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

Transparent planar indium tin oxide for a thermo-photovoltaic selective emitter

Open Access Open Access

Abstract

Designing an efficient emitter design is an important step for achieving a highly efficient TPV conversion process. Wavelength-selective emissivity, spectra match between the emitter and TPV cells, and high thermal stability are three main characteristics that must be considered before implementing the emitter. In this work, an indium tin oxide (ITO)/sapphire emitter structure is investigated for TPV application over the temperature range from 200°C to 1000°C. A 1-µm-thick ITO layer is deposited on a 650-µm-thick sapphire substrate. In addition, 50-nm-thick SiO2 is deposited on top of the ITO to enhance the performance of emitter at high temperatures. High-temperature emissivity and absorptivity measurement of the emitter samples are obtained using FTIR and a Hitachi U-4100 spectrophotometer, respectively. The resultant SiO2/ITO/sapphire/stainless-steel planar emitter structure has selective emission with high emissivity of ∼0.8 in the 1–1.6 µm wavelength regime at 1000°C. This emission range lies at the bandgap edge of silicon TPV cells and thus can be used to harness the true potential for making a low-cost thermophotovoltaic system.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Thermophotovoltaic (TPV) system, first introduced in 1956 [1], is an efficient energy conversion system that operates at a high-temperature and converts thermal radiative energy to electrical energy [14]. TPVs became the extensive research area because of its ability to exceed the efficiency of the conventional Shockley-Queisser limit for a photovoltaic cell [5,6]. In the TPV system, photovoltaic (PV) cells are irradiated with high thermal emission from a local heat source, instead of solar energy, to generate electrical power. The basic components of the modern TPV system include (i) a combustor or a heat source, (ii) spectrally selective emitter (SE), and (iii) an array of a low bandgap PV cell. For an efficient TPV system, ideally, an emitter should have these characteristics: high emissivity, spectrally tailored long and short wavelength emission, and high thermal stability.

High operating temperature is beneficial for the TPV system to procure a high radiative power, as stated by the Stefan-Boltzmann law (radiated power ∝ T4). An emitter is heated (>800°C) by using any heat source such as solar energy [7], combustible fossil fuel [8], radiative isotopes [9], waste heat [10], etc. The heated emitter at elevated temperatures radiates photons having a spectrum over a broadband range. These radiated thermal photons will be collected by the PV cell to generate charge carriers to produce electric power. Most of the radiated photons have long-wavelength having energies less than the PV cell’s bandgap is transparent to the PV cell. These photons will eventually be absorbed by the surrounding, e.g., packaging, and will increase the operating temperature of the PV cell and thus decrease its external quantum efficiency. These photons that may rather be unusable are suppressed by using a spectrally selective emitter [11]. Also, selective emitter must be able to tailor its emissivity peak slightly above the wavelength (λEg) corresponding to the bandgap energy (Eg) of the PV cell to ensure the maximum absorption since at bandgap energy the density of states is zero. Finally, emission of the photons with energies much higher than the bandgap of the PV cell also needs to be suppressed to minimize the thermalization loss [12] and hence enhances the efficiency of the TPV system [13]. The total power emitted by the emitter should be equal to the ideal blackbody spectrum multiplied by the emitter’s selective emissivity. On the other hand, the TPVs overall efficiency is an amalgamation of the various component’s individual efficiency, such as combustion, spectral, radiant, filter, view-factor, PV cell, and inverter efficiency [14]. Therefore, the overall thermophotovoltaic system’s efficiency can also be summarized as an integration of the radiant, spectral, and photovoltaic cell efficiencies [15].

Designing high-performance selective emitters for practical application has great challenges, and various efforts have been made in the past for the experimental realization of the TPV emitters [1624]. Selective emitters need to maintain the vital functional properties under the extremely high-temperatures with a strong thermally induced stress in conjunction with an oxidizing surrounding. Thus, designing an effective absorber/emitter structure for tuning of the spectral response is the crucial step for the overall optimization of an efficient TPV system [2527]. For practical applications, selective emitters must be able to spectral tailor the emission wavelengths and sustain at the high-temperatures. It should have low emissivity for low energy photons (usually lower than the bandgap of PV cell) and high emissivity for photons having energies just above the bandgap of the PV cell [2830]. Several materials have been used earlier for fabricating efficient emitter, including rare-earth materials such as erbium-doped aluminum garnet (ErAG) emitter material based double side filtering structure [31], Er2O3/Yb2O3 [32], photonic crystals (PhCs) [27,33,34], plasmonic metamaterials [30,35], and frequency selective surfaces [36]. Although these efforts are remarkable, further improvements can be made in the aspects of alleviating surface deformation at elevated temperatures, reducing layer thickness and layer number, and eliminating the requirement of lithography and deep plasma etching. The improvements can further enhance the stability, cost-effectiveness, and feasibility of TPV technology.

Here, we present a planar indium tin oxide (ITO) on a sapphire transparent emitter structure for the TPV application, as shown in Fig. 1. ITO, a transparent conducting oxide (TCO), is a wide-bandgap (Eg>4.1 eV) semiconductor that possesses high transmittance (>85%) in the visible spectrum with relatively low electrical resistivity (10−4 Ω-cm) [37,38]. It is mostly used for semiconductor devices such as TPV cells, gas sensors, flat panel displays, anti-reflection coating, etc. [39]. Some previous research works have been published specifying the thermal stability of ITO thin films in the temperature range from 350°C-1250°C [4043]. ITO deposition on high thermal conductivity substrate, such as sapphire, has improved optical lifetime performance and damage thresholds [44]. Also, the sapphire substrate has decent interface stability with ITO [45] at high temperatures. Furthermore, to avoid major mechanisms that degrade the emitter structure, i.e., oxidation [46], surface diffusion, and nanostructure deformation [47,48] of TPV emitters at high-temperatures, ITO/sapphire structure can be beneficial. Additionally, SiO2/ITO/Sapphire/Stainless-steel planar emitter structure has a selective emission with high emissivity in the 1–1.6 µm wavelength regime at 1000°C. This selective emission is ideal for Si (Eg = 1.1 eV) and GaSb (Eg = 0.72 eV) based TPV systems. Moreover, having a planar design for the device eliminates the use of lithography and etching processes [49]. In this work, SiO2 is deposited on top of the ITO to enhance the performance at high temperatures [50]. Stainless steel (type 304, one side polished) is used as a metallic backplate to avoid transmittance from the transparent device at a short wavelength. It will be shown experimentally that ITO/Sapphire transparent structure can provide high emittance at near-infrared while maintaining a visible spectrum without any backplate. Some of the initial results can be found in our previous papers [24,49] and student theses [5153].

 figure: Fig. 1.

Fig. 1. (a) The emitter structure in this study with tSiO2 = 50 nm, tITO = 1 µm, tSapphire = 650 µm, tsteel = 0.8 mm, (b) emitter sample without steel before high temperature measurements, and (c) emitter sample without steel after high temperature measurements.

Download Full Size | PDF

ITO and other TCOs have been widely used for near-field TPV [5459] using fluctuation electrodynamic due to their near-infrared surface plasmonic properties. However, here the ITO/sapphire emitter structure is more relevant to the far-field TPV system. Near-field TPV does not follow the ideal Planck law for blackbody radiation [57] and can emit at near-infrared radiation at a lower temperature. These theoretical results have been very encouraging, yet, the experimental realization of these near-field TPV systems is quite difficult because of the ∼µm TPV cell-emitter spacing that hampers the TPV cell operation. Thus, a vacuum environment is inevitable to prevent heat conduction and convection from the emitter to the TPV cells [60].

Moreover, ITO/sapphire structure is transparent at the visible spectrum, which can be incorporated with the micro combustors and impart the advantage of monitoring the combustor interior structure stability and flame quenching effect at elevated temperatures [61]. Also, an ITO on the sapphire structure is more chemically stable at high-temperature than solely ITO due to interfacial Al-In bonds [45]. Usually, operating temperature for a typical TPV lies in the range of 800-1200°C, and to date, only a few studies have actually provided the emissivity/absorption measurement [59,6268]. In this regard, our planarized emitter structure possesses high stable thermal radiation property due to the superior interfacial property of the planar ITO-sapphire interface.

2. Methods

2.1. Sample fabrication

A 650 µm thick double-sided polished sapphire is bought from the Summit-tech, a Taiwan based company. First, a 1 µm ITO layer is deposited on one side of the sapphire substrate via DC sputtering process. For the entire process, argon and oxygen, precursor gases are maintained at 200 °C. A flow rate of oxygen is maintained at 0.5 sccm, and the vapor partial pressure is kept at ∼60%. The deposition is done at 300 V with a DC power and current of 0.4 KW and 1.28 A, respectively. The vacuum pressure is kept at ∼1×10−5 torr.

Then, 50 nm thick SiO2 is deposited on top of the ITO by using PECVD method. The flow rate of the precursor gases SiH4, N2O, and N2 is maintained at 8.5 sccm, 710 sccm, and 161.5 sccm, respectively. The RF power, temperature, and pressure for the whole deposition period are maintained at 20 W, 300 °C, and 1000 mtorr, respectively.

2.2. Sample measurement

2.2.1. Room-temperature (RT) measurement

Hitachi U-4100 Spectrophotometer and Bruker’s FTIR machine is used to measure Reflectance (R), Transmittance (T), and Absorption (A) at room temperature. Hitachi U-4100 spectrophotometer with a built-in integration sphere is used for measuring the reflectance (R) and transmission (T) in the ultra-violet, visible, and near-infrared wavelength regime (UV-VIS-NIR) ranging from 300 nm to 2600 nm. Absorption (A) can be calculated by using the formula:

$$A = 1 - R - T$$

The detailed information of the measurement setup is mentioned in [69]. Room-temperature (RT) Fourier transform infrared spectrophotometer (FTIR) has been used to measure the R, T, and A of the ITO/sapphire structure. The wavelength range of the RT FTIR is from 1µm–20 µm. At room temperature, Bruker FTIR uses a silicon carbide (SiC) glow bar as a light source. For detection, a pyroelectric detector using deuterated lanthanum α alanine doped triglycine sulfate (DLATGS) with KBr is used. For the RT FTIR measurements, Bruker IFS66 V/S is used. Wavenumber ranges from 7500 to 160 cm−1 with IFS66 V/S stepper scanner PAD 28a card. The mid-IR time-resolved spectra resolution is 5 ns, and the spectral resolution is 5cm−1. The aperture size is 5 mm. Cold Field Emission Scanning Electron Microscope (Hitachi SU8010) and Energy Dispersive Spectrometer are used for obtaining SEM images and EDS analysis of the elements. The electron source is a cathode electron gun, and the voltage is 0.1kV∼30 kV. The resolution is 10 Å (at 15 kV) or 13 Å (at 1 kV). The elemental range that can be detected is from Be4 to U92.

2.2.2. High-temperature (HT) measurement

According to Kirchhoff’s law of thermal radiation, under thermal equilibrium, the absorptivity and emissivity of a surface are equal at a specific temperature and wavelength. Thermo Fisher’s FTIR machine is used to measure the emissivity (ε) at high temperatures. The high-temperature emissivity measurement system contains a sample heater, mirror, blackbody oven, and FTIR. The sample and the blackbody are heated to the same temperature. With the help of a mirror and an external port, radiation intensities from the sample and reference blackbody are directed inside the FTIR machine for measurements. The detector is DTGS/KBr, and the wavelength range is from 1µm to 18µm. The model of HT FTIR is Thermo Fisher Scientific Inc. Nicolet iS50 FTIR. For the rotation setup and resolution, the mirror is controlled by a step motor, and the resolution of the rotation angle is 0.12°, OMNIC software is used to control the FTIR. The setting of FTIR in the wavelength band (400cm−1 ∼ 10420 cm−1), as shown in Table 1, does not require segmental adjustment. When measuring, each spectrum is averaged after 32 scans by FTIR, and the resolution of the measurement is 0.482 cm−1.

Tables Icon

Table 1. Settings of high-temperature FTIR.

In the sample heater, a heating coil is used to heat the ceramic surface, and the ceramic surface then transfers the heat to the sample. The ceramic surface is behind the sample holder made of stainless steel, and the radiation of the sample comes out from the radiation hole of ZnSe glass. The ceramic surface also emits radiation, but the direction is opposite to the radiation hole, and thus it does not affect the measurement. The sample heater is connected with a PID (proportional-integral-derivative) temperature controller (TAIE, FY900). A thermocouple (Highlight Tech Corp.) is used to estimate the temperature of the sample while operating. The whole emissivity measurement from 200°C to 1000°C takes ∼6 hours (5 hours for heating and 1 hour for measurement). The whole system, including the sample holder, is in a vacuum having a pressure of 1 torr during the measurement. The complete setup of the FTIR machine and the heater are shown in Fig. 2. The emissivity is defined as the sample radiation intensity divided by the blackbody radiation intensity, and its value is between 0 and 1. The formula is

$$\mathrm{\varepsilon} (T,\theta ) = \frac{{{I_{\textrm{sample, }\mathrm{\lambda }}}(T,\theta )}}{{{I_{\textrm{blackbody, }\mathrm{\lambda }}}(T)}}$$
where ε is emissivity, I is radiation intensity, T is temperature, θ is emission angle, λ means dependent on wavelength. To eliminate the background or surrounding noise (Isur) during measurement, we have to subtract the radiation intensity released by the background from the numerator and the denominator in Eq. (2), and we arrive at
$${\mathrm{\varepsilon} }(T,\mathrm{\theta }) = \frac{{{I_{\textrm{sample, }\mathrm{\lambda }}}(T,\mathrm{\theta }) - {I_{\textrm{sur, }\mathrm{\lambda }}}({T_{\textrm{sur}}},\mathrm{\theta })}}{{{I_{\textrm{blackbody, }\mathrm{\lambda }}}(T) - {I_{\textrm{sur, }\mathrm{\lambda }}}({T_{\textrm{sur}}},\mathrm{\theta })}}$$
where subscript sur means surrounding or environmental condition.

 figure: Fig. 2.

Fig. 2. System setup (a) Diagram of the high-temperature emissivity measurement system, (b) 3D design of the sample heater.

Download Full Size | PDF

3. Results and discussion

The cross-sectional SEM images of the emitter sample before and after emissivity measurement are shown in Fig. 3. The 1µm thick ITO layer is sandwiched between a top protective layer of SiO2 and a bottom thick sapphire layer of thickness 50 nm and 650µm, respectively. Simulation results of ITO-based emitters can be found in our previous work [49]. Figure 3(a) is the SEM image taken before 1000°C vacuum measurement, and Fig. 3(c) is taken after high-temperature emissivity measurement at 1000°C in vacuum.

 figure: Fig. 3.

Fig. 3. (a) Cross-sectional SEM image of emitter sample before annealing along with EDS analysis of (b) oxygen, aluminum, indium, and tin elements, measured at room temperature before heating, (c) Cross-sectional SEM image of emitter sample after annealing along with EDS analysis of (d) oxygen, aluminum, indium, and tin elements, measured at room temperature after emissivity measurement.

Download Full Size | PDF

3.1. Room-temperature results

EDS analysis of elements distribution in emitter sample before heating is shown in Fig. 3(b). The whole section of ITO and sapphire is selected for the line scan. Indium (blue), tin (purple), and oxygen (red) elements are confirmed in the ITO layer as expected. EDS analysis also confirmed the diffusion of indium and tin atoms into the sapphire layer.

The prerequisite for an emitter to have high emissivity is to have high optical absorption, as stated by the Kirchhoff’s law. Room temperature R, T, and A measurement of the sample measured by RT FTIR for the wavelength range of 1.2-10 µm is shown in Fig. 4(a) and (b) [49]. It is evident from the measurement that the emitter structure without stainless steel has very low transmittance (∼60%) at a wavelength near 1 µm, and transmittance of the sample with stainless steel is zero. Transmittance decreases to zero for both of the samples at a wavelength of ∼2 µm, and this corresponds to the application where the see-through-emitter is not desired. For both emitter structure, high absorption is observed at λ∼1.7 µm and after that absorptivity decrease sharply for λ∼2-4 µm. High absorption of ∼93% is observed for ITO/sapphire with the inclusion of the steel as compared to the ∼88% of the sample without stainless steel. This peak in NIR region in as-deposited 1 µm thick ITO is because of the free charge carrier absorption phenomenon. The carrier concentration (1020−1021cm−3) in as-deposited ITO is lower than metals. Therefore, ITO exhibits plasma frequency in the NIR region [59] below the fundamental absorption edge at the bandgap, where the semiconductor should normally be transparent. This free charge carrier absorption can be explained by using Drude’s free electron model [7072]:

$${{\varepsilon} _\textrm{c}} = {{\varepsilon} _1} - {{\varepsilon} _2} = {{\varepsilon} _\infty } - \frac{{\omega _p^2}}{{{\omega ^2} + \textrm{ i}\gamma \omega }}$$
$$\omega _p^2 = \frac{{N{e^2}}}{{{{\varepsilon} _0}{m^\ast }}}$$
where εc is a dielectric function, ωp is the plasma frequency, γ is damping constant, ε is high-frequency limit, N is free carrier density, e is the electronic charge, m* is the effective mass of electron in the conduction band, ε0 is the space permittivity, ε1 and ε2 are real and imaginary dielectric function. Drude model explains that plasma frequency is directly dependent on the carrier concentration, and it implies that plasma wavelength peak will shift to shorter and longer wavelength with increased or decreased carrier concentration. At frequencies above plasma frequency, the presence of free charge carrier concentration leads to the absorption in metals and doped semiconductors. The absorption coefficient (α) of the free carrier absorption is also proportional to carrier density and varies inversely with frequency, as shown below:
$${\alpha _{free\textrm{ }carrier}} = \frac{{N{e^2}}}{{{m^\ast }{{\varepsilon} _0}nc\tau {\omega ^2}}}$$
where n is the index of refraction, τ is the damping time (reciprocal of the damping rate γ) and ω is the frequency. The free carrier absorption in doped semiconductors corresponds to the excitation of the electron from an occupied state below the fermi level to an unoccupied state above the Fermi level [73]. 1µm thick ITO provides high absorption at near-infrared (NIR) and high reflectance after λ=2.5µm. Thinner ITO makes the effect of the substrate more pronounced, which is undesirable, while a thicker ITO layer leads to higher raw material and deposition cost.

 figure: Fig. 4.

Fig. 4. Reflectance (R), Transmittance (T), and Absorption (A) measurement of the emitter sample at room temperature using (a) FTIR measurement of emitter sample without stainless steel, and (b) with stainless steel, (c) UV-VIS-NIR spectrophotometer measurement of emitter sample without stainless steel, and (d) with stainless steel.

Download Full Size | PDF

Hitachi Spectrophotometer R, T, A measurement at room temperature of the sample for the UV-VIS-NIR wavelength ranging from 300 nm to 2600 nm is shown in Fig. 4(c) and (d) [49]. High transmission at the visible range can be observed for the sample without stainless steel (Fig. 4(c)) because of the transparency of the ITO and sapphire in the visible spectrum. The high absorption curve is observed at the range of λ∼1.6-2 µm for both samples. Likewise, RT FTIR measurement, spectrophotometer measurement also indicates the peak absorption intensity for both of the emitter samples, w/ or w/o stainless steel can be observed at λ∼1.7 µm. Figure 4(b) and (d) shows the enhancement in peak absorption intensity due to the inclusion of the steel backplate. In addition, the absorptivity decreases sharply after λ∼2 µm. Hence, the long wavelength emissivity suppression capability of the sample is verified, which is beneficial for the TPV application. Also, long-wavelength absorption/emission suppression of the sample is evident from both measurements at room temperature.

3.2. High-temperature results

EDS analysis of elements distribution in the emitter sample after heating the sample in vacuum at 1000°C is shown in Fig. 3(d). Even after high-temperature vacuum measurement, the element distribution is quite consistent and shows no excess diffusion of elements in the sample.

The normal working temperature of a typical TPV emitter is at around 800°C -1200°C, and therefore, high temperatures emissivity measurement of ITO/sapphire emitter sample will verify its ability to radiate at high emissivity with long-wavelength emission suppression. Emissivity measurement of stainless steel for different temperatures at the normal emission angle is shown in Fig. 5. The temperature is gradually varied from 200°C to 1000°C, as mentioned above. For this measurement, the emissivity value increases slightly with increased temperature. For the normal emission angle measurement, the emissivity of steel gradually decreases in the wavelength range from 1–6µm and maintains consistency from 6–10µm. We show the emissivity data from low (200°C) to high temperature (1000°C) to see the gradually changed emissivity to gain confidence on the high temperature data. Low-temperature emssivity is, in general, easier to measure. If we do not see gradual transition from the low to high temperature, high-temperature data cannot be confirmed.

 figure: Fig. 5.

Fig. 5. Detailed emissivity measurement of stainless steel at normal emission angle at different temperatures. The shaded zone indicates a noisy portion where data is not usable.

Download Full Size | PDF

Similarly, the emissivity measurement of the ITO/sapphire emitter structure without and with stainless steel at the normal angle is shown in Fig. 6. The signal of the thermal emission has temperature and wavelength dependence according to blackbody radiation and Wien’s displacement law. The short wavelength (λ=1-3µm) portion of the emissivity is thus quite noisy at low temperature. Fortunately, the TPV operating temperature is at 800°C -1200°C, so we are mainly interested at high temperature emission spectrum. We use a method to determine the wavelength limitation for emissivity measurement at different temperatures, which depends on the ratio of the background and the sample signals:

$$\frac{{{I_{\textrm{sur,}\mathrm{\lambda }}}({\textrm{T}_{\textrm{sur}}},\mathrm{\theta })}}{{{I_{\textrm{sample,}\mathrm{\lambda }}}(\textrm{T},\mathrm{\theta })}} \ge \textrm{0}\textrm{.5}$$
In this method, we divide the background signal by the sample signal. If the ratio is greater than 0.5, we discard the portion of the data. In the following measurement, we mark the wavelength range where Eq. (7) is greater than 0.5. It can be observed from Fig. 6 that the ITO/sapphire emitters with or without steel have a high emissivity in the wavelength range of 1-1.6µm for temperatures 800°C-1000°C with a gradual cutoff at long wavelength. This emission wavelength regime is ideal for Si (Eg = 1.1 eV) and GaSb (Eg = 0.72 eV) based TPV cells for TPV applications at elevated temperatures. Again, suppressed emissivity at long wavelengths is also evident here, thereby validating the applicability of our structure as an efficient TPV emitter at high temperatures.

 figure: Fig. 6.

Fig. 6. Detailed emissivity measurement of ITO/sapphire emitter sample (a) without stainless steel at normal emission angle, (b) with stainless steel at normal emission angle, at different temperatures. The shaded zone indicates a noisy portion where data is not usable.

Download Full Size | PDF

In Fig. 6, we observe the high emissivity zone in short wavelength at 800-1000°C. In addition, the suppressed emissivity at long wavelength (>2.5µm) is also observed. The successfully demonstrated selective emission using ITO/sapphire emitter ensure the application in high efficiency TPV systems. It is also observed that the emissivity/absorptivity peak at 1000°C shifts toward shorter wavelength, compared to the room-temperature absorption peak at λ=1.7µm. It should be emphasized that the n-k at high temperature is, in general, deviated from the n-k at room temperature. This is common in most of the materials, especially when temperature goes up to 1000°C [67]. Thus while n-k is already changed at 1000°C, it is not very surprising that the plasmon absorption peak shifts. From Drude model in Eq. (4) and Eq. (5), it can be inferred that the carrier concentration increases when the plasmonic absorption peak has blue shift. The free carrier concentration N, according to Eq. (5), can be rearranged in terms of plasma wavelength λp :

$$N = \frac{{4{\pi ^2}{c^2}{{\varepsilon} _\infty }{{\varepsilon} _0}{m^\ast }}}{{\lambda _\textrm{p}^2{e^2}}}$$

The equation shows the inverse relation between N and square of plasma wavelength λp, which clearly depicts that the plasmonic emission/absorption peak will shift to shorter wavelength with an increased carrier concentration. Therefore, we can observe from Fig. 6 that the emissivity peak gradually shifts from room-temperature absorption peak at λ=1.7µm to 1000°C emissivity peak at λ=1µm. The corresponding bandgap matching to TPV cell changes from GaSb (Eg = 0.72 eV) to Si (Eg = 1.1 eV). This strengthens our emitter structure because Si-based photovoltaics are cheap and mature in technology as compared to GaSb based TPV system. As for the physical reasons for increased carrier concentration at elevated temperature, it can be explained by dopant ionization phenomenon, i.e., incomplete dopant ionization at room temperature. In addition, the increased intrinsic carrier concentration at high temperature due to Fermi Dirac distribution and bandgap narrowing can also lead to increased carrier concentration. In general, it is a common phenomenon that carrier concentration is increased with increased temperature in semiconductors. Since there is very few literature on the n-k values, material properties, or the bandstructure calculations of ITO at 1000°C, it can be difficult to judge at this point which physical mechanisms are the dominant one leading to the increased carrier concentration at 1000°C, and more research on ITO high-temperature material property is required. From room temperature measurement in Fig. 4 it can be seen that samples w/ or w/o stainless steel backplate all lead to selective emission with long-wavelength suppression. It is also observed in room temperature data that there is residual transmittance at short wavelengths (<2µm). While this residual transmittance is beneficial for visible, transparent TPV emitters, it can be somehow problematic in high-temperature emissivity measurement. This is because in the high-temperature emissivity measurement, the apparatus is different from room temperature FTIR absorptivity measurement. In emissivity measurement, photons are emitted from the sample, and there is no external light source. Without a polished stainless steel backplate, the sample is placed on a rough, unpolished stainless steel sample stage. This can lead to lower emissivity since the random light scattering by the unpolished sample stage leads to radiation loss. As a result, we use a polished stainless steel backplate to eliminate the effect of the sample stage. The polished stainless steel backplate enhance the thermal emission, as can be seen in Fig. 6.

In order to have a sustainable emitter structure, the reproducibility of the emissivity measurement is conducted to ensure the feasibility of the SiO2/ITO/sapphire structure. To achieve this, the new emitter sample is fabricated and placed in an air oven and heated up to 1000°C for 6 hours in ambience under atmospheric pressure, and then the temperature is lowered to the room temperature. This procedure is repeated three times, and then the sample is placed back to the vacuum chamber and heated again to 1000°C to perform the high-temperature emissivity measurement. The resultant emissivity of the new emitter sample is shown in Fig. 7(a). It is evident from Fig. 7(a) and (b) that 1000°C heating of the emitter sample in the ambient environment has minimal effect on the spectral emissivity of the ITO/sapphire emitter, and the measured emissivity at 1000°C is stable and the will not change with further heating.

 figure: Fig. 7.

Fig. 7. Detailed emissivity measurement of ITO/sapphire emitter sample with stainless steel at normal emission angle at different temperatures. (a) The measurement taken after heating the emitter sample in the ambience for 6 hours and cooling down to room temperature, this step is repeated 3 times. (b) is the same as Fig. 6(b). The shaded zone indicates a noisy portion where data is not usable.

Download Full Size | PDF

As the final note, it worth pointing out that in our vacuum sample furnace measurement, the effect of vacuum (∼1 torr) annealing and air oven annealing is almost the same. This is because the pressure of our vacuum sample furnace is ∼1 torr, which is low vacuum, and thus there is significant oxygen residual inside it. The reason to pump down the sample furnace to 1 torr is to reduce noise, especially at 1000°C measurement, but we do not need ultra-high vacuum for this purpose.

Conclusively, it can also be observed in our sample that the high-temperature emissivity measurement data is quite consistent with room temperature FTIR absorption measurement, validating the Kirchhoff’s law of thermal radiation. A slight change in enhanced peak emissivity can be observed due to the difference between material properties, i.e., dielectric responses, at high temperatures and room temperature. Therefore, our emitter structure is appropriate and satisfies all of the critical properties that are expected from an ideal TPV emitter. It is worth to mention that because of the planar ITO structure in this work, there are no surface plasmonic effects and has an ideal spectral emission at far-field because of the bulk dielectric response of the ITO. Finally, having a planar design for device eliminates the use of lithography and etching process, this infuses the advantage of large emitter area scaling, lower cost, and the prevention of nanostructure deformation at high temperature.

4. Conclusion

In this work, the experimental work demonstrated the spectral characteristics of the planar ITO/sapphire emitter sample using mainly the bulk dielectric response of ITO. High emissivity (>80%) can be observed in the wavelength range of 1-1.6µm at temperature 800-1000°C. This emission range resembles well with the bandgap of Si or GaSb TPV cells. Besides cut-off, highly suppressed emissivity at long wavelengths can also be observed in this work due to the innate bulk dielectric response of the ITO layer at long wavelengths. This, in turn, enables less energy wastage by mid-IR heat and thus enhances the TPV system performance. In addition, SiO2 is deposited on top of the ITO to prevent oxidation at high temperatures [50]. Additionally, ITO deposition on high thermal conductivity substrate, such as sapphire, proved to have improved optical lifetime performance and damage thresholds [44]. The sapphire substrate has decent interface stability with ITO at high temperatures, and thus at ITO/sapphire structure is beneficial for high-temperature TPV applications. Due to the above-mentioned properties, our planar ITO-sapphire selective emitter with the top SiO2 thin layer eradicates the fast oxidation of TPV emitters and the deformation of the photonic nanostructures at high temperatures. This facilitates the use of ITO-sapphire emitter in ambient conditions instead of in vacuum environments.

Funding

Ministry of Science and Technology, Taiwan (MOST 106-2628-E-007-006-MY3, MOST 106-2628-E-009 -010 -MY3, MOST 108 2218 E 007 009).

Acknowledgements

We acknowledge the assistance from Yen-liang Tu, Guan-wen Lai, and Tzu-hsiang Huang in experiments.

Disclosures

The authors declare no conflicts of interest.

References

1. R. E. Nelson, “A brief history of thermophotovoltaic development,” Semicond. Sci. Technol.18(5), S141–S143 (2003). [CrossRef]  

2. H. Daneshvar, R. Prinja, and N. P. Kherani, “Thermophotovoltaics: fundamentals, challenges and prospects,” Appl. Energy 159, 560–575 (2015). [CrossRef]  

3. P. A. Davies and A. Luque, “Solar thermophotovoltaics: brief review and a new look,” Sol. Energy Mater. Sol. Cells 33(1), 11–22 (1994). [CrossRef]  

4. W. R. Chan, P. Bermel, R. C. N. Pilawa-Podgurski, C. H. Marton, K. F. Jensen, J. J. Senkevich, J. D. Joannopoulos, M. Soljačić, and I. Celanovic, “Toward high-energy-density, high-efficiency, and moderate-temperature chip-scale thermophotovoltaics,” Proc. Natl. Acad. Sci. 110(14), 5309–5314 (2013). [CrossRef]  

5. W. Shockley and H. J. Queisser, “Detailed balance limit of efficiency of p-n junction solar cells,” J. Appl. Phys. 32(3), 510–519 (1961). [CrossRef]  

6. Z. Zhou, Q. Chen, and P. Bermel, “Prospects for high-performance thermophotovoltaic conversion efficiencies exceeding the Shockley–Queisser limit,” Energy Convers. Manage. 97, 63–69 (2015). [CrossRef]  

7. V. Andreev, A. Vlasov, V. Khvostikov, O. Khvostikova, P. Gazaryan, S. Sorokina, and N. Sadchikov, “Solar thermophotovoltaic converters based on tungsten emitters,” J. Sol. Energy Eng. 129(3), 298–303 (2007). [CrossRef]  

8. A. Heinzel, J. Luther, G. Stollwerck, and M. Zenker, “Efficiency and power density potential of thermophotovoltaic systems using low bandgap photovoltaic cells,” AIP Conf. Proc. 460, 103–112 (1999). [CrossRef]  

9. B. Wernsman, R. G. Mahorter, R. Siergiej, S. D. Link, R. J. Wehrer, S. J. Belanger, P. Fourspring, S. Murray, F. Newman, D. Taylor, and T. Rahmlow, “Advanced thermophotovoltaic devices for space nuclear power systems,” AIP Conference Proceedings746, 1441–1448 (2005).

10. T. J. Coutts, “A review of progress in thermophotovoltaic generation of electricity,” Renewable Sustainable Energy Rev. 3(2-3), 77–184 (1999). [CrossRef]  

11. D. M. Bierman, A. Lenert, W. R. Chan, B. Bhatia, I. Celanović, M. Soljačić, and E. N. Wang, “Enhanced photovoltaic energy conversion using thermally based spectral shaping,” Nat. Energy 1(6), 16068 (2016). [CrossRef]  

12. S. Molesky, C. J. Dewalt, and Z. Jacob, “High temperature epsilon-near-zero and epsilon-near-pole metamaterial emitters for thermophotovoltaics,” Opt. Express 21(S1), A96–A110 (2013). [CrossRef]  

13. P. Würfel and U. Würfel, Physics of Solar Cells: From Basic Principles to Advanced Concepts (John Wiley & Sons, 2016).

14. C. Ferrari, F. Melino, M. Pinelli, P. R. Spina, and M. Venturini, “Overview and status of thermophotovoltaic systems,” Energy Procedia 45, 160–169 (2014). [CrossRef]  

15. B. Bitnar, W. Durisch, J. C. Mayor, H. Sigg, and H. R. Tschudi, “Characterisation of rare earth selective emitters for thermophotovoltaic applications,” Sol. Energy Mater. Sol. Cells 73(3), 221–234 (2002). [CrossRef]  

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

17. 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]  

18. S. Lin, J. Moreno, and J. Fleming, “Three-dimensional photonic-crystal emitter for thermal photovoltaic power generation,” Appl. Phys. Lett. 83(2), 380–382 (2003). [CrossRef]  

19. T. Inoue, M. D. Zoysa, T. Asano, and S. Noda, “Single-peak narrow-bandwidth mid-infrared thermal emitters based on quantum wells and photonic crystals,” Appl. Phys. Lett. 102(19), 191110 (2013). [CrossRef]  

20. Z. Wang, T. S. Luk, Y. Tan, D. Ji, M. Zhou, Q. Gan, and Z. Yu, “Tunneling-enabled spectrally selective thermal emitter based on flat metallic films,” Appl. Phys. Lett. 106(10), 101104 (2015). [CrossRef]  

21. Z. Zhou, O. Yehia, and P. Bermel, “Integrated photonic crystal selective emitter for thermophotovoltaics,” J. Nanophotonics 10(1), 016014 (2016). [CrossRef]  

22. M. Chirumamilla, G. V. Krishnamurthy, K. Knopp, T. Krekeler, M. Graf, D. Jalas, M. Ritter, M. Störmer, A. Y. Petrov, and M. Eich, “Metamaterial emitter for thermophotovoltaics stable up to 1400 °C,” Sci. Rep. 9(1), 7241 (2019). [CrossRef]  

23. R. Biswas, D. Zhou, I. Puscasu, E. Johnson, A. Taylor, and W. Zhao, “Sharp thermal emission and absorption from conformally coated metallic photonic crystal with triangular lattice,” Appl. Phys. Lett. 93(6), 063307 (2008). [CrossRef]  

24. P. P. D.-H. Wu, H.-M. Chou, Y.-S. Lin, Y.-C. Lai, P. Yu, Y.-B. Chen, and A. Lin, “A planar indium-tin-oxide thermophotovoltaic emitter with high-temperature verification,” in 35th European Photovoltaic Solar Energy Conference and Exhibition (2018), pp. 92–95.

25. M. Farhat, T.-C. Cheng, K. Q. Le, M. M.-C. Cheng, H. Bağcı, and P.-Y. Chen, “Mirror-backed dark alumina: a nearly perfect absorber for thermoelectronics and thermophotovotaics,” Sci. Rep. 6(1), 19984 (2016). [CrossRef]  

26. J. K. Tong, W.-C. Hsu, Y. Huang, S. V. Boriskina, and G. Chen, “Thin-film ‘thermal well’ emitters and absorbers for high-efficiency thermophotovoltaics,” Sci. Rep. 5(1), 10661 (2015). [CrossRef]  

27. Y. Nam, Y. X. Yeng, A. Lenert, P. Bermel, I. Celanovic, M. Soljačić, and E. N. Wang, “Solar thermophotovoltaic energy conversion systems with two-dimensional tantalum photonic crystal absorbers and emitters,” Sol. Energy Mater. Sol. Cells 122, 287–296 (2014). [CrossRef]  

28. A. Gombert, “An overview Of TPV emitter technologies,” AIP Conference Proceedings653, 123–131 (2003).

29. S. M. Fu, Y. K. Zhong, M. H. Tu, B. R. Chen, and A. Lin, “A planarized thermophotovoltaic emitter with idealized selective emission,” IEEE Photonics J. 8(4), 1–9 (2016). [CrossRef]  

30. C. Wu, B. Neuner Iii, J. John, A. Milder, B. Zollars, S. Savoy, and G. Shvets, “Metamaterial-based integrated plasmonic absorber/emitter for solar thermo-photovoltaic systems,” J. Opt. 14(2), 024005 (2012). [CrossRef]  

31. E. S. Sakr, Z. Zhou, and P. Bermel, “High efficiency rare-earth emitter for thermophotovoltaic applications,” Appl. Phys. Lett. 105(11), 111107 (2014). [CrossRef]  

32. E. M. Wong, J. P. Hickey, G. A. Holmquist, P. N. Uppal, and C. H. Waldman, “Multiband spectral emitters matched to MBE grown photovoltaic cells,” AIP Conference Proceedings358, 278–289 (1996).

33. V. Rinnerbauer, A. Lenert, D. M. Bierman, Y. X. Yeng, W. R. Chan, R. D. Geil, J. J. Senkevich, J. D. Joannopoulos, E. N. Wang, M. Soljačić, and I. Celanovic, “Metallic photonic crystal absorber-emitter for efficient spectral control in high-temperature solar thermophotovoltaics,” Adv. Energy Mater. 4(12), 1400334 (2014). [CrossRef]  

34. A. Lenert, D. M. Bierman, Y. Nam, W. R. Chan, I. Celanović, M. Soljačić, and E. N. Wang, “A nanophotonic solar thermophotovoltaic device,” Nat. Nanotechnol. 9(2), 126–130 (2014). [CrossRef]  

35. S. A. Maier, “Plasmonics: Fundamentals And Applications,” (2007), p. 245.

36. D. Woolf, J. Hensley, J. G. Cederberg, D. T. Bethke, A. D. Grine, and E. A. Shaner, “Heterogeneous metasurface for high temperature selective emission,” Appl. Phys. Lett. 105(8), 081110 (2014). [CrossRef]  

37. J. Xu, Z. Yang, H. Wang, H. Xu, and X. Zhang, “Effect of growth temperature and coating cycles on structural, electrical, optical properties and stability of ITO films deposited by magnetron sputtering,” Mater. Sci. Semicond. Process. 21, 104–110 (2014). [CrossRef]  

38. W. Tang, Y. Chao, X. Weng, L. Deng, and K. Xu, “Optical property and the relationship between resistivity and surface roughness of indium tin oxide thin films,” Phys. Procedia 32, 680–686 (2012). [CrossRef]  

39. A. Pokaipisit, M. Horprathum, and P. Limsuwan, “Vacuum and air annealing effects on properties of indium tin oxide films prepared by ion-assisted electron beam evaporation,” Jpn. J. Appl. Phys. 47(6), 4692–4695 (2008). [CrossRef]  

40. H. Mohamed, “Effect of substrate temperature on physical properties of In2O3 :Sn films deposited by e-beam technique,” International Journal of the Physical Sciences 7, (2012).

41. P. Troyan, Y. Zhidik, and E. Zhidik, “Investigation of temperature stability of ITO films characteristics,” MATEC Web of Conferences143, 03010 (2018).

42. Y.-N. Kim, H.-G. Shin, J.-K. Song, D.-H. Cho, H.-S. Lee, and Y.-G. Jung, “Thermal degradation behavior of indium tin oxide thin films deposited by radio frequency magnetron sputtering,” J. Mater. Res. 20(6), 1574–1579 (2005). [CrossRef]  

43. Y. Liu, W. Ren, P. Shi, D. Liu, M. Liu, W. Jing, B. Tian, Z. Ye, and Z. Jiang, “Preparation and thermal volatility characteristics of In2O3/ITO thin film thermocouple by RF magnetron sputtering,” AIP Adv. 7(11), 115025 (2017). [CrossRef]  

44. J.-H. Yoo, M. Matthews, P. Ramsey, A. C. Barrios, A. Carter, A. Lange, J. Bude, and S. Elhadj, “Thermally ruggedized ITO transparent electrode films for high power optoelectronics,” Opt. Express 25(21), 25533–25545 (2017). [CrossRef]  

45. O. J. Gregory, Q. Luo, and E. E. Crisman, “High temperature stability of indium tin oxide thin films,” Thin Solid Films 406(1-2), 286–293 (2002). [CrossRef]  

46. 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]  

47. D. Peykov, Y. X. Yeng, I. Celanovic, J. D. Joannopoulos, and C. A. Schuh, “Effects of surface diffusion on high temperature selective emitters,” Opt. Express 23(8), 9979–9993 (2015). [CrossRef]  

48. M. Shimizu, A. Kohiyama, and H. Yugami, “Evaluation of thermal stability in spectrally selective few-layer metallo-dielectric structures for solar thermophotovoltaics,” J. Quant. Spectrosc. Radiat. Transfer 212, 45–49 (2018). [CrossRef]  

49. Y.-H. Li, P. Parashar, K.-H. Pen, P. C. Yu, and A. Lin, “A thermophotovoltaic (TPV) micro-combustor using selective emitters,” in 36th European Photovoltaic Solar Energy Conference and Exhibition (Marseille Chanot Convention and Exhibition Centre,2019), pp. 13–17.

50. S. Y. Joo, C. Loka, Y. W. Jo, M. Reddyprakash, S. W. Moon, Y. Choi, S. E. Lee, G. S. Cho, and K.-S. Lee, “ITO/SiO2/ITO structure on a sapphire substrate using the oxidation of ultra-thin Si films as an insulating layer for one-glass-solution capacitive touch-screen panels,” Coatings 10(2), 134 (2020). [CrossRef]  

51. D.-H. Wu, “Construction of numerical model for obtaining optical constants and development of automatic measurement system for high temperature directional emissivity,” (National Tsing-Hua University, 2018).

52. Y.-S. Lin, “Reliability study of TPV emitter with SiO2 passivation,” in Institute of Electronics Engieering (National Chiao-Tung University, 2018).

53. H.-M. Chou, “A study of thermophotovoltaic (TPV) emitter and integrating filter to increase efficiency,” in Institute of Electronics Engineering (National Chiao-Tung University, 2018).

54. T. J. Bright, L. P. Wang, and Z. M. Zhang, “Performance of near-field thermophotovoltaic cells enhanced with a backside reflector,” J. Heat Transfer 136(6), 062701 (2014). [CrossRef]  

55. A. Karalis and J. D. Joannopoulos, “Transparent and ‘opaque’ conducting electrodes for ultra-thin highly-efficient near-field thermophotovoltaic cells,” Sci. Rep. 7(1), 14046 (2017). [CrossRef]  

56. R. S. DiMatteo, P. Greiff, S. L. Finberg, K. A. Young-Waithe, H. K. H. Choy, M. M. Masaki, and C. G. Fonstad, “Enhanced photogeneration of carriers in a semiconductor via coupling across a nonisothermal nanoscale vacuum gap,” Appl. Phys. Lett. 79(12), 1894–1896 (2001). [CrossRef]  

57. O. Ilic, M. Jablan, J. D. Joannopoulos, I. Celanovic, and M. Soljačić, “Overcoming the black body limit in plasmonic and graphene near-field thermophotovoltaic systems,” Opt. Express 20(S3), A366–A384 (2012). [CrossRef]  

58. M. Laroche, R. Carminati, and J.-J. Greffet, “Near-field thermophotovoltaic energy conversion,” J. Appl. Phys. 100(6), 063704 (2006). [CrossRef]  

59. T. D. Dao, A. T. Doan, D. H. Ngo, K. Chen, S. Ishii, A. Tamanai, and T. Nagao, “Selective thermal emitters with infrared plasmonic indium tin oxide working in the atmosphere,” Opt. Mater. Express 9(6), 2534–2544 (2019). [CrossRef]  

60. L. Hu, A. Narayanaswamy, X. Chen, and G. Chen, “Near-field thermal radiation between two closely spaced glass plates exceeding Planck’s blackbody radiation law,” Appl. Phys. Lett. 92(13), 133106 (2008). [CrossRef]  

61. Y. Fan, Y. Suzuki, and N. Kasagi, “Experimental study of micro-scale premixed flame in quartz channels,” Proc. Combust. Inst. 32(2), 3083–3090 (2009). [CrossRef]  

62. V. Rinnerbauer, Y. X. Yeng, W. R. Chan, J. J. Senkevich, J. D. Joannopoulos, M. Soljačić, and I. Celanovic, “High-temperature stability and selective thermal emission of polycrystalline tantalum photonic crystals,” Opt. Express 21(9), 11482–11491 (2013). [CrossRef]  

63. C. Schlemmer, J. Aschaber, V. Boerner, and J. Luther, “Thermal stability of micro-structured selective tungsten emitters,” AIP Conf. Proc. 653, 164–173 (2003). [CrossRef]  

64. J.-M. Kim, K.-H. Park, D.-S. Kim, B.-y. Hwang, S.-K. Kim, H.-M. Chae, B.-K. Ju, and Y.-S. Kim, “Design and fabrication of spectrally selective emitter for thermophotovoltaic system by using nano-imprint lithography,” Appl. Surf. Sci. 429, 138–143 (2018). [CrossRef]  

65. V. Stelmakh, W. R. Chan, J. D. Joannopoulos, M. Soljacic, I. Celanovic, and K. Sablon, “Improved thermal emitters for thermophotovoltaic energy conversion,” in ASME 2016 5th International Conference on Micro/Nanoscale Heat and Mass Transfer (2016).

66. M. Shimizu, M. Suzuki, F. Iguchi, and H. Yugami, “High-temperature solar selective absorbers using transparent conductive oxide coated metal,” Energy Procedia 57, 418–426 (2014). [CrossRef]  

67. Y. S. Touloukian, D. P. DeWitt, and Hernicz, “Thermophysical properties of matter,” Volume 9 in the TPRC data series, Thermal radiative properties - coatings (Reannouncement). Data book (1972).

68. 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]  

69. S. M. Fu, Y. K. Zhong, M. H. Tu, B. R. Chen, and A. Lin, “A fully functionalized metamaterial perfect absorber with simple design and implementation,” Sci. Rep. 6(1), 36244 (2016). [CrossRef]  

70. J. Lau and B. Wong, “Indium tin oxide-based selective emitter for nano-gap thermophotovoltaic applications,” J. Nano Res. 49, 127–148 (2017). [CrossRef]  

71. H. Khamh, E. Sachet, K. Kelly, J.-P. Maria, and S. Franzen, “As good as gold and better: conducting metal oxide materials for mid-infrared plasmonic applications,” J. Mater. Chem. C 6(31), 8326–8342 (2018). [CrossRef]  

72. C.-W. Chen, Y.-C. Lin, C.-H. Chang, P. Yu, J.-M. Shieh, and C.-L. Pan, “Frequency-dependent complex conductivities and dielectric responses of indium tin oxide thin films from the visible to the far-infrared,” IEEE J. Quantum Electron. 46(12), 1746–1754 (2010). [CrossRef]  

73. M. Fox, Optical Properties of Solids (OUP Oxford, 2001).

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) The emitter structure in this study with tSiO2 = 50 nm, tITO = 1 µm, tSapphire = 650 µm, tsteel = 0.8 mm, (b) emitter sample without steel before high temperature measurements, and (c) emitter sample without steel after high temperature measurements.
Fig. 2.
Fig. 2. System setup (a) Diagram of the high-temperature emissivity measurement system, (b) 3D design of the sample heater.
Fig. 3.
Fig. 3. (a) Cross-sectional SEM image of emitter sample before annealing along with EDS analysis of (b) oxygen, aluminum, indium, and tin elements, measured at room temperature before heating, (c) Cross-sectional SEM image of emitter sample after annealing along with EDS analysis of (d) oxygen, aluminum, indium, and tin elements, measured at room temperature after emissivity measurement.
Fig. 4.
Fig. 4. Reflectance (R), Transmittance (T), and Absorption (A) measurement of the emitter sample at room temperature using (a) FTIR measurement of emitter sample without stainless steel, and (b) with stainless steel, (c) UV-VIS-NIR spectrophotometer measurement of emitter sample without stainless steel, and (d) with stainless steel.
Fig. 5.
Fig. 5. Detailed emissivity measurement of stainless steel at normal emission angle at different temperatures. The shaded zone indicates a noisy portion where data is not usable.
Fig. 6.
Fig. 6. Detailed emissivity measurement of ITO/sapphire emitter sample (a) without stainless steel at normal emission angle, (b) with stainless steel at normal emission angle, at different temperatures. The shaded zone indicates a noisy portion where data is not usable.
Fig. 7.
Fig. 7. Detailed emissivity measurement of ITO/sapphire emitter sample with stainless steel at normal emission angle at different temperatures. (a) The measurement taken after heating the emitter sample in the ambience for 6 hours and cooling down to room temperature, this step is repeated 3 times. (b) is the same as Fig. 6(b). The shaded zone indicates a noisy portion where data is not usable.

Tables (1)

Tables Icon

Table 1. Settings of high-temperature FTIR.

Equations (8)

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

A = 1 R T
ε ( T , θ ) = I sample,  λ ( T , θ ) I blackbody,  λ ( T )
ε ( T , θ ) = I sample,  λ ( T , θ ) I sur,  λ ( T sur , θ ) I blackbody,  λ ( T ) I sur,  λ ( T sur , θ )
ε c = ε 1 ε 2 = ε ω p 2 ω 2 +  i γ ω
ω p 2 = N e 2 ε 0 m
α f r e e   c a r r i e r = N e 2 m ε 0 n c τ ω 2
I sur, λ ( T sur , θ ) I sample, λ ( T , θ ) 0 .5
N = 4 π 2 c 2 ε ε 0 m λ p 2 e 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.