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

Multi-band infrared camouflage compatible with radiative cooling and visible colors via a simple multilayer film structure

Open Access Open Access

Abstract

Conventional infrared camouflage materials are typically confined to special spectral ranges. To address challenges posed by the combination of both active and passive detection operating in different bands, a multi-band compatible camouflage material is highly required. Combining rigorous coupled-wave analysis and genetic algorithm, a 9-layer film structure composed of 3 thermally stable materials (SiO2, Ge and TiO2) on a quartz substrate is designed, which achieves multi-band infrared camouflage compatible with radiative cooling and visible colors. Specifically, the structure exhibits low average emissivity of 0.12 and 0.21 in the two atmospheric window bands of 3∼5 µm and 8∼14 µm, while it shows spectral emissivity favorable of radiative cooling (an average of 0.67) in the non-atmospheric window band of 5∼8 µm. The structure also achieves high absorptivity at two widely applied laser wavelengths of 1.06 µm (Nd:YAG lidar) and 10.6 µm (CO2 lidar). In addition, by manipulating the thickness of the top layer, the structure exhibits tunable visible colors covering green, blue and yellow, et al., without much influence on the above selective absorption and emission in the infrared bands. The multi-band selective absorption/emission characteristic is also applicable to large incident/emission angles. The underlying physics are analyzed in detail. The selective emission in the mid-infrared band is mainly caused by the forbidden bands formed by the multilayer film, while the high absorption at both laser wavelengths is attributed to the Fabry-Perot resonance supported in the structure and the intrinsic absorption of the materials. Finally, the overall camouflage performance of the structure is evaluated to make the results more intuitive. The significance of this work lies in the balance between the multi-band compatibility and structural simplicity of camouflage materials.

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

1. Introduction

Camouflage is an adaptive strategy involving various physical and behavioral techniques to conceal an object in its environment in order to enhance the survival chances or success of the object [14]. To confront the threats posed by detectors operating both actively and passively in different spectral ranges, a pressing need exists for multi-band compatible camouflage [57]. Modern multi-band camouflage technologies involve modulation of infrared (IR) radiation in the mid-infrared (MIR) band [8,9], color control in the visible band [10,11], and suppression of reflective signals for lasers and microwaves [1214]. Notably, due to the prevalence of IR detection, IR camouflage research has emerged as a focus area [15,16]. According to the Stefan-Boltzmann law, the way to modulate the IR signal of targets is either controlling the temperature or tuning the target emissivity [8,17]. Conventional metal-based camouflage materials (Al, Au, Ag, etc.) have broadband low IR emissivity, but undesirably block the radiative cooling of high temperature targets [18,19]. Consequently, contemporary MIR camouflage materials require low emission in the detection bands and high emission in the non-detection band to cope with radiative heat dissipation [20,21]. Along with passive detection in the MIR band, active detection lidars, predominantly the Nd:YAG lidar operating at 1.06 µm and CO2 lidar at 10.6 µm, should also be considered into the design of camouflage materials [22,23]. Further compatibility with the visible band necessitates that camouflage materials embody various structural colors like green, blue and yellow in order to match typical backgrounds [24,25]. Therefore, the four requirements for the design of camouflage materials are: (i) low averaged emissivity (smaller than 0.3 [7,13,25]) in two MIR atmospheric window bands where the atmosphere is nearly transparent, i.e. the 3∼5 µm mid-wavelength-infrared (MWIR) and the 8∼14 µm long-wavelength-infrared (LWIR) bands [2629]. The MWIR and LWIR bands are the operating bands for IR detection, so targets with low averaged emissivity in these bands can avoid being detected; (ii) relatively high averaged emissivity in the non-atmospheric window band (5∼8 µm) [3032]. While satisfying the premise of IR stealth in the atmospheric windows, a relatively high averaged emissivity in the non-atmospheric window (5∼8 µm) can radiate more thermal energy into the atmosphere to create a cooling effect; (iii) high absorptivity at the operating wavelengths of lidar (most frequently being 1.06 µm and 10.6 µm) [6,12,33,34]; (iv) tunable structural colors in the visible over a wide color gamut [3538].

However, to date, no reported surface structures simultaneously satisfy all the four camouflage requirements mentioned above. Metasurfaces are often employed in the design of multi-band compatible camouflage. Pan et al. [33], Kang et al. [34], and Yu et al. [22] designed camouflage metasurfaces for the MIR dual-windows and the 10.6 µm CO2 laser, concurrently facilitating radiative cooling. Wu et al. [39] and Lee et al. [40] designed metasurfaces achieving thermal camouflage and radiative thermal management in MIR, while maintaining optical transparency in the visible band. The metasurfaces designed for camouflage typically have complicated structures. In general impression, the structure complexity increases with incorporation of more operating camouflage bands. In contrast, multilayer films receive great concern in multi-band camouflage due to their simple geometries. Qi et al. [35,36] and Zhang et al. [37,38] implemented MIR camouflage and radiative thermal management based on one-dimensional photonic crystals, which achieved structural color control by varying the thickness of the top layer. Jiang et al. [13] designed a 7-layer metamaterial, achieving IR camouflage, radiative thermal management and stealth for 1.06 µm, 1.55 µm and 10.6 µm lasers. The research on camouflage materials closest to fulfilling the four requirements is conducted by Zhu et al. [7], who designed a Ge/ZnS 12-layer film structure on a microwave metasurface that is compatible for visible, infrared, lasers (1.55 µm and 10.6 µm) and microwave camouflage with radiative cooling. However, the 1.55 µm laser is much less applied in practice than the 1.06 µm laser. Additionally, ZnS might be unstable under humid circumstances.

Therefore, camouflage materials reaching a balance between multi-band compatibility and structural simplicity is still in lack. In this work, we report a simple 9-layer film structure on quartz substrate which achieves compatible camouflage obeying the above four requirements [see Fig. 1(a)]. Combining rigorous coupled-wave analysis (RCWA) [41] and genetic algorithm (GA) [42], we optimize multilayered films with no predefined number of layers. The materials of each layer are chosen from SiO2, Ge and TiO2, which are chemically stable and high-temperature resistant. The designed structure achieves low emissivity in the MWIR and LWIR bands, high emissivity in the non-atmospheric window band, high absorptivity at the wavelengths of 1.06 µm and 10.6 µm, and tunable structural colors in the visible band. The underlying physics are analyzed and the overall camouflage performance (OCP) is evaluated.

 figure: Fig. 1.

Fig. 1. (a) The application scenarios of camouflage for visible, infrared and lasers with radiative cooling and the schematic diagram of multilayer film structure (MFS) composed of SiO2, Ge and TiO2 films; (b) Schematic diagram of the workflow of GA; (c) Schematic diagram of the optimized MFS; (d) The ideal emission spectra (red line), the emission/absorption spectra of the optimized MFS (blue line) and the atmospheric transmission spectra calculated by MODTRAN (light blue area).

Download Full Size | PDF

2. Model and methods

The multilayer film structure (MFS) is illustrated in Fig. 1(a). The MFS is composed of SiO2, Ge, and TiO2 layers stacked alternately on a quartz substrate. The main reasons for selecting SiO2, Ge, and TiO2 as layer materials of the MFS are as follows [43,44]: (i) they are stable at high temperatures, so the performance will not deteriorate significantly with increasing temperature of the object; (ii) they have distinct refractive indices and different IR transparent zones (i.e. 0.9∼8 µm for SiO2, 1.7∼23 µm for Ge and in 0.4∼10 µm for TiO2), which will lay the foundation for the MIR selective absorption/transmission of the structure [43,44]. In addition, quartz is chosen as the substrate material due to its high absorptivity/emissivity in the MWIR band [7,44]. The MFS above the substrate will be designed to selectively transmit the thermal emission of the substrate in the band of 5∼8 µm. By fixing the thickness of the quartz substrate at 2000µm, it is possible to absorb almost completely the transmitted electromagnetic waves in the 5∼8 µm band, which is equivalent to a high emission characteristic in the band. Thinner quartz substrates do not satisfy the condition well in the 5∼8 µm band. The radiation properties of quartz films of different thicknesses are discussed in the Supplemental material.

The ideal spectra of the MFS are shown as the red line in Fig. 1(d). Considering the peak shift resulting from processing errors, the ideal spectra should have a relatively broader peak at 1.06 µm [45]. The narrow peak centered at 10.6 µm can reduce the radiation intensity in the LWIR band as much as possible.

Recently, deep learning becomes very popular in nanophotonics design [4548]. However, the radiation characteristics of MFS can be calculated with high speed by RCWA. Therefore, a combination of RCWA and GA to optimize the number of layers, thickness and material of each layer in the MFS is relatively more efficient and accurate. As shown in Fig. 1(b), GA starts with the initialization of the population and each layer contains two attributes, i.e., material and layer thickness, which are randomly selected from the corresponding library. After the initialization of the population, GA enters the process of fitness calculation. If the termination criteria are not reached, GA will perform selection, crossover, and mutation (the genetic operator) to produce individuals of the next generation and reenter the fitness evaluation. The optimal individual is finally obtained by selecting from generation to generation. In the application of GA, we define a fitness function to evaluate the individuals as:

$$F = 3 \cdot {\bar{\alpha }_{\textrm{1}\mathrm{.0\sim 1}\mathrm{.1\mu m}}} + 3 \cdot (1 - {\bar{\varepsilon }_{\textrm{MWIR}}}) + 4 \cdot {\bar{\varepsilon }_{\mathrm{5\sim 8\mu m}}} + 3 \cdot {\alpha _{\textrm{10}\mathrm{.6\mu m}}} + 5 \cdot {\bar{\varepsilon }_{\textrm{LWIR}}}$$
where ${\mathrm{\bar{\alpha }}_{\textrm{1}\mathrm{.0\sim 1}\mathrm{.1\mu m}}}$ is the arithmetic averaged absorptivity in 1∼1.1 µm, α10.6µm is the absorptivity at 10.6 µm, ${\mathrm{\bar{\varepsilon }}_{\textrm{MWIR}}}$, ${\mathrm{\bar{\varepsilon }}_{\mathrm{5\sim 8\mu m}}}$ and ${\mathrm{\bar{\varepsilon }}_{\textrm{LWIR}}}$ represent the averaged emissivity under blackbody radiation weighting in the MWIR, 5∼8 µm and LWIR bands, respectively. The single-wavelength spectral absorptivity at 10.6 µm is used, since on the one hand the optimized MFS must have large absorption of the 10.6 µm laser and on the other hand it must ensure that the averaged emissivity in the LWIR band (8∼14 µm) is low. These absorptivities and emissivities are normal that are calculated under normal incident condition. The five parameters correspond to 1.06 µm Nd:YAG laser stealth, MIR atmospheric window stealth, radiative cooling, and 10.6 µm CO2 laser stealth, respectively. The calculation of the averaged emissivity involves the length of the bands. As a result, weights should vary with the length of the bands during the optimization process. After the validation with several sets of weights, we set the weights of averaged emissivity in the three bands as 3, 4 and 5 according to the length of the bands. Meanwhile, we set the weights of the absorptivity at the two laser wavelengths to 3, i.e., equal to the weight of the averaged emissivity in the MWIR band. It is noted that the emissivity of the structure is always equal to its absorptivity according to the Kirchhoff's law. In the fitness function, we use emissivities and absorptivities separately to clearly denote the thermal emission and the absorption of external lasers by the structure. The averaged emissivity in the MWIR, 5∼8 µm and LWIR bands can be calculated by the following equation:
$${\bar{\varepsilon }_{[{\lambda _1},{\lambda _2}]}} = {{\int_{{\lambda _1}}^{{\lambda _2}} {\varepsilon (\lambda ){E_\textrm{b}}(\lambda ,T)d\lambda } } \left/ {\int_{{\lambda _1}}^{{\lambda _2}} {{E_\textrm{b}}(\lambda ,T)d\lambda } }\right.}$$
where ε(λ) is the spectral emissivity of the MFS, Eb(λ,T) is the blackbody spectral radiation intensity at temperature T which is set to 523 K (the normal operating temperature of some military objectives). Thus, a higher value of the fitness function indicates that the spectra of the MFS are closer to the ideal spectra. In the iteration of GA, the maximum number of layers was set to be 12, and the thickness of each layer was limited to 10∼1500 nm. Additionally, the material of each layer was randomly selected from SiO2, Ge and TiO2. Note that GA may produce two adjacent layers of the same material, in which case the two adjacent layers should be merged into one layer. In the process of optimization, the number of individuals in the population is set to be 200 and the number of parents-mating is set to be 40. In order to get the optimal result, the probabilities of crossover and mutation are set to 0.8 and 0.2, respectively, and the number of iterations is set to be 1000. In addition, the optimization process is conducted in parallel for 20 cases and an optimal result is chosen among the 20 tries.

3. Results and analysis

3.1 Multi-band camouflage

After 1000 iterations of GA, an individual with the fitness value of 15.14 was finally selected, generating a 9-layer MFS. The corresponding values of ${\mathrm{\bar{\alpha }}_{\textrm{1}\mathrm{.0\sim 1}\mathrm{.1}\mathrm{\mu}\textrm{m}}}$, ${\mathrm{\bar{\varepsilon }}_{\textrm{MWIR}}}$, ${\mathrm{\bar{\varepsilon }}_{\mathrm{5\sim 8}\mathrm{\mu}\textrm{m}}}$, α10.6µm and ${\mathrm{\bar{\varepsilon }}_{\textrm{LWIR}}}$ are 0.98, 0.12, 0.66, 0.99, and 0.21, respectively. The geometric structure of the optimized MFS is shown in Fig. 1(c), and the thickness of each layer is given in Table 1. Note that the layers are numbered from top to bottom.

Tables Icon

Table 1. The thickness and material of each layer of the optimized MFS (from top to bottom)

We first inspect the IR emission/absorption properties of the optimized MFS. Figure 1(d) shows the IR emission/absorption spectra of the optimized MFS, along with the ideal spectra and the transmission spectra of the atmosphere calculated by MOTRAN. As shown in Fig. 1(d), the designed MFS exhibits low emissivity in the MWIR and LWIR bands, while displaying a broad emission peak in the non-atmospheric window band from 5 to 8 µm. At the temperature of 523 K, for instance, the averaged emissivities of the MFS in MWIR, non-atmospheric window and LWIR bands are 0.12, 0.67 and 0.21 [see Eq. (2)]. At room temperature (300 K), the averaged emissivities are 0.1, 0.6, and 0.21, respectively. At the Nd:YAG lidar wavelength of 1.06 µm, an absorptivity as high as 0.99 is observed by the optimized MFS. Moreover, high absorptivity is realized at wavelengths around 1.06 µm. At the CO2 lidar wavelength of 10.6 µm, the MFS has narrowband absorptivity as high as 0.99 with a bandwidth of 0.31 µm (the bandwidth is defined as the full width at half-maximum). As a result, the optimized MFS not only has high absorption of the 10.6 µm CO2 lidar light, but also remains low averaged emissivity in the LWIR band. Thus, the optimized MFS satisfies the requirements for camouflage in the multiple IR bands. Furthermore, practical MIR stealth requires low averaged emissivity of the structure in the two atmospheric window bands (smaller than 0.3 [7,13,25]). Although the spectral emissivity in 8∼10 µm of the optimized MFS is not very small, the overall low averaged emissivity in 8∼14 µm can satisfy the requirement of MIR stealth (${\mathrm{\bar{\varepsilon }}_{\textrm{LWIR}}}$=0.21). In addition, the small emission peak of the MFS in 9∼10 µm is attributed to the intrinsic absorption of SiO2 in the layers. As for the radiative cooling effect in non-atmospheric window band, we note that notable radiative cooling effect were experimentally demonstrated in previous works like Refs. [7,13,34], the averaged emissivity in the non-atmospheric band of our optimized structure is comparable to or even larger than those in the previous works. Thus, the MFS, if properly fabricated, is also expected to have a notable radiative cooling effect.

In addition to the IR bands, the MFS can also possess various structural colors in the visible band. In fact, the top two layers (TiO2 and Ge) are responsible for the visible structural colors. Since the Ge layer with a thickness of 0.693 µm is nearly opaque in the visible band, the spectral reflectivity of the MFS in the visible is determined by the thickness of the top TiO2 layer (hTiO2). Meanwhile, the top two layers are highly transparent at IR wavelengths, thus having minimal impact on the thermal emission from the multiple layers below. We calculated the visible-band spectral reflectivity of the MFS with different hTiO2 (ranging from 70 to 200 nm with 10 nm intervals) and converted them into CIE-1931 chromaticity coordinates, as shown in Figs. 2(a) and (b). The spectral reflectivity refers specifically to the hemispherical reflectivity for normal incident wave. With the variation of hTiO2, we demonstrate the different structural colors generated by the MSF from red to green. Figure 2(c) presents the IR emission spectra of the designed MFS for different colors obtained solely by altering the thicknesses of the top layer of the MFS. We can see that while the variation of the thickness of the top layer leads to different colors of the MFS, it does not affect the IR spectral emissivity/absorptivity much. The thickness of the top layer only has non-negligible effect on the absorptivity around 1.06 µm. To see the effect concretely, Fig. 2(d) illustrates the structural colors for different hTiO2 and the corresponding absorptivity of the MFS at the wavelength of 1.06 µm. When hTiO2 varies from 70 to 200 nm, the absorptivity of the MFS at 1.06 µm varies in the range of 0.57∼0.99. The averaged absorptivity at 1.06 µm across the different structural colors reaches 0.82. Thus, by adjusting the thickness of the first layer, the MFS can exhibit diverse colors while keeping good performance in the IR bands. These distinct colors can be used for visible camouflage in different environments, e.g. a thickness of 70/170 nm of the top layer can be suitable for ocean or sky, 120 nm for desert or mountains, and 200 nm for forest or meadow.

 figure: Fig. 2.

Fig. 2. Compatibility of visible structural colors with IR emission/absorption. (a) Reflected structural colors generated by the designed MFS for varying thickness of the top TiO2 layer hTiO2 in the CIE-1931 color space. The dashed line represents route of the color change as hTiO2 varies from 70 to 200 nm with 10 nm intervals; (b) The visible reflection spectra of the designed MFS as hTiO2 varies from 70 nm to 200 nm, the color of the line represents the structural color of the designed MFS with the corresponding hTiO2; (c) The IR emission/absorption spectra of the designed MFS as hTiO2 varies from 70 nm to 200 nm; (d) The structural colors generated by the designed MFS for different hTiO2 and the corresponding absorptivity at 1.06 µm.

Download Full Size | PDF

The angular dependence of the emission and absorption properties of MFS is also an important factor for camouflage. Figure 3 demonstrates the angular dependence of the emissivity/absorptivity in the IR bands. For lasers stealth, the key point is that the MFS can achieve relatively high absorptivity at the laser wavelengths around the normal directions. For large incident angles, the signal will be mainly reflected to other directions so that the detector will not receive the reflected signal. At 1.06 µm and 10.6 µm, particularly, as shown in Fig. 3 (a) and (b), the designed MFS has relatively large absorptivity for incident angles up to 60° (the averaged absorptivity is 0.91 at 1.06 µm and 0.52 at 10.6 µm for incident angle of 60°). For both TM and TE polarizations, as shown in Fig. 3(c) and (d), the emissivities in the two atmosphere window bands remain low while that in the non-atmospheric window band keeps high even at larger emission angles. In addition, the absorption peak of the optimized MFS at the laser wavelength shifts as the incident angle rises. Meanwhile, the MFS maintains a high absorptivity at 1.06 µm due to the wide absorption peak around the wavelength. Conversely, the absorption peak of the MFS at 10.6 µm is very narrow, and hence the absorptivity decreases rapidly when the absorption peak shifts. The absorptivity of the optimized MFS at 10.6 µm shows a peak again around 75°. This is related to the Brewster angle of the MFS.

 figure: Fig. 3.

Fig. 3. The spectral absorptivity for TM- and TE-polarizations, and the averaged value over the two polarizations as a function of incident angle at the wavelengths of (a) 1.06 µm and (b) 10.6 µm. Contour plot of the emissivity with respect to the wavelength and the incident angle for (c) TM-polarization and (d) TE-polarization.

Download Full Size | PDF

3.2 Physical mechanism

To reveal the underlying physics for the selective IR emission/absorption, the spectral reflectivity, transmissivity and absorptivity (R, T and A) of the 9 layers constituting of the designed MFS without quartz substrate are presented in Fig. 4(a). The distribution of the normalized electric field with respect to the wavelength and z-coordinate is presented in Fig. 4(b). As shown, the 9 layers form two distinct forbidden bands in the MWIR and LWIR, characterized by low transmissivity. In other words, most of the electromagnetic radiation within the forbidden bands are reflected out, and the 9 layers absorb little in these bands. In the non-atmospheric window band (5∼8 µm), however, a small proportion of the electromagnetic radiation is absorbed by the 9 layers, while a large portion can transmit through them. It can also be seen from the normalized electric field map distribution that the electric field intensity within the MWIR and LWIR bands is universally low, forming two distinct forbidden bands. The reason for the appearance of such forbidden bands is the destructive interference effect caused by the stack of dielectrics with different refractive indexes [44]. Consequently, when combined with a quartz substrate, the thermal emission from the quartz substrate can efficiently transmit the upper 9 layers only in the non-atmospheric band, the thermal emission in the forbidden bands is blocked. (see Section 1 of Supplement 1 for quartz’s radiation characteristics in different thicknesses). It can be observed from Fig. 4(b) that the electric field intensity within the 9 layers is indeed stronger in the non-forbidden band. Therefore, the combination of the 9 layers and the thick quartz substrate achieves selective emission in the MWIR and LWIR bands.

 figure: Fig. 4.

Fig. 4. (a) The spectral reflectivity, transmissivity and absorptivity of the upper 9 layers without the quartz substrate; (b) The normalized electric field intensity distribution as a function of wavelength from 3 µm to 14 µm in the upper 9 layers.

Download Full Size | PDF

As to the high absorptivity of the designed MFS at 1.06 µm, it is mainly caused by the intrinsic absorption of the second Ge layer (see Section 2 of Supplement 1 for a discussion of the materials’ optical constants) [49]. Figure 5(a) and (c) present the distribution of the electric field intensity and the normalized electromagnetic power loss density at 1.06 µm. As shown, the electric field intensity decays sharply in the second Ge layer and the electromagnetic losses are mainly located inside this layer. At the CO2 laser wavelength of 10.6 µm, the narrowband high absorptivity can be attributed to two mechanisms: (i) the symmetric Fabry-Perot resonance in the cavity formed between the second and fourth layer [50]; (ii) the intrinsic absorption of the first and fifth TiO2 layers. As shown in Figs. 5(b) and (d), the electric field at the wavelength of 10.6 µm is mainly confined in the third SiO2 layer, indicating a localization of the electromagnetic power by a typical Fabry-Perot resonance in the cavity formed by adjacent layers. The Fabry-Perot resonance in the cavity causes a high aggregation of electric field intensity and large electromagnetic loss density. The electromagnetic power at 10.6µm is mainly dissipated in the first and fifth TiO2 layer due to the intrinsic absorption of TiO2. In short, the upper 5 layers mainly contribute to the formation of the Fabry-Perot resonance cavity and the intrinsic absorption of incident waves. Meanwhile, it can be seen from Fig. 4(a) that the upper 9-layer structure without the substrate has non-negligible absorption in the 5∼8 µm band, and we have verified that the bottom 4 layers can enhance the absorption in this band. Therefore, the repetitive layers are all physically significant.

 figure: Fig. 5.

Fig. 5. Distribution of the electric field intensity in the plane of y = 0 at the wavelengths of (a) 1.06 µm and (b) 10.6 µm. Distribution of the normalized electromagnetic power loss density in the plane of y = 0 at at the wavelengths of (c) 1.06 µm and (d) 10.6 µm.

Download Full Size | PDF

3.3 Overall camouflage performance

To better evaluate the overall camouflage performance (OCP) of the designed MFS, the OCP of the optimized MSF is evaluated through [33]:

$$\textrm{OCP} = \frac{{P_{\mathrm{5\sim 8}\mathrm{\mu}\textrm{m}}^2}}{{{P_{\textrm{MWIR}}} \cdot {P_{\textrm{LWIR}}}}} \cdot \frac{{{\alpha _{\textrm{10}\mathrm{.6}\mathrm{\mu}\textrm{m}}}}}{{{{\bar{\varepsilon }}_{\textrm{LWIR}}}}} \cdot {\alpha _{\textrm{1}\mathrm{.06}\mathrm{\mu}\textrm{m}}}$$
$${P_{[{\lambda _1},{\lambda _2}]}} = \int_{{\lambda _1}}^{{\lambda _2}} {\varepsilon (\lambda ){E_\textrm{b}}(\lambda ,T)d\lambda }$$
where PMWIR, P5∼8µm and PLWIR are the radiation intensities of the MFS in MWIR, 5∼8 µm and LWIR bands, respectively [calculated by Eq. (4)]; α10.6µm and α1.06µm represent the absorptivity at 10.6 µm and 1.06 µm; ${\mathrm{\bar{\varepsilon }}_{\textrm{LWIR}}}$ is the averaged emissivity in the LWIR band. The evaluation equation for OCP can be divided into three main parts: i) P25-8µm /PMWIRPLWIR indicates the combined assessment of the non-atmospheric window's radiative cooling capability and the atmospheric windows’ IR stealth; ii) α10.6µm/${\mathrm{\bar{\varepsilon }}_{\textrm{LWIR}}}$ represents the compatibility of stealth for the 10.6 µm CO2 laser. The higher the absorptivity at 10.6 µm and the narrower the absorption peak, the better the compatibility; iii) α1.06µm means the stealth capability for the 1.06 µm Nd:YAG laser. The higher the OCP, the better the camouflage performance of the material. Note that OCP is temperature-dependent due to the variation of blackbody radiation intensity with temperature.

As shown in Fig. 6(a), the OCP of the designed MFS decreases as the temperature rising from 293 K to 573 K. Notably, the highest OCP corresponds to earthy yellow of the MFS, indicating that the structure is most effective for stealth in the environment of desert or mountains. In addition, the MFS’s OCP is much higher than conventional full-band low-emissivity metal-based surfaces’ OCP (e.g. Au, Ag, Al, etc.) indicated by the gray dotted line in Fig. 6 (a). Figure 6(b) shows the radiation intensity of the MFS in the MWIR, 5∼8 µm and LWIR bands at different temperatures and their ratios included in Eq. (3). As the temperature increases, the radiation intensity of the designed MFS rises more sharply in 5∼8 µm compared with both MWIR and LWIR bands. This can be attributed to the peak-shift of the blackbody radiation intensity with temperature. The spectral radiation intensity of the MFS and the blackbody at the temperature of 300 K and 523 K are shown in Fig. 6(c). As temperature increases, the emission peak of the blackbody shifts towards shorter wavelengths. The designed MFS has higher emissivity in the non-atmospheric band compared with ordinary metal-substrate materials, which will lead to radiative cooling effect. Obviously, the radiated power of the designed MFS is spectrally selective.

 figure: Fig. 6.

Fig. 6. Overall camouflage performance (OCP) evaluation. (a) OCP of the designed MFS as a function of the temperature for different hTiO2; (b) The radiation intensity calculated by Eq. (4) of the designed MFS in the MWIR, 5∼8 µm and LWIR bands and their ratio in Eq. (3) for temperatures varying from 293 K to 573 K; (c) The spectral radiation intensity of blackbody and the designed MFS in the MIR band at temperatures of 523 K and 300 K; (d) The signal reduction rate (SRR) calculated by Eqs. (5)∼(7) of the designed MFS in the MWIR and LWIR bands and the radiation intensity in the 5∼8 µm band as temperature varies from 293 K to 1093 K.

Download Full Size | PDF

MIR camouflage performance refers to the stealth capability of the structure in the passive detection bands or the two atmospheric windows. To further assess the MIR camouflage performance of the MFS, we evaluate the IR signal reduction rate (SRR) by the following equations [17]:

$$\textrm{SRR} = \left( {1 - \frac{{\alpha (T)}}{{\beta (T)}}} \right) \times 100\%$$
$$\alpha (T) = \int_{{\lambda _1}}^{{\lambda _2}} {\varepsilon (\lambda ){E_\textrm{b}}(\lambda ,T)} d\lambda$$
$$\beta (T) = \int_{{\lambda _1}}^{{\lambda _2}} {{E_\textrm{b}}(\lambda ,T)d\lambda }$$
where α(T) and β(T) are the integrals of the MFS and blackbody’s spectral radiation intensity over the λ1λ2 µm band. Figure 6(d) presents the SRR of the MFS in the MWIR and LWIR bands from 293 K to 1093 K and the radiation intensity in the spectral range of 5∼8 µm. The SRR varies slightly with temperature and keeps a high rate of about 80%. The radiation intensity in the 5∼8 µm band increases rapidly with temperature. It indicates that the designed MFS can effectively reduce the target IR signal while it achieves efficient radiative cooling in the non-atmospheric window. The OCP and SRR of the designed MFS for different incident angles are discussed in Section 3 of Supplement 1.

Herein, we list some recently reported results in the field of multi-band camouflage in Table 2. The majority of the reported works dose not achieve the compatibility of all the four camouflage bands. The simple 9-layer structure in this work, however, is compatible with all the four camouflage requirements and provides excellent performance in each band.

Tables Icon

Table 2. Comparison of the recent reported results for multi-band camouflage with this work

4. Conclusions

In conclusion, a MFS composed of SiO2, Ge and TiO2 is designed to achieve multi-band camouflage for the IR, the visible and lasers with efficient radiative cooling by combining the methods of GA and RCWA. In terms of visible camouflage, the designed MFS shows tunable structural colors to match different backgrounds. For IR camouflage, low emissivities at the two atmospheric windows (${\mathrm{\bar{\varepsilon }}_{\textrm{MWIR}}}$=0.12 and ${\mathrm{\bar{\varepsilon }}_{\textrm{LWIR}}}$=0.21) are achieved. For lasers camouflage, high absorptivity can be achieved at both the 1.06 µm and 10.6 µm wavelengths. Meanwhile, the MFS has a broad emission peak in the non-atmospheric window, with an averaged value of 0.67. The camouflage characteristics of the designed MFS are applicable to large emission/incident angles. The underlying physics are the forbidden bands, the FP resonance formed by the multilayered structure together with the extra contribution from the intrinsic absorption by the materials. In addition, the overall camouflage performance of the designed MFS is evaluated, demonstrating its superior capabilities in camouflage applications.

Funding

National Natural Science Foundation of China (51906128); Natural Science Foundation of Shandong Province (ZR2019BEE011, ZR2020QE194).

Disclosures

The authors declare no conflicts of interest.

Data availability

All calculation data that support this work are available upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. J. Yang, X. Zhang, X. Zhang, L. Wang, W. Feng, and Q. Li, “Beyond the Visible: Bioinspired Infrared Adaptive Materials,” Adv. Mater. 33(14), 2004754 (2021). [CrossRef]  

2. C. Xu, G. T. Stiubianu, and A. A. Gorodetsky, “Adaptive infrared-reflecting systems inspired by cephalopods,” Science 359(6383), 1495–1500 (2018). [CrossRef]  

3. S. A. Morin, R. F. Shepherd, S. W. Kwok, A. A. Stokes, A. Nemiroski, and G. M. Whitesides, “Camouflage and display for soft machines,” Science 337(6096), 828–832 (2012). [CrossRef]  

4. J. Teyssier, S. V. Saenko, D. Van Der Marel, and M. C. Milinkovitch, “Photonic crystals cause active colour change in chameleons,” Nat. Commun. 6(1), 6368 (2015). [CrossRef]  

5. J. Kim, K. Han, and J. W. Hahn, “Selective dual-band metamaterial perfect absorber for infrared stealth technology,” Sci. Rep. 7(1), 6740 (2017). [CrossRef]  

6. L. Zhao, H. Liu, Z. He, and S. Dong, “All-metal frequency-selective absorber/emitter for laser stealth and infrared stealth,” Appl. Opt. 57(8), 1757–1764 (2018). [CrossRef]  

7. H. Zhu, Q. Li, C. Tao, Y. Hong, Z. Xu, W. Shen, S. Kaur, P. Ghosh, and M. Qiu, “Multispectral camouflage for infrared, visible, lasers and microwave with radiative cooling,” Nat. Commun. 12(1), 1805 (2021). [CrossRef]  

8. L. Peng, D. Liu, H. Cheng, S. Zhou, and M. Zu, “A Multilayer Film Based Selective Thermal Emitter for Infrared Stealth Technology,” Adv. Opt. Mater. 6(23), 1801006 (2018). [CrossRef]  

9. C. Xu, S. Qu, Y. Pang, J. Wang, M. Yan, J. Zhang, Z. Wang, and W. Wang, “Metamaterial absorber for frequency selective thermal radiation,” Infrared Phys. Technol. 88, 133–138 (2018). [CrossRef]  

10. E. Buhara, A. Ghobadi, and E. Ozbay, “Adaptive visible and short-wave infrared camouflage using a dynamically tunable metasurface,” Opt. Lett. 46(19), 4777–4780 (2021). [CrossRef]  

11. J. Pikul, S. Li, H. Bai, R. Hanlon, I. Cohen, and R. F. Shepherd, “Stretchable surfaces with programmable 3D texture morphing for synthetic camouflaging skins,” Science 358(6360), 210–214 (2017). [CrossRef]  

12. X. Feng, X. Xie, M. Pu, X. Ma, Y. Guo, X. Li, and X. Luo, “Hierarchical metamaterials for laser-infrared-microwave compatible camouflage,” Opt. Express 28(7), 9445–9453 (2020). [CrossRef]  

13. X. Jiang, H. Yuan, X. He, T. Du, H. Ma, X. Li, M. Luo, Z. Zhang, H. Chen, Y. Yu, G. Zhu, P. Yan, J. Wu, Z. Zhang, and J. Yang, “Implementing of infrared camouflage with thermal management based on inverse design and hierarchical metamaterial,” Nanophotonics 12(10), 1891–1902 (2023). [CrossRef]  

14. T. Kim, J. Y. Bae, N. Lee, and H. H. Cho, “Hierarchical Metamaterials for Multispectral Camouflage of Infrared and Microwaves,” Adv. Funct. Mater. 29(10), 1807319 (2019). [CrossRef]  

15. S. Chandra, D. Franklin, J. Cozart, A. Safaei, and D. Chanda, “Adaptive multispectral infrared camouflage,” ACS Photonics 5(11), 4513–4519 (2018). [CrossRef]  

16. Z. Wang, Y. Cheng, Y. Nie, X. Wang, and R. Gong, “Design and realization of one-dimensional double hetero-structure photonic crystals for infrared-radar stealth-compatible materials applications,” J. Appl. Phys. 116(5), 054905 (2014). [CrossRef]  

17. N. Lee, T. Kim, J. S. Lim, I. Chang, and H. H. Cho, “Metamaterial-Selective Emitter for Maximizing Infrared Camouflage Performance with Energy Dissipation,” ACS Appl. Mater. Interfaces 11(23), 21250–21257 (2019). [CrossRef]  

18. K.-K. Du, Q. Li, Y.-B. Lyu, J.-C. Ding, Y. Lu, Z.-Y. Cheng, and M. Qiu, “Control over emissivity of zero-static-power thermal emitters based on phase-changing material GST,” Light: Sci. Appl. 6(1), e16194 (2016). [CrossRef]  

19. Y. Huang, M. Pu, Z. Zhao, X. Li, X. Ma, and X. Luo, “Broadband metamaterial as an “invisible” radiative cooling coat,” Opt. Commun. 407, 204–207 (2018). [CrossRef]  

20. H. Zhu, Q. Li, C. Zheng, Y. Hong, Z. Xu, H. Wang, W. Shen, S. Kaur, P. Ghosh, and M. Qiu, “High-temperature infrared camouflage with efficient thermal management,” Light: Sci. Appl. 9(1), 60 (2020). [CrossRef]  

21. Y. Qu, Q. Li, L. Cai, M. Pan, P. Ghosh, K. Du, and M. Qiu, “Thermal camouflage based on the phase-changing material GST,” Light: Sci. Appl. 7(1), 26 (2018). [CrossRef]  

22. K. Yu, W. Zhang, M. Qian, P. Shen, and Y. Liu, “Multiband metamaterial emitters for infrared and laser compatible stealth with thermal management based on dissipative dielectrics,” Photonics Res. 11(2), 290–298 (2023). [CrossRef]  

23. C. Zhang, X. Wu, C. Huang, J. Peng, C. Ji, J. Yang, Y. Huang, Y. Guo, and X. Luo, “Flexible and transparent microwave–infrared bistealth structure,” Adv. Mater. Technol. 4(8), 1900063 (2019). [CrossRef]  

24. Y. Huang, Y. Zhu, B. Qin, Y. Zhou, R. Qin, P. Ghosh, M. Qiu, and Q. Li, “Hierarchical visible-infrared-microwave scattering surfaces for multispectral camouflage,” Nanophotonics 11(16), 3613–3622 (2022). [CrossRef]  

25. J. Kim, C. Park, and J. W. Hahn, “Metal–Semiconductor–Metal Metasurface for Multiband Infrared Stealth Technology Using Camouflage Color Pattern in Visible Range,” Adv. Opt. Mater. 10(6), 2101930 (2022). [CrossRef]  

26. L. Peng, D. Liu, and H. Cheng, “Ultrathin continuous silver film for efficient infrared emission,” Mater. Today Commun. 20, 100592 (2019). [CrossRef]  

27. N. Lee, B. Yoon, T. Kim, J. Y. Bae, J. S. Lim, I. Chang, and H. H. Cho, “Multiple Resonance Metamaterial Emitter for Deception of Infrared Emission with Enhanced Energy Dissipation,” ACS Appl. Mater. Interfaces 12(7), 8862–8869 (2020). [CrossRef]  

28. M. Li, D. Liu, H. Cheng, L. Peng, and M. Zu, “Manipulating metals for adaptive thermal camouflage,” Sci. Adv. 6(22), eaba3494 (2020). [CrossRef]  

29. Z. Deng, Y. Su, W. Gong, X. Wang, and R. Gong, “Temperature characteristics of Ge/ZnS one-dimension photonic crystal for infrared camouflage,” Opt. Mater. 121, 111564 (2021). [CrossRef]  

30. N. N. Shi, C.-C. Tsai, F. Camino, G. D. Bernard, N. Yu, and R. Wehner, “Keeping cool: Enhanced optical reflection and radiative heat dissipation in Saharan silver ants,” Science 349(6245), 298–301 (2015). [CrossRef]  

31. Z. Zhan and Y. Han, “Efficient thermal management of the target surface through introducing selective metasurface,” Infrared Phys. Technol. 117, 103814 (2021). [CrossRef]  

32. Y. Zhao and F. Fang, “Thermochromism Frequency-Selective Emitter for Infrared Stealth Application,” ACS Appl. Electron. Mater. 3(6), 2694–2702 (2021). [CrossRef]  

33. M. Pan, Y. Huang, Q. Li, H. Luo, H. Zhu, S. Kaur, and M. Qiu, “Multi-band middle-infrared-compatible camouflage with thermal management via simple photonic structures,” Nano Energy 69, 104449 (2020). [CrossRef]  

34. Q. Kang, D. Li, W. Wang, K. Guo, and Z. Guo, “Multiband tunable thermal camouflage compatible with laser camouflage based on GST plasmonic metamaterial,” J. Phys. D: Appl. Phys. 55(6), 065103 (2022). [CrossRef]  

35. D. Qi, X. Wang, Y. Cheng, R. Gong, and B. Li, “Design and characterization of one-dimensional photonic crystals based on ZnS/Ge for infrared-visible compatible stealth applications,” Opt. Mater. 62, 52–56 (2016). [CrossRef]  

36. D. Qi, F. Chen, X. Wang, H. Luo, Y. Cheng, X. Niu, and R. Gong, “Effective strategy for visible-infrared compatible camouflage: surface graphical one-dimensional photonic crystal,” Opt. Lett. 43(21), 5323–5326 (2018). [CrossRef]  

37. J. K. Zhang, R. H. Liu, D. P. Zhao, H. Wang, Y. Q. Zhang, C. M. Wang, and J. M. Shi, “Design, fabrication and characterization of a thin infrared-visible bi-stealth film based on one-dimensional photonic crystal,” Opt. Mater. Express 9(1), 195–202 (2019). [CrossRef]  

38. J. K. Zhang, D. P. Zhao, Z. S. Chen, Y. Liu, H. Wang, Z. Q. Lin, and J. M. Shi, “One dimensional photonic crystal based multilayer film with low IR and visible signatures,” Opt. Mater. 91, 261–267 (2019). [CrossRef]  

39. Y. Wu, J. Luo, M. Pu, B. Liu, J. Jin, X. Li, X. Ma, Y. Guo, Y. Guo, and X. Luo, “Optically transparent infrared selective emitter for visible-infrared compatible camouflage,” Opt. Express 30(10), 17259–17269 (2022). [CrossRef]  

40. N. Lee, J.-S. Lim, I. Chang, D. Lee, and H. H. Cho, “Transparent Metamaterials for Multispectral Camouflage with Thermal Management,” Int. J. Heat Mass Transfer 173, 121173 (2021). [CrossRef]  

41. V. Liu and S. Fan, “S4 : A free electromagnetic solver for layered periodic structures,” Comput. Phys. Commun. 183(10), 2233–2244 (2012). [CrossRef]  

42. A. F. Gad, “Pygad: An intuitive genetic algorithm python library,” arXiv, arXiv preprint arXiv:2106.06158 (2021).

43. W. Zhang and D. Lv, “Preparation and characterization of Ge/TiO2//Si/SiO2 one-dimensional heterostructure photonic crystal with infrared spectrally selective low emissivity,” Opt. Mater. 96, 109333 (2019). [CrossRef]  

44. L. Wang, Y. Yang, X. Tang, B. Li, Y. Hu, Y. Zhu, and H. Yang, “Combined multi-band infrared camouflage and thermal management via a simple multilayer structure design,” Opt. Lett. 46(20), 5224–5227 (2021). [CrossRef]  

45. X. Liu, P. Wang, C. Xiao, L. Fu, J. Xu, D. Zhang, H. Zhou, and T. Fan, “Compatible Stealth Metasurface for Laser and Infrared with Radiative Thermal Engineering Enabled by Machine Learning,” Adv. Funct. Mater. 33(11), 2212068 (2023). [CrossRef]  

46. L. Ma, K. Hu, C. Wang, J. Y. Yang, and L. Liu, “Prediction and Inverse Design of Structural Colors of Nanoparticle Systems via Deep Neural Network,” Nanomaterials 11(12), 3339 (2021). [CrossRef]  

47. L. Hu, L. Ma, C. Wang, and L. Liu, “Machine learning-assisted design of polarization-controlled dynamically switchable full-color metasurfaces,” Opt. Express 30(15), 26519–26533 (2022). [CrossRef]  

48. L. Wang, J. Dong, W. Zhang, C. Zheng, and L. Liu, “Deep Learning Assisted Optimization of Metasurface for Multi-Band Compatible Infrared Stealth and Radiative Thermal Management,” Nanomaterials 13(6), 1030 (2023). [CrossRef]  

49. M. A. Kats, R. Blanchard, P. Genevet, and F. Capasso, “Nanometre optical coatings based on strong interference effects in highly absorbing media,” Nat. Mater. 12(1), 20–24 (2013). [CrossRef]  

50. M. A. Kats and F. Capasso, “Optical absorbers based on strong interference in ultra-thin films,” Laser Photonics Rev. 10(5), 735–749 (2016). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       See Supplement for the effect of quartz thickness, the material optical constants, the OCP and SRR, and the GA convergence.

Data availability

All calculation data that support this work are available upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. (a) The application scenarios of camouflage for visible, infrared and lasers with radiative cooling and the schematic diagram of multilayer film structure (MFS) composed of SiO2, Ge and TiO2 films; (b) Schematic diagram of the workflow of GA; (c) Schematic diagram of the optimized MFS; (d) The ideal emission spectra (red line), the emission/absorption spectra of the optimized MFS (blue line) and the atmospheric transmission spectra calculated by MODTRAN (light blue area).
Fig. 2.
Fig. 2. Compatibility of visible structural colors with IR emission/absorption. (a) Reflected structural colors generated by the designed MFS for varying thickness of the top TiO2 layer hTiO2 in the CIE-1931 color space. The dashed line represents route of the color change as hTiO2 varies from 70 to 200 nm with 10 nm intervals; (b) The visible reflection spectra of the designed MFS as hTiO2 varies from 70 nm to 200 nm, the color of the line represents the structural color of the designed MFS with the corresponding hTiO2; (c) The IR emission/absorption spectra of the designed MFS as hTiO2 varies from 70 nm to 200 nm; (d) The structural colors generated by the designed MFS for different hTiO2 and the corresponding absorptivity at 1.06 µm.
Fig. 3.
Fig. 3. The spectral absorptivity for TM- and TE-polarizations, and the averaged value over the two polarizations as a function of incident angle at the wavelengths of (a) 1.06 µm and (b) 10.6 µm. Contour plot of the emissivity with respect to the wavelength and the incident angle for (c) TM-polarization and (d) TE-polarization.
Fig. 4.
Fig. 4. (a) The spectral reflectivity, transmissivity and absorptivity of the upper 9 layers without the quartz substrate; (b) The normalized electric field intensity distribution as a function of wavelength from 3 µm to 14 µm in the upper 9 layers.
Fig. 5.
Fig. 5. Distribution of the electric field intensity in the plane of y = 0 at the wavelengths of (a) 1.06 µm and (b) 10.6 µm. Distribution of the normalized electromagnetic power loss density in the plane of y = 0 at at the wavelengths of (c) 1.06 µm and (d) 10.6 µm.
Fig. 6.
Fig. 6. Overall camouflage performance (OCP) evaluation. (a) OCP of the designed MFS as a function of the temperature for different hTiO2; (b) The radiation intensity calculated by Eq. (4) of the designed MFS in the MWIR, 5∼8 µm and LWIR bands and their ratio in Eq. (3) for temperatures varying from 293 K to 573 K; (c) The spectral radiation intensity of blackbody and the designed MFS in the MIR band at temperatures of 523 K and 300 K; (d) The signal reduction rate (SRR) calculated by Eqs. (5)∼(7) of the designed MFS in the MWIR and LWIR bands and the radiation intensity in the 5∼8 µm band as temperature varies from 293 K to 1093 K.

Tables (2)

Tables Icon

Table 1. The thickness and material of each layer of the optimized MFS (from top to bottom)

Tables Icon

Table 2. Comparison of the recent reported results for multi-band camouflage with this work

Equations (7)

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

F = 3 α ¯ 1 .0 1 .1 μ m + 3 ( 1 ε ¯ MWIR ) + 4 ε ¯ 5 8 μ m + 3 α 10 .6 μ m + 5 ε ¯ LWIR
ε ¯ [ λ 1 , λ 2 ] = λ 1 λ 2 ε ( λ ) E b ( λ , T ) d λ / λ 1 λ 2 E b ( λ , T ) d λ
OCP = P 5 8 μ m 2 P MWIR P LWIR α 10 .6 μ m ε ¯ LWIR α 1 .06 μ m
P [ λ 1 , λ 2 ] = λ 1 λ 2 ε ( λ ) E b ( λ , T ) d λ
SRR = ( 1 α ( T ) β ( T ) ) × 100 %
α ( T ) = λ 1 λ 2 ε ( λ ) E b ( λ , T ) d λ
β ( T ) = λ 1 λ 2 E b ( λ , T ) d λ
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.