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

Selective thermal emitters with infrared plasmonic indium tin oxide working in the atmosphere

Open Access Open Access

Abstract

In this paper, we prove that indium tin oxide (ITO) can be a material of choice for wavelength-selective thermal emitters that enable operation in ambient air. The wavelength-selective thermal emission is achieved by plasmonic ITO perfect absorbers (IPAs) composed of ITO – Al2O3 – ITO tri-layers where the top ITO layer is a hexagonal array of ITO disks. Simulated optical spectra of the IPAs reveal nearly perfect absorptivities (i.e. 0.99) at obvious resonances, which can be easily tuned just by changing the size of the ITO disks. The fabricated IPAs that followed the pre-designed simulation also exhibit excellented resonant absorptivities (i.e. 0.92), as well as good resonance tunability, which indicate the appreciable performance of IPAs. Furthermore, we show that IPAs are also feasible to serve as highly-efficient thermal emitters that can operate in ambient air without the need for strict vacuum or inert gas conditions, providing another platform for IR plasmonic devices.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Engineering the thermal emission using wavelength-selective perfect absorbers [1] covers a wide range of applications in infrared (IR) photonics and thermal managements ranging from thermophotovoltaics [25] to non-dispersive IR (NDIR) gas sensors [6,7] and radiative coolers [811]. For the practical applications, plasmonic materials such as tungsten [12,13], molybdenum [14] or metal nitrides and metal carbides [1519] have been proposed; however, they are not robust against oxidation if they are used at elevated temperature in air. Hence, the practical applications of thermal metallic emitters are limited to the strict vacuum or inert gas ambient conditions. Thus, serious demand exists for IR plasmonic materials that are cost-effective and robust at high temperature. In this regard, transparent conductive oxides (TCOs) are already in oxidized state and thus inherently robust in ambient air high-temperature operations. They exhibit screening plasma frequencies in the near-IR region owing to their lower carrier concentrations (1020–1022 cm-3) and thus leading to low-loss nature compared to the elemental metals (1023 cm-3), which render them as promising candidate of IR plasmonic materials.

In the past decades, commonly utilized TCOs including indium tin oxide (ITO) [2024], fluorine doped tin oxide (FTO) [25,26] or doped zinc oxide (AZO and GZO) [2729] have been found to be attractive in IR plasmonics [30]. Among them, ITO has shown significant advantages compared to other TCOs owing to its robustness in moist air, surface smoothness, mechanical stability and simple fabrication process [31]. Thus, ITO has been intensively applied in many studies subjected to localized plasmonic nanoparticles [3234], optical metasurfaces [35,36], antenna arrays for surface-enhanced IR absorption spectroscopy [3739], active tunable plasmonic devices [4043]. Although ITO is a favourable IR plasmonic material for wavelength-selective perfect absorbers and thermal emitters with the chemical and electrical stability at elevated temperature [44], it has not been used for these applications so far.

In this paper, to our best knowledge, wavelength-selective thermal emitters made of IR plasmonic ITO operated in ambient air were demonstrated for the first time. The proposed ITO perfect absorbers (IPAs) were composed of tri-layered ITO – Al2O3 – ITO wherein the top ITO disk resonators array was separated to the bottom ITO film by the Al2O3 insulator. The IPA structures were numerically simulated to optimize the geometrical parameters and then realized using a scalable colloidal lithography process. The fabricated IPAs exhibited nearly perfect resonant absorptions (i.e. 0.92) with polarization independence and facile tunability. We demonstrated that the IPA thermal emitters can efficiently operate in ambient air with desirable wavelength selectivity, high emissivity and high stability against the temperature changes. The proposed IPAs presented in this work can be used for practical IR devices such as surface-enhanced IR absorption spectroscopy, NDIR sensors, as well as visible-transparent – IR wavelength-selective heaters for drying furnace or radiative cooling for smart window. Although this work demonstrates ITO thermal emitters in the IR region, the resonance can extend to the THz region, and the concept is also applicable to other TCOs.

The remainder of the paper is organized as follows. The second section describes the structural design and simulated optical properties of the proposed IPAs. The third section provides a detailed fabrication procedure to realize scalable IPAs. The thermal emission and emissivity spectra of IPAs are demonstrated in the fourth section. Finally, the paper is summarized in the fifth section.

2. Proposed IPA structure and optical properties

Prior to performing numerical simulation of the proposed IPAs, spectroscopic ellipsometry was conducted for the ITO film fabricated by sputtering to obtain the actual optical property of ITO for the accurate numerical simulations and further fabrications. Herein, a 200 nm ITO film was deposited on a Si (100) substrate (p-type, 10 Ω·cm) by RF sputtering (i-Miller CFS-4EP-LL, Shibaura) using an ITO target (In2O3/SnO2 90/10 wt %). The sputtered ITO film was then annealed at 400 °C in vacuum condition for 1 hour to improve the conductivity and optical property of ITO. The film was optically characterized by two spectroscopic ellipsometers (SENTECH, SE 850 DUV for UV – NIR region and SENDIRA for mid-IR region). The complex permittivity ($\varepsilon = {\varepsilon _1} + i{\varepsilon _2}$) over a wide wavelength range, from 0.19 μm to 25 μm, derived from the ellipsometry measurements of the ITO film is shown in Fig. 1(a) (red curves). As seen, the real part of the sputtered ITO’s permittivity approaches negative values beyond 1.8 μm in wavelength, which evidences the plasmonic behavior of ITO in the IR region. For a comparison, the complex permittivities of two typical metals that are commonly used in IR perfect absorbers and thermal emitters; a noble metal – Au (orange curves) and a refractory metal – W (blue curves) taken from literature [45] are also plotted in Fig. 1(a). We also plot the ratio (${{ - {\varepsilon _1}} \mathord{\left/ {\vphantom {{ - {\varepsilon_1}} {{\varepsilon_2}}}} \right.} {{\varepsilon _2}}}$) between the real part and imaginary part of permitivities of the three plasmonic metals (Fig. (1b)), which reflect their figure-of-merit (FoM) values to determine the field enhancement of localized surface plasmon resonance (LSPR). As seen in Fig. 1(b), although both real part and imaginary part of the sputtered ITO film are much smaller than those of Au and W, nevertheless, the FoM value of ITO is almost comparable to that of Au and W in the mid-IR region; it is slightly smaller than the FoM of Au but higher than the FoM of W. This indicates that ITO can be a good material for IR plasmonic devices utilizing LSPR, especially for IR plasmonic absorbers.

 figure: Fig. 1.

Fig. 1. (a) A comparison between measured permittivity of the sputtered ITO film (red curves) with permitivities of Au (orange curves) and W (blue curves). Negative real part of ITO’s permittivity clearly reveals plasmonic behavior of ITO in the IR region. The complex permitivities of Au and W taken from the literature [45] were divided by a factor of 10 for a better comparison. (b) Figure-of-merit ($- {{{\varepsilon _1}} \mathord{\left/ {\vphantom {{{\varepsilon_1}} {{\varepsilon_2}}}} \right.} {{\varepsilon _2}}}$, solid curves) and skin depths (dashed curves) of ITO (red), Au (orange) and W (blue).

Download Full Size | PDF

Here, we applied IR plasmonic ITO for IR perfect absorbers by adopting a metal – insulator – metal configuration wherein the top layer made of a hexagonal array of ITO disks is isolated from the ITO bottom mirror via an Al2O3 film (Fig. 2(a)). Employing the rigorous coupled-wave analysis (RCWA) simulation (DiffractMOD, Synopsys’ RSoft) with the measured ITO permittivity, whereas the permitivities of Al2O3 and Si were taken from the literature [46], we demonstrated that with a proper design and optimal parameters, perfect absorptivities could be achieved. Figure 2(b) plots simulated spectra of an IPA array with a periodicity – p of 3 μm, diameter – d of 2.01 μm and insulator thickness – t of 0.4 μm. It is worth noting that because of a larger skin depth – $\delta$ ($\delta = {\lambda \mathord{\left/ {\vphantom {\lambda {({2\pi \kappa } )}}} \right.} {({2\pi \kappa } )}}$, with $\kappa$ is the extinction coefficient) compared to elemental metals (dashed curves in Fig. 1(b)), the thickness of ITO resonators and ITO film designed for absorbers are certainly thicker than those of elemental metals. In this work, the thicknesses of ITO disks and ITO films are optimized and kept unchanged at 0.15 μm and 0.4 μm, respectively. As seen in Fig. 2(b), IPA exhibits two main resonant bands at normal incidence; the first one locating at the shorter wavelength region (∼ 4.0 μm), namely M1, behaves like a dual-band resonance with high absorptivities (i.e. 0.93) whereas the second one at the longer wavelength region (8.4 μm), namely M2, reveals a nearly perfect absorptivity (i.e. 0.99). Interestingly, M2 can be easily tuned by changing the diameter – d of ITO disk [1,47,48]. As shown in Fig. 2(c), M2 readily tunes from 5 μm to 10 μm when the ITO diameter changes from 1.0 μm to 2.7 μm. In addition, the resonant peak of M2 is directly proportional to the disk diameter, indicating that M2 behaves like the dipole antenna resonance. In contrast, M1 acts as a hybrid resonance that splits into two peaks when the diameter of ITO disk increases.

 figure: Fig. 2.

Fig. 2. (a) Scheme of IPA configured by tri-layered ITO – Al2O3 – ITO with a hexagonal lattice ITO disk resonator array arranged on the top of Al2O3-ITO layered films. Dashed rectangle indicates the unit cell defined in the simulation. (b) Simulated optical spectra including absorptivity – A (red curve), transmittance – T (blue curve) and reflectance – R (black curve) of IPA with a periodicity – p of 3 μm, diameter – d of 2.01 μm and insulator thickness – t of 0.4 μm. (c) Resonance tunability of IPA by changing diameter – d of ITO disk resonator while keeping the other parameters unchanged (p = 3 μm, t = 0.4 μm). The thicknesses of ITO disk and ITO film are kept unchanged at 0.15 μm and 0.4 μm, respectively. In the simulation, the incident electromagnetic field propagated along the –z-axis, the electric field oscillated along the x-axis.

Download Full Size | PDF

To further verify the optical properties of IPA, full wave simulations based on the finite-difference time-domain (FDTD) method (FullWAVE, Synopsys’ RSoft) were also performed to calculate distributions of electric fields (Ex and Ez) and magnetic field (Hy) in IPA excited at both M1 (Fig. 3(a)) and M2 (Fig. 3(b)). As seen, the induced electric fields Ez clearly indicate that the excited electric dipoles at the top and bottom metals oscillate in the inverse phases, resulting in significantly enhanced magnetic fields in the insulating layer between the top and bottom electric dipoles. The resonances are the so-called magnetic resonances; the first-order (fundamental) magnetic resonance at M2 and the third-order magnetic resonance at M1. In addition, we observe that the magnetic dipoles in each unit cell at M1 couple to the others along the incident plane, which suggests that M1 arises from the coupling between the third-order magnetic resonance with the photonic property of the hexagonal lattice. For a plasmonic hexagonal lattice, the angular-dependent dispersion relation of the surface plasmon polariton (SPP) resonance excited by an electromagnetic wave with a wavelength of $\lambda$ under an incidence of $\theta$ (see Fig. 2(a)) is given by

$$\frac{{{\varepsilon _\text{m}}{\varepsilon _\text{d}}}}{{{\varepsilon _\text{m}}{} + {\varepsilon _\text{d}}}} = {\sin ^2}\theta + \frac{2}{p}({i + j} )\lambda \sin \theta + \frac{4}{{3{p^2}}}({{i^2} + ij + {j^2}} ){\lambda ^2}$$
where ${\varepsilon _\text{m}}$ and ${\varepsilon _\text{d}}$ are complex permitivities of metal and dielectric, respectively; p is the lattice periodicity; i and j are integers. To further elucidate the coupling behavior of M1, we simulated the angle-resolved absorptivity (0° – 85°), and the result is plotted in Fig. 3(c). Herein, while M2 resonated as the first-order magnetic dipole is slightly blue-shifted when the incident angle increases, M1 reveals a hybridization of SPPs (dashed curves in Fig. 3(c)) in a plasmonic hexagonal lattice, mainly at mode (-1,0) of the SPP at the ITO-air interface, coupled to the third order of the resonant magnetic dipole. In addition, when the incident angle increases above 10°, the higher order magnetic resonance also couples to the mode (-1,0) of the SPP at the ITO-Al2O3 interface. In the other hand, it was found that M1 is also associated with the cavity mode, which relies on the insulator thickness – t (Fig. 3(d)). By tuning the Al2O3 thickness, the resonant absorptivities at both M1 and M2 can be maximized.

 figure: Fig. 3.

Fig. 3. Simulated electric fields (Ex and Ez) and magnetic field (Hy) distributions of the IPA (d = 2.01 μm, p = 3 μm, t = 0.4 μm) excited at two resonance modes: (a) at M1 (wavelength 4.0 μm) and (b) at M2 (8.4 μm). (c) Simulated angle-dependent absorptivity of the IPA with geometrical parameters of d = 2.01 μm, p = 3 μm, t = 0.4 μm. The dashed curves indicate SPPs at metal-dielectric interfaces of ITO disk hexagonal lattice (see Eq. (1)). (d) Simulated insulator thickness dependence of IPA’s absorptivity while keeping d and p unchanged at 2.01 μm and at 3 μm, respectively. In the simulation, the incident electromagnetic field propagated along the –z-axis, the electric field oscillated along the x-axis, and the incident amplitudes were normalized to 1.

Download Full Size | PDF

3. Fabrication of IPAs

The proposed IPAs were fabricated following the geometrical parameters optimized by the pre-designed simulations. Here we used a colloidal lithography combined with the reactive-ion etching (RIE) process developed in our previous works to fabricate large-area IPAs with the size of 5×28 mm2 [48,49]. Figure 4(a) describes the fabrication process. Herein, tri-layered ITO (0.15 μm) – Al2O3 (0.4 μm) – ITO (0.4 μm) films followed the optimal parameters taken from the pre-designed simulations were deposited on 5×28 mm2 Si substrates by utilizing three RF sputtering steps. To fabricate the top ITO disk array, a monolayer of polystyrene (PS) spheres with diameter of 3.0 μm (Polybead Polystyrene Microspheres – Polysciences) was deposited on top of each tri-layered film as RIE mask. To reduce the size of the PS spheres, oxygen plasma etching (ULVAC CE-300I) with an etching rate of 2 nm/second was applied. The etching time on each sample was precisely controlled to achieve different PS sizes corresponding to the pre-designed ITO disks. Then, another RIE step was also employed to etch ITO with PS mask using the following recipe: mixture gases of BCl­3 (10 sccm) and Cl2 (10 sccm); APC pressure of 0.3 Pa; RF antenna power of 200 W; RF bias power of 10 W; etching rate of 0.23 nm/second. IPAs were finally achieved by removing the PS mask in toluene and ethanol.

 figure: Fig. 4.

Fig. 4. (a) Schematic diagrams illustrate scalable fabrication process of IPAs using colloidal lithography combined two steps of RIE. (b) SEM image (top) and photo (bottom) of the fabricated IPA. Arrows indicate typical defects of periodic ITO disk array. (c) SEM images of two fabricated IPAs having diameters of 1.62 μm (S1) and 2.01 μm (S2) with the same other parameters (p = 3 μm, t = 0.4 μm). (d) Simulated absorptivities and (e) measured absorptivities of S1 (blue curves) and S2 (red curves).

Download Full Size | PDF

Figure 4(b) represents a typical top-view scanning electron microscope (SEM) image (top) and a photo (bottom) of the fabricated IPAs. Despite some defects (white arrows in Fig. 4(b)) in the fabricated ITO disk lattice due to the non-uniformity and misalignments of PS array used as mask for RIE, the ITO disk hexagonal array was well-constructed as the pre-designed structure with grain sizes in the range of few tens to hundred microns scale. This result proves that the concise fabrication technique is applicable for realizing large-area IPAs. In this work, we fabricated two IPAs with the pre-designed geometrical parameters of p = 3 μm, t = 0.4 μm and different diameters obtained by simulations, namely S1 with d = 1.62 μm, and S2 with d = 2.01 μm. SEM images of the two fabricated absorbers are shown in Fig. 4(c). The transmittivity (T) and the reflectivity (R) spectra of the two fabricated IPAs were measured using a Fourier transform infrared (FTIR) spectrometer (Thermo Nicolet NEXUS 670, MCT detector, KBr beam splitter), then the absorptivity (A) spectra was achieved using the below relation

$$A = 1 - T - R$$
The simulated and measured absorptivity spectra of S1 (blue curves) and S2 (red curves) at normal incidence are plotted in Fig. 4(d) and Fig. 4(e), respectively. As shown, both S1 and S2 exhibit two main absorption bands; the SPP coupled third-order magnetic resonances at the shorter wavelength region are 3.5 μm (S1) and 4.0 μm (S2) and the fundamental magnetic resonances at the longer wavelength region are 7.5 μm (S1) and 8.4 μm (S2). The mode splitting at M1 resulted from the coupling between the third-order magnetic resonance and SPP (-1,0) mode of metallic hexagonal lattice is also clearly observed in both simulation and experiment, especially when the ITO disk diameter increases from 1.62 μm (S1) to 2.01 μm (S2), which is consistent with the above discussion (Fig. 2(c)). In particular, the measured absorptivity spectra agree well with the simulated absorptivity spectra, especially the resonant peaks and shapes, indicating the appropriate fabrication process used in this work. Furthermore, because of the high-rotational symmetry, the ITO periodic array absorber shows no dependence on the polarization as proven by both simulation (Fig. 5(a)) and measurement (Fig. 5(b)). This certainly makes the absorber more substantial in the practical applications such as highly efficient thermal emitters and other IR photonics devices with unpolarized light sources.

 figure: Fig. 5.

Fig. 5. Polarization-independent absorptivity of IPA with geometrical parameters of d = 2.01 μm, p = 3 μm, t = 0.4 μm: (a) Simulation and (b) measurement.

Download Full Size | PDF

4. Selective thermal emission of IPAs

A promising application of resonant perfect absorbers is spectrally selective thermal emitters. According to the Kirchhoff's law of thermal radiation, the thermal emissivity of an arbitrary body equals to its absorptivity at the thermodynamic equilibrium. Thus, a perfect absorber can inevitably serve as a highly-efficient emitter. Although most wavelength selective thermal emitters are made of metallic structures, only Au- and Pt-based emitters can operate in ambient air. Other metal-based emitters are not compatible with high-temperature ambient in the presence of oxygen gas or water vapor. In the following section, we demonstrate that ITO-based thermal emitters can work in air without any degradation in their performances and do not require any strict vacuum or inert gas ambient conditions. The fabricated S1 and S2 perfect absorbers were heated up using a heating plate and the emission spectra were characterized by a Fourier-transform infrared (FTIR) spectrometer (Thermo Nicolet NEXUS 670). The emitter can also be heated directly by applying an electrical current through the bottom ITO film. Figure 6(a) depicts the thermal emission measurement setup. As shown in this figure, the emission from IPAs are collected and guided to an FTIR spectrometer using a pair of concave metal coated mirrors. Emission spectra at normal angle of S2 IPA operated in ambient air at different temperatures varying from 178 °C to 307 °C measured at the surface of the emitter are described in Fig. 6(b). It is worth noting that in each measurement, the emitter was kept at a stabilized temperature and the emission spectrum was averaged for 100 spectra. Interestingly, the ITO emitter – S2 emits thermal radiation that has peaks at exactly the same resonant features of the IPA. In particular, the emission peaks almost did not change when the temperature increases from 178 °C to 307 °C, even kept at 307 °C for few hours, indicating the stability against the change of temperature and the robustness in ambient air of this wavelength-selective ITO emitter. Although the ITO emitter has shown good stability up to 307 °C after each time use, and the microstructure could retain the morphology at high temperatures; the device performance degraded gradually at higher temperatures above 400 °C (Appendix, Fig. 7). The ITO emitter can be another candidate besides gold emitter to work in air ambient; however, for the high temperature applications, refractory tungsten [13,50] and metal nitrides [15,16] are favorable materials since they have shown good thermal stability at very high temperatures.

 figure: Fig. 6.

Fig. 6. (a) Schematic diagrams of emissivity measurements. (b) Measured emission spectra of IPA – S2 at different temperatures ranging from 178 °C to 307 °C. Dashed black curve plots black body emission measured at 307 °C. (c) Simulated emissivity and (d) measured emissivity of S1 and S2.

Download Full Size | PDF

To verify the emissivity of the IPAs, the emission spectrum from a black body paint (JSC-3, Japan Sensor Corp.) at 307 °C was also taken for the emissivity determination (dotted curve in Fig. 6(b)). The emissivity spectra of S1 and S2 were therefore calculated by dividing their emission spectra taken at 307 °C to the black body’s emission spectrum taken above. Figures 6(c)–(d) plot the simulated and measured emissivity spectra of S1 (blue curves) and S2 (red curves), respectively. It should be noted that the simulated emissivity is identical to the simulated absorptivity according to the Kirchhoff's law. Interestingly, the measured emissivity spectra of the fabricated IPAs agree well with the simulated spectra of the pre-designed devices with high emissivity (above 0.86), and they also match perfectly to their measured absorptivity spectra shown in Fig. 4(e). Thus, the proposed IPA and fabrication process are realizable and can be used for various practical applications. For examples, by designing the resonances of the IPA matching to the molecular vibrations, the IPA can be used as the IR source for NDIR gas sensors or wavelength-selective heaters for food and material drying. Because the IPA composed of ITO and Al2O3, which are transparency in the visible region, the IPA devices can be applied for visible-transparent – IR wavelength-selective emitters for transparent heaters or shelf-cooling smart windows by radiative cooling.

5. Summary

In summary, we have demonstrated that ITO is a promising plasmonic material of choice for IR wavelength-selective perfect absorbers and their application in thermal emitters. The proposed IPAs composed of an ITO disk hexagonal lattice isolated to the bottom ITO mirror via an Al2O3 insulator. The IPA structures were numerically simulated to verify their optical properties, and then successfully fabricated using a scalable colloidal lithography fabrication process. The fabrication recipe proposed in this work precisely embodied the pre-designed IPA devices. The fabricated IPAs exhibited desired performance as high-performance thermal emitters, such as polarization independence, near-perfect absorptivity, and resonance tunability. Furthermore, the fabricated IPA emitters stably work in ambient air up to 307 °C without any vacuum or inert gas ambient conditions, providing another platform that can reduce the cost and be merged into practical IR devices.

Appendix

Optical property of IPA was stable after operating at 307 °C but it was dramatically degraded after operating at 590 °C even though the microstructure could retain the morphology.

 figure: Fig. 7.

Fig. 7. (a) Absorptivity of IPA (S2) after operating at 307 °C (top) and 590 °C (bottom). (b) Top-view and (c) Cross-sectional view SEM images of IPA after operating at 590 °C. (d) – (g) EDX elemental maps of IPA corresponding to SEM image in (c).

Download Full Size | PDF

Funding

Japan Society for the Promotion of Science (JSPS) (16F16315, 16H06364, 17K19045); Core Research for Evolutional Science and Technology (CREST) (JPMJCR13C3) from Japan Science and Technology Agency.

Acknowledgements

The authors would like to thank all the staffs at Namiki Foundry in NIMS for their kind supports, especially to Mr. Katsumi Ohno. Thang. D. Dao would like to thank the fellowship program (P16315) from JSPS.

References

1. X. Liu, T. Tyler, T. Starr, A. F. Starr, N. M. Jokerst, and W. J. Padilla, “Taming the Blackbody with Infrared Metamaterials as Selective Thermal Emitters,” Phys. Rev. Lett. 107(4), 045901 (2011). [CrossRef]  

2. E. Rephaeli and S. Fan, “Absorber and emitter for solar thermo-photovoltaic systems to achieve efficiency exceeding the Shockley-Queisser limit,” Opt. Express 17(17), 15145–15159 (2009). [CrossRef]  

3. L. P. Wang and Z. M. Zhang, “Wavelength-selective and diffuse emitter enhanced by magnetic polaritons for thermophotovoltaics,” Appl. Phys. Lett. 100(6), 063902 (2012). [CrossRef]  

4. B. Zhao, L. Wang, Y. Shuai, and Z. M. Zhang, “Thermophotovoltaic emitters based on a two-dimensional grating/thin-film nanostructure,” Int. J. Heat Mass Transfer 67, 637–645 (2013). [CrossRef]  

5. J.-Y. Chang, Y. Yang, and L. Wang, “Tungsten nanowire based hyperbolic metamaterial emitters for near-field thermophotovoltaic applications,” Int. J. Heat Mass Transfer 87, 237–247 (2015). [CrossRef]  

6. H. T. Miyazaki, T. Kasaya, M. Iwanaga, B. Choi, Y. Sugimoto, and K. Sakoda, “Dual-band infrared metasurface thermal emitter for CO2 sensing,” Appl. Phys. Lett. 105(12), 121107 (2014). [CrossRef]  

7. A. Lochbaum, Y. Fedoryshyn, A. Dorodnyy, U. Koch, C. Hafner, and J. Leuthold, “On-Chip Narrowband Thermal Emitter for Mid-IR Optical Gas Sensing,” ACS Photonics 4(6), 1371–1380 (2017). [CrossRef]  

8. E. Rephaeli, A. Raman, and S. Fan, “Ultrabroadband Photonic Structures To Achieve High-Performance Daytime Radiative Cooling,” Nano Lett. 13(4), 1457–1461 (2013). [CrossRef]  

9. A. P. Raman, M. A. Anoma, L. Zhu, E. Rephaeli, and S. Fan, “Passive radiative cooling below ambient air temperature under direct sunlight,” Nature 515(7528), 540–544 (2014). [CrossRef]  

10. L. Zhu, A. Raman, K. X. Wang, M. A. Anoma, and S. Fan, “Radiative cooling of solar cells,” Optica 1(1), 32 (2014). [CrossRef]  

11. T. Liu and J. Takahara, “Ultrabroadband absorber based on single-sized embedded metal-dielectric-metal structures and application of radiative cooling,” Opt. Express 25(12), A612 (2017). [CrossRef]  

12. I. Celanovic, N. Jovanovic, and J. Kassakian, “Two-dimensional tungsten photonic crystals as selective thermal emitters,” Appl. Phys. Lett. 92(19), 193101 (2008). [CrossRef]  

13. K. A. Arpin, M. D. Losego, A. N. Cloud, H. Ning, J. Mallek, N. P. Sergeant, L. Zhu, Z. Yu, B. Kalanyan, G. N. Parsons, G. S. Girolami, J. R. Abelson, S. Fan, and P. V. Braun, “Three-dimensional self-assembled photonic crystals with high temperature stability for thermal emission modification,” Nat. Commun. 4(1), 2630 (2013). [CrossRef]  

14. T. Yokoyama, T. D. Dao, K. Chen, S. Ishii, R. P. Sugavaneshwar, M. Kitajima, and T. Nagao, “Spectrally Selective Mid-Infrared Thermal Emission from Molybdenum Plasmonic Metamaterial Operated up to 1000° C,” Adv. Opt. Mater. 4(12), 1987–1992 (2016). [CrossRef]  

15. U. Guler, A. Boltasseva, and V. M. Shalaev, “Refractory Plasmonics,” Science 344(6181), 263–264 (2014). [CrossRef]  

16. W. Li, U. Guler, N. Kinsey, G. V. Naik, A. Boltasseva, J. Guan, V. M. Shalaev, and A. V. Kildishev, “Refractory Plasmonics with Titanium Nitride: Broadband Metamaterial Absorber,” Adv. Mater. 26(47), 7959–7965 (2014). [CrossRef]  

17. J. Liu, U. Guler, A. Lagutchev, A. Kildishev, O. Malis, A. Boltasseva, and V. M. Shalaev, “Quasi-coherent thermal emitter based on refractory plasmonic materials,” Opt. Mater. Express 5(12), 2721 (2015). [CrossRef]  

18. M. Kumar, N. Umezawa, S. Ishii, and T. Nagao, “Examining the Performance of Refractory Conductive Ceramics as Plasmonic Materials: A Theoretical Approach,” ACS Photonics 3(1), 43–50 (2016). [CrossRef]  

19. R. P. Sugavaneshwar, S. Ishii, T. D. Dao, A. Ohi, T. Nabatame, and T. Nagao, “Fabrication of Highly Metallic TiN Films by Pulsed Laser Deposition Method for Plasmonic Applications,” ACS Photonics 5(3), 814–819 (2018). [CrossRef]  

20. P. F. Robusto and R. Braunstein, “Optical Measurements of the Surface Plasmon of Indium-Tin Oxide,” Phys. Status Solidi A 119(1), 155–168 (1990). [CrossRef]  

21. S. H. Brewer and S. Franzen, “Indium Tin Oxide Plasma Frequency Dependence on Sheet Resistance and Surface Adlayers Determined by Reflectance FTIR Spectroscopy,” J. Phys. Chem. B 106(50), 12986–12992 (2002). [CrossRef]  

22. S. A. Knickerbocker and A. K. Kulkarni, “Estimation and verification of the optical properties of indium tin oxide based on the energy band diagram,” J. Vac. Sci. Technol., A 14(3), 757–761 (1996). [CrossRef]  

23. C. Rhodes, S. Franzen, J.-P. Maria, M. Losego, D. N. Leonard, B. Laughlin, G. Duscher, and S. Weibel, “Surface plasmon resonance in conducting metal oxides,” J. Appl. Phys. 100(5), 054905 (2006). [CrossRef]  

24. A. Tamanai, T. D. Dao, M. Sendner, T. Nagao, and A. Pucci, “Mid-infrared optical and electrical properties of indium tin oxide films,” Phys. Status Solidi A 214(3), 1600467 (2017). [CrossRef]  

25. L. Dominici, F. Michelotti, T. M. Brown, A. Reale, and A. D. Carlo, “Plasmon polaritons in the near infrared on fluorine doped tin oxide films,” Opt. Express 17(12), 10155–10167 (2009). [CrossRef]  

26. F. Khalilzadeh-Rezaie, I. O. Oladeji, J. W. Cleary, N. Nader, J. Nath, I. Rezadad, and R. E. Peale, “Fluorine-doped tin oxides for mid-infrared plasmonics,” Opt. Mater. Express 5(10), 2184 (2015). [CrossRef]  

27. R. Buonsanti, A. Llordes, S. Aloni, B. A. Helms, and D. J. Milliron, “Tunable Infrared Absorption and Visible Transparency of Colloidal Aluminum-Doped Zinc Oxide Nanocrystals,” Nano Lett. 11(11), 4706–4710 (2011). [CrossRef]  

28. G. V. Naik, J. Liu, A. V. Kildishev, V. M. Shalaev, and A. Boltasseva, “Demonstration of Al: ZnO as a plasmonic component for near-infrared metamaterials,” Proc. Natl. Acad. Sci. 109(23), 8834–8838 (2012). [CrossRef]  

29. J. Kim, G. V. Naik, A. V. Gavrilenko, K. Dondapati, V. I. Gavrilenko, S. M. Prokes, O. J. Glembocki, V. M. Shalaev, and A. Boltasseva, “Optical Properties of Gallium-Doped Zinc Oxide—A Low-Loss Plasmonic Material: First-Principles Theory and Experiment,” Phys. Rev. X 3(4), 041037 (2013). [CrossRef]  

30. P. R. West, S. Ishii, G. V. Naik, N. K. Emani, V. M. Shalaev, and A. Boltasseva, “Searching for better plasmonic materials,” Laser Photonics Rev. 4(6), 795–808 (2010). [CrossRef]  

31. G. V. Naik, V. M. Shalaev, and A. Boltasseva, “Alternative Plasmonic Materials: Beyond Gold and Silver,” Adv. Mater. 25(24), 3264–3294 (2013). [CrossRef]  

32. M. Kanehara, H. Koike, T. Yoshinaga, and T. Teranishi, “Indium Tin Oxide Nanoparticles with Compositionally Tunable Surface Plasmon Resonance Frequencies in the Near-IR Region,” J. Am. Chem. Soc. 131(49), 17736–17737 (2009). [CrossRef]  

33. T. Wang and P. V. Radovanovic, “Free Electron Concentration in Colloidal Indium Tin Oxide Nanocrystals Determined by Their Size and Structure,” J. Phys. Chem. C 115(2), 406–413 (2011). [CrossRef]  

34. M. Xi and B. M. Reinhard, “Localized Surface Plasmon Coupling between Mid-IR-Resonant ITO Nanocrystals,” J. Phys. Chem. C 122(10), 5698–5704 (2018). [CrossRef]  

35. S. Q. Li, P. Guo, L. Zhang, W. Zhou, T. W. Odom, T. Seideman, J. B. Ketterson, and R. P. H. Chang, “Infrared Plasmonics with Indium–Tin-Oxide Nanorod Arrays,” ACS Nano 5(11), 9161–9170 (2011). [CrossRef]  

36. S. Shrestha, Y. Wang, A. C. Overvig, M. Lu, A. Stein, L. D. Negro, and N. Yu, “Indium Tin Oxide Broadband Metasurface Absorber,” ACS Photonics 5(9), 3526–3533 (2018). [CrossRef]  

37. M. Abb, Y. Wang, N. Papasimakis, C. H. de Groot, and O. L. Muskens, “Surface-Enhanced Infrared Spectroscopy Using Metal Oxide Plasmonic Antenna Arrays,” Nano Lett. 14(1), 346–352 (2014). [CrossRef]  

38. K. Chen, P. Guo, T. D. Dao, S.-Q. Li, S. Ishii, T. Nagao, and R. P. H. Chang, “Protein-Functionalized Indium-Tin Oxide Nanoantenna Arrays for Selective Infrared Biosensing,” Adv. Opt. Mater. 5(17), 1700091 (2017). [CrossRef]  

39. F. D’apuzzo, M. Esposito, M. Cuscunà, A. Cannavale, S. Gambino, G. E. Lio, A. De Luca, G. Gigli, and S. Lupi, “Mid-Infrared Plasmonic Excitation in Indium Tin Oxide Microhole Arrays,” ACS Photonics 5(6), 2431–2436 (2018). [CrossRef]  

40. F. Yi, E. Shim, A. Y. Zhu, H. Zhu, J. C. Reed, and E. Cubukcu, “Voltage tuning of plasmonic absorbers by indium tin oxide,” Appl. Phys. Lett. 102(22), 221102 (2013). [CrossRef]  

41. P. Guo, R. D. Schaller, J. B. Ketterson, and R. P. H. Chang, “Ultrafast switching of tunable infrared plasmons in indium tin oxide nanorod arrays with large absolute amplitude,” Nat. Photonics 10(4), 267–273 (2016). [CrossRef]  

42. X. Liu, J.-H. Kang, H. Yuan, J. Park, S. J. Kim, Y. Cui, H. Y. Hwang, and M. L. Brongersma, “Electrical tuning of a quantum plasmonic resonance,” Nat. Nanotechnol. 12(9), 866–870 (2017). [CrossRef]  

43. X. Liu, J.-H. Kang, H. Yuan, J. Park, Y. Cui, H. Y. Hwang, and M. L. Brongersma, “Tuning of Plasmons in Transparent Conductive Oxides by Carrier Accumulation,” ACS Photonics 5(4), 1493–1498 (2018). [CrossRef]  

44. T. Minami, T. Miyata, and T. Yamamoto, “Stability of transparent conducting oxide films for use at high temperatures,” J. Vac. Sci. Technol., A 17(4), 1822–1826 (1999). [CrossRef]  

45. A. D. Rakić, A. B. Djurišić, J. M. Elazar, and M. L. Majewski, “Optical properties of metallic films for vertical-cavity optoelectronic devices,” Appl. Opt. 37(22), 5271–5283 (1998). [CrossRef]  

46. E. D. Palik, Handbook of Optical Constants of Solids, 3rd ed. (Academic Press, 1998).

47. M. Pu, C. Hu, M. Wang, C. Huang, Z. Zhao, C. Wang, Q. Feng, and X. Luo, “Design principles for infrared wide-angle perfect absorber based on plasmonic structure,” Opt. Express 19(18), 17413–17420 (2011). [CrossRef]  

48. T. D. Dao, K. Chen, S. Ishii, A. Ohi, T. Nabatame, M. Kitajima, and T. Nagao, “Infrared Perfect Absorbers Fabricated by Colloidal Mask Etching of Al–Al2O3–Al Trilayers,” ACS Photonics 2(7), 964–970 (2015). [CrossRef]  

49. T. D. Dao, S. Ishii, T. Yokoyama, T. Sawada, R. P. Sugavaneshwar, K. Chen, Y. Wada, T. Nabatame, and T. Nagao, “Hole Array Perfect Absorbers for Spectrally Selective Mid-Wavelength Infrared Pyroelectric Detectors,” ACS Photonics 3(7), 1271–1278 (2016). [CrossRef]  

50. P. N. Dyachenko, S. Molesky, A. Y. Petrov, M. Störmer, T. Krekeler, S. Lang, M. Ritter, Z. Jacob, and M. Eich, “Controlling thermal emission with refractory epsilon-near-zero metamaterials via topological transitions,” Nat. Commun. 7(1), 11809 (2016). [CrossRef]  

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 (7)

Fig. 1.
Fig. 1. (a) A comparison between measured permittivity of the sputtered ITO film (red curves) with permitivities of Au (orange curves) and W (blue curves). Negative real part of ITO’s permittivity clearly reveals plasmonic behavior of ITO in the IR region. The complex permitivities of Au and W taken from the literature [45] were divided by a factor of 10 for a better comparison. (b) Figure-of-merit ($- {{{\varepsilon _1}} \mathord{\left/ {\vphantom {{{\varepsilon_1}} {{\varepsilon_2}}}} \right.} {{\varepsilon _2}}}$, solid curves) and skin depths (dashed curves) of ITO (red), Au (orange) and W (blue).
Fig. 2.
Fig. 2. (a) Scheme of IPA configured by tri-layered ITO – Al2O3 – ITO with a hexagonal lattice ITO disk resonator array arranged on the top of Al2O3-ITO layered films. Dashed rectangle indicates the unit cell defined in the simulation. (b) Simulated optical spectra including absorptivity – A (red curve), transmittance – T (blue curve) and reflectance – R (black curve) of IPA with a periodicity – p of 3 μm, diameter – d of 2.01 μm and insulator thickness – t of 0.4 μm. (c) Resonance tunability of IPA by changing diameter – d of ITO disk resonator while keeping the other parameters unchanged (p = 3 μm, t = 0.4 μm). The thicknesses of ITO disk and ITO film are kept unchanged at 0.15 μm and 0.4 μm, respectively. In the simulation, the incident electromagnetic field propagated along the –z-axis, the electric field oscillated along the x-axis.
Fig. 3.
Fig. 3. Simulated electric fields (Ex and Ez) and magnetic field (Hy) distributions of the IPA (d = 2.01 μm, p = 3 μm, t = 0.4 μm) excited at two resonance modes: (a) at M1 (wavelength 4.0 μm) and (b) at M2 (8.4 μm). (c) Simulated angle-dependent absorptivity of the IPA with geometrical parameters of d = 2.01 μm, p = 3 μm, t = 0.4 μm. The dashed curves indicate SPPs at metal-dielectric interfaces of ITO disk hexagonal lattice (see Eq. (1)). (d) Simulated insulator thickness dependence of IPA’s absorptivity while keeping d and p unchanged at 2.01 μm and at 3 μm, respectively. In the simulation, the incident electromagnetic field propagated along the –z-axis, the electric field oscillated along the x-axis, and the incident amplitudes were normalized to 1.
Fig. 4.
Fig. 4. (a) Schematic diagrams illustrate scalable fabrication process of IPAs using colloidal lithography combined two steps of RIE. (b) SEM image (top) and photo (bottom) of the fabricated IPA. Arrows indicate typical defects of periodic ITO disk array. (c) SEM images of two fabricated IPAs having diameters of 1.62 μm (S1) and 2.01 μm (S2) with the same other parameters (p = 3 μm, t = 0.4 μm). (d) Simulated absorptivities and (e) measured absorptivities of S1 (blue curves) and S2 (red curves).
Fig. 5.
Fig. 5. Polarization-independent absorptivity of IPA with geometrical parameters of d = 2.01 μm, p = 3 μm, t = 0.4 μm: (a) Simulation and (b) measurement.
Fig. 6.
Fig. 6. (a) Schematic diagrams of emissivity measurements. (b) Measured emission spectra of IPA – S2 at different temperatures ranging from 178 °C to 307 °C. Dashed black curve plots black body emission measured at 307 °C. (c) Simulated emissivity and (d) measured emissivity of S1 and S2.
Fig. 7.
Fig. 7. (a) Absorptivity of IPA (S2) after operating at 307 °C (top) and 590 °C (bottom). (b) Top-view and (c) Cross-sectional view SEM images of IPA after operating at 590 °C. (d) – (g) EDX elemental maps of IPA corresponding to SEM image in (c).

Equations (2)

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

εmεdεm+εd=sin2θ+2p(i+j)λsinθ+43p2(i2+ij+j2)λ2
A=1TR
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.