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

Near-field thermal rectification devices using phase change periodic nanostructure

Open Access Open Access

Abstract

We theoretically analyze two near-field thermal rectification devices: a radiative thermal diode and a thermal transistor that utilize a phase change material to achieve dynamic control over heat flow by exploiting metal-insulator transition of VO2 near 341 K. The thermal analogue of electronic diode allows high heat flow in one direction while it restricts the heat flow when the polarity of temperature gradient is reversed. We show that with the introduction of 1-D rectangular grating, thermal rectification is dramatically enhanced in the near-field due to reduced tunneling of surface waves across the interfaces for negative polarity. The radiative thermal transistor also works around phase transition temperature of VO2 and controls heat flow. We demonstrate a transistor-like behavior wherein heat flow across the source and the drain can be greatly varied by making a small change in gate temperature.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Thermal analogues of electronic rectification devices such as thermal diodes [1], thermal transistors [2] and memory elements [3] have been studied during past decade. During early years, conduction based devices were discussed [6–8], while more recently radiative thermal rectification devices have gathered considerable attention [4,5, 9–12]. In this work, we analyse radiative thermal rectification devices viz. thermal diode and thermal transistor. Thermal rectification has a wide variety of potential applications such as thermal management and energy storage, thermal circuits [13], thermal logic gates [14] and information processing [15].

Analogous to an electrical diode, a thermal diode has two terminals where a temperature bias is applied. The thermal diode exhibits a high degree of asymmetry in the magnitude of heat flow depending on the applied temperature bias. The goal is to increase this asymmetry as much as possible. Exploiting temperature dependent properties of phase changing materials such as La0.7Ca0.15Sr0.15MnO3 (LCSMO) and vanadium dioxide (VO2) [10, 16, 17] is a common theme to achieve radiative thermal rectification. In the far-field, thermal rectification is achieved by change in emissivity/reflectivity of a phase change structure. In the near-field, difference in the degree of tunneling of surface waves between structures dictates thermal rectification. Several studies focus on modulation of radiative heat transfer in near-field regime, while many others deal with far-field thermal radiation. In general it has been observed that a higher rectification can be achieved in the near-field regime. Following Ben-Abdallah and Biehs’s work on a VO2 based simple far-field radiative thermal diode, numerous studies on thermal rectification [16–19], thermal transistors [20], negative differential conductance [21] have been published. While materials such as SiC, AIST (an alloy of Ag, In, Sb, and Te) have been investigated, VO2 is arguably the most commonly used material in the context of radiative thermal rectification. Modulation in near-field thermal radiative heat transfer upon phase change of VO2 has also been experimentally demonstrated [22, 23].

A thermal transistor is a device that has three terminals: source, drain and gate. A temperature bias is applied across the the source and the drain while gate temperature is varied. By controlling the temperature of gate one can control the heat flow across the other two terminals. Li et al. [2] introduced a phonon based thermal analogue of electronic transistor. Later, concepts of thermal logic gates [14] and thermal memory devices [15] was also demonstrated. More recently, concepts of radiative thermal transistors in the near-field and far-field regimes were also presented. [5, 24].

Present study focuses on rectification in a near-field thermal diode and a near-field thermal transistor that employ VO2 as phase-transition material. VO2 can be reversibly switched in a very short amount of time (∼ 100 fs) from an insulating state to a metallic state [25]. To enhance the rectification, one may use several different materials. It has been observed that metamaterials such as dielectric mixtures [26] and grating structures [27] can manipulate radiative transfer in near-field and can be considered for enhancement of thermal rectification. Previous studies on radiative thermal rectification have been devoted mainly to bulk materials and thin films. Authors’ previous work [28] dealt with concepts of near-field thermal diode with a detailed discussion about effect of bulk VO2, thin film, rectangular and triangular gratings on thermal rectification at various separations. In the present study, grating-enhanced near-field thermal diode introduced in [28] is revisited and the concept is extended to near-field thermal transistor. Our calculations indicate that thermal rectification in thermal diode and heat modulation in thermal transistor can be raised significantly by using 1-D surface gratings of VO2.

A typical near-field thermal diode consists of two planar structures at a separation distance less than thermal wavelength [16]. One structure (hereafter referred to as active terminal) has a phase change material and its counterpart has fixed material properties (passive terminal). Figure 1(a) introduces the concepts of thermal diode proposed in this study that has two structures at a distance of L = 100 nm. The active terminal contains top layer of phase change material VO2 at temperature T1 = 341 K + ΔT. On the passive side, structure 2 has its temperature T2 = 341 K − ΔT. Mean temperature is chosen to be the phase transition temperature of VO2 at 341 K. When T1 > T2 (referred to as forward bias), VO2 layer is in metallic phase; when T1 < T2 (reverse bias), VO2 layer is in insulator phase with its optical axis aligned along the distance between them. This concept can be extended to thermal transistor as shown in Fig. 1(b). It has three terminals. The source and the drain are same as the passive side of thermal diode. The gate has same structure on both faces as the active side of thermal diode. In present study, source and drain temperatures are fixed above and below the phase transition temperature of VO2, while gate temperature is varied. Proposed configurations will be discussed in more detail along with results.

 figure: Fig. 1

Fig. 1 Schematics of (a) near-field thermal diode and (b) near-field thermal transistor. Active side of the diode and both the faces of the gate of the transistor have top layer of 1-D rectangular grating made of VO2 of height h, width w, period Λ and filling ratio ϕ on a gold layer deposited on a substrate. The passive side of thermal diode and the source and the drain of the transistor has 1 μm layer of boron nitride (BN) on 1 μm layer of gold on a substrate.

Download Full Size | PDF

Although 341 K is said to be the phase transition temperature of VO2. the phase transition is not homogeneous nor does it happen instantaneously [18, 25, 29]. We calculate thermal rectification values for the thermal diode at a minimal temperature difference of 20 K (ΔT = ±10 K), as it reflects intrinsic properties of the proposed device around its intended temperature of use. Choice of these temperatures assure that VO2 is in either completely metallic or completely insulator phase. For thermal transistor calculations, we considered a gradual change in optical properties of VO2 from 341 K to 345 K rather than modelling it as a sudden change. This provides a better insight into transistor-like behavior of proposed device.

2. Theoretical fundamentals

A well known formula derived from dyadic Green’s function formalism [30] has been employed to calculate near-field radiative transfer between planar surfaces of thermal diode and thermal transistor.

Q12(T1,T2,L)=0dω2π[Θ(ω,T1)Θ(ω,T2)]T12(ω,L)
where Θ(ω, T) = (ℏω/2) coth(ℏω/2kBT) is the energy of harmonic oscillator at frequency ω and temperature T, is the reduced Planck constant, and kB is the Boltzmann constant. The function T1→2(ω, L) is known as the spectral transmissivity in radiative transfer between media 1 and 2 separated of distance L [18, 26, 27, 30]. The function T1→2(ω) corresponds to the spectral transmissivity in radiative transfer between media 1 and 2 separated by distance L and is expressed as
T12(ω)=0kρdkρ2πξ(ω,kρ)

Here, kρ is the parallel component of wavevector and the integrand is known as energy transmission coefficient and can be defined as

ξ(ω,kρω/c)=μ=s,p(1|R˜1(μ)|2)(1|R˜2(μ)|2)|1R˜1(μ)R˜2(μ)e2jkzL|2
ξ(ω,kρ>ω/c)=μ=s,p4J(R˜1μ)J(R˜2(μ))e2|kz|L|1R˜1(μ)R˜2(μ)e2|kz|L|2
where R˜1(μ) and R˜2(μ) are polarization dependent reflection coefficients of the two half spaces, μ = s (or p) refers to transverse electric (or magnetic) polarization, and kz is the z-component of wavevector in vacuum. kρω/c and kρ > ω/c correspond to propagating and evanescent modes, respectively.

As our proposed designs involve 1-D grating structure of VO2 in vacuum, we use second order approximation of effective medium theory to obtain the dielectric properties given by the expressions [31–33]

εTE,2=εTE,0[1+π23(Λλ)2ϕ2(1ϕ)2(εAεB)2εTE,0]
εTM,2=εTM,0[1+π23(Λλ)2ϕ2(1ϕ)2(εAεB)2εTB,0(εTM,0εAεB)2]
where εA and εB are dielectric functions of the two materials (VO2 and vacuum) in surface gratings, λ is the wavelength, Λ is grating period and filling ratio ϕ = w/Λ where w is the width of VO2 segment. The expressions for zeroth order effective dielectric functions εTE,0 and εTM,0 are given by [31,34]
εTE,0=ϕεA+(1ϕ)εB
εTM,0=(ϕεA+1ϕεB)1

In this study, temperatures involved are around 341 K and grating period (50 nm) is much less than the thermal wavelength (∼ 8.5 μm). Thus, the condition for effective medium approximation [33] is satisfied. For the near-field radiative transfer pertaining to periodic gratings, effective medium theory is valid for gaps larger than the grating period [35]. If the condition for EMT is not satisfied, more rigorous numerical methods such as discussed in Lussange et al. [36] and Guérout et al. [37] should be used.

The insulating state of VO2 (below 341 K) is anisotropic and supports several phonon modes. Such anisotropic optical response can be described using separate classical oscillator formula ε(ω)=ε+i=1NSiωi2ωi2jγiωω2 for ordinary and extraordinary mode of dielectric function. In our case, ordinary mode εO is along the plane perpendicular to the optical axis (xy plane) and extraordinary mode εE is along the optical axis z axis. Note that, for both the thermal diode and thermal transistor configuration, the optical axis is assumed to be along the distance between the bodies. Barker et al. [38] provides experimental values of high-frequency constant ε, phonon frequency ωi, scattering rate γi and oscillator strength Si. In the metallic state, VO2 is isotropic and the dielectric function follows Drude model [38] given by ε(ω)=ωp2εω2jωΓ. The dielectric function for BN is taken from Palik [39] and is of the form ε(ω)=ε(ω2ωLO2+jωγ)(ω2ωTO2+jωγ). Here ωTO and ωLO are known as transverse and longitudinal optical phonon frequencies respectively, and γ is the damping constant. The values of ε, ωTO, ωLO and γ for BN are 4.46, 0.1309 eV, 0.1616 eV and 6.55 × 10−4 eV respectively. Dielectric properties of gold can be found in Johnson and Christy [40].

3. Results

First, we will consider the thermal diode concept shown in Fig. 1(a). To quantify rectification, the rectification ratio R can be defined as R = (QfQr)/Qr where Qf and Qr refer to forward and reverse heat flux respectively [41]. Alternatively, the rectification coefficient can be defined as η = (QfQr)/max(Qr, Qf). The goal is to increase R as much as possible so that QfQr and rectification is reasonably high to be utilized in a practical device. Fig. 2 highlights thermal rectification characteristics of thermal diodes. Heat flux is plotted against temperature difference between active and passive sides. Two cases are considered, the first is a simple diode with bulk VO2 and bulk BN while the second is the configuration shown in Fig. 1. 1-D rectangular grating has a period (Λ) of 50 nm, filling ratio (ϕ) of 0.3 and height (h) of 0.5 μm. In both the cases, the separation between the two sides is 100 nm. Owing to different optical properties of insulator VO2 and metallic VO2, weak rectification exists in the bulk thermal diode device as can be seen from Fig. 2. For second case, when 1-D grating is used, significant enhancement in rectification is seen and rectification value is around 14 at the gap of 100 nm and temperature difference of 20 K (ΔT =10 K).

 figure: Fig. 2

Fig. 2 Rectification characteristics of the proposed thermal diode highlighted by heal flux plot against temperature difference. Positive temperature difference corresponds to forward bias while negative temperature difference corresponds to reverse bias.

Download Full Size | PDF

In order to understand near-field thermal rectification across proposed thermal diode, we plot energy transmission coefficient ξ(ω, kρ) across the interfaces of the configuration considered (Figs. 3(a)) at a gap of 100 nm. Here kρc/ω is normalized parallel wavevector. In the forward bias configuration, the transmission coefficient is close to unity well beyond light line (kρc/ω = 1). The two prominent frequencies (highlighted by dashed lines) seen where transmission coefficient is high are close to characteristic wavelengths of BN (7.6 μm and 9.8 μm). Since metallic VO2 does not support surface phonon polariton in the infrared region [18], the symmetric and antisymmetric surface phonons supported by the BN layer (kρc/ω ≫ 1 in Fig. 3) are primary modes of near-field radiative transfer in forward bias. In addition, there exists a secondary contribution due to Fabry-Perot modes and frustrated modes (kρc/ω ≈ 1) in the near-field regime. High energy transmission in forward bias is due to tunneling of surface waves across interfaces. In reverse bias, although both BN and insulator VO2 support surface phonon modes, they do not overlap and near-field radiative transfer is dominated by non-resonant surface waves. Note that although metallic VO2 does not support surface phonons, surface phonons of BN exists in a range where metallic VO2 has high extinction coefficient. On the contrary, surface phonons of insulator VO2 and BN occur in frequency range where the opposite side has low extinction coefficient (κ ≈ 0). This results in a mismatch of phonon frequencies. As a result, tunneling between BN and insulator VO2 is much weaker than that between BN and metallic VO2, that leads to thermal rectification. Furthermore, when bulk VO2 is replaced by a 1-D rectangular grating, the transmission coefficient for reverse bias is reduced due to the presence of grating which suppresses the tunneling of surface waves supported by insulator VO2 and BN. Tunneling between BN and metallic VO2 however is relatively unchanged. Consequently, a higher rectification is achieved. Resulting difference in spectral heat fluxes can also be observed. More detailed analysis of effect of 1-D gratings on thermal rectification when compared to bulk and thin films can be found in Ghanekar et al. [19].

 figure: Fig. 3

Fig. 3 Coefficient of energy transmission ξ(ω, kρ) across the the two interfaces of thermal diode with 1-D rectangular grating plotted against angular frequency ω and normalized parallel wavevector kρc/ω for (a) forward bias, (b) reverse bias cases.

Download Full Size | PDF

Next, we consider an extension of thermal diode concept that is, a thermal transistor. Only a specific case of thermal transistor is analysed. It has a structure similar to the thermal diode discussed earlier. The source and the drain has 1 μm layer of BN over 1 μm layer of gold. The gate has two faces and each face has a 0.5 μm thick 1-D rectangular grating of VO2 with period Λ = 50 nm and filling ratio ϕ = 0.3. The gate is separated by 100 nm from the source and the drain. The optical axis of insulating VO2 lies along the separation. Source temperature TS and drain temperature TD are fixed, while gate temperature is varied. To realistically model properties of VO2 during phase transition, effective medium approximation is used. As the temperature is increased, metallic puddles are formed whose volume fraction gradually increased to 1 at 345 K [25]. Alternatively, Bruggeman approximation can also be used. As seen in Ben-Abdallah and Biehs [5], both EMT and Bruggeman approximation lead to very similar results.

Figure 4 displays transistor characteristics of the proposed thermal transistor for two different cases of source and drain temperature. Source and drain temperatures are kept constant while gate temperature is varied. Heat flux across the source (QS), the drain (QD) and the gate (QG) are plotted against gate temperature TG. Under steady state operation, QS = QD + QG. In Fig. 4(a), source temperature TS and drain temperature TD are 317 K and 321 K, respectively. Upon phase transition of VO2, a small change in gate heat flux QG and a large change in QD is seen. Before and after the phase transistor region, the heat transfer coefficient for QS and QD (e.g. QS/(TSTG) for QS) is nearly constant, owing to linear temperature dependence of radiative heat transfer for small temperature differences. During phase transition however, the heat transfer coefficient varies with TG. This is the transistor-like behavior. Note that earlier studies Ben-Abdallah and Biehs [5] and Prod’homme et al. [20] show similar transistor characteristics with exactly opposite trend. For instance, increase in gate temperature leads to decrease in heat flux during phase transition while in present study it increases. This transistor-like behavior is simply inherited from corresponding thermal diodes that would also show an opposite trend. Present transistor behaves more like an amplifier than a switch. It can be observed that, by changing the temperatures TS and/or TD, one can change the sensitivity of the transistor. In Fig. 4(a), QG is positive meaning heat is flowing out of the gate. This setting is not desirable as cooling is more complex than heating. Hence in Fig, 4(b), we consider a case where heat is flowing into the gate. This configuration is achieved simply because TSTG < TGTD. Here TS and TG are chosen to be 361 K and 311 K, respectively. As seen in Prod’homme et al. [20], for some combination of TD and TS, QG varies monotonically (no local maxima). However, for any combination of TS and TD, local maxima of QG if exists, must occur within the transition zone of VO2. Transistor-like amplification can be characterized by amplification factor α defined as [24],

α=|dQDdQG|

 figure: Fig. 4

Fig. 4 Heat flux across source, drain and gate is plotted against gate temperature while source and drain temperatures are fixed. (a) TS = 371 K and TD = 321 K, (b) TS = 361 K and TD = 311 K.

Download Full Size | PDF

Note that, when QG reaches a maxima, δQG ≈ 0. This produces extremely high value of α which is not a true representative of the overall transistor behavior. Therefore we define an alternative way to quantify amplification,

α=|ΔQDΔQG|
where, ΔQG is the maximum change in QG during phase transition and ΔQD is the corresponding change in QD. The points corresponding to ΔQG and ΔQD are highlighted in Fig. 4. For Fig. 4(a) and 4(b), α using this definition is 2.24 and 2.36, respectively. The fundamental difference between the thermal diode and the thermal transistor is that the transistor would behave the same way even when polarity of source and drain is reversed. If TS < TG < TD, transistor characteristics are essentially unchanged, meaning source and drain can be flipped.

4. Conclusion

Thus we have theoretically demonstrated that an enhanced thermal rectification can be achieved in a thermal diode and a thermal transistor by using 1-D grating of phase change material VO2. For the near-field thermal diode considered in the study, we predict a reasonably high value of rectification ratio (∼ 14) that can be obtained at a gap of 100 nm. Improved rectification is attributed to reduced tunneling of surface waves across the interfaces for reverse bias. The concept of thermal diode was extended to thermal transistor with a source, a drain and a gate. Source and drain temperatures are fixed above and below the phase transition temperature of VO2 respectively, while gate temperature is varied. The device displays a typical transistor-like behaviour wherein a small change in gate heat flux leads to a large change in drain heat flux. Essentially, the gate acts like a tap that controls heat flow across the source and the drain.

Funding

National Science Foundation (1655221); Institutional Development Award (IDeA) Network for Biomedical Research Excellence from the National Institute of General Medical Sciences of the National Institutes of Health (P20GM103430); and Rhode Island Foundation Research Grant (20164342).

References and links

1. B. Li, L. Wang, and G. Casati, “Thermal diode: Rectification of heat flux,” Phys. Rev. Lett. 93, 184301 (2004). [CrossRef]   [PubMed]  

2. B. Li, L. Wang, and G. Casati, “Negative differential thermal resistance and thermal transistor,” Appl. Phys. Lett. 88, 143501 (2006). [CrossRef]  

3. L. Wang and B. Li, “Thermal memory: a storage of phononic information,” Phys. Rev. Lett. 101, 267203 (2008). [CrossRef]  

4. P. Ben-Abdallah and S.-A. Biehs, “Contactless heat flux control with photonic devices,” AIP Advances 5, 053502 (2015). [CrossRef]  

5. P. Ben-Abdallah and S.-A. Biehs, “Near-field thermal transistor,” Phys. Rev. Lett. 112, 044301 (2014). [CrossRef]   [PubMed]  

6. W. Kobayashi, Y. Teraoka, and I. Terasaki, “An oxide thermal rectifier,” Appl. Phys. Lett. 95, 171905 (2009). [CrossRef]  

7. D. Sawaki, W. Kobayashi, Y. Moritomo, and I. Terasaki, “Thermal rectification in bulk materials with asymmetric shape,” arXiv preprint arXiv:1102.4182 (2011).

8. W. Kobayashi, D. Sawaki, T. Omura, T. Katsufuji, Y. Moritomo, and I. Terasaki, “Thermal rectification in the vicinity of a structural phase transition,” Applied Physics Express 5, 027302 (2012). [CrossRef]  

9. P. Ben-Abdallah and S.-A. Biehs, “Phase-change radiative thermal diode,” Appl. Phys. Lett. 103, 191907 (2013). [CrossRef]  

10. C. R. Otey, W. T. Lau, and S. Fan, “Thermal rectification through vacuum,” Phys. Rev. Lett. 104, 154301 (2010). [CrossRef]   [PubMed]  

11. Z. Chen, C. Wong, S. Lubner, S. Yee, J. Miller, W. Jang, C. Hardin, A. Fong, J. E. Garay, and C. Dames, “A photon thermal diode,” Nat. Commun. 5, 5446 (2014). [CrossRef]  

12. L. Tang and M. Francoeur, “Photonic thermal diode enabled by surface polariton coupling in nanostructures,” arXiv preprint arXiv:1703.09862 (2017).

13. L. Wang and B. Li, “Phononics gets hot,” Phys. World 21, 27–29 (2008). [CrossRef]  

14. L. Wang and B. Li, “Thermal logic gates: computation with phonons,” Phys. Rev. Lett. 99, 177208 (2007). [CrossRef]   [PubMed]  

15. N. Li, J. Ren, L. Wang, G. Zhang, P. Hänggi, and B. Li, “Colloquium: Phononics: Manipulating heat flow with electronic analogs and beyond,” Reviews of Modern Physics 84, 1045 (2012). [CrossRef]  

16. Y. Yang, S. Basu, and L. Wang, “Radiation-based near-field thermal rectification with phase transition materials,” Appl. Phys. Lett. 103, 163101 (2013). [CrossRef]  

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

18. Y. Yang, S. Basu, and L. Wang, “Vacuum thermal switch made of phase transition materials considering thin film and substrate effects,” Journal Quant. Spectrosc. Rad. Trans. 158, 69–77 (2015). [CrossRef]  

19. A. Ghanekar, G. Xiao, and Y. Zheng, “High contrast far-field radiative thermal diode,” Sci. Rep. 7, 6339 (2017). [CrossRef]   [PubMed]  

20. H. Prod’homme, J. Ordonez-Miranda, Y. Ezzahri, J. Drevillon, and K. Joulain, “Optimized thermal amplification in a radiative transistor,” J. Appl. Phys. 119, 194502 (2016). [CrossRef]  

21. P. Van Zwol, K. Joulain, P. Ben-Abdallah, J.-J. Greffet, and J. Chevrier, “Fast nanoscale heat-flux modulation with phase-change materials,” Phys. Rev. B 83, 201404 (2011). [CrossRef]  

22. P. Van Zwol, L. Ranno, and J. Chevrier, “Tuning near field radiative heat flux through surface excitations with a metal insulator transition,” Phys. Rev. Lett. 108, 234301 (2012). [CrossRef]   [PubMed]  

23. F. Menges, M. Dittberner, L. Novotny, D. Passarello, S. Parkin, M. Spieser, H. Riel, and B. Gotsmann, “Thermal radiative near field transport between vanadium dioxide and silicon oxide across the metal insulator transition,” Appl. Phys. Lett. 108, 171904 (2016). [CrossRef]  

24. K. Joulain, Y. Ezzahri, J. Drevillon, and P. Ben-Abdallah, “Modulation and amplification of radiative far field heat transfer: Towards a simple radiative thermal transistor,” Appl. Phys. Lett. 106, 133505 (2015). [CrossRef]  

25. M. Qazilbash, M. Brehm, G. Andreev, A. Frenzel, P.-C. Ho, B.-G. Chae, B.-J. Kim, S. J. Yun, H.-T. Kim, A. Balatsky, and O. Shpyrko, “Infrared spectroscopy and nano-imaging of the insulator-to-metal transition in vanadium dioxide,” Phys. Rev. B 79, 075107 (2009). [CrossRef]  

26. A. Ghanekar, L. Lin, J. Su, H. Sun, and Y. Zheng, “Role of nanoparticles in wavelength selectivity of multilayered structures in the far-field and near-field regimes,” Opt. Express 23, A1129–A1139 (2015). [CrossRef]   [PubMed]  

27. S.-A. Biehs, F. S. Rosa, and P. Ben-Abdallah, “Modulation of near-field heat transfer between two gratings,” Appl. Phys. Lett. 98, 243102 (2011). [CrossRef]  

28. A. Ghanekar, J. Ji, and Y. Zheng, “High-rectification near-field thermal diode using phase change periodic nanostructure,” Appl. Phys. Lett. 109, 123106 (2016). [CrossRef]  

29. A. Frenzel, M. M. Qazilbash, M. Brehm, B.-G. Chae, B.-J. Kim, H.-T. Kim, A. Balatsky, F. Keilmann, and D. Basov, “Inhomogeneous electronic state near the insulator-to-metal transition in the correlated oxide vo 2,” Phys. Rev. B 80, 115115 (2009). [CrossRef]  

30. A. Narayanaswamy and Y. Zheng, “A Green’s function formalism of energy and momentum transfer in fluctuational electrodynamics,” J. Quant. Spectrosc. Rad. Trans. 132, 12–21 (2014). [CrossRef]  

31. D. H. Raguin and G. M. Morris, “Antireflection structured surfaces for the infrared spectral region,” Appl. Opt. 32, 1154–1167 (1993). [CrossRef]   [PubMed]  

32. P. Lalanne and D. Lemercier-Lalanne, “On the effective medium theory of subwavelength periodic structures,” J. Mod. Opt. 43, 2063–2085 (1996). [CrossRef]  

33. Y.-B. Chen, Z. Zhang, and P. Timans, “Radiative properties of patterned wafers with nanoscale linewidth,” J. Heat Trans. 129, 79–90 (2007). [CrossRef]  

34. E. Glytsis and T. K. Gaylord, “High-spatial-frequency binary and multilevel stairstep gratings: polarization-selective mirrors and broadband antireflection surfaces,” Appl. Opt. 31, 4459–4470 (1992). [CrossRef]   [PubMed]  

35. X. Liu, T. Bright, and Z. Zhang, “Application conditions of effective medium theory in near-field radiative heat transfer between multilayered metamaterials,” J. Heat Trans. 136, 092703 (2014). [CrossRef]  

36. J. Lussange, R. Guérout, F. S. Rosa, J.-J. Greffet, A. Lambrecht, and S. Reynaud, “Radiative heat transfer between two dielectric nanogratings in the scattering approach,” Phys. Rev. B 86, 085432 (2012). [CrossRef]  

37. R. Guérout, J. Lussange, F. Rosa, J.-P. Hugonin, D. Dalvit, J.-J. Greffet, A. Lambrecht, and S. Reynaud, “Enhanced radiative heat transfer between nanostructured gold plates,” in “Journal of Physics: Conference Series,” vol. 395 (IOP Publishing, 2012), p. 012154.

38. A. Barker Jr, H. Verleur, and H. Guggenheim, “Infrared optical properties of vanadium dioxide above and below the transition temperature,” Phys. Rev. Lett. 17, 1286 (1966). [CrossRef]  

39. E. D. Palik, Handbook of Optical Constants of Solids, vol. 3 (Academic Press, 1998).

40. P. B. Johnson and R.-W. Christy, “Optical constants of the noble metals,” Phys. Rev. B 6, 4370 (1972). [CrossRef]  

41. B. Song, A. Fiorino, E. Meyhofer, and P. Reddy, “Near-field radiative thermal transport: From theory to experiment,” AIP Advances 5, 053503 (2015). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (4)

Fig. 1
Fig. 1 Schematics of (a) near-field thermal diode and (b) near-field thermal transistor. Active side of the diode and both the faces of the gate of the transistor have top layer of 1-D rectangular grating made of VO2 of height h, width w, period Λ and filling ratio ϕ on a gold layer deposited on a substrate. The passive side of thermal diode and the source and the drain of the transistor has 1 μm layer of boron nitride (BN) on 1 μm layer of gold on a substrate.
Fig. 2
Fig. 2 Rectification characteristics of the proposed thermal diode highlighted by heal flux plot against temperature difference. Positive temperature difference corresponds to forward bias while negative temperature difference corresponds to reverse bias.
Fig. 3
Fig. 3 Coefficient of energy transmission ξ(ω, kρ) across the the two interfaces of thermal diode with 1-D rectangular grating plotted against angular frequency ω and normalized parallel wavevector kρc/ω for (a) forward bias, (b) reverse bias cases.
Fig. 4
Fig. 4 Heat flux across source, drain and gate is plotted against gate temperature while source and drain temperatures are fixed. (a) TS = 371 K and TD = 321 K, (b) TS = 361 K and TD = 311 K.

Equations (10)

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

Q 1 2 ( T 1 , T 2 , L ) = 0 d ω 2 π [ Θ ( ω , T 1 ) Θ ( ω , T 2 ) ] T 1 2 ( ω , L )
T 1 2 ( ω ) = 0 k ρ d k ρ 2 π ξ ( ω , k ρ )
ξ ( ω , k ρ ω / c ) = μ = s , p ( 1 | R ˜ 1 ( μ ) | 2 ) ( 1 | R ˜ 2 ( μ ) | 2 ) | 1 R ˜ 1 ( μ ) R ˜ 2 ( μ ) e 2 j k z L | 2
ξ ( ω , k ρ > ω / c ) = μ = s , p 4 J ( R ˜ 1 μ ) J ( R ˜ 2 ( μ ) ) e 2 | k z | L | 1 R ˜ 1 ( μ ) R ˜ 2 ( μ ) e 2 | k z | L | 2
ε T E , 2 = ε T E , 0 [ 1 + π 2 3 ( Λ λ ) 2 ϕ 2 ( 1 ϕ ) 2 ( ε A ε B ) 2 ε T E , 0 ]
ε T M , 2 = ε T M , 0 [ 1 + π 2 3 ( Λ λ ) 2 ϕ 2 ( 1 ϕ ) 2 ( ε A ε B ) 2 ε T B , 0 ( ε T M , 0 ε A ε B ) 2 ]
ε T E , 0 = ϕ ε A + ( 1 ϕ ) ε B
ε T M , 0 = ( ϕ ε A + 1 ϕ ε B ) 1
α = | d Q D d Q G |
α = | Δ Q D Δ Q G |
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.