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

Visible and infrared dual-band anti-counterfeiting with self-assembled photonic heterostructures

Open Access Open Access

Abstract

Self-assembled photonic structures have greatly expanded the paradigm of optical materials due to their ease of access, the richness of results offered and the strong interaction with light. Among them, photonic heterostructure shows unprecedent advances in exploring novel optical responses that only can be realized by interfaces or multiple components. In this work, we realize visible and infrared dual-band anti-counterfeiting using metamaterial (MM) - photonic crystal (PhC) heterostructures for the first time. Sedimentation of TiO2 nanoparticles in horizontal mode and polystyrene (PS) microspheres in vertical mode self-assembles a van der Waals interface, connecting TiO2 MM to PS PhC. Difference of characteristic length scales between two components support photonic bandgap engineering in the visible band, and creates a concrete interface at mid-infrared to prevent interference. Consequently, the encoded TiO2 MM is hidden by structurally colored PS PhC and visualized either by adding refractive index matching liquid or by thermal imaging. The well-defined compatibility of optical modes and facility in interface treatments further paves the way for multifunctional photonic heterostructures.

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

1. Introduction

Photonic structure with characteristic length at operational wavelength scale, e.g. photonic crystals (PhC), or subwavelength scale, e.g. metamaterials (MMs), can interact strongly with light. Various remarkable optical responses ranging from photonic bandgap [13], slow light [47], to negative refraction [812] and other phenomena [13,14], are thus intrigued to beneficial many practical applications including lasing [1518], fiber [19,20], sensing [2123] and optical circuits [2426]. These responses can be finely designed and tuned by the types and parameters of their structures. Being more specific, PhCs and MMs are two fundamental types of photonic structures developed from distinct light-matter interactions. As first proposed by Yablonovitch [27] and John [28] at 1987, PhCs are periodic dielectric structures that possess photonic band structure. They are functionalized by Bragg’s diffraction law, to confine, manipulate, and guide light. For MMs, quasi-static approximation allows the structures to be regarded as effective media with averaged parameters that can be engineered to modify the propagation of electromagnetic waves.

As a matter of fact, photonic heterostructures combining the merits from each component are well-known. Among them, self-assembled photonic heterostructures show great potential in application, due to their ease of fabrication, low cost, and arbitrary shape designer [29,30]. The most reported self-assembled photonic heterostructures are PhC heterostructures composed of various PhC components that are developed as a photonic analog to the semiconductor heterostructures [31]. The compatibility of optical modes and facility in interface treatments are well-guaranteed as working and fabricating in the same paradigm. Therefore, self-assembled PhC heterostructures with enlarged photonic bandgap and minimized propagation losses, have been explored for wideband optical mirror [32,33], low-loss photonic crystal waveguides [34], and etc [3538]. Further expanding the interface and components, the combination of PhCs with other distinct photonic structures gives rise to more interesting optical responses. For example, self-assembled PhC-diffraction grating heterostructure has been proposed to propagate more light through breaking the sharp angular selectivity of the ordinary diffraction grating [39]. On the other hand, self-assembled MMs have been widely studied to achieve exotic optical parameters determined by elements and structures. Beyond the scope of PhCs [4042], these MMs could enable a plethora of exciting application in the light trapping [4345], subwavelength imaging [46,47], and wideband absorber [48,49]. Considering the rich physics and the extensive studies in self-assembled MMs that shows fabrication compatibility to self-assembled PhC, self-assembled MM-PhC or PhC-MM heterostructures are fundamentally and applicably promising.

PhCs matched in visible band would produce structural colors due to Bragg’s law of diffraction. Structural color derived from PhCs has attracted increasing interest in the field of anti-counterfeiting label in visible band [5052]. Yao Meng et al. [53] achieved the anti-counterfeiting application of banknotes by using a 3D inverse opal structure. Lin Chu, et al. [54] realized the anti-counterfeiting of banknotes through hollow silica opals/cellulose acetate nanocomposite films with structural colors. Self-assembly of mono-dispersed colloidal microspheres is favorable approach to form periodic 3D PCs, based on a variety of assembly technologies [5557]. The vertical deposition self-assembly method has been widely used in the fabrication of 3D colloidal microsphere photonic crystals due to the advantages of controllable crystal layer thickness, simple equipment, economic, short preparation period [5861]. In addition to using 3D photonic crystals to realize the application of anti-counterfeiting labels in visible band, thermal radiation labels with customized infrared emissivity realized by MMs have received important research attention as a promising anti-counterfeiting candidate [6265]. These thermal radiation labels are imperceptible to the human eyes and can only be revealed using thermographic equipment. Daniel Franklin, et al. [66] discussed a cavity-coupled plasmonic system with resonances that are tunable across the 3–5 or 8–14 µm infrared bands while retaining near-invariant spectral properties in the visible domain. Through experiments, they realized the covert infrared image encoding. Joong Hoon Lee, et al. [67] proposed a colored, covert infrared display with visible and thermally concurrently encoded data using perfect spectral division based on a hybrid planar-plasmonic cavities. Although there are many researches on visible light anti-counterfeiting label and infrared thermal radiation label, the previous researches have focused on a single function in a single band. So far, there is no research on visible and infrared dual-band anti-counterfeiting labels.

In this work, we realize photonic MM-PhC heterostructures for the first time by optical modes segmentation in design and two-step self-assembly in fabrication to demonstrate dual-band anti-counterfeiting labels. The proposed TiO2 MMs and polystyrene (PS) PhCs have characteristic length, i.e. self-assembled particle sizes, in nanometer (∼ 25 nm) scale and sub-micrometer (∼ 120 nm) scale respectively. Close packing of particles in three dimensions combines them together to form MM-PhC heterostructures. The difference in compositional length scale supports photonic bandgap engineering in the visible bands, and creates a concrete interface at mid-infrared to prevent interference. To demonstrate the dual band function of our heterostructures, a new type of anti-counterfeiting labels is introduced, where coding patterns are hidden by structural coloring and visualized by either adding refractive index matching liquid or thermal imaging. Our design and fabrication thus pave the way for multi-functional photonic heterostructures composed of distinct components.

2. Theory

To analyze the working mechanism of the proposed heterostructures, characteristics of the PhCs, MM, and MM-PhC heterostructures are compared through theoretical calculations and numerical simulations in this section. Multiphysics simulation software (COMSOL) is used to obtain photonic band structure of the PS PhCs via eigenfrequency solver, and spectral characteristics of all the proposed structures via frequency domain solver.

2.1 PS PhC

Self-assembled PhC is formed by PS microspheres of radius r = 110 nm in 3D face-centered cubic (FCC) lattice in air (see Fig. 1(a) for a unit cell) with lattice constant a = r / 0.3535. Complex permittivity of the PS microsphere is shown in Fig. S1 (a). The corresponding photonic band structure (see Fig. 1(b)) opens no complete band gap due to high symmetry of FCC structure. [68] However, a directional band gap (493 nm ∼ 525 nm) appears from LG (crystal plane (111) direction) as reported in [69], which leads to a square shaped high reflection spectrum in Fig. 1(c). This reflection band of reflection coefficient over 0.95 from 498.7 nm to 531.3 nm gives a structural color of green to the PS PhC (see the chromaticity diagram in Fig. S2). The center wavelength (λgap) of the bandgap can be verified by Bragg’s diffraction law for the (111) plane of FCC structure:

$${{\lambda }_{\textrm{gap}}}{ = 2}{{d}_{{111}}}{{n}_{{eff}}}. $$
where ${{\rm d}_{{111}}}$. is (111) plane spacing (${{d}_{{111}}} { = D} \sqrt{2} / \sqrt{3} $., with D = 2r the diameter of PS microspheres) and neff is the effective refractive index of the PS PhC. The effective parameter is related to filling ratio fps of the PS microseres by:
$${{n}_{{eff}}}{ = }{{f}_{{ps}}}{{n}_{{ps}}}{ + (1 - }{{f}_{{ps}}}{)}{{n}_{m}}$$
nps and nm the refractive index of PS microspheres and the medium they filled, respectively. Here, fps = 0.74, nps = 1.59 and nm = 1 for air, so the value of neff is 1.4366. Following Eq. (1), λgap is 516.11 nm, close to the simulated value, i.e. 515 nm. Note that the structural color disappears once the air gaps in PS PhC are filled by a refractive index matching media, e.g. alcohol. The weak refractive index contrast between the two components prevents the formation of photonic band structure [70]. Consequently, the PS PhC becomes transparent in alcohol. At infrared wavelength, the proposed PS PhC is highly transparent in addition to the weak absorption caused by the intrinsic loss of PS at 13.2 µm (see Fig. 1(d)).

 figure: Fig. 1.

Fig. 1. Numerical study of self-assembled PS PhC. (a) Schematic of the 3D FCC PhC unit cell (r/a = 0.3535, r = 110 nm). (b) Photonic band structure of the PS PhC. Inset: diagram of the First Brillouin zone. (c) The visible spectrum of the PS PhC in air and alcohol medium. Here, R / T means reflection / transmission. The dark green line indicates the center wavelength of reflection band. Inset: schematic of the simulated PhC structure with a thickness of 41 × (61/2/3) × (2r) + (2r), i.e. total 42 layers of PS microspheres. The colored arrow indicates the normal incidence of light. (d) The infrared spectrum of the PS PhC in air. The same configuration of (c). Here, A means absorption.

Download Full Size | PDF

Next, we study the PS PhC with r = 130 nm (other structural and material parameters remain unchanged), which is shown in Fig. S3. Compared with the previous case, the band gap red-shifts about 93 nm for the center wavelength, and the reflection spectrum red-shifts about 89 nm. This phenomenon conforms to Bragg’s diffraction law as the increase in r causes the increase in λgap. The calculated λgap = 609.9 nm agrees with the simulated one (604 nm), promising a structural color of red. The radius of PS microspheres has negligible influence on other properties of the PS PhC, i.e. the invalidation of structural coloring in alcohol and the IR spectral characteristics.

2.2 TiO2 MM

Optical response in visible band for self-assembled TiO2 MM is mainly determined by the Mie scattering of nanoparticles [71]. According to Mie theory, the efficiency factor for scattering Qsca is calculated by [72]:

$${{Q}_{{sca}}}{ = }\frac{{{{C}_{{sca}}}}}{{G}}{ = (}\mathop \smallint \nolimits_{0}^{{2\pi }} \mathop \smallint \nolimits_{0}^{{2\pi }} \frac{{{{|{X} |}^{2}}}}{{{{k}^{2}}}}{sin\theta d\theta d\varphi )/G\ =\ }{\smallint _{{4\pi }}}\frac{{{{|{X} |}^{2}}}}{{{{k}^{2}}}}{d\Omega /G}.$$
where |X|2/k2 is sometimes called the differential scattering cross section, and G is the particle cross-sectional area projected onto a plane perpendicular to the incident beam (G = πr2 for a sphere of radius r). Numerical result unveils that a single TiO2 nanosphere of r = 25 nm without absorption has a relatively strong Qsca in the visible frequencies, due to the electric quadrupole resonance (see Fig. 2(a-b)). For TiO2 MM, the compositional nanoparticles have radii ranging from ∼15 nm to 35 nm, and hence form a relatively rough surface of diffuse reflection. The high reflection spans the visible spectrum (in Fig. 2(c)) and colors the MM in bright white.

The infrared spectrum of TiO2 MM shows a high absorption at 9 µm - 14 µm due to the intrinsic loss of the material (in Fig. 2(d)). According to Kirchhoff's law [73], the absorptivity for a material body in thermodynamic equilibrium is equal to its emissivity. The TiO2 MM thus has high emissivity at 9 µm - 14 µm. Like [74,75], self-assembled TiO2 MM here is treated as a thin film structure to obtain the simulated spectra. The effective permittivity of the thin film is calculated using effective medium theory, i.e. using Eq. (2) but with TiO2 replacing PS. Here, the filling ratio (f) remains at 0.74 due to the close packing of TiO2 nanospheres in fcc lattice.

 figure: Fig. 2.

Fig. 2. Numerical study of the self-assembled TiO2 MM. (a) A diagram of light scattering by a single dielectric spherical nanoparticle due to Mie resonance. The color indicates visible band. (b) The scattering efficiency of spherical TiO2 nanoparticle with radius of 25 nm. Inset: electric field distribution at the center cross section of nanoparticles at 500 nm wavelength. (c) The visible spectrum of TiO2 MM. Inset: schematic of the simulated TiO2 MM structure as an equivalent thin film with thickness 10 µm. (d) The infrared simulation spectrum of TiO2 MM. The same structure of (c). Here, R / T / A means reflection / transmission / absorption.

Download Full Size | PDF

2.3 TiO2 MM - PS PhC heterostructure

In the visible spectra, the self-assembled TiO2 MM-PS PhC heterostructures on a Si substrate host the reflection bands of the upper PS PhCs (see Fig. 3 (a, c)), as their photonic band-gap prohibits light transmission. Beyond the bandgap region, light transmits through the whole structure. Only weak diffraction peaks are observed, while the main energy attenuates in the highly loss Si (see Fig. S4). It is the MM-PS interface that smooths the highly reflectivity of TiO2 MMs and allows the shielding of their white color by structural coloring. The center wavelength of reflection band red-shifts about 90 nm when the radius of PS microspheres changes from 110 nm to 130 nm.

 figure: Fig. 3.

Fig. 3. Numerical study of the self-assembled TiO2 MM-PS PhC heterostructure on Si substrate. (a) The visible spectrum of TiO2 MM- PS (r = 110 nm) PhC heterostructure. Inset: schematic of the simulated TiO2 MM-PS PhC heterostructure. Here, the thickness of PhC, MM and Si components are 20 × (61/2/3) × (2r) + (2r), i.e. total 21 layers of PS microspheres, 2 µm and 2 µm, respectively. The colored arrow indicates the normal incidence of light. (b) The infrared simulation spectrum of TiO2 MM- PS (r = 110 nm) PhC heterostructure. (c) The visible spectrum of TiO2 MM- PS (r = 130 nm) PhC heterostructure. (d) The infrared simulation spectrum of TiO2 MM- PS (r = 130 nm) PhC heterostructure. (b-d) has same structure of (a). Here, R / T / A means reflection / transmission / absorption.

Download Full Size | PDF

The infrared spectra of the heterostructures mainly show high absorptivity in the range of 9 µm - 14 µm (see Fig. 3(b, d)). Except for a slight increase in absorption at 13.2 µm due to the intrinsic loss of PS material, they are very similar to the infrared spectrum of TiO2 MM. Therefore, PS PhCs have little influence on the infrared response of TiO2 MM, corresponding to the high transmission of PS PhCs (see Fig. 1 (d) and Fig. S3 (d)). The heterostructures thus achieve independent control of optical modes in visible and infrared bands.

3. Results and discussion

In experiments, the PhCs, MMs, and MM-PhC heterostructures are fabricated and analyzed. To optimize the dual band performance, thickness of each structure, as well as PhC and MM components in heterostructures, are much larger than that proposed in simulation studies. Note that, neither photonic bandgap engineering nor infrared imaging are thickness-dependent. Therefore, the proposed mechanism, i.e. the simulated optical modes, are highly consistent with the experimental observations as discussed below.

3.1 Properties of PS PhCs

To fabricate PS PhCs, the hydrophilic silicon substrate is vertically inserted into a beaker containing PS microspheres dispersion solution. Then, the beaker is placed in an oven at 50 °C for vertical deposition self-assembly. The schematic diagram is shown in Fig. S5 (also see fabrication and measurement details there). It is clear that PS microspheres are closely packed in a 3D fcc lattice at thicknesses about 23 µm (for r = 110 nm, in Fig. 4(a-b)) or 17 µm (for r = 130 nm, in Fig. 4(e-f)). The center wavelength of photonic bandgap is measured by the spectrometer, i.e. at 519 nm / 606 nm for PS PhCs with r = 110 nm / 130 nm respectively (in Fig. 4(c) and (g)). Meanwhile, the shape of spectral peak is so much narrower and sharper in experiment than in simulation. It is because that he photonic crystal prepared in experiment is not perfect and has defects. In fact, there is substrate material in the samples [78]. At infrared, the PS PhCs are highly transparent except for weak absorption caused by the intrinsic loss of the PS at 13.2 µm (see Fig. 4(d) and (h)), comparing to the bare Si substrate.

 figure: Fig. 4.

Fig. 4. Experimental study of self-assembled PS PhCs. (a) SEM image of PS PhC (r = 110 nm) surface. Inset: zoom-in picture. (b) SEM image of PS PhC (r = 110 nm) cross-section. The thicknesses of PS PhCs are show in Fig. S6. (c-d) The visible and infrared spectrum of Si substrate with and without PS PhC (r = 110 nm) in air. (e) SEM image of PS PhC (r = 130 nm) surface. Inset: zoom-in picture. (f) SEM image of PS PhC (r = 130 nm) cross-section. (g-h) The visible and infrared spectrum of Si substrate with and without PS PhC (r = 130 nm) in air.

Download Full Size | PDF

3.2 Properties of TiO2 MM

Unlike PS PhCs, TiO2 MM are self-assembled via horizontal sedimentation on a Si substrate. The TiO2 nanoparticles have uneven size distribution in range of ∼ 15 nm to 35 nm, that usually cause large crack spacing in vertical drying nanoparticle film [76]. Horizontal sedimentation helps to ease the effect. Moreover, the dense film formed by this method shows relatively rough surface (see Fig. 5(a-b)), that means good hydrophilicity [77]. It is beneficial to the subsequent vertical deposition of PS PhCs in the next session. Comparing to the bare Si substrate, the visible spectrum of self-assembled TiO2 MM shows a moderately increased reflection (see Fig. 5(c)). At infrared, the TiO2 MM has high absorption (emissivity) at 9 µm -14 µm.

 figure: Fig. 5.

Fig. 5. Experimental study of the self-assembled TiO2 MM. (a) SEM image of TiO2 MM surface. (b) SEM image of TiO2 MM cross-section. Inset: zoom-out of the film sample to mark the thickness. (c-d) The visible and infrared spectrum of Si substrate with and without TiO2 MM.

Download Full Size | PDF

3.3 Properties of TiO2 MM- PS PhC heterostructures

The TiO2 MM - PS PhC (r = 110 nm and 130 nm) heterostructures are self-assembled via the combination of horizontal and vertical sedimentation methods in our experiment. The thickness of each component in our heterostructures (see Fig. S7) is basically consistent with that of the individual structures in sections 3.1 and 3.2. A van der Waals interface pastes TiO2 MM and PS PhCs together concretely (see Fig. 6(b) and (f)). Fabrication compatibility of two photonic structures are well-established at this interface, by virtue of the prominent size-tolerance in particle packing and hydrophilicity of TiO2 MM surface as discussed above. Compared with the visible spectra of PS PhCs (see Fig. 4(c) and (g)), the visible spectra of the heterostructures (see Fig. 6(c) and (g)) are basically the same. At infrared, the heterostructures have high absorption at 9 µm - 14 µm (in Fig. 6(d) and (h)), basically consistent with that of TiO2 MM.

 figure: Fig. 6.

Fig. 6. Experimental studies of the self-assembled TiO2 MM-PS PhC heterostructures. (a) SEM image of TiO2 MM-PS (r = 110 nm) PhC heterostructure surface. (b) SEM image of TiO2 MM-PS (r = 110 nm) PhC heterostructure cross-section. (c-d) The visible and infrared spectrum of Si substrate with and without TiO2 MM-PS (r = 110 nm) PhC heterostructure. (e) SEM image of TiO2 MM-PS (r = 130 nm) PhC heterostructure surface. (f) SEM image of TiO2 MM-PS (r = 130 nm) PhC heterostructure cross-section. (g-h) The visible and infrared spectrum of Si substrate with and without TiO2 MM-PS (r = 130 nm) PhC heterostructure. The thickness of PS PhC is about 22.4 µm (for r = 110 nm) or 18.6 µm (r = 130 nm) and the thickness of TiO2 MM is about 14 µm (see Fig. S7).

Download Full Size | PDF

3.4 Properties of anti-counterfeiting label

In order to demonstrate the visible and infrared dual-band anti-counterfeiting effect, the self-assembled TiO2 MM-PS PhC heterostructures are prepared on pattern-engraved Si substrates. The thicknesses of PS PhC components in the obtained samples are the same as that in section 3.1. While the thicknesses of TiO2 MM components in the obtained samples are dependent on the engraving depth (see Fig. S8, large than that in section 3.2).

For the first sample, PS PhC with radius of 110 nm can hide the smiley face coding pattern containing TiO2 MM by green color (Fig. 7(a-b)). When the sample is immersed in alcohol medium, the underlying pattern can be clearly observed (Fig. 7(c)). This is because that PS PhC is highly transparent to visible light in alcohol medium (See Fig. 1(c)). After the alcohol volatilizes, the coding pattern can be hidden again (see Fig. 7(d)). Therefore, our heterostructure can realize the visible band anti-counterfeiting. Meanwhile, the underlying coding pattern emerges under FLIR thermal imaging camera (see Fig. 7(e-h)).

 figure: Fig. 7.

Fig. 7. Visible imaging results. (a) The smiley face coding pattern containing TiO2 MM in air. (b) TiO2 MM-PS PhC heterostructure in air. (c) TiO2 MM-PS PhC heterostructure in alcohol. (d) TiO2 MM-PS PhC heterostructure after alcohol volatilizes. Infrared imaging results. (e) The smiley face coding pattern containing TiO2 MM. (f) TiO2 MM-PS PhC heterostructure in air. (g) TiO2 MM-PS PhC heterostructure in alcohol. (h) TiO2 MM-PS PhC heterostructure after alcohol volatilizes.

Download Full Size | PDF

Next, we study another coding pattern (UESTC) based on TiO2 MM-PS PhC (r = 130 nm) heterostructure (see Fig. 8(a-h)). The change of coding pattern and PS microsphere’s radius does not affect the function of the designed heterostructure. However, the displayed patterns are black in Fig. 7(c) and white in Fig. 8 (c), while the unpattern regions are gold and pink respectively. This is due to the thickness difference of TiO2 MM and PS PhC in two samples.

 figure: Fig. 8.

Fig. 8. Visible imaging results. (a) The “UESTC” coding pattern containing TiO2 MM in air. (b) TiO2 MM-PS PhC heterostructure in air. (c) TiO2 MM-PS PhC heterostructure in alcohol. (d) TiO2 MM-PS PhC heterostructure after alcohol volatilizes. Infrared imaging results. (e) The “UESTC” coding pattern containing TiO2 MM. (f) TiO2 MM-PS PhC heterostructure in air. (g) TiO2 MM-PS PhC heterostructure in alcohol. (h) TiO2 MM-PS PhC heterostructure after alcohol volatilizes.

Download Full Size | PDF

4. Conclusion

In this study, we realize photonic MM-PhC heterostructures and demonstrate the function integration in a dual-band manner. Our design and fabrication methods make full use of large difference in characteristic lengths between MM and PhC components: theoretically, effective media approximation and photonic bandgap engineer allows different optical modes, and a refractive-index-matched interface moderates unwanted interferences; experimentally, uneven surfaces formed by dense packing of small particles facilitate large particle packing on them, making mechanically stable interfaces.

Explicitly, the self-assembled TiO2 MM-polystyrene (PS) PhC heterostructures are prepared by vertical and subsequent horizontal sedimentations. A van der Waals interface pastes the two distinct photonic structures. High emission TiO2 MMs filled in the substrate slit creates light paths in the visible and thermal imaging patterns at mid-infrared due to deep subwavelength-size of TiO2. Bandgap engineering of PS PhCs allows structural coloring in the visible due to wavelength-size of PSs. At mid-infrared, PS PhCs recede to a high transparent effective media to pass imaging lights.

Due to this kind of optical modes segmentation and interface matching, a new type of dual-band anti-counterfeiting labels is implemented by our heterostructures. We have also paved the way for design and fabrication of multi-functional photonic heterostructures and their application in photonics.

Funding

Regional Joint Key Fund of National Natural Science Foundation of China (U19A2096); National Natural Science Foundation of China (52021001, 52022018).

Acknowledgments

Peiheng Zhou ang Jianliang Xie thank the National Natural Science Foundation of China for help identifying collaborators for this work.

Disclosures

The authors declare no conflicts of interest.

Data availability

No data were generated or analyzed in the presented research.

Supplemental document

See Supplement 1 for supporting content.

References

1. T. Tajiri, S. Takahashi, A. M. Harteveld C, Y. Arakawa, S. Iwamoto, and L. Vos W, “Reflectivity of three-dimensional GaAs photonic band-gap crystals of finite thickness,” Phys. Rev. B 101(23), 235303 (2020). [CrossRef]  

2. M. Adhikary, R. Uppu, A. M. Harteveld C, A. Grishina D, and L. Vos W, “Experimental probe of a complete 3D photonic band gap,” Opt. Express 28(3), 2683–2698 (2020). [CrossRef]  

3. J. Ma, Z. G. Wang, X. J. Liu, S. Q. Zhang, Y. Liang, and X. Y. Wu, “Photonic band gap structure properties of two-dimensional function photonic crystals,” Physica E: Low-dimensional Systems and Nanostructures 89, 61–66 (2017). [CrossRef]  

4. M. C. Zhang, Y. Xu, Y. Liu, Y. Li, and D. Wen, “Numerical study on slow-light effects in an ultra-compact grating-based Fabry-Perot cavity,” Opt. Quantum Electron. 52(4), 210 (2020). [CrossRef]  

5. J. H. Chu, O. Voskoboynikov, and C. P. Lee, “Slow light in photonic crystals,” Microelectron. J. 36(3-6), 282–284 (2005). [CrossRef]  

6. L. Thévenaz, I. Dicaire, and S. Chin, “Enhancing the light-matter interaction using slow light: towards the concept of dense light,” Advances in Slow and Fast Light V 8273, 82731D (2012). [CrossRef]  

7. H. Yoshimi, T. Yamaguchi, R. Katsumi, Y. Ota, Y. Arakawa, and S. Iwamoto, “Experimental demonstration of topological slow light waveguides in valley photonic crystals,” Opt. Express 29(9), 13441–13450 (2021). [CrossRef]  

8. A. Shelby R, R. Smith D, and S. Schultz, “Experimental verification of a negative index of refraction,” Science 292(5514), 77–79 (2001). [CrossRef]  

9. J. Valentine, S. Zhang, T. Zentgraf, E. Ulin-Avila, D. A. Genov, G. Bartal, and X. Zhang, “Three-dimensional optical metamaterial with a negative refractive index,” Nature 455(7211), 376–379 (2008). [CrossRef]  

10. R. Smith D, B. Pendry J, and C. Wiltshire M, “Metamaterials and negative refractive index,” Science 305(5685), 788–792 (2004). [CrossRef]  

11. B. Pendry J, “Negative refraction index makes perfect lens,” Phys. Rev. Lett. 85(18), 3966–3969 (2000). [CrossRef]  

12. E. Cubukcu, K. Aydin, E. Ozbay, S. Foteinopoulou, and M. Soukoulis C, “Subwavelength resolution in a two-dimensional photonic-crystal-based superlens,” Phys. Rev. Lett. 91(20), 207401 (2003). [CrossRef]  

13. D. Lu, J. Kan J, E. Fullerton E, and Z. Liu, “Enhancing spontaneous emission rates of molecules using nanopatterned multilayer hyperbolic metamaterials,” Nat. Nanotechnol. 9(1), 48–53 (2014). [CrossRef]  

14. M. Kang, J. Chen, X. Cui H, N. Li Y, and T. Wang H, “Asymmetric transmission for linearly polarized electromagnetic radiation,” Opt. Express 19(9), 8347–8356 (2011). [CrossRef]  

15. S. Wuestner, M. Hamm J, A. Pusch, F. Renn, L. Tsakmakidis K, and O. Hess, “Control and Dynamic Competition of Bright and Dark Lasing States in Active Nanoplasmonic Metamaterials,” Phys. Rev. B 85, 201406 (2012). [CrossRef]  

16. S. Droulias, T. Koschny, M. Kafesaki, and M. Soukoulis C, “On loss compensation, amplification and lasing in metallic metamaterials,” Nanomaterials and Nanotechnology 9, 184798041881794 (2019). [CrossRef]  

17. O. Painter, K. Lee R, A. Scherer, A. Yariv, D. O’Brien J, D. Dapkus P, and I. Kim, “Two-dimensional photonic band-gap defect mode laser,” Science 284(5421), 1819–1821 (1999). [CrossRef]  

18. S. Iadanza, A. Liles A, M. Butler S, P. Hegarty S, and L. O’Faolain, “Photonic crystal lasers: from photonic crystal surface emitting lasers (PCSELs) to hybrid external cavity lasers (HECLs) and topological PhC lasers,” Opt. Mater. Express 11(9), 3245–3274 (2021). [CrossRef]  

19. P. Russell, “Photonic crystal fibers,” Science 299(5605), 358–362 (2003). [CrossRef]  

20. N. Mahnot, S. Maheshwary, and R. Mehra, “Photonic crystal fiber-an overview,” Int. J. Sci. Eng. Res. 6, 45–53 (2015).

21. G. Pitruzzello and F. Krauss T, “Photonic crystal resonances for sensing and imaging,” J. Opt. 20(7), 073004 (2018). [CrossRef]  

22. C. Fenzl, T. Hirsch, and S. Wolfbeis O, “Photonic crystals for chemical sensing and biosensing,” Angew. Chem. Int. Ed. 53(13), 3318–3335 (2014). [CrossRef]  

23. T. Chen and L. S. Y. Sun, “Metamaterials application in sensing,” Sensors 12(3), 2742–2765 (2012). [CrossRef]  

24. N. Engheta, “Circuits with light at nanoscales: optical nanocircuits inspired by metamaterials,” Science 317(5845), 1698–1702 (2007). [CrossRef]  

25. L. Min and L. Huang, “All-semiconductor metamaterial-based optical circuit board at the microscale,” J. Appl. Phys. 118(1), 013104 (2015). [CrossRef]  

26. H. Rong Y, Y. Bin G, and Y. Yu W, “Study of photonic crystal slow-wave circuits,” J. Infrared, Millimeter, Terahertz Waves 30(9), 982–993 (2009). [CrossRef]  

27. E. Yablonovitch, “Inhibited spontaneous emission in solid-state physics and electronics,” Phys. Rev. Lett. 58(20), 2059–2062 (1987). [CrossRef]  

28. S. John, “Strong localization of photons in certain disordered dielectric superlattices,” Phys. Rev. Lett. 58(23), 2486–2489 (1987). [CrossRef]  

29. A. Baron, A. Aradian, V. Ponsinet, and P. Barois, “Self-assembled optical metamaterials,” Opt. Laser Technol. 82, 94–100 (2016). [CrossRef]  

30. J. Fontana, J. Naciri, R. Rendell, and R. Ratna B, “Macroscopic self-assembly and optical characterization of nanoparticle-ligand metamaterials,” Adv. Opt. Mater. 1(1), 100–106 (2013). [CrossRef]  

31. E. Istrate and H. Sargent E, “Photonic crystal heterostructures and interfaces,” Rev. Mod. Phys. 78(2), 455–481 (2006). [CrossRef]  

32. A. Sharkawy, Y. Shi S, and W. Prather D, “Heterostructure photonic crystals theory and applications,” Appl. Opt. 41(34), 7245–7253 (2002). [CrossRef]  

33. X. Wang, X. Hu, Y. Li, W. Jia, C. Xu, X. Liu, and J. Zi, “Enlargement of omnidirectional total reflection frequency range in one-dimensional photonic crystals by using photonic heterostructures,” Appl. Phys. Lett. 80(23), 4291–4293 (2002). [CrossRef]  

34. T. H. Pei , and Y. T. Huang, “The heterostructure photonic crystal waveguide splitter,” IEEE Photonics Technol. Lett. 23(16), 1145–1147 (2011). [CrossRef]  

35. B. Céline, H. P. Ngoc, D. M. Willy, C. Laurent, M. Agnès, and S. Catherine, “Manipulation of the fluorescence of nanocrystals by opal-based heterostructures,” J. Mater. Chem. C 3(37), 9734–9739 (2015). [CrossRef]  

36. E. Eftekhari, X. Li, H. Kim T, Z. Gan, S. Cole I, D. Zhao, D. Kielpinski, M. Gu, and Q. Li, “Anomalous fluorescence enhancement from double heterostructure 3D colloidal photonic crystals-a multifunctional fluorescence-based sensor platform,” Sci. Rep. 5(1), 14439 (2015). [CrossRef]  

37. E. Eftekhari, W. Wang, X. Li, Z. Wu, R. Klein, S. Cole I, and Q. Li, “Picomolar reversible Hg (II) solid-state sensor based on carbon dots in double heterostructure colloidal photonic crystals,” Sensors and Actuators B: Chemical 240, 204–211 (2017). [CrossRef]  

38. M. Djavid, A. Ghaffari, F. Monifi, and S. Abrishamian M, “Heterostructure photonic crystal channel drop filters using mirror cavities,” J. Opt. A: Pure Appl. Opt. 10(5), 055203 (2008). [CrossRef]  

39. A. Haldar, G. Kumar, and R. Vijaya, “Optical nanopatterned heterostructure with dual features of broad spectral distribution and wide spatial diffraction for beam steering in photonic devices,” ACS Appl. Nano Mater. 2(10), 6655–6661 (2019). [CrossRef]  

40. L. Li, L. Sun, S. Gomez-Diaz J, L. Hogan N, P. Lu, F. Khatkhatay, W. Zhang, J. Jian, J. Huang, Q. Su, M. Fan, C. Jacob, J. Li, and H. Wang, “Self-assembled epitaxial au-oxide vertically aligned nanocomposites for nanoscale metamaterials,” Nano Lett. 16(6), 3936–3943 (2016). [CrossRef]  

41. Z. Chen, X. Wang, Y. Qi, S. Yang, A. Soares J, A. Apgar B, R. Gao, R. Xu, Y. Lee, X. Zhang, J. Yao, and W. Martin L, “Self-assembled, nanostructured, tunable metamaterials via spinodal decomposition,” ACS Nano 10(11), 10237–10244 (2016). [CrossRef]  

42. S. Misra, D. Zhang, Z. Qi, D. Li, J. Lu, T. Chen H, and H. Wang, “Morphology control of self-assembled three-phase Au-BaTiO3-ZnO hybrid metamaterial for tunable optical properties,” Cryst. Growth Des. 20(9), 6101–6108 (2020). [CrossRef]  

43. P. Johnson N, L. Tsakmakidis K, O. Hess, E. Özbay, I. Zheludev N, and W. Ziolkowski R, “Slow and stopped light in metamaterials: the trapped rainbow,” Metamaterials III 6987, 698702 (2008). [CrossRef]  

44. L. Tsakmakidis K, D. Boardman A, and O. Hess, “‘Trapped rainbow’ storage of light in metamaterials,” Nature 450(7168), 397–401 (2007). [CrossRef]  

45. J. Xu, S. Xiao, P. He, Y. Wang, Y. Shen, L. Hong, Y. Luo, and B. He, “Realization of broadband truly rainbow trapping in gradient-index metamaterials,” Opt. Express 30(3), 3941–3953 (2022). [CrossRef]  

46. A. Salandrino and N. Engheta, “Far-field subdiffraction optical microscopy using metamaterial crystals: theory and simulations,” Phys. Rev. B 74(7), 075103 (2006). [CrossRef]  

47. L. Paldi R, X. Wang, X. Sun, Z. He, Z. Qi, X. Zhang, and H. Wang, “Vertically aligned AgxAu1-x alloyed nanopillars embedded in ZnO as nanoengineered low-loss hybrid plasmonic metamaterials,” Nano Lett. 20(5), 3778–3785 (2020). [CrossRef]  

48. X. Wang, J. Jian, Z. Zhou, C. Fan, Y. Dai, L. Li, J. Huang, J. Sun, A. Donohue, P. Bermel, X. Zhang, T. Chen H, and H. Wang, “Self-assembled Ag-TiN hybrid plasmonic metamaterial: tailorable tilted nanopillar and optical properties,” Adv. Opt. Mater. 7(3), 1801180 (2019). [CrossRef]  

49. Q. Chen, J. Gu, P. Liu, J. Xie, J. Wang, Y. Liu, and W. Zhu, “Nanowire-based ultra-wideband absorber for visible and ultraviolet light,” Opt. Laser Technol. 105, 102–105 (2018). [CrossRef]  

50. S. Wu, B. Liu, X. Su, and S. Zhang, “Structural Color Patterns on Paper Fabricated by Inkjet Printer and Their Application in Anticounterfeiting,” J. Phys. Chem. Lett. 8(13), 2835–2841 (2017). [CrossRef]  

51. Y. Meng, J. Qiu, S. Wu, B. Ju, and S. Zhang, “Biomimetic Structural Color Films with Bilayer Inverse Heterostructure for Anti-Counterfeiting Application,” ACS Appl. Mater. Interfaces 10(44), 38459–38465 (2018). [CrossRef]  

52. Y. Qi, L. Chu, W. Niu, B. Tang, and S. Zhang, “New Encryption Strategy of Photonic Crystals with Bilayer Inverse Heterostructure Guided from Transparency Response,” Adv. Funct. Mater. 29(40), 1903743 (2019). [CrossRef]  

53. M. Yao, F. Liu, M. Umair M, B. Ju, S. Zhang, and B. Tang, “Patterned and Iridescent Plastics with 3D Inverse Opal Structure for Anticounterfeiting of the Banknotes,” Adv. Opt. Mater. 6(8), 1701351 (2018). [CrossRef]  

54. L. Chu, X. Zhang, W. Niu, S. Wu, W. Ma, B. Tang, and S. Zhang, “Hollow silica opals/cellulose acetate nanocomposite films with structural colors for anti-counterfeiting of banknotes,” J. Mater. Chem. C 7(24), 7411–7417 (2019). [CrossRef]  

55. Y. Wang, S. Dou, L. Shang, P. Zhang, X. Yan, K. Zhang, J. Zhao, and Y. Li, “Effects of Microsphere Size on the Mechanical Properties of Photonic Crystals,” Crystals 8(12), 453 (2018). [CrossRef]  

56. X. Ye, H. Jin, Y. Zeng, L. Sun, F. Ceng, H. Liu, F. Wang, X. Jiang, W. Wu, and W. Zheng, “Monolayer Colloidal Crystals by Modified Air-Water Interface Self-Assembly Approach,” Nanomaterials 7(10), 291 (2017). [CrossRef]  

57. A. S. Sinitskii, V. V. Abramova, and Y. D. Tretyakov, “Structural and optical properties of titania photonic crystal films prepared by a sol–gel method,” Mendeleev Commun. 17(1), 1–3 (2007). [CrossRef]  

58. B. Hatton, L. Mishchenko, S. Davis, H. Sandhage K, and J. Aizenberg, “Assembly of large-area, highly ordered, crack-free inverse opal films,” Proc. Natl. Acad. Sci. 107(23), 10354–10359 (2010). [CrossRef]  

59. F. Farzaneh, C. Pouria, S. Allahveisi, S. Ahmadi-Kandjani, and M.-R. Rashidi, “Investigation of optical and physical property in opal films prepared by colloidal and freeze-dried microspheres,” Colloids and Surfaces A: Physicochem. Eng. Aspects 611, 125842 (2021). [CrossRef]  

60. H. Wu, G. Niu, W. Ren, L. Jiang, O. Liang, J. Zhao, Y. Liu, and Y. Xie, “Crucial Impact of Hydrophilicity on the Self-Assembled 2D Colloidal Crystals Using Langmuir–Blodgett Method,” Langmuir 36(34), 10061–10068 (2020). [CrossRef]  

61. S. Matsushita, T. Takagi, K. Kamimura, T. Kasaya, and T. Miyazaki H, “Enhanced light diffraction from self-assembled double-layer colloidal crystals,” J. Appl. Phys. 110(6), 2059 (2011). [CrossRef]  

62. R. J. H. Ng, R. V. Krishnan, H. Wang, and K. W. Yang J, “Erratum to: Darkfield colors from multi-periodic arrays of gap plasmon resonators,” Nanophotonics 9(6), 1553 (2020). [CrossRef]  

63. A. Tittl, A. U. Michel, M. Schferling, X. Yin, B. Gholipour, C. Long, M. Wuttig, T. Taubner, F. Neubrech, and H. Giessen, “A Switchable Mid-Infrared Plasmonic Perfect Absorber with Multispectral Thermal Imaging Capability,” Adv. Mater. 27(31), 4597–4603 (2015). [CrossRef]  

64. G. Bakan, S. Ayas, M. Serhatlioglu, C. Elbuken, and A. Dana, “Invisible Thin-Film Patterns with Strong Infrared Emission as an Optical Security Feature,” Adv. Opt. Mater. 6(21), 1800613 (2018). [CrossRef]  

65. C. Sayan, F. Daniel, C. Jared, S. Alireza, and C. Debashis, “Adaptive Multispectral Infrared Camouflage,” ACS Photonics 5(11), 4513–4519 (2018). [CrossRef]  

66. D. Franklin, S. Modak, A. Vázquez-Guardado, A. Safaei, and D. Chanda, “Covert infrared image encoding through imprinted plasmonic cavities,” Light: Sci. Appl. 7(6), 93 (2018). [CrossRef]  

67. J H Lee, Y J Kim, Y J Yoo, S Chang, G J Lee, J H Ko, K M Kang, D Chanda, and Y M. Song, “Colored, Covert Infrared Display through Hybrid Planar-Plasmonic Cavities [J],” Adv. Opt. Mater. 9(17), 2100429 (2021). [CrossRef]  

68. Y. Chen, L. Li, Q. Xu, W. Chen, Y. Dong, J. Fan, and D. Ma, “Recent advances in opal/inverted opal photonic crystal photocatalysts [J],” Sol. RRL 5(6), 2000541 (2021). [CrossRef]  

69. L. V. Mikhnev, E. A. Bondarenko, О. М. Chapura, A. A. Skomorokhov, and A. A. Kravtsov, “Influence of annealing temperature on optical properties of the photonic-crystal structures obtained by self-organization of colloidal microspheres of polystyrene and silica,” Opt. Mater. 75, 453–458 (2018). [CrossRef]  

70. D. Ge, L. Yang, G. Wu, and S. Yang, “Spray coating of superhydrophobic and angle-independently colored films,” Chem. Commun. 50(19), 2469–2472 (2014). [CrossRef]  

71. I. Kuznetsov A, E. Miroshnichenko A, L. Brongersma M, S. Kivshar Y, and B. Luk’yanchuk, “Optically resonant dielectric nanostructures,” Science 354(6314), 2472 (2016). [CrossRef]  

72. C F Bohren and D R. Huffman, “Absorption and scattering of light by small particle,” Wiley: USA2004.

73. A. Resnick, C. Persons, and G. Lindquist, “Polarized emissivity and Kirchhoff's law,” Appl. Opt. 38(8), 1384–1387 (1999). [CrossRef]  

74. D. Sikdar and A. Kornyshev A, “Theory of tailorable optical response of two-dimensional arrays of plasmonic nanoparticles at dielectric interfaces,” Sci. Rep. 6(1), 33712 (2016). [CrossRef]  

75. A. Kornyshev A, M. Urbakh, and J. Paget, “Reflection of light by metal nanoparticles at electrodes,” Phys. Chem. Chem. Phys. 14(6), 1850–1859 (2012). [CrossRef]  

76. H. Cao, D. Lan, and Y. Wang, “Crack and size-dependence of shear modulus in a drying particulate film,” Sci. China Phys. Mech. Astron. 55(6), 1093–1097 (2012). [CrossRef]  

77. R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima, A. Kitamura, M. Shimohigoshi, and T. Watanabe, “Light-induced amphiphilic surfaces,” Nature 388(6641), 431–432 (1997). [CrossRef]  

78. F. Ming, L. Deng, A. Zhao, Y. Wang, and D. He, “Fabrication and optical properties of Alq3 doped PMMA microsphere arrays templated by ZnO inverse opal structure,” Opt. Mater. 32(9), 1210–1215 (2010). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       supplemental figures

Data availability

No data were generated or analyzed in the presented research.

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 (8)

Fig. 1.
Fig. 1. Numerical study of self-assembled PS PhC. (a) Schematic of the 3D FCC PhC unit cell (r/a = 0.3535, r = 110 nm). (b) Photonic band structure of the PS PhC. Inset: diagram of the First Brillouin zone. (c) The visible spectrum of the PS PhC in air and alcohol medium. Here, R / T means reflection / transmission. The dark green line indicates the center wavelength of reflection band. Inset: schematic of the simulated PhC structure with a thickness of 41 × (61/2/3) × (2r) + (2r), i.e. total 42 layers of PS microspheres. The colored arrow indicates the normal incidence of light. (d) The infrared spectrum of the PS PhC in air. The same configuration of (c). Here, A means absorption.
Fig. 2.
Fig. 2. Numerical study of the self-assembled TiO2 MM. (a) A diagram of light scattering by a single dielectric spherical nanoparticle due to Mie resonance. The color indicates visible band. (b) The scattering efficiency of spherical TiO2 nanoparticle with radius of 25 nm. Inset: electric field distribution at the center cross section of nanoparticles at 500 nm wavelength. (c) The visible spectrum of TiO2 MM. Inset: schematic of the simulated TiO2 MM structure as an equivalent thin film with thickness 10 µm. (d) The infrared simulation spectrum of TiO2 MM. The same structure of (c). Here, R / T / A means reflection / transmission / absorption.
Fig. 3.
Fig. 3. Numerical study of the self-assembled TiO2 MM-PS PhC heterostructure on Si substrate. (a) The visible spectrum of TiO2 MM- PS (r = 110 nm) PhC heterostructure. Inset: schematic of the simulated TiO2 MM-PS PhC heterostructure. Here, the thickness of PhC, MM and Si components are 20 × (61/2/3) × (2r) + (2r), i.e. total 21 layers of PS microspheres, 2 µm and 2 µm, respectively. The colored arrow indicates the normal incidence of light. (b) The infrared simulation spectrum of TiO2 MM- PS (r = 110 nm) PhC heterostructure. (c) The visible spectrum of TiO2 MM- PS (r = 130 nm) PhC heterostructure. (d) The infrared simulation spectrum of TiO2 MM- PS (r = 130 nm) PhC heterostructure. (b-d) has same structure of (a). Here, R / T / A means reflection / transmission / absorption.
Fig. 4.
Fig. 4. Experimental study of self-assembled PS PhCs. (a) SEM image of PS PhC (r = 110 nm) surface. Inset: zoom-in picture. (b) SEM image of PS PhC (r = 110 nm) cross-section. The thicknesses of PS PhCs are show in Fig. S6. (c-d) The visible and infrared spectrum of Si substrate with and without PS PhC (r = 110 nm) in air. (e) SEM image of PS PhC (r = 130 nm) surface. Inset: zoom-in picture. (f) SEM image of PS PhC (r = 130 nm) cross-section. (g-h) The visible and infrared spectrum of Si substrate with and without PS PhC (r = 130 nm) in air.
Fig. 5.
Fig. 5. Experimental study of the self-assembled TiO2 MM. (a) SEM image of TiO2 MM surface. (b) SEM image of TiO2 MM cross-section. Inset: zoom-out of the film sample to mark the thickness. (c-d) The visible and infrared spectrum of Si substrate with and without TiO2 MM.
Fig. 6.
Fig. 6. Experimental studies of the self-assembled TiO2 MM-PS PhC heterostructures. (a) SEM image of TiO2 MM-PS (r = 110 nm) PhC heterostructure surface. (b) SEM image of TiO2 MM-PS (r = 110 nm) PhC heterostructure cross-section. (c-d) The visible and infrared spectrum of Si substrate with and without TiO2 MM-PS (r = 110 nm) PhC heterostructure. (e) SEM image of TiO2 MM-PS (r = 130 nm) PhC heterostructure surface. (f) SEM image of TiO2 MM-PS (r = 130 nm) PhC heterostructure cross-section. (g-h) The visible and infrared spectrum of Si substrate with and without TiO2 MM-PS (r = 130 nm) PhC heterostructure. The thickness of PS PhC is about 22.4 µm (for r = 110 nm) or 18.6 µm (r = 130 nm) and the thickness of TiO2 MM is about 14 µm (see Fig. S7).
Fig. 7.
Fig. 7. Visible imaging results. (a) The smiley face coding pattern containing TiO2 MM in air. (b) TiO2 MM-PS PhC heterostructure in air. (c) TiO2 MM-PS PhC heterostructure in alcohol. (d) TiO2 MM-PS PhC heterostructure after alcohol volatilizes. Infrared imaging results. (e) The smiley face coding pattern containing TiO2 MM. (f) TiO2 MM-PS PhC heterostructure in air. (g) TiO2 MM-PS PhC heterostructure in alcohol. (h) TiO2 MM-PS PhC heterostructure after alcohol volatilizes.
Fig. 8.
Fig. 8. Visible imaging results. (a) The “UESTC” coding pattern containing TiO2 MM in air. (b) TiO2 MM-PS PhC heterostructure in air. (c) TiO2 MM-PS PhC heterostructure in alcohol. (d) TiO2 MM-PS PhC heterostructure after alcohol volatilizes. Infrared imaging results. (e) The “UESTC” coding pattern containing TiO2 MM. (f) TiO2 MM-PS PhC heterostructure in air. (g) TiO2 MM-PS PhC heterostructure in alcohol. (h) TiO2 MM-PS PhC heterostructure after alcohol volatilizes.

Equations (3)

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

λ gap = 2 d 111 n e f f .
n e f f = f p s n p s + ( 1 f p s ) n m
Q s c a = C s c a G = ( 0 2 π 0 2 π | X | 2 k 2 s i n θ d θ d φ ) / G   =   4 π | X | 2 k 2 d Ω / G .
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.