3525–3531, 2022. https://10.1021/acs.nanolett.1c04175 10.1021/acs.nanolett.1c04175Suche in Google Scholar PubMed PubMed Central
[30] Y. Liao, Y. Fan, and D. Lei, “Thermally tunable binary-phase VO2 metasurfaces for switchable holography and digital encryption,” Nanophotonics, vol. 13, no. 7, pp. 1109–1117, 2024. https://10.1515/nanoph-2023-0824 10.1515/nanoph-2023-0824Suche in Google Scholar PubMed PubMed Central
[31] C. Wan, et al.., “Switchable induced-transmission filters enabled by vanadium dioxide,” Nano Lett., vol. 22, no. 1, pp. 6–13, 2021. https://10.1021/acs.nanolett.1c02296 10.1021/acs.nanolett.1c02296Suche in Google Scholar PubMed
[32] H. Ma, et al.., “A flexible, multifunctional, active terahertz modulator with an ultra-low triggering threshold,” J. Mater. Chem. C, vol. 8, no. 30, pp. 10213–10220, 2020. https://10.1039/d0tc02446e 10.1039/D0TC02446ESuche in Google Scholar
[33] J. H. Shin, S. P. Han, M. Song, and H. C. Ryu, “Gradual tuning of the terahertz passband using a square-loop metamaterial based on A W-doped VO2 thin film,” Appl. Phys. Express, vol. 12, no. 3, 2019, Art. no. 032007. https://10.7567/1882-0786/ab0395 10.7567/1882-0786/ab0395Suche in Google Scholar
[34] Z. Zhu, P. G. Evans, R. F. Haglund, and J. G. Valentine, “Dynamically reconfigurable metadevice employing nanostructured phase-change materials,” Nano Lett., vol. 17, no. 8, pp. 4881–4885, 2017. https://10.1021/acs.nanolett.7b01767 10.1021/acs.nanolett.7b01767Suche in Google Scholar PubMed
[35] L. Liu, L. Kang, T. S. Mayer, and D. H. Werner, “Hybrid metamaterials for electrically triggered multifunctional control,” Nat. Commun., vol. 7, no. 1, 2016, Art. no. 13236. https://10.1038/ncomms13236 10.1038/ncomms13236Suche in Google Scholar PubMed PubMed Central
[36] Y. Kim, et al.., “Phase modulation with electrically tunable vanadium dioxide phase-change metasurfaces,” Nano Lett., vol. 19, no. 6, pp. 3961–3968, 2019. https://10.1021/acs.nanolett.9b01246 10.1021/acs.nanolett.9b01246Suche in Google Scholar PubMed
[37] A. Crunteanu, et al.., “Voltage- and current-activated metal–insulator transition in VO2-based electrical switches: a lifetime operation Analysis,” Sci. Technol. Adv. Mater., vol. 11, no. 6, 2016. https://10.1088/1468-6996/11/6/065002.10.1088/1468-6996/11/6/065002Suche in Google Scholar PubMed PubMed Central
[38] T. Wang, D. Torres, F. E. Fernández, C. Wang, and N. Sepúlveda, “Maximizing the performance of photothermal actuators by combining smart materials with supplementary advantages,” Sci. Adv., vol. 3, no. 4, p. e1602697, 2017. https://10.1126/sciadv.1602697 10.1126/sciadv.1602697Suche in Google Scholar PubMed PubMed Central
[39] K. Liu, C. Cheng, Z. Cheng, K. Wang, R. Ramesh, and J. Wu, “Giant-amplitude, high-work density microactuators with phase transition activated nanolayer bimorphs,” Nano Lett., vol. 12, no. 12, pp. 6302–6308, 2012. https://10.1021/nl303405g 10.1021/nl303405gSuche in Google Scholar PubMed
[40] F. Z. Shu, et al.., “Dynamic plasmonic color generation based on phase transition of vanadium dioxide,” Adv. Opt. Mater., vol. 6, no. 7, 2018, Art. no. 1700939. https://10.1002/adom.201700939 10.1002/adom.201700939Suche in Google Scholar
[41] X. Li, et al.., “Self-rolling of vanadium dioxide nanomembranes for enhanced multi-level solar modulation,” Nat. Commun., vol. 13, no. 1, p. 7819, 2022. https://10.1038/s41467-022-35513-w 10.1038/s41467-022-35513-wSuche in Google Scholar PubMed PubMed Central
[42] K. Yujie, et al.., “Two-dimensional SiO2/VO2 photonic crystals with statically visible and dynamically infrared modulated for smart window deployment,” ACS Appl. Mater. Interfaces, vol. 8, no. 48, pp. 33112–33120, 2016. https://10.1021/acsami.6b12175 10.1021/acsami.6b12175Suche in Google Scholar PubMed
[43] S. Chen, et al.., “Gate-controlled VO2 phase transition for high-performance smart windows,” Sci. Adv., vol. 5, no. 3, p. eaav6815, 2019. https://10.48550/arXiv.1810.00942 10.1126/sciadv.aav6815Suche in Google Scholar PubMed PubMed Central
[44] L. Xiao, et al.., “Fast adaptive thermal camouflage based on flexible VO2/graphene/CNT thin films,” Nano Lett., vol. 15, no. 12, pp. 8365–8370, 2015. https://10.1021/acs.nanolett.5b04090 10.1021/acs.nanolett.5b04090Suche in Google Scholar PubMed
[45] K. Yin, et al.., “Advanced liquid crystal devices for augmented reality and virtual reality displays: principles and applications,” Light: Sci. Appl., vol. 11, no. 1, p. 161, 2022. https://10.1038/s41377-022-00851-3 10.1038/s41377-022-00851-3Suche in Google Scholar PubMed PubMed Central
[46] A. Libanori, G. Chen, X. Zhao, Y. Zhou, and J. Chen, “Smart textiles for personalized healthcare,” Nat. Electron., vol. 5, no. 3, pp. 142–156, 2022. https://10.1038/s41928-022-00723-z 10.1038/s41928-022-00723-zSuche in Google Scholar
[47] H. W. Choi, et al.., “Smart textile lighting/display system with multifunctional fibre devices for large scale smart home and iot applications,” Nat. Commun., vol. 13, no. 1, p. 814, 2022. https://10.1038/s41467-022-28459-6 10.1038/s41467-022-28459-6Suche in Google Scholar PubMed PubMed Central
[48] Y. Luo, et al.., “Technology roadmap for flexible sensors,” ACS Nano, vol. 17, no. 6, pp. 5211–5295, 2023. https://10.1021/acsnano.2c12606 Suche in Google Scholar
[49] K. Meng, et al.., “Wearable pressure sensors for pulse wave monitoring,” Adv. Mater., vol. 34, no. 21, 2022, Art. no. 2109357. https://10.1002/adma.202109357 10.1002/adma.202270158Suche in Google Scholar
[50] M. Bai, et al.., “General synthesis of layered rare-earth hydroxides (RE = Sm, Eu, Gd, Tb, Dy, Ho, Er, Y) and direct exfoliation into monolayer nanosheets with high color purity,” J. Phys. Chem. Lett., vol. 12, no. 41, pp. 10135–10143, 2021. https://10.1021/acs.jpclett.1c03047 10.1021/acs.jpclett.1c03047Suche in Google Scholar PubMed
[51] J. Zhao, et al.., “Defining deep‐subwavelength‐resolution, wide‐color‐gamut, and large‐viewing‐angle flexible subtractive colors with an ultrathin asymmetric fabry–perot lossy cavity,” Adv. Opt. Mater., vol. 7, no. 23, 2019, Art. no. 1900646. https://10.1002/adom.201900646 10.1002/adom.201900646Suche in Google Scholar
[52] J. Zheng, et al.., “Nonvolatile electrically reconfigurable integrated photonic switch enabled by a silicon PIN diode heater,” Adv. Mater., vol. 32, no. 31, 2020, Art. no. 2001218. https://10.1002/adma.202001218 10.1002/adma.202001218Suche in Google Scholar PubMed
Supplementary Material
This article contains supplementary material (https://doi.org/10.1515/nanoph-2024-0546).
© 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
- Editorial
- Special issue: “Metamaterials and Plasmonics in Asia”
- Reviews
- All-optical analog differential operation and information processing empowered by meta-devices
- Metasurface-enhanced biomedical spectroscopy
- Topological guided-mode resonances: basic theory, experiments, and applications
- Letter
- Ultrasensitive circular dichroism spectroscopy based on coupled quasi-bound states in the continuum
- Research Articles
- Data-efficient prediction of OLED optical properties enabled by transfer learning
- Semimetal–dielectric–metal metasurface for infrared camouflage with high-performance energy dissipation in non-atmospheric transparency window
- Deep-subwavelength engineering of stealthy hyperuniformity
- Tunable structural colors based on grayscale lithography and conformal coating of VO2
- A general recipe to observe non-Abelian gauge field in metamaterials
- Free-form catenary-inspired meta-couplers for ultra-high or broadband vertical coupling
- Enhanced photoluminescence of strongly coupled single molecule-plasmonic nanocavity: analysis of spectral modifications using nonlocal response theory
- Spectral Hadamard microscopy with metasurface-based patterned illumination
- Tunneling of two-dimensional surface polaritons through plasmonic nanoplates on atomically thin crystals
- Highly sensitive microdisk laser sensor for refractive index sensing via periodic meta-hole patterning
- Scaled transverse translation by planar optical elements for sub-pixel sampling and remote super-resolution imaging
- Hyperbolic polariton-coupled emission optical microscopy
- Broadband perfect Littrow diffraction metasurface under large-angle incidence
- Role of complex energy and momentum in open cavity resonances
- Are nanophotonic intermediate mirrors really effective in enhancing the efficiency of perovskite tandem solar cells?
- Tunable meta-device for large depth of field quantitative phase imaging
- Enhanced terahertz magneto-plasmonic effect enabled by epsilon-near-zero iron slot antennas
- Baseline-free structured light 3D imaging using a metasurface double-helix dot projector
- Nanophotonic device design based on large language models: multilayer and metasurface examples
- High-efficiency generation of bi-functional holography with metasurfaces
- Dielectric metasurfaces based on a phase singularity in the region of high reflectance
- Exceptional points in a passive strip waveguide
Artikel in diesem Heft
- Frontmatter
- Editorial
- Special issue: “Metamaterials and Plasmonics in Asia”
- Reviews
- All-optical analog differential operation and information processing empowered by meta-devices
- Metasurface-enhanced biomedical spectroscopy
- Topological guided-mode resonances: basic theory, experiments, and applications
- Letter
- Ultrasensitive circular dichroism spectroscopy based on coupled quasi-bound states in the continuum
- Research Articles
- Data-efficient prediction of OLED optical properties enabled by transfer learning
- Semimetal–dielectric–metal metasurface for infrared camouflage with high-performance energy dissipation in non-atmospheric transparency window
- Deep-subwavelength engineering of stealthy hyperuniformity
- Tunable structural colors based on grayscale lithography and conformal coating of VO2
- A general recipe to observe non-Abelian gauge field in metamaterials
- Free-form catenary-inspired meta-couplers for ultra-high or broadband vertical coupling
- Enhanced photoluminescence of strongly coupled single molecule-plasmonic nanocavity: analysis of spectral modifications using nonlocal response theory
- Spectral Hadamard microscopy with metasurface-based patterned illumination
- Tunneling of two-dimensional surface polaritons through plasmonic nanoplates on atomically thin crystals
- Highly sensitive microdisk laser sensor for refractive index sensing via periodic meta-hole patterning
- Scaled transverse translation by planar optical elements for sub-pixel sampling and remote super-resolution imaging
- Hyperbolic polariton-coupled emission optical microscopy
- Broadband perfect Littrow diffraction metasurface under large-angle incidence
- Role of complex energy and momentum in open cavity resonances
- Are nanophotonic intermediate mirrors really effective in enhancing the efficiency of perovskite tandem solar cells?
- Tunable meta-device for large depth of field quantitative phase imaging
- Enhanced terahertz magneto-plasmonic effect enabled by epsilon-near-zero iron slot antennas
- Baseline-free structured light 3D imaging using a metasurface double-helix dot projector
- Nanophotonic device design based on large language models: multilayer and metasurface examples
- High-efficiency generation of bi-functional holography with metasurfaces
- Dielectric metasurfaces based on a phase singularity in the region of high reflectance
- Exceptional points in a passive strip waveguide