Research article Special Issues

Thermodynamic performance of near-field electroluminescence and negative electroluminescent refrigeration systems

  • Electroluminescent (EL) and negative electroluminescent (NEL) devices are radiative thermoelectric energy converters that use electric power for refrigeration. For the EL system, we apply a forward bias to the emitter that we want to cool, whereas a reverse bias voltage is applied to the hot absorber for the NEL system. In this work, we derive the thermodynamic limits of the cooling power density and coefficient of performance (COP) of near-field EL and NEL refrigeration systems based on entropy analysis that considers near-field effects. We show numerically that operating the EL and NEL systems in the near-field regime could increase the cooling power density and the COP bounds to a certain extent. As the vacuum gap decrease from 1000 to 10 nm, the near-field effects improve the performance of the NEL system all the time, but the performance of the EL system increases to the optimal value and then decreases. In addition, the increase in temperature difference weakens the performance of both refrigeration systems greatly. Moreover, we also investigate the effects of the absence of sub-bandgap thermal radiation on the performance of the EL and NEL systems. Our work indicates significant opportunities for evaluating the performance of near-field radiative thermoelectric energy converters from the perspective of thermodynamic limits. Meanwhile, these results establish the targets for cooling power density and COP of the near-field EL and NEL systems.

    Citation: Bowen Li, Qiang Cheng, Jinlin Song, Kun Zhou, Lu Lu, Zixue Luo. Thermodynamic performance of near-field electroluminescence and negative electroluminescent refrigeration systems[J]. AIMS Energy, 2021, 9(3): 465-482. doi: 10.3934/energy.2021023

    Related Papers:

    [1] Thiago M. Vieira, Ézio C. Santana, Luiz F. S. Souza, Renan O. Silva, Tarso V. Ferreira, Douglas B. Riffel . A novel experimental procedure for lock-in thermography on solar cells. AIMS Energy, 2023, 11(3): 503-521. doi: 10.3934/energy.2023026
    [2] Xiaofeng Zhang, Qing Ai, Kuilong Song, Heping Tan . Bidirectional monte carlo method for thermal radiation transfer in participating medium. AIMS Energy, 2021, 9(3): 603-622. doi: 10.3934/energy.2021029
    [3] Caliot Cyril, Flamant Gilles . Pressurized Carbon Dioxide as Heat Transfer Fluid: In uence of Radiation on Turbulent Flow Characteristics in Pipe. AIMS Energy, 2014, 1(2): 172-182. doi: 10.3934/energy.2014.2.172
    [4] Hong-Yu Pan, Chuang Sun, Xue Chen . Transient thermal characteristics of infrared window coupled radiative transfer subjected to high heat flux. AIMS Energy, 2021, 9(5): 882-898. doi: 10.3934/energy.2021041
    [5] Hong-Wei Chen, Fu-Qiang Wang, Yang Li, Chang-Hua Lin, Xin-Lin Xia, He-Ping Tan . Numerical design of dual-scale foams to enhance radiation absorption. AIMS Energy, 2021, 9(4): 842-853. doi: 10.3934/energy.2021039
    [6] Chao Wei, Gabriel Alexander Vasquez Diaz, Kun Wang, Peiwen Li . 3D-printed tubes with complex internal fins for heat transfer enhancement—CFD analysis and performance evaluation. AIMS Energy, 2020, 8(1): 27-47. doi: 10.3934/energy.2020.1.27
    [7] Yasong Sun, Jiazi Zhao, Xinyu Li, Sida Li, Jing Ma, Xin Jing . Prediction of coupled radiative and conductive heat transfer in concentric cylinders with nonlinear anisotropic scattering medium by spectral collocation method. AIMS Energy, 2021, 9(3): 581-602. doi: 10.3934/energy.2021028
    [8] Cheng Ziming, Lin Bo, Shi Xuhang, Wang Fuqiang, Liang Huaxu, Shuai Yong . Influences of atmospheric water vapor on spectral effective emissivity of a single-layer radiative cooling coating. AIMS Energy, 2021, 9(1): 96-116. doi: 10.3934/energy.2021006
    [9] Lifita N. Tande, Valerie Dupont . Autothermal reforming of palm empty fruit bunch bio-oil: thermodynamic modelling. AIMS Energy, 2016, 4(1): 68-92. doi: 10.3934/energy.2016.1.68
    [10] Abdullah Kadhlm Ali, Ahmed Qassem Mohammed, Qasim Selah Mahdi . Experimental study of a natural draft hybrid (wet/dry) cooling tower with a splash fill type. AIMS Energy, 2022, 10(4): 648-664. doi: 10.3934/energy.2022031
  • Electroluminescent (EL) and negative electroluminescent (NEL) devices are radiative thermoelectric energy converters that use electric power for refrigeration. For the EL system, we apply a forward bias to the emitter that we want to cool, whereas a reverse bias voltage is applied to the hot absorber for the NEL system. In this work, we derive the thermodynamic limits of the cooling power density and coefficient of performance (COP) of near-field EL and NEL refrigeration systems based on entropy analysis that considers near-field effects. We show numerically that operating the EL and NEL systems in the near-field regime could increase the cooling power density and the COP bounds to a certain extent. As the vacuum gap decrease from 1000 to 10 nm, the near-field effects improve the performance of the NEL system all the time, but the performance of the EL system increases to the optimal value and then decreases. In addition, the increase in temperature difference weakens the performance of both refrigeration systems greatly. Moreover, we also investigate the effects of the absence of sub-bandgap thermal radiation on the performance of the EL and NEL systems. Our work indicates significant opportunities for evaluating the performance of near-field radiative thermoelectric energy converters from the perspective of thermodynamic limits. Meanwhile, these results establish the targets for cooling power density and COP of the near-field EL and NEL systems.



    In the near-field, the radiative heat transfer could be enhanced by several orders of magnitude as compared to the Planck's blackbody limit due to the presence of evanescent waves [1,2,3,4,5,6,7,8,9]. In recent years, this enhancement has been experimentally demonstrated [8,9,10,11,12,13,14]. Such enhancement has been widely studied in many applications like energy conversion systems [15,16,17,18,19,20,21,22]. Moreover, there has been growing interest in active control of near-field heat transfer like electroluminescent (EL) [23,24,25] and negative electroluminescent (NEL) refrigeration [26,27,28].

    In most of the previous work on the near-field radiative heat transfer, it is assumed that the chemical potential of the object involved is zero. However, when photons are in quasiequilibrium with semiconductors under external bias, they can have a chemical potential [29]. One approach to operating a semiconductor p-n junction as an EL refrigeration device is to apply a forward bias to make the body emit excess photons and cool down below ambient temperature [30].

    The first prediction of EL refrigeration was made by Tuac in 1957 [31], then this evidence was achievable by Dousmanis et al. in 1963 [32]. Compared with traditional solid-state thermoelectric refrigerators, this refrigeration approach has many special superiorities, such as higher efficiency, easier integration with other optoelectronic devices, and wider operating temperatures [24]. In the past few decades, there have been some studies expounding on the potential of EL refrigeration [33,34,35,36,37,38]. Recently, the electroluminescence refrigeration effect was experimentally demonstrated for narrow bandgap emitters under an ultralow applied bias [39].

    Due to the quite low cooling power density of far-field EL cooling devices, the concept of near-field electromagnetic heat transfer can be used to enhance the performance of EL refrigeration. In 2012, Guha et al. [40] experimentally demonstrated the efficient EL refrigeration and theoretically proved that it is feasible to realize very strong cooling in the near-field regime. Chen et al. [23] first theoretically investigated the near-field EL refrigeration considering contributions of evanescent waves and phonon modes in 2015. They predicted that refrigeration effects can occur between two semiconductor bodies when the gap spacing ranges from tens to hundreds of nanometers. Liu and Zhang [24] developed a multilayer model to consider the nonuniform distribution of the chemical potential of photons and nanoscale radiative transfer.

    In addition to the EL effect of the forward-biased p-n junction, the reverse-biased junction can also achieve refrigeration through the NEL effect [41,42,43,44]. This effect was demonstrated experimentally by Ashley et al. [45] for reverse-biased InSb and mercury cadmium telluride (MCT) diodes. Chen et al. [26] analyzed the near-field enhanced NEL, for an ideal narrow-band-gap semiconductor, they show that power density can be increased significantly and the efficiency can be close to the Carnot limit in the near-field regime. Then Zhu et al. [46] firstly reported an experimental demonstration of near-field NEL refrigeration using a nanocalorimetric device and a photodiode. Very recently, Zhou et al. [28] theoretically analyze the performance of a near-field NEL refrigeration system consisting of a Mie-metamaterial emitter.

    With the consideration of the near-field effect, the influence on the maximum active work and energy conversion efficiency of the EL and NEL systems should be concerned. Moreover, entropy plays a key role in the determination of the thermodynamic limit of the maximum work and the energy conversion efficiency [47,48,49,50,51,52]. In the near-field regime, Dorofeyev first studied the energy and entropy density due to near-field effects in equilibrium at the interface [53]. The maximum work and the upper bound for the efficiency of thermal radiation have been obtained by Perez-Madrid et al. [54,55,56]. Moreover, in consideration of both near-field effects and the condition of thermal nonequilibrium, Narayanaswamy and Zheng [57] derived a method to obtain the maximum work and the energy conversion efficiency limit in near-field thermal radiation. Recently, this method was applied to the performance evaluation of near-field thermophotovoltaic systems [58].

    In the present work, we derive thermodynamic bounds for the cooling power densities and the coefficient of performance (COP) of the near-field EL and NEL systems by calculating the radiative heat flux and entropy flux that considers near-field effects. Then we present the performance of the near-field EL and NEL systems at different vacuum gaps. Furthermore, the effects of the temperature difference and sub-bandgap thermal radiation on the thermodynamic performance are thoroughly investigated.

    The considered configurations of the near-field EL and NEL systems to be investigated in this study are schematically shown in Figure 1. A near-field EL energy conversion system consists of a heat source (cooling target), an EL p-n device as an emitter, and an absorber. Where the vacuum gap spacing between the emitter and absorber is denoted by d. The emitter and absorber are enclosed within Au substrates acting as perfect mirrors at the top and bottom boundaries respectively, as shown by the yellow layers. The structure of a NEL system is similar to that of an EL system, as shown in Figure 1(b). The difference is that the NEL device acts as an absorber. This work focuses on the near-field radiative transfer so that we assume ideal thermal contact between the different layers.

    Figure 1.  Schematic and energy flow diagram of a near-field (a) EL and (b) NEL system. The EL system consists of three bodies, i.e., heat source, EL device (emitter), and absorber. The NEL system consists of an emitter, NEL device (absorber), and heat sink. The emitter and absorber are enclosed within two Au substrates.

    Then we briefly review the concept of EL and NEL refrigeration. In the case where the EL device is colder than the absorber. By supplying with electrical work as delivered by an external forward voltage VEL, a near-field EL device emits an increased number of photons as compared to the same material under equilibrium conditions, and some of this photon energy is derived from the thermal energy of the lattice. Under an adequate high operating voltage, such an increased photon emission can result in a net heat flux pumped from a cold EL device to a hot absorber in spite of the fact that the EL device has a lower temperature. As for a NEL device under a reverse bias voltage (i.e., VNEL < 0), the photon emission decreases, it can extract a net radiative heat flux from an emitter (cooling target) with a lower temperature. Hence the resulting structures shown in Figure 1 can be used for refrigeration purposes.

    In the calculation, the thicknesses of emitter and absorber denoted as d1 and d2 are chosen to be 5 μm to ensure significant emission and absorption. For a fair comparison between the EL and NEL systems, the temperatures of the two heat reservoirs are fixed. The cooling target temperatures (that is, the temperatures of the heat source of the EL system and the emitter of the NEL system) are maintained at a low-temperature T1, whereas the absorber of the EL system and the heat sink of the NEL system are maintained at a high-temperature T2 = 300 K. In the near-field EL system, we choose InAs as the semiconductor for both the absorber and EL device, whose bandgap energy Eg is 0.354 eV at room temperature [59]. For the near-field NEL system, we use a narrow band-gap semiconductor MCT as the NEL device and emitter. It has a bandgap of Eg = 0.169 eV [26]. The optical constants of InAs and MCT are taken from Ref. [60]. Moreover, the dielectric function of Au as a function of angular frequency ω, is approximated by the Drude model [61]: εAu(ω) = 1ω2p/(ω2+iγω) with a plasma frequency ωp = 1.37 × 1016 rad/s and damping rate γ = 7.31 × 1013 rad/s.

    The energy flows (solid arrows) and entropy flows (dashed arrows) of the near-field EL and NEL systems are shown in Figure 1. An EL device absorbs net outflux of heat QC and entropy SC from the cold heat source in the form of heat conduction. Then by the injected electric power W, the EL device pumps heat flux (E1 - E2) and entropy flux (S1 - S2) to the hot absorber via near-field radiation. As for a NEL system, the biased NEL device absorbs heat flux (E1 - E2) and entropy flux (S1 - S2) from the cold emitter and rejects heat flux QC and entropy flux SC to the hot heat sink.

    The polarized radiative heat flux (E1 - E2) between two bodies has contributions both from sub-bandgap frequency components (E1, sE2, s) due to phonon-polariton excitations, and from above-bandgap frequency components (E1, aE2, a) due to electronic excitations. Using the fluctuation-dissipation theorem [62,63,64], the radiative heat flux can be expressed as

    E(j)1, sE(j)2, s=ωg/c0ωdk0{k002π0[n(j)1(0)υ(j)1n(j)2(0)υ(j)2]k0/|kz0|dkρkρdϕ+k02π0[n(j)1(0)υ(j)1n(j)2(0)υ(j)2]β2z0+k2ρ/|βz0|dkρkρdϕ}, (1a)
    E(j)1, aE(j)2, a=+ωg/cωdk0{k002π0[n(j)1(V1)υ(j)1n(j)2(V2)υ(j)2]k0/|kz0|dkρkρdϕ+k02π0[n(j)1(V1)υ(j)1n(j)2(V2)υ(j)2]β2z0+k2ρ/|βz0|dkρkρdϕ}. (1b)

    Here, j = s, p accounts for the polarization states, ωg is the band-gap frequency of the semiconductor defined as ωg=Eg/, is the reduced Planck constant, V denotes the applied voltage. For the EL system, V1>0 and V2=0; For the NEL system, V1=0 and V2<0. We note that Eq (1a) and (1b) contain triple integrations over the azimuthal angle ϕ, the component of the wavevector on the x-y plane kρ, and the free-space wavevector k0 = ω/c with c being the speed of light in vacuum. The wavevectors component in the z-direction for propagating waves (PW) and evanescent waves (EW) are expressed as kz0=k20k2ρ and βz0=k2ρk20, respectively. Moreover, n(j)h(V) is the number of emitted photons with h = 1, 2, and υ(j)h is the z component of the local velocity of the energy transmission. Further, n(j)h(V) and υ(j)h are respectively defined as

    n(j)h(V)={sgn(kz0)8π3fh(V)(1|R(j)h|2)[1+|R(j)¯h|2+k2ρk202Re(R(j)¯hei2kz0d¯h)]|1R(j)hR(j)¯hei2kz0d|2forPW, sgn(βz0)8π3fh(V)k0β2z0+k2ρ2Im(R(j)h)e2βz0d|1R(j)hR(j)¯he2βz0d|2×[2Re(R(j)¯h)+k2ρk20(e2βz0d¯h+e2βz0d¯h|R(j)¯h|2)]forEW,  (2)
    υ(j)h={ckz0k01|R(j)¯h|21+|R(j)¯h|21+k2ρk202Re(R(j)¯hei2kz0d¯h)1+|R(j)¯h|2 for PW, cβz0k0k2ρ2Im(R(j)¯h)e2βz0d¯h1+|R(j)¯h|2e4βz0d¯h1+k20k2ρ2Re(R(j)¯h)e2βz0d¯h1+|R(j)¯h|2e4βz0d¯h     for EW, (3)

    where dh = |zzh|, ¯h = 2 if h = 1, and vice versa. fh(V) = [e(ωqV)/kBTh1]1 is the photon distribution function, kB is the Boltzmann constant and q is the magnitude of the electron's charge. We define R(j)h as the reflection coefficients from vacuum to body h for j polarization with Au substrates [65,66].

    The entropy flux carried by the related energy flux [Eq (1)] takes the form

    S(j)1, sS(j)2, s=ωg/c0dω{k002π0[s(j)1(0)υ(j)1s(j)2(0)υ(j)2]k0/|kz0|dkρkρdϕ+k02π0[s(j)1(0)υ(j)1s(j)2(0)υ(j)2]β2z0+k2ρ/|βz0|dkρkρdϕ}, (4a)
    S(j)1, aS(j)2, a=+ωg/cdω{k002π0[s(j)1(V1)υ(j)1s(j)2(V2)υ(j)2]k0/|kz0|dkρkρdϕ+k02π0[s(j)1(V1)υ(j)1s(j)2(V2)υ(j)2]β2z0+k2ρ/|βz0|dkρkρdϕ}, (4b)

    with the near-field entropy density sh given by [57]

    s(j)h(V)=kBρ(j)a[(1+n(j)h(V)ρ(j)a)ln(1+n(j)h(V)ρ(j)a)n(j)h(V)ρ(j)aln(n(j)h(V)ρ(j)a)]. (5)

    Here, ρa expresses the local density of the accessible microscopic states, which is determined by

    ρ(j)a={sgn(kz0)8π3Re[(1+R(j)1R(j)2ei2kz0d+k2ρk20R(j)2ei2kz0d2+k2ρk20R(j)1ei2kz0d1)1R(j)1R(j)2ei2kz0d]forPW, sgn(βz0)8π3k0β2z0+k2ρ×Im[(1+R(j)1R(j)2e2βz0d+k2ρk20R(j)2e2βz0d2+k2ρk20R(j)1e2βz0d1)1R(j)1R(j)2e2βz0d]forEW. (6)

    To evaluate the performance of the near-field EL and NEL systems, we assume that the EL device and the heat source have the same temperature (TEL = T1) since the flux of heat conduction is usually much larger than that of thermal radiation, especially when the temperature is below 300 K. Similarly, the temperature of the NEL device is the same as the temperature of the heat sink (TNEL = T2). Then we get the thermodynamic bounds of the cooling power densities P and the COP of the EL and NEL systems when the systems are assumed to be operating in the ultimate case and there is no entropy generated in the EL or NEL devices (Sgen = 0), therefore SC = S1 - S2.

    For the EL system, PEL and COP bound can be expressed as

    PEL=QC=T1×SC=T1×(S1S2). (7)
    ηEL=PELW=QC(E1E2)QC=T1×SC(E1E2)T1×SC=T1×(S1S2)(E1E2)T1×(S1S2). (8)

    For the NEL system, PNEL and COP bound are determined by

    PNEL=E1E2. (9)
    ηNEL=PNELW=E1E2QC(E1E2)=E1E2T2×SC(E1E2)=E1E2T2×(S1S2)(E1E2). (10)

    It is worth to note that for the EL device, the main part of the input entropy is carried by the thermal conduction, but the output entropy is purely carried by the thermal radiation, as shown in Figure 1. And for the NEL device, the entropy is input by the thermal radiation and output by the thermal conduction. However, the entropy to energy ratio of thermal conduction (1/T) is different from that of thermal radiation [67]. Thus, our results do show the deviation from the Carnot limit.

    We first focus on the performance of the near-field EL and NEL systems at different vacuum gaps for the heat source of the EL system and the emitter of the NEL system at a cooling target temperature T1 = 290 K. In Figure 2, we plot the spectral radiative energy flux for an applied voltage of 0.2 V in the EL system and - 0.2 V in the NEL system, respectively. For both systems, it can be seen that in the region where the frequency is below ωg, the radiative heat transfer is negative due to the temperature difference when the chemical potential of photons below the bandgap is zero. In the high-frequency region, the non-zero chemical potential provided by the applied voltage of the EL and NEL devices produces a positive heat transfer, thus resulting in a refrigeration effect, as shown in Figure 2. Moreover, the absolute heat flux increases greatly as the vacuum gap reduces in the entire frequency region for both EL and NEL systems. This effect is typical for near-field radiative heat transfer where the transferred heat flux can be significantly enhanced as the gap distance between the two bodies decreases. The corresponding spectral entropy flux curves show a similar trend with the spectral heat flux curves, but they are more than two orders of magnitude lower, as shown in Figure 3.

    Figure 2.  (a) Spectral radiative heat flux for an applied voltage of 0.2 V in the EL system. (b) Spectral radiative heat flux for an applied voltage of - 0.2 V in the NEL system. The cooling target temperature T1 = 290 K and the red, green, and blue curves are for d = 10, 100, and 1000 nm, respectively.
    Figure 3.  (a) Spectral radiative entropy flux for an applied voltage of 0.2 V in the EL system. (b) Spectral radiative energy flux for an applied voltage of -0.2 V in the NEL system. The cooling target temperature T1 = 290 K and the red, green, and blue curves are for d = 10, 100, and 1000 nm, respectively.

    The physical origins of the spectral radiative energy flux can be better illustrated by examining the energy transmission coefficient as a function of ω and β. Figure 4(a) displays the contour plot of the transmission coefficient of the EL system for d = 10 nm. Note that due to the sum of both s and p polarizations the maximum transmission coefficient is 2. The bright color indicates a high transmission coefficient. In the above-bandgap frequency range, InAs exhibits significant absorption since it is a direct bandgap material. In the frequency range below the InAs bandgap, the surface phonon polaritons (SPhPs) of InAs clearly cause a heat flux peak around 0.45 × 1014 rad/s as can be identified in Figure 4(a). Since InAs is a polar material, the real part of the dielectric function Re(ε) changes from positive to negative around the optical phonon frequency. The SPhPs can be excited and contribute significantly to radiative heat transfer at the frequency where Re(ε) is - 1. The contour plot of the transmission coefficient of the NEL system for d = 10 nm is shown in Figure 4(b), it is found that no surface mode could be excited between two MCT films in the frequency range below the bandgap.

    Figure 4.  Contour plots of the energy transmission coefficient of (a) the EL system and (b) the NEL system, respectively, when d = 10 nm. Wavevector β is normalized by β0 = ω0/c with ω0 = 1014 rad/s.

    In Figure 5(a) we plot the integrated cooling power density of the EL system as a function of the applied voltage for vacuum gap d = 10, 100, 1000 nm, respectively. When the applied voltage VEL = 0, the cooling power density PEL < 0 at every vacuum gap d, which means that there is net heat flux from the hot absorber to the cold EL device because of the negative thermal radiation caused by the temperature difference. With the applied voltage VEL increases, the heat flux from the EL device increases so slowly. As VEL continues to increase, PEL increases approximately exponentially. At d = 10 nm, we see that the cooling power density PEL stays negative at any VEL. It can be observed from Figure 2(a), in the near-field, the sub-bandgap thermal radiation accounts for the main part of the heat transfer between the EL device and the absorber, and hence the cooling effect does not appear in the range of applied voltage we considered. Whereas for an intermediate vacuum gap d = 100 nm, as VEL increases beyond a threshold voltage of 0.136 V, the cooling power density PEL becomes positive and the EL device appears a net outflow of heat and hence cooling. The maximum PEL reaches 1289 W/m2 at VEL = 0.2 V. As the vacuum gap further increases to 1000 nm, a much lower cooling power density can also be achieved, with a maximum PEL of 106 W/m2 at VEL = 0.2 V.

    Figure 5.  The cooling power density as a function of the applied voltage VEL or the magnitude of voltage |VNEL| for different vacuum gaps d in (a) the EL system and (b) the NEL system at T1 = 290 K. The red, green, and blue curves are for d = 10, 100, and 1000 nm, respectively.

    Then we plot the cooling power density of the NEL system in Figure 5(b). For each vacuum gap d, PNEL is negative at VNEL = 0, then it increases dramatically and becomes positive as the magnitude of applied reverse voltage |VNEL| increases due to the lower bandgap of the MCT which leads to the predominant above-bandgap heat transfer. Unlike the EL system, the cooling power density of the NEL system slows down gradually and finally reaches saturation and no longer increases markedly, because when |VNEL| is large enough, the heat flux from the biased NEL device is insignificant compared to the heat flux from the emitter. Besides, as d decreases, PNEL can also be greatly enhanced due to the near-field effects, we note that the maximum PNEL increases about 22 times, from 16 to 347 W/m2 as the vacuum gap decreases from 1000 to 10 nm at VNEL = - 0.2 V.

    The COP bounds versus the voltage of the EL and NEL systems are given in Figure 6(a) and (b), respectively. We note that the vacuum gap plays an important role in the COP bounds of both refrigeration systems. For the EL system, the COP bound is zero when the applied voltage VEL is small since the cooling power density PEL is negative. As VEL increases to exceed the threshold voltage, the COP bound quickly increases to a maximum value and then decreases as VEL further increases. The maximum COP bound reaches 1.19 at VEL = 0.176 V for d = 100 nm and 1.10 at = 0.182 V for d = 1000 nm.

    Figure 6.  COP bound as a function of the applied voltage VEL or the magnitude of voltage |VNEL| at different vacuum gaps d in (a) the EL system and (b) the NEL system at T1 = 290 K. The red, green, and blue curves are for d = 10, 100, and 1000 nm, respectively.

    For the NEL system, the COP bound is zero at the magnitude of applied voltage |VNEL| below the threshold voltage since the applied voltage is too weak to suppress the thermal radiation from the semiconductor NEL device. After that, it rapidly increases to the maximum value at |VNEL| slightly above the threshold voltage and then decreases as |VNEL| further increases. When the vacuum gap d = 1000 and 100 nm, the maximum COP bound is 3.3 and 7.9 respectively, which is much larger than that of the EL system under the same conditions. Further reducing the vacuum gap to 10 nm, the maximum COP bound reaches 8.76 at VNEL = - 0.023 V. Moreover, the saturation of the COP bound of the NEL system is 5.0 at d = 10 nm, 4.8 at d = 100 nm, and 2.9 at d = 1000 nm.

    Figure 7.  The cooling power density and the COP bound of the EL system as a function of the vacuum gaps d for applied voltage VEL = 0.2 V at T1 = 290 K.

    Based on the above analysis, for the EL system, it can be found that the cooling power density PEL and the COP bound do not always increase with decreasing the vacuum gap d. Thus, there should be an optimized d between 10 and 1000 nm that maximizes PEL and the COP bound of the EL system. Figure 7 gives the cooling power density and the COP bound as a function of the vacuum gap at an applied voltage of 0.2 V. The cooling power density PEL is negative at vacuum gap d < 24 nm since the sub-bandgap thermal radiation dominates over the above-bandgap thermal radiation. With the increase of d, PEL increases rapidly and reaches the maximum of 1398 W/m2 at d = 63.1 nm. Further increasing d, PEL decreases and stabilizes gradually. The COP bound also increases quickly first, and then it decreases as d increases. The maximum COP bound is 1.1 and is found at d = 200 nm.

    In the above section, the cooling target temperatures T1 is set to 290 K, whereas high-temperature T2 is assumed to be fixed at 300 K. Of course, the temperature difference ΔT = T2 - T1 could be higher by decreasing T1. Figure 8 compares the cooling power densities of near-field EL and NEL systems at temperature difference ΔT = 10, 20, and 30 K. We set the vacuum gap d = 100 nm, which is easier to achieve in practice. At the same time, the EL system shows an acceptable cooling power density and COP at this vacuum gap.

    Figure 8.  (a) The cooling power density as a function of the applied voltage VEL at different temperature differences in the EL system. (b) The cooling power density as a function of the magnitude of voltage |VNEL| at different temperature differences in the NEL system. The vacuum gap d = 100 nm.

    In terms of Eq (2), as temperature difference ΔT increases or cooling target temperature T1 decreases, the photons emitted from the EL device of the EL system and the emitter of the NEL system decrease. As a result, it can be observed that the cooling power density decreases in both systems. As ΔT increases from 10 to 30 K, for the EL system, the maximum cooling power density PEL for d = 100 nm at VEL = 0.2 V decreases from 1289 to 414 W/m2. On the other hand, for the NEL system, maximum PNEL at VNEL = - 0.2 V decreases from 205 to 83 W/m2 as ΔT increases from 10 to 20 K. In addition, the NEL system fails to achieve any cooling effect at |VNEL| < 0.2 V when ΔT increases to 30 K.

    In Figure 9(a) and (b), we plot the COP bounds of near-field EL and NEL systems as a function of the applied voltage for various temperature differences ΔT, respectively. Similar to the cooling power density, we note that the COP bound also decreases in both EL and NEL systems as the temperature difference ΔT increases. For the near-field EL system, as ΔT increases from 10 to 30 K, the maximum COP bound decreases from 1.19 to 0.6. On the other hand, the COP bound of the NEL system shows a more drastic reduction as compared to that of the EL system. The saturation of the COP bound decreases from 4.8 to 2.3 as ΔT increases from 10 to 20 K. We also mention that the threshold voltage for achieving the refrigeration effect (positive P) increases as the temperature difference ΔT goes up in both systems because a higher chemical potential is needed to overcome the increased temperature differences.

    Figure 9.  (a) COP bound as a function of the applied voltage VEL at different temperature differences in the EL system. (b) COP bound as a function of the magnitude of voltage |VNEL| at different temperature differences in the NEL system. The vacuum gap d = 100 nm.

    By comparing the EL and NEL systems from Figures 8 and 9, it is found that the EL system can reach higher cooling power density and temperature difference. However, the NEL system can achieve a refrigeration effect at a lower voltage and temperature difference with a higher COP.

    In the above sections, we demonstrate that the EL and NEL systems shown in Figure 1 can work as cooling devices with considerable cooling power density and COP bound. However, in this near-field regime, the presence of sub-bandgap thermal radiation can become very substantial. For both EL and NEL cooling systems, such thermal radiation represents a disadvantage since it results in undesired and detrimental heat flux from the hot heat reservoir to the cold cooling target. In order to evaluate the adverse effects of the sub-bandgap thermal radiation and provide suggestions for system performance improvement. In this section, we evaluate the performance of the ideal EL and NEL refrigeration systems in the absence of sub-bandgap heat transfer.

    Figure 10 shows the integrated cooling power density as a function of applied voltage VEL or the magnitude of negative applied voltage |VNEL| for different vacuum gaps d = 10, 100, and 1000 nm, respectively. For both the EL and NEL system, the absence of sub-bandgap thermal radiation enhances the cooling power density for all three cases significantly. Completely different from Figure 5(a), the heat flux PEL at d = 10 nm becomes positive and the maximum value reaches 2536 W/m2 at VEL = 0.2 V. And the stationary PNEL increases by 28% at d = 10 nm and 123% at d = 1000 nm without the consideration of sub-bandgap thermal radiation. We also find that the cooling power density of an EL system increases significantly as the vacuum gap d decreases. The maximum cooling power density PEL is 122 W/m2 at the applied voltage of 0.2 V for d = 1000 nm. For comparison, we note that the maximum PEL at d = 10 nm for this system can be increased by about 21 times to 2536 W/m2 at the same applied voltage. Besides, for the NEL system, the maximum PNEL increases about 12 times, from 36 to 445 W/m2 as the vacuum gap decreases from 1000 to 10 nm at VNEL = - 0.2 V.

    Figure 10.  (a) The cooling power density as a function of the applied voltage VEL at different vacuum gaps d in the EL system. (b) The cooling power density as a function of the magnitude of voltage |VNEL| at different vacuum gaps d in the NEL system. The cooling target temperature T1 = 290 K and the red, green, and blue curves are for d = 10, 100, and 1000 nm, respectively.

    Then we show the COP bound of the near-field EL system as a function of applied voltage for three vacuum gaps d in Figure 11(a). For all vacuum gaps, we observe that the COP bounds of the EL systems exhibit nearly identical curves. This phenomenon is because the near-field effects can enhance both the heat flux and entropy flux, according to Eq (8), the vacuum gap plays little role in the COP. The COP bound is zero when the applied voltage VEL is small since the cooling power density PEL is negative. Then the COP bound increases rapidly and reaches a maximum at VEL slightly above the threshold voltage. It is found that an ideal EL system exhibits a maximum COP bound of 26.2 at VEL = 0.014 V when the vacuum gap d = 10 nm, which is close to 29 of the Carnot limit (T1/(T2T1)). Finally, the COP bound gradually decreases and approaches zero as VEL further increases.

    Figure 11.  (a) COP bound as a function of the applied voltage VEL at different vacuum gaps d in the EL system. (b) COP bound as a function of the magnitude of voltage |VNEL| at different vacuum gaps d in the NEL system. The cooling target temperature T1 = 290 K and the red, green, and blue curves are for d = 10, 100, and 1000 nm, respectively. The horizontal dashed lines represent the Carnot efficiency limit.

    The obtained COP bound for the NEL system is shown in Figure 11(b). Similar to that in Figure 10(a), the vacuum gap almost makes no difference to the COP bound of NEL systems. For all three vacuum gaps, the COP bound is zero at the magnitude of applied voltage |VNEL| < 0.008 V since the applied voltage is too weak to suppress the thermal radiation from the semiconductor NEL device. After that, it rapidly increases to the maximum value at |VNEL| slightly above the threshold voltage and then decreases as |VNEL| further increases. The maximum COP bound is 21, which is lower than that of the EL system. Unlike the COP bound of the EL system, it can be observed that as |VNEL| further increases, the COP bound of the NEL system tends to stabilize at 6.1.

    Based on the above results, it can be found that the sub-bandgap heat transfer weakens the cooling power densities and COP bounds of the EL and NEL systems tremendously. Therefore, to improve the performance of these two refrigeration systems, it is necessary to suppress the sub-bandgap heat transfer through optimization of structures and materials.

    In summary, this study presents the thermodynamic limits for the performance of near-field EL and NEL refrigeration systems. It is shown how the cooling power density and COP bounds can be obtained from the formulation of thermodynamics by using the radiative heat flux and entropy flux considering near-field effects. The results show that the near-field effects are positive for the performance of the NEL systems. Indeed, we observe an enhancement of the maximum cooling power density by a factor of 22 for the NEL system as the vacuum gap decreases from 1000 nm to 10 nm. But the near-field effects are not always beneficial for the EL system, the cooling power density achieves the maximum of 1398 W/m2 at d = 63.1 nm and the maximum COP bound reaches 1.1 at d = 200 nm in the EL system. Moreover, as the temperature difference increases or low cooling target temperature, the cooling power densities and the COP bounds of the EL and NEL systems both decrease significantly. Finally, the absence of the sub-bandgap thermal radiation enhances the performance of both ideal refrigeration systems greatly. The COP bounds of the EL and NEL systems can be close to the Carnot limit. This study could pave the way toward providing fundamental bounds on the cooling power density and COP of near-field EL and NEL systems. And the results obtained here would present valuable guidance for improving the performance of these two systems.

    This work was mainly supported by the National Natural Science Foundation of China (51776078, 51676077, 51827808, 51806070), the Fundamental Research Funds for the Central Universities (2019kfyXKJC033), the China Postdoctoral Science Foundation (2018M632849).

    The authors declare that there are no conflicts of interest regarding the publication of this paper.



    [1] Polder D, Van Hove M (1971) Theory of Radiative Heat Transfer between Closely Spaced Bodies. Phys Rev B 4: 3303-3314. doi: 10.1103/PhysRevB.4.3303
    [2] Loomis JJ, Maris HJ (1994) Theory of heat transfer by evanescent electromagnetic waves. Phys Rev B 50: 18517-18524. doi: 10.1103/PhysRevB.50.18517
    [3] Pendry JB (1999) Radiative exchange of heat between nanostructures. J Phys-Condens Mat 11: 6621-6633. doi: 10.1088/0953-8984/11/35/301
    [4] Fu CJ, Zhang ZM (2006) Nanoscale radiation heat transfer for silicon at different doping levels. Int J Heat Mass Tran 49: 1703-1718. doi: 10.1016/j.ijheatmasstransfer.2005.09.037
    [5] Volokitin AI, Persson BNJ (2007) Near-field radiative heat transfer and noncontact friction. Rev Mod Phys 79: 1291-1329. doi: 10.1103/RevModPhys.79.1291
    [6] Hu L, Narayanaswamy A, Chen X, et al. (2008) Near-field thermal radiation between two closely spaced glass plates exceeding Planck's blackbody radiation law. Appl Phys Lett 92: 133106. doi: 10.1063/1.2905286
    [7] Basu S, Zhang ZM, Fu CJ (2009) Review of near-field thermal radiation and its application to energy conversion. Int J Energ Res 33: 1203-1232. doi: 10.1002/er.1607
    [8] Rousseau E, Siria A, Jourdan G, et al. (2009) Radiative heat transfer at the nanoscale. Nat Photonics 3: 514-517. doi: 10.1038/nphoton.2009.144
    [9] Shen S, Narayanaswamy A, Chen G (2009) Surface Phonon Polaritons Mediated Energy Transfer between Nanoscale Gaps. Nano Lett 9: 2909-2913. doi: 10.1021/nl901208v
    [10] Qu W, Mudawar I (2002) Experimental and numerical study of pressure drop and heat transfer in a single-phase micro-channel heat sink. Int J Heat Mass Tran 45: 2549-2565. doi: 10.1016/S0017-9310(01)00337-4
    [11] Kittel A, Muller-Hirsch W, Parisi J, et al. (2005) Near-field heat transfer in a scanning thermal microscope. Phys Rev Lett 95: 224301. doi: 10.1103/PhysRevLett.95.224301
    [12] Ottens RS, Quetschke V, Wise S, et al. (2011) Near-field radiative heat transfer between macroscopic planar surfaces. Phys Rev Lett 107: 014301. doi: 10.1103/PhysRevLett.107.014301
    [13] St-Gelais R, Guha B, Zhu L, et al. (2014) Demonstration of strong near-field radiative heat transfer between integrated nanostructures. Nano Lett 14: 6971-6975. doi: 10.1021/nl503236k
    [14] Song B, Fiorino A, Meyhofer E, et al. (2015) Near-field radiative thermal transport: From theory to experiment. Aip Adv 5: 053503. doi: 10.1063/1.4919048
    [15] Laroche M, Carminati R, Greffet JJ (2006) Near-field thermophotovoltaic energy conversion. J Appl Phys 100: 063704. doi: 10.1063/1.2234560
    [16] Svetovoy VB, Palasantzas G (2014) Graphene-on-Silicon near-field thermophotovoltaic cell. Phys Rev Appl 2: 034006 doi: 10.1103/PhysRevApplied.2.034006
    [17] Chang JY, Yang Y, Wang LP (2015) Tungsten nanowire based hyperbolic metamaterial emitters for near-field thermophotovoltaic applications. Int J Heat Mass Tran 87: 237-247. doi: 10.1016/j.ijheatmasstransfer.2015.03.087
    [18] Lenert A, Bierman DM, Nam Y, et al. (2014) A nanophotonic solar thermophotovoltaic device. Nat Nanotechnol 9: 126-130. doi: 10.1038/nnano.2013.286
    [19] Fiorino A, Zhu L, Thompson D, et al. (2018) Nanogap near-field thermophotovoltaics. Nat Nanotechnol 13: 806-811. doi: 10.1038/s41565-018-0172-5
    [20] Lin C, Wang B, Teo KH, et al. (2017) Near-field enhancement of thermoradiative devices. J Appl Phys 122: 143102. doi: 10.1063/1.5007036
    [21] Wang B, Lin C, Teo KH, et al. (2017) Thermoradiative device enhanced by near-field coupled structures. J. Quant Spectros Radiat Transfer 196: 10-16. doi: 10.1016/j.jqsrt.2017.03.038
    [22] Liao T, Zhang X, Yang Z, et al. (2019) Near-Field Thermoradiative Electron Device. IEEE Trans Electron Devices 66: 3099-3102. doi: 10.1109/TED.2019.2915765
    [23] Chen K, Santhanam P, Sandhu S, et al. (2015) Heat-flux control and solid-state cooling by regulating chemical potential of photons in near-field electromagnetic heat transfer. Phys Rev B 91: 134301. doi: 10.1103/PhysRevB.91.134301
    [24] Liu X, Zhang ZM (2016) High-performance electroluminescent refrigeration enabled by photon tunneling. Nano Energy 26: 353-359. doi: 10.1016/j.nanoen.2016.05.049
    [25] Liao T, Tao C, Chen X, et al. (2019) Parametric optimum design of a near-field electroluminescent refrigerator. J Phys D: Appl Phys 52: 325108. doi: 10.1088/1361-6463/ab2341
    [26] Chen K, Santhanam P, Fan S (2016) Near-Field enhanced negative luminescent refrigeration. Phys Rev Appl 6: 024014. doi: 10.1103/PhysRevApplied.6.024014
    [27] Lin C, Wang B, Teo KH, et al. (2018) A coherent description of thermal radiative devices and its application on the near-field negative electroluminescent cooling. Energy 147: 177-186. doi: 10.1016/j.energy.2018.01.005
    [28] Zhou C, Zhang Y, Qu L, et al. (2020) Near-field negative electroluminescent cooling via nanoparticle doping. J Quant Spectrosc Radiat Transfer 245: 106889-106897. doi: 10.1016/j.jqsrt.2020.106889
    [29] Wurfel P (1982) The chemical potential of radiation. J Phys C Solid State Phys 15: 3967. doi: 10.1088/0022-3719/15/18/012
    [30] Tervo E, Bagherisereshki E, Zhang Z (2017) Near-field radiative thermoelectric energy converters: a review. Front Energy 12: 5-21. doi: 10.1007/s11708-017-0517-z
    [31] Tauc J (1957) The share of thermal energy taken from the surroundings in the electro-luminescent energy radiated from a p-n junction. Cechoslovackij Fiziceskij Zurnal 7: 275-276.
    [32] Dousmanis G, Mueller C, Nelson H, et al. (1964) Evidence of refrigerating action by means of photon emission in semiconductor diodes. Phys Rev 133: A316-A318.
    [33] Berdahl P (1985) Radiant refrigeration by semiconductor diodes. J Appl Phys 58: 1369-1374. doi: 10.1063/1.336309
    [34] Mal'Shukov A, Chao K (2001) Opto-thermionic refrigeration in semiconductor heterostructures. Phys Rev Lett 86: 5570. doi: 10.1103/PhysRevLett.86.5570
    [35] Han P, Jin KJ, Zhou YL, et al. (2006) Analysis of optothermionic refrigeration based on semiconductor heterojunction. J Appl Phys 99: 074504. doi: 10.1063/1.2188249
    [36] Yen ST, Lee KC (2010) Analysis of heterostructures for electroluminescent refrigeration and light emitting without heat generation. J Appl Phys 107: 054513. doi: 10.1063/1.3326944
    [37] Oksanen J, Tulkki J (2010) Thermophotonic heat pump—a theoretical model and numerical simulations. J Appl Phys 107: 093106. doi: 10.1063/1.3419716
    [38] Santhanam P, Huang D, Gray Jr DJ, et al. (2013) Electro-luminescent cooling: light emitting diodes above unity efficiency. Laser Refrigeration of Solids VI, 863807.
    [39] Santhanam P, Gray DJ Jr, Ram RJ (2012) Thermoelectrically pumped light-emitting diodes operating above unity efficiency. Phys Rev Lett 108: 097403. doi: 10.1103/PhysRevLett.108.097403
    [40] Guha B, Otey C, Poitras CB, et al. (2012) Near-field radiative cooling of nanostructures. Nano Lett 12: 4546-4550. doi: 10.1021/nl301708e
    [41] Bewley WW, Jurkovic MJ, Felix CL, et al. (2001) HgCdTe photodetectors with negative luminescent efficiencies > 80%. Appl Phys Lett 78: 3082-3084. doi: 10.1063/1.1370539
    [42] Nash GR, Ashby MK, Lindle JR, et al. (2003) Long wavelength infrared negative luminescent devices with strong Auger suppression. J Appl Phys 94: 7300-7304. doi: 10.1063/1.1625094
    [43] Ashley T, Gordon NT, Nash GR, et al. (2001) Long-wavelength HgCdTe negative luminescent devices. Appl Phys Lett 79: 1136-1138. doi: 10.1063/1.1395521
    [44] Hoffman D, Hood A, Wei Y, et al. (2005) Negative luminescence of long-wavelength InAs∕GaSb superlattice photodiodes. Appl Phys Lett 87: 201103. doi: 10.1063/1.2130536
    [45] Ashley T, Elliott CT, Gordon NT, et al. (1995) Negative luminescence from In1−xAlxSb and CdxHg1−xTe diodes. Infrared Phys Technol 36: 1037-1044. doi: 10.1016/1350-4495(95)00043-7
    [46] Zhu L, Fiorino A, Thompson D, et al. (2019) Near-field photonic cooling through control of the chemical potential of photons. Nature 566: 239-244. doi: 10.1038/s41586-019-0918-8
    [47] Rawat R, Lamba R, Kaushik SC (2017) Thermodynamic study of solar photovoltaic energy conversion: An overview. Renewable Sustainable Energy Rev 71: 630-638. doi: 10.1016/j.rser.2016.12.089
    [48] Ruan XL, Rand SC, Kaviany M (2007) Entropy and efficiency in laser cooling of solids. Phys Rev B 75: 214304. doi: 10.1103/PhysRevB.75.214304
    [49] Hsu WC, Tong JK, Liao B, et al. (2016) Entropic and near-field improvements of thermoradiative cells. Sci Rep 6: 34837. doi: 10.1038/srep34837
    [50] Gribik J, Osterle J (1984) The second law efficiency of solar energy conversion. J Sol Energy Eng 106: 16-21. doi: 10.1115/1.3267555
    [51] Pusch A, Gordon JM, Mellor A, et al. (2019) Fundamental efficiency bounds for the conversion of a radiative heat engine's own emission into work. Phys Rev Appl 12: 064018. doi: 10.1103/PhysRevApplied.12.064018
    [52] Li W, Buddhiraju S, Fan S (2020) Thermodynamic limits for simultaneous energy harvesting from the hot sun and cold outer space. Light Sci Appl 9: 68. doi: 10.1038/s41377-020-0296-x
    [53] Dorofeyev I (2011) Thermodynamic functions of fluctuating electromagnetic fields within a heterogeneous system. Phys Scr 84: 055003. doi: 10.1088/0031-8949/84/05/055003
    [54] Perez-Madrid A, Lapas LC, Rubi JM (2009) Heat exchange between two interacting nanoparticles beyond the fluctuation-dissipation regime. Phys Rev Lett 103: 048301. doi: 10.1103/PhysRevLett.103.048301
    [55] Pérez-Madrid A, Rubí JM, Lapas LC (2008) Heat transfer between nanoparticles: Thermal conductance for near-field interactions. Phys Rev B 77: 155417 doi: 10.1103/PhysRevB.77.155417
    [56] Latella I, Pérez-Madrid A, Lapas LC, et al. (2014) Near-field thermodynamics: Useful work, efficiency, and energy harvesting. J Appl Phys 115: 124307. doi: 10.1063/1.4869744
    [57] Narayanaswamy A, Zheng Y (2013) Theory of thermal nonequilibrium entropy in near-field thermal radiation. Phys Rev B 88: 075412. doi: 10.1103/PhysRevB.88.075412
    [58] Li B, Cheng Q, Song J, et al. (2020) Evaluation of performance of near-field thermophotovoltaic systems based on entropy analysis. J Appl Phys 127: 063103. doi: 10.1063/1.5135729
    [59] Ravindra NM, Srivastava VK (1979) Temperature dependence of the energy gap in semiconductors. J Phys Chem Solids 40: 791-793. doi: 10.1016/0022-3697(79)90162-8
    [60] Palik ED (1985) Handbook of Optical Constants of Solids, Orlando: Academic Press.
    [61] Johnson PB, Christy RW (1972) Optical Constants of the Noble Metals. Phys Rev B 6: 4370-4379. doi: 10.1103/PhysRevB.6.4370
    [62] Kruger M, Emig T, Kardar M (2011) Nonequilibrium electromagnetic fluctuations: heat transfer and interactions. Phys Rev Lett 106: 210404. doi: 10.1103/PhysRevLett.106.210404
    [63] Otey C, Fan S (2011) Numerically exact calculation of electromagnetic heat transfer between a dielectric sphere and plate. Phys Rev B 84: 245431. doi: 10.1103/PhysRevB.84.245431
    [64] Chew WC (1995) Waves and Fields in Inhomogeneous Media, New York: IEEE.
    [65] Basu S, Lee BJ, Zhang ZM (2010) Infrared radiative properties of heavily doped silicon at room temperature. J Heat Transfer 132: 023301. doi: 10.1115/1.4000171
    [66] Smith GB (1990) Theory of angular selective transmittance in oblique columnar thin films containing metal and voids. Appl Opt 29: 3685-3693. doi: 10.1364/AO.29.003685
    [67] Zhang Z (2007) Nano/microscale heat transfer, New York: McGraw-Hill.
  • This article has been cited by:

    1. A. N. M. Fuhadul Islam, S. Mostafa Ghiaasiaan, Zhuomin M. Zhang, Thermodynamic Limit of Electroluminescent Refrigeration Devices, 2025, 27, 1099-4300, 496, 10.3390/e27050496
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3334) PDF downloads(193) Cited by(1)

Figures and Tables

Figures(11)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog