Tungsten nanowire based hyperbolic metamaterial emitters for near-field thermophotovoltaic applications

Tungsten nanowire based hyperbolic metamaterial emitters for near-field thermophotovoltaic applications

International Journal of Heat and Mass Transfer 87 (2015) 237–247 Contents lists available at ScienceDirect International Journal of Heat and Mass T...

3MB Sizes 2 Downloads 60 Views

International Journal of Heat and Mass Transfer 87 (2015) 237–247

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer journal homepage: www.elsevier.com/locate/ijhmt

Tungsten nanowire based hyperbolic metamaterial emitters for near-field thermophotovoltaic applications Jui-Yung Chang, Yue Yang, Liping Wang ⇑ School for Engineering of Matter, Transport & Energy, Arizona State University, Tempe, AZ 85287, USA

a r t i c l e

i n f o

Article history: Received 11 October 2014 Received in revised form 23 March 2015 Accepted 30 March 2015

Keywords: Hyperbolic metamaterial Near-field radiation Thermophotovoltaic Effective medium Thin films

a b s t r a c t Recently, near-field radiative heat transfer enhancement across nanometer vacuum gaps has been intensively studied between two hyperbolic metamaterials (HMMs) due to unlimited wavevectors and high photonic density of state. In this work, we theoretically analyze the energy conversion performance of a thermophotovoltaic (TPV) cell made of In0.2Ga0.8Sb when paired with a HMM emitter composed of tungsten nanowire arrays embedded in Al2O3 host at nanometer vacuum gaps. Fluctuational electrodynamics integrated with effective medium theory and anisotropic thin-film optics is used to calculate the near-field radiative heat transfer. It is found that the spectral radiative energy is enhanced by the epsilon-near-pole and hyperbolic modes at different polarizations. As a result, the power output from a semi-infinite TPV cell is improved by 2.15 times with the nanowire HMM emitter over that with a plain tungsten emitter at a vacuum gap of 20 nm. Moreover, by using a thin TPV cell with 10 lm thickness, the conversion efficiency can be greatly improved from 17.7% to 31.1% without affecting the power generation, due to the total internal reflection occurring at the bottom cell interface that minimizes the subbandgap spectral radiative energy. Furthermore, the effects of a TPV cell and a nanowire emitter with finite thicknesses are also studied. The result shows that the maximum efficiency of 31.8% is achieved with an optimal cell thickness of 3 lm while the nanowire HMM emitter should be thick enough to be opaque. The fundamental understanding and insights obtained here will facilitate the design and application of novel materials in enhancing near-field TPV energy conversion. Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction In the past few years, near-field thermal radiation has been demonstrated to exceed the Planck’s blackbody limit by orders of magnitude [1–5]. The enhanced radiative heat transfer across nanoscale vacuum gaps has great potential in improving power generation for thermophotovoltaic (TPV) applications [6–11]. Coupled surface plasmon polariton (SPP) and surface phonon polariton (SPhP) have been proven to be the dominating resonance enhancement mechanisms in near-field thermal radiation [1,2]. SPP or SPhP coupling across the vacuum gap requires similar materials for the emitter and receiver that support SPP or SPhP at close frequencies such that the coupling strength is strong and resulting near-field heat transfer could be greatly enhanced. However, the TPV receivers are usually photovoltaic cells made of semiconductors, while the TPV emitters require high-temperature materials such as tungsten. Therefore, it has been a daunting challenge to

⇑ Corresponding author. Tel.: +1 480 727 8615. E-mail address: [email protected] (L. Wang). http://dx.doi.org/10.1016/j.ijheatmasstransfer.2015.03.087 0017-9310/Ó 2015 Elsevier Ltd. All rights reserved.

enhance near-field TPV performance with coupled SPPs across the nanometer vacuum gap due to the inherent material mismatch. Recently, graphene has been proposed to enhance the performance of TPV cells [12,13]. Other physical mechanisms are highly desired for improving near-field TPV energy conversion. Hyperbolic metamaterials (HMMs), such as metallic nanowires embedded in dielectric materials [14,15], have shown the potential in numerous applications for far-field radiation [16–21], due to the unlimited length of propagating wave vector (i.e., high-k mode) and broadband enhancement on photonic density of state (PDOS) [16,17,22–28] in addition to the epsilon-near-zero (ENZ) and epsilon-near-pole (ENP) modes [29]. Recently, nanowire-based HMMs have also been intensively studied for enhancing near-field thermal radiation by several times or even orders of magnitude over that with bulk materials [15,22–24,30–34]. The broadband resonance-free enhancement with HMMs in near-field radiative transport suggests their great potential in improving near-field TPV energy conversion. However, most of previous studies considered near-field radiative transfer between two identical HMMs across the vacuum gap, while the near-field TPV emitter and cell are usually made of dissimilar materials. A detailed analysis and

238

J.-Y. Chang et al. / International Journal of Heat and Mass Transfer 87 (2015) 237–247

Nomenclature a c D d E e f g h J0 Jph kB kx,y,z PEl PR q00 q00x qk r T Voc

period of structure, m light velocity in vacuum, 2.998  108 m s1 diameter of nanowire, m gap distance, m electrical field electric charge, C filling ratio depolarization factor Planck constant dark current, A short circuit current, A Boltzmann constant, J/K x, y, z component of wavevector electrical power, W radiative power, W heat flux, W/m2 spectral heat flux, W/(m2 rad/s) spectral heat flux, W/(m2 m) reflection coefficient temperature, K open circuit voltage, V

Greek symbols b parallel component of wavevector c vertical component of wavevector e dielectric function x angular frequency, rad/s H Planck oscillator n transmission coefficient

understanding of a HMM emitter on the conversion performance of a near-field TPV cell is still lacking. The present work studies the effect of a HMM emitter made of a tungsten nanowire array embedded in aluminum oxide (Al2O3) on the near-field TPV energy conversion. The influence of the volumetric filling ratio of the HMM emitter and the vacuum gap distance on near-field radiative transfer will be thoroughly investigated on the radiative transport, power generation and conversion efficiency. Possible enhancement on the near-field radiation due to the ENP and hyperbolic modes of the HMM emitter will be investigated. Moreover, the effect of a thin-film TPV cell with finite thicknesses will be studied aiming to further improve the conversion efficiency. Finally, the thickness effects of both a thin film TPV cell and a HMM emitter with vacuum substrate on both power output and conversion efficiency will be explored to optimize the cell performance.

h k

g gq

reduced Planck constant wavelength, m conversion efficiency quantum efficiency

Subscripts 0, 1, 2, 3, 4 layer 0, 1, 2, 3, 4 \ vertical component k parallel component A acceptor Al2O3 aluminum oxide D donor El electrical e electron eff effective medium evan evanescent wave h hole i, j medium i, j in intrinsic prop propagating wave R radiative W tungsten k spectral Superscripts a state of polarization s s polarization p p polarization

fabrication for further experimental demonstration. The Al2O3 nanohole array can be first fabricated in large areas by anodic aluminum oxide (AAO) with pore diameters down to 15 nm [36]. Tungsten nanowires can be then grown in the nanoholes by different electrodepositing methods depending on the geometric

2. Theoretical methods 2.1. Effective medium theory Fig. 1 shows the configuration of the near-field TPV system to be investigated in this study, which consists of a HMM emitter (Layer 1) made of tungsten nanowire arrays embedded in the Al2O3 host, and a TPV receiver (Layer 3) made of In0.2Ga0.8Sb [35] separated by a vacuum gap (Layer 2) with distance d. The HMM emitter and the TPV receiver respectively have a finite thickness h and t, and vacuum is assumed to be the substrates (Layer 0 and Layer 4). The volumetric filling ratio of the nanowire array is f = pD2/(4a2), where a is the period and D is the nanowire diameter, respectively. Here, the materials for the HMM emitter are chosen not only because of the high-temperature stability but also due to the possibility of

Fig. 1. Schematic of a near-field TPV system consisting of a tungsten nanowire based HMM emitter and a TPV cell with finite thicknesses in a 5-layer structure: vacuum substrate (Layer 0), emitter (Layer 1), vacuum gap (Layer 2), receiver (Layer 3), vacuum substrate (Layer 4).

239

J.-Y. Chang et al. / International Journal of Heat and Mass Transfer 87 (2015) 237–247

features [37]. Note that, though AAOs generally have hexagonal unit cells, while square ones can also be fabricated by pre-imprinting the aluminum [38]. Since the HMM emitter is an inhomogeneous medium made of tungsten and Al2O3, effective medium theory (EMT), which is a homogenization approach based on field average, is employed here to obtain effective dielectric functions of the emitter by approximating it as a homogeneous medium. Maxwell–Garnett (MG) method, which approximates the effective dielectric properties by treating one constituent as the embedded filler (i.e., tungsten nanowire) and all other constituents as host (i.e., Al2O3), is applied in this work. Note that the MG approximation assumes that fillers are independent with each other, which is an appropriate assumption for small filling ratios. Therefore, only the HMM emitter with f < 0.5 is investigated in this study. By setting depolarization factors g\ = 0 (when E field is perpendicular to the interface) and gk = 0.5 (when E field is parallel with the interface) for nanowires with high aspect ratios h/D > 20 [39], the MG theory can be expressed as [40,41]:

ek;eff ¼ ed

ðem þ ed Þ þ f ðem  ed Þ ðem þ ed Þ  f ðem  ed Þ

e?;eff ¼ ed þ f ðem  ed Þ

and

ð1Þ ð2Þ

where subscripts m and d represent tungsten (i.e., filler material) and Al2O3 (i.e., host material), respectively. The subscript k and \ denote the ordinary and extraordinary component of dielectric function. Clearly, the effective properties are highly dependent on the filling ratio, filler and host materials. Current nanofabrication techniques could synthesize nanowire arrays with period a on the order of 50 nm and even smaller. A vacuum gap of 20 nm is chosen in this study since the EMT is applicable only if the feature size of the nanowire array is much smaller than the vacuum gap and thermal wavelength, i.e., a < pd and a  kth = ⁄c/kBT (i.e., 1.15 lm for T = 2000 K) [34]. The optical properties of tungsten and alumina are obtained from Palik’s tabular data and assumed to be independent of temperature [42].

After the effective dielectric functions of the HMM emitter are obtained by the MG theory, the emitter can be treated as a homogeneous but uniaxial medium with a dielectric function in a tensor form as:

ek;eff

0

e ¼ 6 4 0 0

0

0

3

7 0 5

ek;eff

2

2

e?;eff

2

þ

kz

ek;eff

¼

Based on thermal stochastic nature of charges or dipoles in media [43], fluctuational electrodynamics can be applied to calculate the near-field radiative transport between objects at different temperatures. Recent work mainly focused on the near-field thermal radiation between isotropic thin films [44] or between semi-infinite uniaxial media [30]. Near-field radiation between two uniaxial thin films, however, is little studied. Here, we extend theoretical framework from Francoeur et al. [45] on the near-field heat transfer between isotropic thin films to that between two uniaxial layers with finite thicknesses, by incorporating anisotropic wave optics. The net near-field heat flux between two uniaxial media with finite thicknesses at different temperatures (T1 = 2000 K for the emitter and T3 = 300 K for the receiver in this study) separated by a vacuum gap of d can be expressed as [1,32,45]:

q00 ¼

Z

1

q00x dx Z 1 Z 1 1 ½  ¼ d x H ð x ; T Þ  H ð x ; T Þ nðx; bÞbdb 1 3 4p2 0 0 0

a Planck oscillator at temperature T. ⁄ and kB denote the reduced Planck constant and the Boltzmann constant, respectively. The energy transmission coefficient n(x, b) has different expressions for propagating waves and evanescent waves as [45]:

   1  jRs1 j2  jT s1 j2 1  jRs3 j2  jT s3 j2

þ

and

j1  Rs1 Rs3 ei2c2 d j2    1  jRp1 j2  jT p1 j2 1  jRp3 j2  jT p3 j2

c

j1  Rp1 Rp3 ei2c2 d j2

nev an ðx; bÞ ¼

ð6aÞ

4ImðRs1 ÞImðRs3 Þe2Imðc2 Þd   1  Rs Rs ei2c2 d 2 1 3

4ImðRp1 ÞImðRp3 Þe2Imðc2 Þd þ   1  Rp Rp ei2c2 d 2

e?;eff

x2

ð5Þ

x where Hðx; TÞ ¼ expðhxh=k represents the spectral mean energy of B TÞ1

ð3Þ

ð6bÞ

1 3

Therefore, with certain conditions the nanowire emitter can exhibit hyperboloid shapes on the isofrequency relation for transverse-magnetic or p-polarized waves:

kx þ ky

2.3. Near-field radiative heat transfer between a HMM emitter and a TPV cell both with finite thicknesses

nprop ðx; bÞ ¼

2.2. Hyperbolic dielectric behavior of tungsten nanowire arrays

2

(PDOS) [27,28]. Both the large wavevector of propagating waves and the high PDOS inside HMMs are beneficial for enhancing near-field radiation heat transfer [14]. Note that though the wavevector is theoretically unbounded inside HMMs, only a limited number of k-modes, whose maximum is dependent on the vacuum gap distance, can be harvested to contribute to the near-field radiative transfer.

ð4Þ

where x is the angular frequency of incident wave, c is the speed of light in vacuum, and kx, ky, and kz are the wavevector components on x, y, and z directions, respectively. Hyperboloid shapes only appear when e?;eff and ek;eff have opposite signs, i.e., ek;eff > 0 and e?;eff < 0 (denoted as type I HMM), or ek;eff < 0 and e?;eff > 0 (denoted as type II HMM). Both types of HMM possess unlimited length of wavevector since hyperboloid shapes are unbounded, and thus can support high-k mode, which allows propagating waves with large wavevectors [14]. Furthermore, this high-k mode also supports a broadband singularity in photonic density of states

where

and

Raj ¼

T aj ¼

r aj1;j þ r aj;jþ1 ei2cj tj 1 þ raj1;j r aj;jþ1 ei2cj tj t aj1;j taj;jþ1 eicj tj

1 þ raj1;j r aj;jþ1 ei2cj tj

ð7aÞ

ð7bÞ

Here, the subscript number j = 0, 1, 2, 3, or 4 is the layer index shown in Fig. 1, which represents the vacuum substrate, emitter, vacuum gap, TPV receiver, and vacuum substrate, respectively. a = s or p denotes the polarization states. rsj1;j ¼ ðcj1  cj Þ= ðcj1 þ cj Þ and r pj1;j ¼ ðejk cj1  ej1;k cj Þ=ðejk cj1 þ ej1;k cj Þ are the Fresnel reflection coefficients for s and p polarizations at the interface of layers j  1 and j, respectively. cj is the z-component wavevector in an uniaxial layer j and can be expressed differently for s and p polarizations [30,34]:

240

J.-Y. Chang et al. / International Journal of Heat and Mass Transfer 87 (2015) 237–247

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi csj ¼ ejk x2 =c2  b2

ð8Þ

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi p j

c ¼

and

ejk x2 =c2 

ejk 2 b ej?

ð9Þ

where b is the parallel component of wavevector kj, and is identical in different layers due to the continuity boundary condition at the interfaces. When the layer is isotropic, cpj ¼ csj is held with ejk ¼ ej? : Note that when the thicknesses of emitter and receiver become infinitely large, the structure reduces to a three-layer configuration with a half-space emitter (Layer 1), vacuum (Layer 2), and a halfspace receiver (Layer 3). In this case, n(x, b) in Eq. (6) can be simplified through replacing the reflection coefficient of a thin film (Raj ) by that at an interface (r aj;j1 or raj;jþ1 ) and setting T aj = 0 [30]. In addition, the near-field radiative transfer in a four-layer thermal rectifier made of vacuum, a thin-film emitter, vacuum gap, and a half-space receiver has been investigated with a similar approach [46].

Here, h is the Planck constant, nin is the intrinsic carrier concentration, ND, and NA are the carrier concentration of donor and acceptor, respectively. Dh, De, sh, and se are the diffusion coefficient and relaxation time of holes and electrons, respectively. Moreover, the opencircuit voltage is given by:

V oc ¼ ðkB T=eÞ lnðJ ph =J 0 þ 1Þ

ð12Þ

Therefore, the electrical power output can be expressed as [9]:

PEl ¼ J ph V oc ð1  1=yÞ½1  lnðyÞ=y

ð13Þ

where

ð14Þ

y ¼ lnðJ ph =J 0 Þ

Finally, by dividing the electrical power output with the total radiative power input, the conversion efficiency of the proposed near-field TPV system with a HMM emitter is [9]:

g ¼ PEl =PR

ð15Þ

2.4. TPV power generation and conversion efficiency Photons with spectral energy above the cell bandgap, i.e., Eg = 0.56 eV (or equivalently 2.2 lm in wavelength) for In0.2Ga0.8Sb, will be absorbed by the TPV cell to generate electron–hole pairs. Electrical currents could be produced with external loads after the photon-generated charges are collected. However, the photons with lower energy cannot excite electron hole pairs and result in low conversion efficiency. The near-field radiative transfer and the charge transport inside the TPV cell is a coupled problem, which has been theoretically modeled for multilayer isotropic media by calculating the charge density distribution due to the number of photons absorbed at different cell location [9]. In fact, the quantum efficiency in response to the near-field thermal radiation is different from that to a far-field source because of the inherent difference between near-field and far-field thermal radiation spectra. The small penetration depth of evanescent waves, especially at resonance modes, significantly affects the photon absorption at different depths away from the cell surface, which would lead to modified quantum efficiency at different frequencies in the near field. However, due to the limitation of the present theoretical model, which cannot predict the spectral energy absorbed at different depths by treating the TPV cell as a multilayer, the quantum efficiency to the near-field thermal radiation between the uniaxial HMM emitter and the TPV cell cannot be thus obtained accurately. Instead of assuming a perfect TPV cell with 100% quantum efficiency, which will overpredict the performance, we use the far-field internal quantum efficiency of In0.2Ga0.8Sb [35] for evaluating the proposed near-field TPV system with a nanowire-based HMM emitter, aiming for more practical insights. We also neglected the thermal impact on the overall TPV performance by simply assuming the nanowire emitter and TPV at respective thermal equilibrium with uniform temperature distribution. For the sake of completeness, the methods for calculating the power generation and conversion efficiency are summarized here. With the near-field spectral heat flux qk(k) calculated from Eq. (5) and the far-field internal quantum efficiency data gq(k), the short-circuit current can be obtained as [9]:

J ph ¼

e hc

Z 0

1

gq ðkÞqk ðkÞkdk

ð10Þ

The dark current can be calculated by [9]:

J0 ¼ e



 n2in Dh n2 De pffiffiffiffiffi þ inpffiffiffiffiffi ND sh N A se

ð11Þ

3. Results and discussion 3.1. Effective dielectric functions of the hyperbolic metamaterial emitter Fig. 2(a) and (b) show the real part of effective dielectric functions for parallel and vertical components of the HMM emitter, respectively. Clearly the tungsten nanowire array behaves as the type I HMM within frequencies at which e?;eff is negative while and ek;eff is always positive within the considered spectral regime. Furthermore, the filling ratio effect on effective dielectric functions can also be observed. Fig. 2(a) illustrates that, with increasing filling ratios the ENP peak for parallel effective dielectric function becomes larger, which indicates stronger ENP enhancement under large filling ratios. On the other hand, Fig. 2(b) demonstrates that, the characteristic frequency at which the vertical effective dielectric function becomes negative will shift to higher frequencies with larger filling ratios. Therefore, since the parallel effective dielectric function is always positive under all filling ratios throughout the whole spectrum, higher filling ratio will result in type I HMM region shifting to higher frequency. Note that at the ENP region shown in Fig. 2(a), plain tungsten has metallic behavior which does not support ENP enhancement. Fig. 2(c) and (d) show the imaginary part of parallel and vertical effective dielectric functions of the HMM emitter, respectively. The effect of filling ratio on the ENP enhancement can also be observed in Fig. 2(c) with higher ENP peak. Similar effect of filling ratio can also be seen in Fig. 2(d) on the vertical effective dielectric function which becomes more lossy at higher filling ratio. Note that all four figures characterize the effective dielectric functions of both parallel and vertical components of the tungsten nanowire based HMM for both s and p polarizations. For s polarization in which only the parallel component dielectric function takes effect, the underlying mechanism is the ENP mode, which would result in the reduction of impedance mismatch and thus higher energy transfer [29]. On the other hand, the HMM behavior will appear at p polarization. The angular frequency region and wavelength band of the type I HMM behavior can be classified by optical phase diagram with respect to different filling ratios as shown in Fig. 2(e). Clearly, the region of type I HMM extends to smaller wavelengths (i.e., higher frequencies) with larger filling ratios. Note that, the tungsten nanowire arrays behave as an effective dielectric with both positive ek;eff and e?;eff outside of the HMM regime.

J.-Y. Chang et al. / International Journal of Heat and Mass Transfer 87 (2015) 237–247

241

Fig. 2. Dielectric functions of the tungsten nanowire HMM with respect to different filling ratios: (a) real part of ek; (b) real part of e\; (c) imaginary part of ek; (d) imaginary part of e\. (e) Optical phase diagram of the tungsten nanowire based HMM emitter with respect to angular frequency or wavelength.

3.2. TPV performance with a semi-infinite nanowire emitter and a semi-infinite PV cell Let us first consider a near-field TPV system made of a semi-infinite tungsten nanowire HMM emitter and a semi-infinite TPV cell. At a vacuum gap of d = 20 nm, the effect of filling ratio on the transmission coefficients, spectral heat flux, power output, and conversion efficiency will be studied subsequently, followed by the effect of vacuum gap distances on the TPV performance. All the calculations in this and following sections are completed with sufficient amount of data points to ensure the accuracy in numerical integration: 1000 data points in the frequency domain

of 3:7  1014 6 x 6 6:2  1015 rad=s, 1000 points for propagating waves with 0 6 bc=x < 1; and 10000 points for evanescent waves with bc=x > 1. The numerical error is checked to be less than 0.1% when the amounts of data points for angular frequency x and dimensionless wavevector bc/x are doubled.

3.2.1. Transmission coefficients at different polarizations Fig. 3(a–c) show the transmission coefficient n(x, b) at s polarization with filling ratio f = 0.1, 0.3, and 0.5, respectively, when the vacuum gap is fixed at d = 20 nm. By comparison, the transmission coefficients in the shaded region are enhanced with more channels

242

J.-Y. Chang et al. / International Journal of Heat and Mass Transfer 87 (2015) 237–247

Fig. 3. n function of the near-field radiative heat transfer from a 3-layer near-field TPV device with the nanowire HMM emitter for s polarization with filling ratios of: (a) 0.1; (b) 0.3; (c) 0.5 and for p polarization with filling ratios of: (d) 0.1; (e) 0.3; (f) 0.5. The vacuum gap distance is d = 20 nm.

of heat transfer (i.e., bc/x) at larger filling ratios, which can be explained by the stronger ENP mode at the same frequency region shown in Fig. 2(a) and (c). Note that, only the parallel component of dielectric function of the HMM emitter ek;eff is involved with s-polarized waves. When the filling ratio increases, higher absorption loss (or emission) takes place at the ENP mode for ek;eff , and results in larger radiative transfer or transmission coefficients from the emitter to the receiver. Fig. 3(d–f) present the transmission coefficient n(x, b) at p polarization with f = 0.1, 0.3, and 0.5, respectively. As the filling ratio increases, the shaded region with large transmission coefficient at low frequencies gets broadened towards higher frequency, which matches well with the type I HMM region as shown in Fig. 2(e). Therefore, the enhancement in n(x, b) at low-frequency region is due to the HMM behavior of the tungsten nanowire emitter. The enhancement at higher frequencies is due to the effective

dielectric behavior of the nanowire emitter, in which the absorption loss becomes greater with larger imaginary parts of both ek;eff and e?;eff with increasing filling ratio, as indicated by Fig. 2(c) and (d). 3.2.2. Spectral radiative heat flux According to Eq. (5), the enhancement of transmission coefficients will result in the enhancement of spectral heat flux. Fig. 4(a) and (b) show the spectral heat fluxes between the semi-infinite nanowire HMM emitter and semi-infinite TPV cell for s and p polarizations, respectively. At s polarization, a steady enhancement on the spectral heat flux appears with the increment of filling ratio. Furthermore, due to the narrow-band ENP enhancement on transmission coefficient n(x, b) demonstrated by Fig. 3(a–c), an enhancement within the ENP region also takes place with larger filling ratios. Note that, even though the spectral heat

J.-Y. Chang et al. / International Journal of Heat and Mass Transfer 87 (2015) 237–247

243

plain tungsten for high filling ratios up to f = 0.5. Plain tungsten has negative real part of dielectric function at angular frequency x < 2  1015 rad/s, which indicates metallic behavior. On the other hand, tungsten nanowires possess positive real part of ek;eff in the same frequency region, and behave as lossy dielectrics. As a result, the spectral heat flux from the nanowire emitters is higher than that from plain tungsten at longer wavelengths, as shown in Fig. 4(a). Note that, the spectral peak with plain tungsten at 1.4 lm is due to its intrinsic bandgap absorption. As shown in Fig. 4(b) for p polarization, the spectral heat flux distribution with the HMM emitters exceeds that with plain tungsten at wavelengths k > 1.6 lm or so for all filling ratios. The enhanced spectral heat flux region matches well with the type I HMM region as shown in Fig. 2(e). The spectral peak also shifts slightly towards lower wavelengths with larger filling ratios, which agrees with the trend observed from the enhanced transmission coefficient region in Fig. 3 for p polarization at different f values. Therefore, the enhanced spectral flux with HMM emitters at longer wavelengths is due to the hyperbolic behavior in tungsten nanowires, which do not exist in plain tungsten. Among different filling ratios, the difference in spectral heat flux is not as apparent as that for s polarization. Furthermore, the spectral heat flux drops abruptly at the wavelength k  2 lm for all filling ratios, which does not appear for s polarization. The abrupt reduction in q(k) at p polarization can be explained by the sudden drop of real part of e?;eff in Fig. 2(b), or equivalently by the sharp increase of the imaginary part of e?;eff in Fig. 2(d). Fig. 4(c) shows the overall spectral heat flux from both s and p polarizations for tungsten nanowire HMM emitters with different filling ratios. The maximum spectral heat flux could be as high as 3.6 MW/m2-lm around k = 1.8 lm with f = 0.5. Note that the TPV cell made of In0.2Ga0.8Sb has a bandgap of Eg = 2.2 lm. The spectral heat flux from the HMM emitters with f > 0.3 is larger than that with plain tungsten at the wavelengths above the cell bandgap Eg, suggesting the potential to enhance power generation and conversion efficiency by replacing plain tungsten emitter with tungsten nanowires. However, the spectral heat flux below the cell bandgap Eg is also enhanced with nanowire HMM emitters, indicating more wasted long-wavelength photons which might decrease the conversion efficiency. In order to gain a better understanding on how the nanowire HMM emitters would impact the TPV performance, the power generation and overall photon-toelectricity conversion efficiency need to be evaluated.

Fig. 4. Spectral heat flux from a semi-infinite nanowire HMM emitter to a semiinfinite TPV cell separated by a vacuum gap d = 20 nm: (a) s polarized waves; (b) p polarized waves; (c) overall summarized from both polarized waves.

flux for s polarization rises with f from 0.1 to 0.5, it does not necessarily indicate that the spectral heat flux for plain tungsten emitter (i.e., f = 1) will become larger than that with nanowire emitters, which in fact is lower than that of nanowires with filling ratio f > 0.3. This can be explained by Fig. 2(a) and (c) where the effective material property of nanowire emitters is not similar to

3.2.3. Power output and conversion efficiency Fig. 5(a) shows the normalized radiative power input and electrical power output of the TPV cell with HMM emitters at different filling ratios over that with plain tungsten at the vacuum gap d = 20 nm. Note that, the radiative heat flux and power generation with a plain tungsten emitter are 2.51 MW/m2 and 0.58 MW/m2 respectively, leading to a conversion efficiency of 23.1%. By replacing the plain tungsten emitter with tungsten nanowire HMMs, the electrical power output increases with larger filling ratios, and exceeds that with plain tungsten under all filling ratios. The enhancement in power output could reach about 2.15 times higher with f = 0.5. On the other hand, the total radiative heat flux from the emitters to the TPV cell is also amplified with nanowire HMMs, and increases with filling ratio. This is because the HMM emitter enhances the spectral energy not only above but also below the bandgap at longer wavelengths as the broadband enhancement shown in Fig. 4(c). As a result, the conversion efficiency with nanowire HMM emitters is between 16.1% and 17.7% with different filling ratios, which is about 5.4% to 7% less compared with that (23.1%) from a plain tungsten emitter as shown in Fig. 5(b), though the electrical power output is enhanced more than two times.

244

J.-Y. Chang et al. / International Journal of Heat and Mass Transfer 87 (2015) 237–247

Fig. 5. (a) The electrical power output and the radiative power input of a 3-layer near-field TPV system with the nanowire HMM emitter (f = 0.5) normalized to those with plain tungsten emitter. (b) The conversion efficiency of a 3-layer near-field TPV system with the nanowire HMM emitter (f = 0.5) and a plain tungsten emitter. The vacuum gap distance is d = 20 nm.

3.2.4. Effect of vacuum gap distance The electrical power output and the TPV conversion efficiency are calculated at different vacuum gap distances for the nanowire HMM emitter with f = 0.5, as shown in Fig. 6(a) and (b), respectively. The power generation decreases monotonically when the HMM emitter and the TPV cell are further apart, due to the reduction of near-field radiative heat transfer. However, the conversion efficiency with the HMM emitter drops first from 19.1% to a minimum of 9.4% when the vacuum gap becomes larger from d = 10 nm to 400 nm or so, and then rises up to 15.3% at a 1-lm vacuum gap. The plain tungsten exhibits similar behaviors in both power output and conversion efficiency but with smaller PEl and a larger efficiency g by 4% or so. The different dependence on the vacuum gap between the electrical power and conversion efficiency can be understood by the effect of vacuum gap on the spectral radiative heat flux, as only that above the cell bandgap would enhance the conversion efficiency. When the gap distance is smaller than 400 nm, small gap distances contribute more on the enhancement of heat flux above the bandgap rather than the waste energy. This can be confirmed by Fig. 6(a) that the power output increases rapidly with the reduction of gap distance when d is less than 400 nm. On the other hand, the heat flux enhancement due to small vacuum gap focused more on longer wavelengths (i.e., energy below the band gap) with gap distances larger than

Fig. 6. The vacuum gap effect on: (a) power output; (b) conversion efficiency of a 3layer near-field TPV system with the nanowire HMM emitter (f = 0.5) and a plain tungsten emitter.

400 nm as the electrical power input remains constant and the conversion efficiency drops. In order to make the nanowire HMM emitter more useful in enhancing the near-field TPV conversion efficiency over the plain tungsten, approaches other than reducing gap distance such as considering the TPV emitter and cell as thin films need to be further explored. 3.3. Performance enhancement from nanowire HMM emitters and TPV cells with finite thicknesses In this section, both the HMM emitter and the TPV cell will be considered as thin films, and the effect of film thicknesses on the near-field radiative heat transfer and the TPV conversion efficiency will be studied, aiming to enhance the TPV performance with the tungsten nanowire HMM emitters. First, the TPV cell will be modeled as a thin film while the emitter is still semi-infinite in a 4-layer near-field TPV design. Then, the thickness effect of both thin TPV cell and the nanowire HMM emitter will be investigated via a 5-layer structure, while the substrate for both the TPV cell and the HMM emitter is taken to be vacuum for simplicity. 3.3.1. Thin film effect of a TPV cell with finite thickness The key to enhance the TPV conversion efficiency is to reduce the spectral radiative energy below the bandgap that is not useful to generate electron–hole pairs in the cell [11]. Here, the semi-infinite

J.-Y. Chang et al. / International Journal of Heat and Mass Transfer 87 (2015) 237–247

245

TPV cell is replaced by a free-standing thin layer with thickness t = 10 lm, while the nanowire HMM emitter is kept semi-infinite, forming a 4-layer near-field TPV system. The filling ratio of the HMM emitter is chosen to be the same as f = 0.5 for comparison with the previous results from the semi-infinite TPV cell. Fig. 7(a) shows the spectral radiative heat flux from the 4-layer TPV system with a cell thickness of t = 10 lm. It can clearly be observed that the spectral heat flux at wavelengths smaller than the cell bandgap is almost the same with that from a semi-infinite cell shown in Fig. 4(c) with relative difference less than 1%, indicating that the spectral radiative energy above the cell bandgap is totally absorbed within the 10-lm TPV cell. However, the spectral radiative energy below the bandgap is reduced to almost zero at longer wavelengths, which is actually due to the total internal reflection occurring at the bottom interface between the TPV cell and vacuum substrate (i.e., layer 3 and 4) because the electromagnetic waves are incident from a dense medium (i.e., TPV cell) with a larger refractive index into vacuum. In fact, only the waves with spectral energy below the bandgap could reach the bottom cell interface, while the energy associated with the evanescent waves will be reflected back since the total internal reflection occurs at bc/x > 1. The significantly reduced long-wavelength radiative energy absorbed by the TPV cell due to the total internal reflection could greatly enhance the TPV conversion efficiency. The normalized radiative power input and electrical power output from the 4-layer TPV system with the nanowire HMM emitter to that with plain tungsten emitter, as shown in Fig. 7(b), clearly shows the desired drop of the radiative power by comparing with Fig. 5(a). Since the spectral radiative energy above the cell bandgap is almost identical, the electrical power generation is almost kept unaffected when the TPV cell becomes a thin layer of 10 lm in thickness. As a result, the TPV conversion efficiency, which is the ratio of power generation to the radiative power input, is improved as expected due to the minimization of long-wavelength radiative transfer below the bandgap. As shown in Fig. 7(c), the improved conversion efficiency of TPV system with a semi-infinite HMM emitter and a thin TPV cell could reach 31.1% or so regardless filling ratios, which is nearly a 4% increase to that (26.7%) from a thin TPV cell of 10 lm thick or even a 8% increase over that (23.1%) from a semi-infinite TPV cell both with a plain tungsten emitter. 3.3.2. Thickness effects of both thin TPV cell and thin HMM emitter As the improved near-field TPV performance was demonstrated with a thin TPV cell of 10 lm in thickness, we now aim to gain a quantitative understanding in the cell thickness effect on the TPV conversion efficiency and the power generation, which is shown in Fig. 8(a). The parameters such as vacuum gap d = 20 nm and emitter filling ratio f = 0.5 are kept the same. As the TPV cell thickness decreases from t = 10 lm, the power output drops monotonically. However, the conversion efficiency slightly rises up to a maximum of 31.8% around t = 3 lm. Within this thickness range, the TPV cell is still thick enough to absorb most of the spectral energy above the cell bandgap but absorbs less spectral energy below the bandgap due to smaller cell thickness (or less material). As a result, the conversion efficiency is slightly increased. When the TPV cell thickness t is further reduced to 10 nm, the conversion efficiency drops dramatically down to 22%, which is apparently because the cell is not thick enough to absorb all the spectral energy above the bandgap. Therefore, the thickness effect indicates that, the TPV cell should be chosen to be thicker than 2 lm to achieve the optimal conversion efficiency. It should be noted that the present conversion analysis on the cell thickness effect does not consider the influences from the electrical losses due to charge recombination and from the thermal losses due to non-uniform temperature distribution inside the TPV cell. In practice, different cell thickness would affect the charge

Fig. 7. A 4-layer near-field TPV system made of a semi-infinite nanowire HMM emitter with different filling ratios and a free-standing thin TPV cell with thickness t = 10 lm separated by a vacuum gap d = 20 nm: (a) spectral heat flux; (b) electrical power output and the radiative power input normalized to that with a plain tungsten emitter; (c) conversion efficiency.

transport as well as the thermal transport. In order to rigorously model both electrical and thermal losses, the spectral energy absorbed at different cell depth has to be calculated [9,10]. However, because of the uniaxial nature of the nanowire emitter, the rigorous calculation requires a near-field radiative transfer model incorporated with multilayer uniaxial wave optics, which is yet to be developed. Therefore, the present study considers the

246

J.-Y. Chang et al. / International Journal of Heat and Mass Transfer 87 (2015) 237–247

the effective material properties of the tungsten nanowire arrays. Therefore, the TPV conversion efficiency decreases significantly from 30.5% at h = 200 nm to 23% with a thin nanowire array of 10 nm in thickness. 4. Conclusions

Fig. 8. Conversion efficiency, radiative power, the electrical power density of a 5layer near-field TPV system made of a thin nanowire HMM emitter and a thin TPV cell, both of which are on vacuum substrates as a function of (a) the cell thickness t with nanowire height h = 10 lm or (b) nanowire height h with cell thickness t = 10 lm. The filling ratio of the nanowire emitter is f = 0.5, while the vacuum gap is d = 20 nm.

TPV cell (i.e., In0.2Ga0.8Sb) with quantum efficiency from the farfield response and neglects the thickness effect on the electrical and thermal losses, aiming to only focus on near-field radiative transfer with the nanowire HMM emitter in terms of the near-field TPV performance. Finally, the thickness effect of the tungsten nanowire emitter on the near-field TPV performance is considered by treating the nanowire array with a finite height h on vacuum substrate, forming a 5-layer near-field TPV system with a thin TPV cell with thickness t = 10 lm. As shown in Fig. 8(b), the radiative and electrical power as well as the resulting conversion efficiency do not change much when the nanowire emitter thickness h reduces from 10 lm to 1 lm, suggesting that 1-lm-long nanowires are thick enough to be opaque. When further reducing the nanowire height down to 200 nm or so, oscillation occurs in both the radiative and electrical power plots as a function of emitter thickness h, due to the wave interference inside the thin-film emitter as the thermal wavelength is comparable with the nanowire thickness. As a result, the conversion efficiency also oscillates with a 1% fluctuation but does not change much. However, when the tungsten nanowire height is further reduced to be h < 200 nm, the radiative power drops dramatically with a much faster rate than the electricity power generation. This can be explained by the less emitting materials as the ENP and the HMM behavior are responsible for the near-field radiative heat transfer enhancement, both of which are simply associated with

In summary, the performance of a near-field TPV system with HMM emitter made of tungsten nanowire embedded in Al2O3 host has been thoroughly investigated. The filling ratio effect on the material properties of near-field radiative heat transfer was analyzed. The results show that a semi-infinite HMM emitter with f = 0.5 can achieve 2.15 times (1.25 MW/m2) of enhancement in electrical power output from semi-infinite TPV cell made of In0.2Ga0.8Sb comparing to that with plain tungsten emitter. The mechanisms for radiative heat transfer enhancement have been illustrated as ENP for s polarization and hyperbolic modes for p polarization. However, it turned out that the conversion efficiency becomes lower because the spectral radiative energy below the cell bandgap is also enhanced with the HMM emitter. In order to improve the conversion efficiency, a thin-film TPV cell with t = 10 lm is considered. Total internal reflection occurs at the bottom interface of TPV cell and thus minimizes the energy absorption below band gap. As a result, the conversion efficiency has been enhanced from 17.7% to 31.1% with the same amount of power output, compared to 26.7% with a plain tungsten emitter. Furthermore, with a TPV cell thickness of t = 3 lm, the conversion efficiency can be further improved to 31.8% where less energy below the bandgap is absorbed, while the nanowire emitter is desired to be opaque. Note that the current theoretical model only considers the nearfield radiative transport between the uniaxial nanowire HMM emitter and the TPV cell. In order to accurately predict the near-field quantum efficiency and fully understand the thermal impact on the overall near-field TPV systems enhanced by nanowire HMMs, a rigorous and complete model for near-field radiative transport with multilayer uniaxial media coupled with charge transport and thermal transport is yet to be developed. The results gained here would facilitate the practical design and demonstration of nearfield TPV devices enhanced by novel metamaterial emitters. Conflict of interest None declared. Acknowledgment The authors are grateful to the support from the New Faculty Startup Fund at ASU. YY and LW would like to thank the support from the National Science Foundation (CBET-1454698). References [1] S. Basu, Z.M. Zhang, C.J. Fu, Review of near-field thermal radiation and its application to energy conversion, Int. J. Energy Res. 33 (2009) 1203–1232. [2] K. Park, Z.M. Zhang, Fundamentals and applications of near-field radiative energy transfer, Front. Heat Mass Transfer 4 (2013) 013001. [3] L. Hu, A. Narayanaswamy, X. Chen, G. Chen, Near-Field thermal radiation between two closely spaced glass plates exceeding Planck’s blackbody radiation law, Appl. Phys. Lett. 92 (2008) 133106. [4] E. Rousseau, A. Siria, G. Jourdan, S. Volz, F. Comin, J. Chevrier, J.-J. Greffet, Radiative heat transfer at the nanoscale, Nat. Photonics 3 (2009) 514–517. [5] S. Shen, A. Narayanaswamy, G. Chen, Surface phonon polaritons mediated energy transfer between nanoscale gaps, Nano Lett. 9 (2009) 2909–2913. [6] A. Narayanaswamy, G. Chen, Surface modes for near field thermophotovoltaics, Appl. Phys. Lett. 82 (2003) 3544–3546. [7] R. DiMatteo, P. Greiff, D. Seltzer, D. Meulenberg, E. Brown, E. Carlen, K. Kaiser, S. Finberg, H. Nguyen, J. Azarkevich, P. Baldasaro, J. Beausang, L. Danielson, M. Dashiell, D. DePoy, H. Ehsani, W. Topper, K. Rahner, Micron-gap thermophotovoltaics (MTPV), AIP Conf. Proc. 653 (2003) 232–240.

J.-Y. Chang et al. / International Journal of Heat and Mass Transfer 87 (2015) 237–247 [8] M. Laroche, R. Carminati, J.-J. Greffet, Near-field thermophotovoltaic energy conversion, J. Appl. Phys. 100 (2006) 063704. [9] K. Park, S. Basu, W.P. King, Z.M. Zheng, Performance analysis of near-field thermophotovoltaic devices considering absorption distribution, J. Quant. Spectrosc. Radiat. Transfer 109 (2008) 305–316. [10] M. Francoeur, R. Vaillon, M.P. Mengüç, Thermal impacts on the performance of nanoscale-gap thermophotovoltaic power generators, IEEE Trans. Energy Convers. 26 (2011) 686–698. [11] T.J. Bright, L.P. Wang, Z.M. Zhang, Performance of near-field thermophotovoltaic cells enhanced with a backside reflector, J. Heat Transfer 136 (2014) 062701. [12] O. Ilic, M. Jablan, J.D. Joannopoulos, I. Celanovic, M. Soljacˇic´, Overcoming the black body limit in plasmonic and graphene near-field thermophotovoltaic systems, Opt. Express 20 (2012) A366–A384. [13] R. Messina, P. Ben-Abdallah, Graphene-based photovoltaic cells for near-field thermal energy conversion, Sci. Rep. 3 (2013) 1383. [14] Z. Jacob, Quantum plasmonics, MRS Bull. 37 (2012) 761–767. [15] S.-A. Biehs, M. Tschikin, P. Ben-Abdallah, Hyperbolic metamaterials as an analog of a blackbody in the near field, Phys. Rev. Lett. 109 (2012) 104301. [16] E.E. Narimanov, H. Li, Yu.A. Barnakov, T.U. Tumkur, M.A. Noginov, Reduced reflection from roughened hyperbolic metamaterial, Opt. Express 21 (2013) 14956–14961. [17] M.A. Noginov, H. Li, Yu.A. Barnakov, D. Dryden, G. Nataraj, G. Zhu, C.E. Bonner, M. Mayy, Z. Jacob, E.E. Narimanov, Controlling spontaneous emission with metamaterials, Opt. Lett. 35 (2010) 1863–1865. [18] X.L. Liu, Z.M. Zhang, Metal-free low-loss negative refraction in the midinfrared region, Appl. Phys. Lett. 103 (2013) 103101. [19] M.A. Noginov, Yu.A. Barnakov, G. Zhu, T. Tumkur, H. Li, E.E. Narimanov, Bulk photonic metamaterial with hyperbolic dispersion, Appl. Phys. Lett. 94 (2009) 151105. [20] E.E. Narimanov, H. Li, Yu A. Barnakov, T.U. Tumkur, M.A. Noginov, Darker than black: radiation-absorbing metamaterial, in: Quantum Electronics and Laser Science Conference, OSA Technical Digest (CD) (Optical Society of America, 2010), 2010, p. QPDA6. [21] J. Elser, R. Wangberg, V.A. Podolskiy, E.E. Narimanov, Nanowire metamaterials with extreme optical anisotropy, Appl. Phys. Lett. 89 (2006) 261102. [22] C.L. Cortes, W. Newman, S. Molesky, Z. Jacob, Quantum nanophotonics using hyperbolic metamaterials, J. Opt. 14 (2012) 063001. [23] Y. Guo, W. Newman, C.L. Cortes, Z. Jacob, Applications of hyperbolic metamaterial substrates, Adv. OptoElectron. 2012 (2012) 1–9. [24] Y. Guo, C.L. Cortes, S. Molesky, Z. Jacob, Broadband super-Planckian thermal emission from hyperbolic metamaterials, Appl. Phys. Lett. 101 (2012) 131106. [25] T. Tumkur, G. Zhu, P. Black, Yu.A. Barnakov, C.E. Bonner, M.A. Noginov, Control of spontaneous emission in a volume of functionalized hyperbolic metamaterial, Appl. Phys. Lett. 99 (2011) 151115. [26] C.J. Zapata-Rodríguez, J.J. Miret, S. Vukovic´, J.A. Sorni, Z. Jakšic´, Designed surface modes propagating along hyperbolic metamaterials, 15th Int. Conf. on Transparent Optical Networks (ICTON) (2013) 1–3.

247

[27] Z. Jacob, J.-Y. Kim, G.V. Naik, A. Boltasseva, E.E. Narimanov, V.M. Shalaev, Engineering photonic density of states using metamaterials, Appl. Phys. B 100 (2010) 215–218. [28] W. Yang, M.A. Fiddy, Photonic density of states in hyperbolic metamaterials with finite feature sizes, in: Proc. of IEEE Southeastcon, 2012. [29] S. Molesky, C.J. Dewalt, Z. Jacob, High temperature epsilon-near-zero and epsilon-near-pole metamaterial emitters for thermophotovoltaics, Opt. Express 21 (2013) A96–A110. [30] S. Basu, L.P. Wang, Near-field radiative heat transfer between doped silicon nanowire arrays, Appl. Phys. Lett. 102 (2013) 053101. [31] B. Liu, J. Shi, K. Liew, S. Shen, Near-field radiative heat transfer for Si based metamaterials, Opt. Commun. 314 (2014) 57–65. [32] X.L. Liu, R.Z. Zhang, Z.M. Zhang, Near-field thermal radiation between hyperbolic metamaterials: graphite and carbon nanotubes, Appl. Phys. Lett. 103 (2013) 213102. [33] B. Liu, S. Shen, Broadband near-field radiative thermal emitter/absorber based on hyperbolic metamaterials: direct numerical simulation by the wiener chaos expansion method, Phys. Rev. B 87 (2013) 115403. [34] S.-A. Biehs, P. Ben-Abdallah, F.S.S. Rosa, K. Joulain, J.-J. Greffet, Nanoscale heat flux between nanoporous materials, Opt. Express 19 (2011) A1088–A1103. [35] S. Basu, Y.-B. Chen, Z.M. Zhang, Microscale radiation in thermophotovoltaic devices—A review, Int. J. Energy Res. 31 (2007) 689–716. [36] H. Asoh, K. Nishio, M. Nakao, A. Yokoo, T. Tamamura, H. Masuda, Fabrication of ideally ordered anodic porous alumina with 63 nm hole periodicity using sulfuric acid, J. Vac. Sci. Technol., B 19 (2001) 569. [37] G. Kartopu, O. Yalçın, Fabrication and applications of metal nanowire arrays electrodeposited in ordered porous templates, Electrodeposited Nanowires and their Applications, 2010. [38] W. Lee, R. Ji, C.A. Ross, U. Gösele, K. Nielsch, Wafer-scale Ni imprint stamps for porous alumina membranes based on interference lithography, Small 2 (2006) 978–982. [39] X.J. Wang, J.L. Abell, Y.-P. Zhao, Z.M. Zhang, Angle-resolved reflectance of obliquely aligned silver nanorods, Appl. Opt. 51 (2012) 1521–1531. [40] T.C. Choy, Effective Medium Theory: Principles and Applications, Oxford University Press, 1999. [41] H. Wang, X.L. Liu, L.P. Wang, Z.M. Zhang, Anisotropic optical properties of silicon nanowire arrays based on the effective medium approximation, Int. J. Therm. Sci. 65 (2013) 62–69. [42] E.D. Palik, Handbook of Optical Constants of Solids, Academic Press, 1998. [43] S.M. Rytov, Yu.A. Krastov, V.I. Tatarskii, ‘‘Principles of Statistical Radiophysics, vol. 3, Springer-Verlag, New York, 1987. [44] S. Basu, M. Francoeur, Maximum near-field radiative heat transfer between thin films, Appl. Phys. Lett. 98 (2011) 243120. [45] M. Francoeur, M.P. Mengüç, R. Vaillon, Spectral tuning of near-field radiative heat flux between two thin silicon carbide films, J. Phys. D Appl. Phys. 43 (2010) 075501. [46] Y. Yang, S. Basu, L.P. Wang, Radiation-based near-field thermal rectification with phase transition materials, Appl. Phys. Lett. 103 (2013) 163101.