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

Temperature assisted reflection control using VO2/Si core-shell nanoparticles

Open Access Open Access

Abstract

Vanadium dioxide (VO2) has attracted interest due to its phase transition from the insulating to the metallic states for potential use in a variety of optical and photonic applications. Here, we propose a VO2/Si core-shell structure to improve switching in band-selective reflection properties of the composite. Mie scattering formulation is used to analyze the structure before and after phase transition to show the impact of resonance quality on the mid-infrared light back-scattering. After investigating the effect of various VO2 nano-sphere radius sizes in both phases on light reflectivity, a VO2/Si core-shell structure is proposed to boost reflectivity and improve light controllability. Randomly distributed nanoparticles are studied to illustrate how these composites have similar behavior to their deterministic-distributed counterpart. Our results indicate that up to two-thirds of incident light power can be controlled by embedding proper core-shells in a polymer host material.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Thermochromic coatings have attracted significant interest from researchers due to their important applications in energy-efficient smart windows [18]. A stimuli-sensitive nanoparticle that resonates with light can control light flow in these coatings [913]. The external stimuli alter the optical refractive index of the particle, which changes the light-scattering properties. As a result, reflection and transmission from the coating can be manipulated by controlling the stimuli. Thermochromic materials can be classified into three types, including inorganic, organic, and liquid crystal types [14]. Inorganic thermochromic materials such as VO2, ZnO, TiO2, Cu2HgI4, etc., exhibit thermochromic behavior ranging from 70-500°C [15]. Among them, vanadium dioxide (VO2), with attractive optical and electrical properties around its transition temperature (around 68°C), is known as the most suitable inorganic thermochromic material in the near IR region [16]. Also, VO2 has been employed as a primary material for thermochromic composites in which temperature variation alternates the optical index in the material [17,18]. Due to these reasons, VO2 was chosen as the phase change material in this study. The temperature dependence of the optical properties of vanadium dioxide has led researchers to investigate this material's morphology [1921]. VO2-based thermochromic composites can be classified into VO2-based layered structures and VO2 nanoparticles [22]. The research focused on controlling and optimizing the optical properties of VO2 thin films and doping with various elements [2325], multilayer structures [26], and manufacturing using various techniques such as sputtering [27]. In addition, the transmittance and absorbance of these coatings has been investigated extensively in the literature [28,29].

VO2 core-shell nanoparticles offer more flexibility for tailoring the optical properties through the geometry and materials of the core or shell structures [3034]. Adding shells to resonant nanoparticles can allow nano-scale control of energy [35,36] and tuning of the localized surface plasmons [37]. The insulator of a resonant localized surface plasmon shell in solar cells enhances the photon absorption rate of the particle and governs its reflectivity to improve light harvesting [38]. Using this method, the reflectivity of the core nanoparticles can be improved.

This paper proposes a VO2/Si core-shell structure to boost the composite reflection response switching. Implementing the proposed structure further reduces the losses in the nanoparticle at room temperature (T =30°C), which shows that 67% of the incident power dynamically can be controlled by switching the composite temperature from 85°C to 30°C. In general, metals are reflective above their plasma wavelength [39]. The metallic VO2 plasma wavelength is around 900 nm [40]; therefore, this work studies VO2 plasmonic resonance above 900 nm. On the other hand, VO2 refractive index switching is not significant in the mid-far infrared regime [40]. Therefore, we focus on the near-infrared regime, where VO2 particles support plasmonic resonances in this wavelength range. VO2 is a lossy metal that absorbs incident light with a core-shell structure; however, high reflectivity is obtained by dielectric resonance. Optical dielectric resonators can be engineered to provide high reflectivity like their radio frequency counterparts [4143]. Switching VO2 core temperature helps switching the quality of this dielectric resonance. When the VO2 core is metal, the resonance quality of the dielectric shell drops. As a result, the reflectivity of the core-shell structure drops when VO2 is metal. In general, the metallic form of VO2 has higher reflectivity in comparison with the dielectric form. The results of the current work demonstrate that the opposite of this situation can be obtained under certain conditions using the dielectric particle resonances and the lossy nature of VO2. Another important contribution of this work is the demonstration of a random distribution of VO2/Si core-shell particles and a comparison of the spectral results of the random distribution with those of the deterministic counterpart. This random distribution can account for possible fabrication imperfections. After studying the structure in a lattice form, randomly distributed core-shells are studied to illustrate that a composite may have behavior similar to its lattice counterpart.

2. Reflectivity of VO2 nano-spheres

Reflection from the surface of a particle array is directly related to the ability of each particle to scatter light. On the other hand, a small particle with dimensions comparable to wavelength can have remarkable scattering properties [44]. A mathematical formulation has been developed for spherical particles [45,46]. This method analyzes isolated dielectric spheres by calculating the particle's electric and magnetic resonance modes and studying their interference's scattering characteristics in the far-field region. Hence, sub micro-meter semiconductors in the infrared regime can be modeled by their first-order mode electric and magnetic polarizabilities neglecting the small contribution from higher-order modes. The back-scattering cross-section of the low-loss dielectric nanosphere particle normalized to its physical cross-section is given as [45]:

$${Q_b} = 9\frac{{{{|{{a_1} - {b_1}} |}^2}}}{{k_0^2{r^2}}}$$

In this equation, a1 and b1 are Mie scattering coefficients corresponding to the transverse magnetic and transverse electric modes, k0 is the wavenumber in the ambient medium, and r is the particle's radius. Equation (1) is valid for an isolated spherical nanoparticle. Coatings involving phase change materials can provide an adaptive response to infrared radiation, which can play an important role in thermal management for problems involving significant radiative thermal load. This is particularly crucial for aerospace applications and extreme environments where the temperatures can be in the 1000 K – 2500 K range, for which the peak of black body radiation corresponds to (1.0 - 3.0 µm) spectral range. If absorbed, the incident radiation can create high levels of heating, which can create structural and stability problems in materials. Therefore, it is essential to achieve surfaces and surface coatings that can reduce the IR radiation load through adaptive reflective surfaces. In this study, we focus on adaptive coatings involving phase change materials VO2 to control the reflectivity of the surfaces in the (1.0 - 3.0 µm) spectral range. To achieve this, VO2 particles must have a considerable back-scattering coefficient at the lower temperature at this wavelength. The inherent VO2 losses at high temperatures will reduce the back-scattering, and therefore, the reflection will decrease. To obtain an optimum particle size for higher back-scattering, Eq. (1) is utilized here. The host material is a polymer with a dispersion-less index of 1.5, and the VO2 index is 3:2 + 0:17j for λ =2000nm. Extracting the scattering coefficients from [47] and applying them to Eq. (1) for various r gives an overview of the Qb dependence on particle radius. Figure 1(a). illustrates this dependency for a range of radius r. Hence Eq. (1) is not valid for metallic VO2 nanoparticles due to higher losses of VO2 at high temperatures starting from T =50°C.

 figure: Fig. 1.

Fig. 1. (a) Normalized back-scattering cross-section of insulator VO2 at λ =2000nm as a function of particle radius. (b) The reflectivity of VO2 arrays for the lattice sizes of a =1000, 2000, and 3000 nm in 1000 <λ< 2500 nm.

Download Full Size | PDF

Figure 1(a) shows that for a small radius, Qb is very small. When the particle size is much smaller than the wavelength, the back-scattering cross-section of the particle is a negligible fraction of its physical cross-section. However, for insulator VO2 this amount increases smoothly up to 64% of the physical cross-section around r =320 nm, then decreases until r =400 nm. Qb goes larger than 0.64 for larger particle sizes, but after the first local minimum at r =400 nm, the higher-order modes get excited, but their scattering is not significant compared to the main mode in r =320 nm. Therefore, an r =320 nm particle can give the maximum backward scattering.

When the nanoparticles are placed in arrays, in addition to the particle size, array lattice size plays an important role in the resonance wavelength. To find out the lattice size effect, reflection from a two-dimensional array is examined in Fig. 1(b) for the spectrum ranging from 1400 to 2500 nm. Array lattice a is selected so that the particle covers a considerable physical cross-section of the array, and the effect of different lattice sizes has been analyzed numerically using a commercial-grade simulator [48] based on the finite-difference time-domain (FDTD) method [49]. The dispersive optical index of VO2 is taken from [50], and the silicon optical index is extracted from the experiment data by Palik [51]. The results are depicted in Fig. 1 (b) for lattice sizes of a =1000, 2000, and 3000 nm. The results in Fig. 1(b) indicate that increasing the lattice size shifts the resonance wavelengths. The collective coupling of the structure has a strong dependency on the lattice size [52]. With a smaller lattice, the electric dipole looks dominant in reflectance; nonetheless, as the lattice size grows, the sharp magnetic resonance impact increases and might exceed the electric dipole as in the case of the 2000nm lattice (red). However, with a bigger lattice size, the incident wave can pass through the particles without interacting with them. As a result, the total reflectance drops significantly. Therefore, the small lattice sizes are more desirable to achieve higher reflectivity levels. Furthermore, shrinking the lattice size to smaller sizes is not viable because VO2 particles may overlap, and overlapped particles do not support Mie scattering.

In Fig. 2, normalized reflected power is given at both T =30°C and T =85°C for the lattice size a =1000 nm, in which the particle takes 32% of the array surface. This figure shows that by switching the temperature from T =30°C to T =85°C, the reflectivity of the array changes due to the insulator to metal phase change in VO2. The difference in reflectivity at two temperatures maximizes around 1900nm, which the insulator VO2 particle resonates by the incident wave. At this wavelength, the VO2 phase transition switches the reflectivity from 0.33 to 0.05. For longer wavelengths starting from 2140 nm, the reflectivity of metallic VO2 at T =85°C exceeds the reflectivity of insulator VO2 at T =30°C.

 figure: Fig. 2.

Fig. 2. Power reflection by a 2-D array of VO2 spheres with r =320 nm and a lattice period of a =1000 nm upon a normal light incident at mid-infrared wavelengths. Increasing ambient temperature from T =30°C to T =85°C changes the optical index of VO2; consequently, the reflectivity of the array switches.

Download Full Size | PDF

Fundamentally, the phase transition of VO2 changes both resonance wavelength and loss factor. At the metallic phase, the particle resonates at longer wavelengths with a relatively higher absorption rate. Therefore, the reflectivity of metallic VO2 particles at longer wavelengths is greater than the insulator state. Basically, in an insulator state, resonance is weaker. Maximum switching in reflectivity response is around 25% of the incident power. This amount is comparable with the reported thermal reflectivity controls in the literature [10]. Note that the resonance of reflection at T =30°C has slightly shifted from the designed wavelength. This shift occurs because of the resonance mode interaction in closely packed coupled particles. The maximum reflection wavelength can be set to the designed value by tuning (in this case, increasing) the particle's radius. As shown in Fig. 2, the maximum normalized reflected power difference at two temperatures occurs around 1900nm, which manipulates approximately 25% of the incident power in this structure. Lattice dimensions shifted the resonance wavelength slightly, and the particle dimensions can be adjusted to fine-tune the resonance wavelength. This amount of reflection control is promising; however, a higher switching capability is needed in applications where controlling power is desired.

3. Reflectivity of VO2/Si core-shells

Even though decreasing the lattice size can increase the reflected power at the low-loss resonances, controlling inter-particle distance is a challenging task [44]. Here an alternative method based on spherical core-shell structures is employed. In this technique, a lower-loss material is utilized as the shell of the VO2 spherical core to diminish the absorbed power in the structure at resonances. Due to the high reflectivity of a pure Si nanosphere and its closer optical index to the insulator form of VO2, Si is selected as the shell material. The combination of the VO2 core and Si shell led to stronger scattering peak resonances and, therefore, higher reflection. Keeping the outer radius of the structure constant and increasing the thickness of the shell, the effect of a low-loss shell can be studied on the ability of the structure to control the power reflectivity. Figure 3 illustrates the power reflectivity for different shell thicknesses at T =30°C and T =85°C. The optical properties of Si are taken from [51].

 figure: Fig. 3.

Fig. 3. Power reflection by a 2-D array of VO2/Si core-shells with r =320 nm and lattice period of a =1000 nm upon a normal light incident at mid-infrared wavelengths for T =30°C to T =85°C and different Si thicknesses.

Download Full Size | PDF

Figure 3 shows that the reflectivity of the core-shell is enhanced by increasing the silicon thickness at room temperature. Although the overall diameter of the core-shell remains constant, the peak of reflectivity response shifts to higher wavelength as the shell thickness grows. The index of silicon, which is slightly bigger than the index of VO2 at T =30°C, causes this shift. At T =85°C, by increasing the thickness of the Si shell up to 0.4r, the reflectivity of the core-shell array increases. By further increasing the shell thickness at T = 85°C, the structure resonates around 2000nm. This resonance increases because the metallic VO2 losses diminish as the volume of VO2 decreases in the structure. As a result, the reflectivity around the resonance increases. It can be understood from the figure that the difference between power reflectivity at T = 30°C and T = 85°C maximizes at Si = 0.4r. The maximum power reflection controllability core-shell reaches 70% of the incident wave, which is bigger than the VO2 spheres of the same size.

Generally, having an array with precise lattice is difficult, and implementing such a structure is expensive. Rather than that, composites with the same core-shell sizes but with random distribution over space can be more affordable. Here core-shells are randomly relocated in a three-layer array of VO2/Si to study the effect of irregular composites. To form the random arrays, Gaussian random points with a median of zero and the standard deviation of 100 nm, which is almost 30% of the nanoparticle radius, are built-in 3D space using MATLAB. The numerical analysis here is limited to a 3*3 array with random relocations. The reflectivity of this composite in T =30°C and T =85°C is shown in Fig. 4 (a). This figure shows the ability of randomly distributed VO2/Si core shells to manipulate the reflected power by shifting the temperature. In addition, Fig. 4 (b) compares the normalized reflected power for some deviation from the curve of Fig. 3 for the same Si thickness for the specific case with the randomly distributed arrays. The results indicate that the overall trend of both curves matches except for small differences around 1600 nm for T =30°C. It can be concluded that this resonance is not an inherent resonance of the core-shell, but it comes from the regular lattice of the structure. By disordering the lattice and switching it to the random array, this resonance disappears too. If the standard deviation of the nanoparticle distribution further increases, the reflection curve can have more fluctuation than in Fig. 4. We anticipate the spectral response will get smoother as the area of the structure expands. We can expect the randomly distributed core-shells of VO2/Si to switch the reflection up to 67% by temperature switching in the structure.

 figure: Fig. 4.

Fig. 4. (a) Power reflectivity by a 3-D array of randomly distributed VO2/Si core-shells at T =30°C and T =85°C. The nanoparticle radius is 320 nm, and Si has 40% of the overall thickness. (b) The comparison of the periodic lattice (a =1000 nm, r =320 nm, tSi =0.4r) with the randomly distributed array of VO2/Si at T =30°C and T =85°C.

Download Full Size | PDF

4. Conclusion

In summary, a core-shell structure of VO2/Si was proposed to manage a more considerable portion of the incident light. To achieve this, first, the metal to insulator phase change of vanadium dioxide was employed to control the back-scattering from nano-spheres by adjusting their temperature. The resonance of the incident light with VO2 nano-spheres causes bigger back-scattering at room temperature, while VO2 has an insulator form. However, when the temperature rises to T =85°C, VO2 reversibly transforms to a lossy metal, resonance quality degrades, and its frequency shifts to larger wavelengths. This switching in back-scattering is used to control reflection from an array of VO2 nano-spheres. It is shown that a quarter of incident light energy can be switched around in the main resonance by reflection control via temperature switching. Then, the optical properties of the proposed VO2/Si core-shells structure were studied numerically for a deterministic and randomly distributed set of particles embedded in a polymer host. The core-shells could switch up to two-thirds of incident light power in reflection form by controlling the resonance via temperature switching. That proves the capability of the proposed structure in implementing switchable mirrors in smart reflector applications.

Funding

Türkiye Bilimsel ve Teknolojik Araştirma Kurumu (115M033).

Disclosures

The authors declare no conflicts of interest.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. Z. Yuxuan, Z. Yunyun, Y. Jianrong, and Z. Xiaoqiang, “Energy-saving performance of thermochromic coatings with different colors for buildings,” Energy Build. 215, 109920 (2020). [CrossRef]  

2. Y. Zhang and X. Zhai, “Preparation and testing of thermochromic coatings for buildings,” Sol. Energy 191, 540–548 (2019). [CrossRef]  

3. N. Numan, B. Mabakachaba, A. Simo, Z. Nuru, and M. Maaza, “VO2-based active tunable emittance thermochromic flexible coatings,” J. Opt. Soc. Am. B 37(11), C45 (2020). [CrossRef]  

4. J. Zhang, J. Wang, C. Yang, H. Jia, X. Cui, S. Zhao, and Y. Xu, “Mesoporous SiO2/VO2 double-layer thermochromic coating with improved visible transmittance for smart window,” Sol. Energy Mater. Sol. Cells 162, 134–141 (2017). [CrossRef]  

5. J. Huang, Q. Li, Z. Zheng, and Y. Xuan, “Thermal rectification based on thermochromic materials,” Int. J. Heat Mass Transfer 67, 575–580 (2013). [CrossRef]  

6. M. Kamalisarvestani, R. Saidur, S. Mekhilef, and F. Javadi, “Performance, materials and coating technologies of thermochromic thin films on smart windows,” Sol. Energy Mater. Sol. Cells 26, 353–364 (2013). [CrossRef]  

7. S. Babulanam, T. Eriksson, G. Niklasson, and C. Granqvist, “Thermochromic VO2 films for energy-efficient windows,” Sol. Energy Mater. 16(5), 347–363 (1987). [CrossRef]  

8. S. Allinikov, “Photochromic-thermochromic coating composition,” U.S. Patent No. 3,744,295 (1973).

9. Y. Li, S. Ji, Y. Gao, H. Luo, and M. Kanehira, “Core-shell VO2@TiO2 nanorods that combine thermochromic and photocatalytic properties for application as energy-saving smart coatings,” Sci. Rep. 3(1), 1–13 (2013). [CrossRef]  

10. Q. Li, D. Fan, and Y. Xuan, “Thermal radiative properties of plasma-sprayed thermochromic coating,” J. Alloys Compd. 583, 516–522 (2014). [CrossRef]  

11. H. Zhou, X. Cao, M. Jiang, S. Bao, and P. Jin, “Surface plasmon resonance tunability in VO2/Au/VO2 thermochromic structure,” Laser Photonics Rev. 8(4), 617–625 (2014). [CrossRef]  

12. V. Nagy, I. Suleimanov, G. Molnár, L. Salmon, A. Bousseksou, and L. Csóka, “Cellulose–spin crossover particle composite papers with reverse printing performance: a proof of concept,” J. Mater. Chem. C 3(30), 7897–7905 (2015). [CrossRef]  

13. C. Lv, L. Hu, Y. Yang, H. Li, C. Huang, and X. Liu, “Waterborne UV-curable polyurethane acrylate/silica nanocomposites for thermochromic coatings,” RSC Adv. 5(33), 25730–25737 (2015). [CrossRef]  

14. Y. Cheng, X. Zhang, C. Fang, J. Chen, and Z. Wang, “Discoloration mechanism, structures and recent applications of thermochromic materials via different methods: A review,” J. Mater. Sci. Technol. 34(12), 2225–2234 (2018). [CrossRef]  

15. A. Hakami, S. S. Srinivasan, P. K. Biswas, A. Krishnegowda, S. L. Wallen, and E. K. Stefanakos, “Review on thermochromic materials: development, characterization, and applications,” JCT Res. 19(2), 377–402 (2022). [CrossRef]  

16. C. Y. Kim, T. Slusar, J. Cho, and H. T. Kim, “Mott switching and structural transition in the metal phase of VO2 nanodomain,” ACS Appl. Electron. Mater. 3(2), 605–610 (2021). [CrossRef]  

17. J. Y. Suh, R. Lopez, L. C. Feldman, and R. Haglund Jr, “Semiconductor to metal phase transition in the nucleation and growth of VO2 nanoparticles and thin films,” J. Appl. Phys. 96(2), 1209–1213 (2004). [CrossRef]  

18. K. C. Kam and A. K. Cheetham, “Thermochromic VO2 nanorods and other vanadium oxides nanostructures,” Mater. Res. Bull. 41(5), 1015–1021 (2006). [CrossRef]  

19. O. Karahan, A. Tufani, S. Unal, I. B. Misirlioglu, Y. Z. Menceloglu, and K. Sendur, “Synthesis and morphological control of VO2 nanostructures via a one-step hydrothermal method,” Nanomaterials 11(3), 752 (2021). [CrossRef]  

20. S. Long, X. Cao, Y. Wang, T. Chang, N. Li, L. Jin, L. Ma, F. Xu, G. Sun, and P. Jin, “Karst landform-like VO2 single layer solution: Controllable morphology and excellent optical performance for smart glazing applications,” Sol. Energy Mater. Sol. Cells 209, 110449 (2020). [CrossRef]  

21. M. Ertas Uslu, R. A. Yalcin, I. B. Misirlioglu, and K. Sendur, “Morphology induced spectral reflectance lineshapes in VO2 thin films,” J. Appl. Phys. 125(22), 223103 (2019). [CrossRef]  

22. F. Xu, X. Cao, H. Luo, and P. Jin, “Recent advances in VO2-based thermochromic composites for smart windows,” J. Mater. Chem. C 6(8), 1903–1919 (2018). [CrossRef]  

23. M. E. Uslu, I. B. Misirlioglu, and K. Sendur, “Crossover of spectral reflectance lineshapes in Ge-doped VO2 thin films,” Opt. Mater. 104, 109890 (2020). [CrossRef]  

24. S.-Y. Li, N. R. Mlyuka, D. Primetzhofer, A. Hallén, G. Possnert, G. A. Niklasson, and C. G. Granqvist, “Bandgap widening in thermochromic Mg-doped VO2 thin films: Quantitative data based on optical absorption,” Appl. Phys. Lett. 103(16), 161907 (2013). [CrossRef]  

25. F. Beteille and J. Livage, “Optical switching in VO2 thin films,” J. Sol-Gel Sci. Technol. 13(1/3), 915–921 (1998). [CrossRef]  

26. G. Sun, H. Zhou, X. Cao, R. Li, M. Tazawa, M. Okada, and P. Jin, “Self-assembled multilayer structure and enhanced thermochromic performance of spinodally decomposed TiO2–VO2 thin film,” ACS Appl. Mater. Interfaces 8(11), 7054–7059 (2016). [CrossRef]  

27. Y. Luo, L. Zhu, Y. Zhang, S. Pan, S. Xu, M. Liu, and G. Li, “Optimization of microstructure and optical properties of VO2 thin film prepared by reactive sputtering,” J. Appl. Phys. 113(18), 183520 (2013). [CrossRef]  

28. S. Chen, H. Ma, X. Yi, T. Xiong, H. Wang, and C. Ke, “Smart VO2 thin film for protection of sensitive infrared detectors from strong laser radiation,” Sens. Actuators, A 115(1), 28–31 (2004). [CrossRef]  

29. J. Zhou, Y. Gao, Z. Zhang, H. Luo, C. Cao, Z. Chen, L. Dai, and X. Liu, “VO2 thermochromic smart window for energy savings and generation,” Sci. Rep. 3(1), 1–5 (2013). [CrossRef]  

30. X. Wu, L. Yuan, X. Weng, L. Qi, B. Wei, and W. He, “Passive Smart Thermal Control Coatings Incorporating CaF2/VO2 Core-Shell Microsphere Structures,” Nano Lett. 21(9), 3908–3914 (2021). [CrossRef]  

31. Z. Wen, Y. Ke, C. Feng, S. Fang, M. Sun, X. Liu, and Y. Long, “Mg-doped VO2@ ZrO2 Core-Shell Nanoflakes for Thermochromic Smart Windows with Enhanced Performance,” Adv. Mater. Interfaces 8(1), 2001606 (2021). [CrossRef]  

32. Y. Chen, X. Zeng, J. Zhu, R. Li, H. Yao, X. Cao, S. Ji, and P. Jin, “High performance and enhanced durability of thermochromic films using VO2@ ZnO core-shell nanoparticles,” ACS Appl. Mater. Interfaces 9(33), 27784–27791 (2017). [CrossRef]  

33. Y. Gao, S. Wang, H. Luo, L. Dai, C. Cao, Y. Liu, Z. Chen, and M. Kanehira, “Enhanced chemical stability of VO2 nanoparticles by the formation of SiO2/VO2 core/shell structures and the application to transparent and flexible VO2-based composite foil with excellent thermochromic properties for solar heat control,” Energy Environ. Sci. 5(3), 6104–6110 (2012). [CrossRef]  

34. S.-Y. Li, G. A. Niklasson, and C.-G. Granqvist, “Nanothermochromics with VO2-based core-shell structures: Calculated luminous and solar optical properties,” J. Appl. Phys. 109(11), 113515 (2011). [CrossRef]  

35. X. Huang, S. Tang, B. Liu, B. Ren, and N. Zheng, “Enhancing the photothermal stability of plasmonic metal Nanoplates by a core-shell architecture,” Adv. Mater. 23(30), 3420–3425 (2011). [CrossRef]  

36. C. S. Levin, C. Hofmann, T. A. Ali, A. T. Kelly, E. Morosan, P. Nordlander, K. H. Whitmire, and N. J. Halas, “Magnetic− plasmonic core-shell nanoparticles,” ACS Nano 3(6), 1379–1388 (2009). [CrossRef]  

37. M. D. Brown, T. Suteewong, R. S. S. Kumar, V. D’Innocenzo, A. Petrozza, M. M. Lee, U. Wiesner, and H. J. Snaith, “Plasmonic dye-sensitized solar cells using core-shell metal-insulator nanoparticles,” Nano Lett. 11(2), 438–445 (2011). [CrossRef]  

38. C. P. Byers, H. Zhang, D. F. Swearer, M. Yorulmaz, B. S. Hoener, D. Huang, A. Hoggard, W.-S. Chang, P. Mulvaney, and E. Ringe, “From tunable core-shell nanoparticles to plasmonic drawbridges: Active control of nanoparticle optical properties,” Sci. Adv. 1(11), e1500988 (2015). [CrossRef]  

39. J.I. Rojas and D. Crespo, “The physics and chemistry of materials,” Wiley, New York757 (2001).

40. D. Fu, K. Liu, T. Tao, K. Lo, C. Cheng, B. Liu, R. Zhang, H.A. Bechtel, and J. Wu, “Comprehensive study of the metal-insulator transition in pulsed laser deposited epitaxial VO2 thin films,” J. Appl. Phys. 113(4), 043707 (2013). [CrossRef]  

41. E. Rahimi and K. Şendur, “Femtosecond pulse shaping by ultrathin plasmonic metasurfaces,” J. Opt. Soc. Am. B 33(2), A1–7 (2016). [CrossRef]  

42. L. Zou, W. Withayachumnankul, C.M. Shah, A. Mitchell, M. Bhaskaran, S. Sriram, and C. Fumeaux, “Dielectric resonator nanoantennas at visible frequencies,” Opt. Express 21(1), 1344–1352 (2013). [CrossRef]  

43. A.I. Kuznetsov, A.E. Miroshnichenko, M.L. Brongersma, Y.S. Kivshar, and B. Luk’yanchuk, “Optically resonant dielectric nanostructures,” Science 354(6314), aag2472 (2016). [CrossRef]  

44. K. L. Kelly, E. Coronado, L. L. Zhao, and G. C. Schatz, “The optical properties of metal nanoparticles: the influence of size, shape, and dielectric environment,” J. Phys. Chem. B 107(3), 668 (2003). [CrossRef]  

45. R. A. Shore, “Scattering of an electromagnetic linearly polarized plane wave by a multilayered sphere: obtaining a computational form of Mie coefficients for the scattered field,” IEEE Antennas Propag. Mag. 57(6), 69–116 (2015). [CrossRef]  

46. H. Du, “Mie-scattering calculation,” Appl. Opt. 43(9), 1951 (2004). [CrossRef]  

47. C. Mätzler, “MATLAB functions for Mie scattering and absorption, version 2,” (2002).

48. FDTD Solutions, Lumerical Solutions, inc. (Vancouver, 2003).

49. A. Taflove, S. C. Hagness, and M. Piket-May, “Computational electromagnetics: the finite-difference time-domain method,” The Electrical Engineering Handbook 3 (CRC Press, 2005).

50. C. Lamsal and N. Ravindra, “Optical properties of vanadium oxides-an analysis,” J. Mater. Sci. 48(18), 6341–6351 (2013). [CrossRef]  

51. E. D. Palik, Handbook of Optical Constants of Solids (Academic Press, 1998).

52. T. Liu, R. Xu, P. Yu, Z. Wang, and J. Takahara, “Multipole and multimode engineering in Mie resonance-based metastructures,” Nanophotonics 9(5), 1115–1137 (2020). [CrossRef]  

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

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

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. (a) Normalized back-scattering cross-section of insulator VO2 at λ =2000nm as a function of particle radius. (b) The reflectivity of VO2 arrays for the lattice sizes of a =1000, 2000, and 3000 nm in 1000 <λ< 2500 nm.
Fig. 2.
Fig. 2. Power reflection by a 2-D array of VO2 spheres with r =320 nm and a lattice period of a =1000 nm upon a normal light incident at mid-infrared wavelengths. Increasing ambient temperature from T =30°C to T =85°C changes the optical index of VO2; consequently, the reflectivity of the array switches.
Fig. 3.
Fig. 3. Power reflection by a 2-D array of VO2/Si core-shells with r =320 nm and lattice period of a =1000 nm upon a normal light incident at mid-infrared wavelengths for T =30°C to T =85°C and different Si thicknesses.
Fig. 4.
Fig. 4. (a) Power reflectivity by a 3-D array of randomly distributed VO2/Si core-shells at T =30°C and T =85°C. The nanoparticle radius is 320 nm, and Si has 40% of the overall thickness. (b) The comparison of the periodic lattice (a =1000 nm, r =320 nm, tSi =0.4r) with the randomly distributed array of VO2/Si at T =30°C and T =85°C.

Equations (1)

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

Q b = 9 | a 1 b 1 | 2 k 0 2 r 2
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.