Highly-efficient radiative thermal rectifiers based on near-field gap variations

Bei Yang ab and Qing Dai *ab
aCAS Key Laboratory of Nanophotonic Materials and Devices, CAS Key Laboratory of Standardization and Measurement for Nanotechnology, CAS Center for Excellence in Nanoscience, National Center for Nanoscience and Technology, Beijing 100190, China. E-mail: daiq@nanoctr.cn.nm
bCenter of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, Beijing 100049, China

Received 9th August 2022 , Accepted 19th October 2022

First published on 21st October 2022


Abstract

Near-field radiative thermal rectifiers (NFRTRs) enabling directional heat transport hold great promise for various applications, including thermal logic computing, thermal management, and energy conversion. Current NFRTR designs rely on dissimilar terminal materials with high contrasts in their temperature-dependent dielectric properties, which in turn hinders the spectral match for radiative heat transfer and thus limits the device's efficiency. Herein, this dilemma is solved by designing heterostructures where a pair of polaritonic layers are separately stacked on a thermally-expanding layer and a rigid substrate, spaced by a vacuum gap. In this scheme, the symmetric polaritonic layers can provide stable near-field radiative channels for heat transfer, while the thermally-expanding layer can modulate the gap size with flipped temperature bias to allow high contrasts in heat flux. In exemplified implementations, the hBN-based design has achieved a record-high thermal rectification factor (TRF, ∼104) even under small thermal gradients (∼20 K), which can be further boosted by polaritonic hybridizations in the graphene/hBN-based design. This study paves the way to design novel NFRTRs with 2D materials, thus providing enriched polaritons to realize higher TRFs.


1. Introduction

Active control of heat transport at the nanoscale is vital for various applications, including thermal logic computing, thermal management, and energy conversion. One key component to realizing thermal control in solid-state devices is the thermal rectifier, which consists of two terminals and enables heat transport in preferential directions.1 To evaluate its performance, a figure of merit termed thermal rectification factor (TRF) is defined as TRF = (QfQr)/Qr, where Qf and Qr denote the heat flux under the forward and reverse temperature bias, respectively. Thermal rectifiers with TRFs greater than ten are required for practical use,2 but this value can hardly be achieved with conventional phonon-based thermal rectifiers. The reason is that their contact heat-conducting channels between the two terminals3–6 suffer from limitations such as low group velocity of acoustic phonons, interfacial phonon scattering, and multiple phonon–phonon interactions.7,8 In contrast, near-field radiative thermal rectifiers (NFRTRs), based on photon tunneling through non-contact terminals, can overcome these limitations to yield much higher efficiency and have gained growing popularity in recent years.

The TRF of a NFRTR is determined by the temperature-dependent electromagnetic (EM) properties of its terminal materials.9 The basic requirement is that they should enable a spectral match to enhance the near-field radiative heat flux under a given forward temperature bias, while the spectral match should diminish due to their varied temperature dependence when the bias is reversed, thus strongly attenuating the heat flux. In this way, an asymmetry in heat flux between the forward and reverse scenarios is created and thus high TRFs can be achieved. To date, numerous NFRTRs by pairing dissimilar terminals among various polar dielectrics,10,11 low-dimensional materials,12,13 and phase change materials14–17 have been explored. However, a contradiction in these designs is that these asymmetrical material pairings with mismatched EM properties significantly hinder the spectral match,18,19 thereby limiting the radiative heat flux. Besides, the heat transport asymmetry in these systems remains weak due to the limited temperature dependence of the materials’ EM properties. Although three-body systems have recently been proposed to strengthen this asymmetry by introducing many-body interactions, enabling thermal rectification even without temperature-dependent EM properties, the obtained TRFs remain low (∼1).20 Therefore, it is still a challenging task to design rational terminal pairings for NFRTRs to achieve TRFs higher than 100.

In the present study, a design scheme for NFRTRs is proposed to decouple the above-mentioned antithetical requirements for terminal materials by using distinct gap-isolated heterostructures, where a pair of layers supporting infrared (IR) polaritons are separately stacked on a thermally-expanding layer and a rigid substrate. The symmetrical polaritonic layers can provide stable near-field radiative channels for heat flux, while the thermally-expanding layer can modulate the gap size with flipped temperature bias to allow high contrasts in heat flux due to its sharp gap-size dependence.18,21–23 A record-high TRF (∼104) has been achieved with the pairing of hBN/PDMS/SiO2 and hBN/SiO2 due to the thermally excited phonon polaritons, and it can be further boosted by the effective polaritonic coupling in graphene/hBN-based heterostructures. These TRFs have demonstrated an improvement of 1–4 orders of magnitude over previous studies (Table S1 in the ESI), indicating the tremendous potential of the proposed scheme to design highly-efficient NFRTRs. It is quite impressive that this scheme depends on the spatial distributions of evanescent fields mediated by surface polaritons (SPs) across the vacuum gap rather than their spectral distributions, thus providing a rectification mechanism distinct from previous studies. Such a mechanism not only improves the performance of a photon-based thermal rectifier but also derives high design flexibility by taking advantage of enriched polaritons in 2D materials.

2. Methodology

2.1 Model systems

The proposed NFRTR design and its working principle are schematically illustrated in Fig. 1. The kernel in this design is the two distinct terminal heterostructures separated by a vacuum gap (d), where a pair of thin polaritonic layers are separately stacked on a thermally-expanding layer and a rigid substrate. Each heterostructure consists of at least two functional layers. As for the hot terminal heterostructure under the forward scenarios, the active layer with a large coefficient of thermal expansion (CTE) can modulate the gap size in response to the temperature changes, while the radiative layer supporting IR polaritons can provide near-field radiative channels for heat flux. More diverse configurations can be obtained by incorporating additional layers into each functional layer. Hence, this scheme is quite general due to its flexible modular designs.
image file: d2nr04350e-f1.tif
Fig. 1 Schematic illustration of the proposed NFRTR design and its working principle.

Assuming that fixed constraints or other feasible measures are applied to the active terminal, directional thermal expansion/contraction and desired gap size variation can be obtained. In the forward scenarios (left panel in Fig. 1), the vacuum gap (df) decreases due to the downward movement of the hot terminal caused by the thermal expansion of the active layer, resulting in greatly enhanced heat flux. When the temperature bias is reversed (right panel in Fig. 1), the vacuum gap (dr) enlarges due to the thermal contraction of the active layer, thus attenuating the heat flux.

In the following sections, specific material combinations are applied to implement this scheme and to demonstrate its thermal rectification potential.

2.2 Computational techniques

Within the theoretical framework of fluctuational electrodynamics,24 the near-field radiative heat flux Q across the vacuum gap d between bi-planar heterostructures (Fig. 1) can be computed as
 
image file: d2nr04350e-t1.tif(1)
where Θ(ω,T) = ħω/[exp(ħω/kBT) − 1] is the mean energy of a Planck oscillator; ω is the angular frequency; β is the magnitude of the in-plane wavevector; Th and Tc are the equilibrated temperatures of the two terminals; and ξ(ω,β) is the energy transmission coefficient for s- and p-polarized waves, which is also termed the photon tunneling probability (PTP) and given as
 
image file: d2nr04350e-t2.tif(2)
where the superscripts 1 and 2 denote the emitter and receiver, respectively; k0 = ω/c is the vacuum wavevector; kz0 is the out-of-plane wavevector given by kz0 = (k02β2)1/2; and Rn is the Fresnel reflection coefficient of each terminal heterostructure, which can be derived for multilayers by using the scattering matrix method.25,26 In eqn (2), the far-field (βk0, propagating modes) contribution to the heat flux is clearly separated from the near-field (β > k0, evanescent modes) one. Due to the exponential damping of evanescent modes, β is limited at a given frequency to a value below the cutoff wavevector kc.27,28 More details about this calculation can be found in the ESI.

2.3 Materials properties

2.3.1 Thermally expanding materials. Thermal expansion is a material's tendency to change its volume in response to temperature variation. When a material is heated, the kinetic energy of molecules increases, causing them to move more and, in most cases, maintain a greater average separation. It is an intrinsic property of a material that can be quantified using the coefficient of thermal expansion (CTE). The required linear CTE can be calculated as
 
image file: d2nr04350e-t3.tif(3)
where L is the linear dimension (length or thickness) and image file: d2nr04350e-t4.tif is the change rate of the linear dimension per unit change in temperature. Polymers have much higher CTEs than inorganic materials and can expand to a greater extent when heated to the same temperature.

Here, polydimethylsiloxane (PDMS) and silica are selected as the active layer and the rigid substrate, respectively. PDMS is a silicone polymer commonly used in biomedical, microfluidic, or lab-on-a-chip applications due to its low cost, rapid manufacturing, and ease of use.29 Furthermore, it has high thermal sensitivity and thus has been widely used in photothermal actuators.30,31 The CTE of PDMS is reported as a constant, ranging from 2.66 × 10−4 to 3.14 × 10−4 (K−1), and it works normally in the temperature range from −45 °C to 200 °C.32–34 Thus, the CTE of PDMS is set constant as 3 × 10−4 K−1 and the operating temperature range (Th) is considered as 300–500 K in this study. Notably, if other high-CTE materials with higher-temperature stability are selected, the operating temperature of the designed NFRTRs can be accordingly expanded to higher values. For a given thickness of the PDMS layer (tPDMS) under a given temperature gradient (ΔT), the thermally-induced gap size variation can be determined as follows:

 
Δd = αtPDMS ΔT = αtPDMS(ThTc).(4)

This means that a gap size reduction of 300 nm can be achieved for a 20 μm-thick PDMS layer with a temperature rise of 50 K. Obviously, larger gap size variations can be obtained by using thicker PDMS layers or larger temperature gradients. In contrast, silica possesses a CTE of about 5 × 10−7 K−1, which causes negligible gap size variations under the same conditions.

2.3.2 Two-dimensional (2D) materials supporting infrared polaritons. In principle, any material enabling enhanced near-field radiative heat transfer can be selected as the radiative layer. Materials supporting IR polariton modes outperform those based on evanescent frustrated modes in this respect due to their more pronounced enhancement and better tunability.22 Herein, 2D van der Waals materials including graphene and hexagonal boron nitride (hBN) are chosen due to their advantages such as single-atom-layer thickness, broadband IR frequency responses, strong field localization in 2D, and highly tunable polaritonic modes. All these advantages make them well-suited for mediating, enhancing, and tailoring near-field radiative heat transfer. Previous research has demonstrated their potential in this regard.35

The dielectric function of graphene with a finite thickness of h can be given as36

 
image file: d2nr04350e-t5.tif(4)
where ε0 is the vacuum permittivity, ω is the angular frequency, and σ is its in-plane conductivity, which can be approximated at room temperature T (kBTEF) as35
 
image file: d2nr04350e-t6.tif(5)
where τ = μcmEF/evF2 is the relaxation time; μcm is the carrier mobility; and EF and vF = 106 m s−1 are the Fermi energy and velocity, respectively. Unless otherwise specified, the parameters for the graphene monolayer used in this study are h = 0.334 nm, μcm ≈ 2000 cm2 V−1 s−1, and EF = 0.3 eV.

hBN is a natural hyperbolic material supporting hyperbolic phonon polaritons (HPhPs) within two reststrahlen bands (RBs), enclosed by transverse-optical (TO) and longitudinal-optical (LO) phonon frequencies. Two types of HPhPs can be excited in the hBN crystal layer (type I: ε > 0, ε < 0 and type II: ε < 0, ε > 0) in the RBs, characterized by a negative real part of the dielectric constant. The in-plane and out-of-plane dielectric constants of hBN can be well approximated by the classic Lorentz model as follows:

 
image file: d2nr04350e-t7.tif(6)
where the subscript m = ⊥ or ‖ denotes the in-plane and out-of-plane components corresponding to the perpendicular or parallel direction to the optical axis of the hBN crystal, respectively; ε∞,m is the high-frequency permittivity; ωLO,m and ωTO,m are the LO and TO phonon frequencies, respectively; and Γm is the damping factor. Here the fitting parameters are given as37,38ε‖,∞ = 2.8, ωLO,‖ = 845 cm−1, ωTO,‖ = 785 cm−1, and Γ = 1 cm−1, while ε⊥,∞ = 3, ωLO,⊥ = 1630 cm−1, ωTO,⊥ = 1395 cm−1, and Γ = 2 cm−1. For all calculations, the hBN thickness is fixed at 20 nm. The dielectric functions of PDMS and silica can be found in the literature.39 The dielectric constants of the four adopted materials have been plotted as a function of angular frequency, as shown in Fig. S2 in the ESI.

3. Results and discussion

3.1 Gigantic TRFs and underlying mechanisms

We now demonstrate that ultrahigh TRFs can be accomplished in the NFRTR designs with gap size variations induced by thermal expansion/contraction of the PDMS layer. As an illustration, we set Tc = 300 K and a tunable Th to achieve the required gap size variations (eqn (4)). It has been well established that heat transport is in the crossover regime between conduction and radiation when the gap size is down to a few nanometers.40,41 To avoid any possible impacts from thermal conduction, the gap size under the forward bias df, i.e., the minimum operating gap size, is fixed at 10 nm to serve as the benchmark. The reverse gap size can be accordingly determined using dr = df + Δd. In the following calculations, tPDMS is fixed at 20 μm, and a 1 mm-thick silica substrate is added beneath both terminal heterostructures to provide the necessary support. Notably, both terminals work at equilibrated temperatures of Th and Tc, respectively, implying that heat transfer inside them is not taken into account.

As shown in Fig. 2a, both greatly enhanced Q and ultra-high TRFs have been achieved even with small temperature gradients for the two distinct NFRTR designs, i.e. hBN- and graphene/hBN-based pairings. For the former design, a heat flux of 4257.15 W m−2, about 692 times higher than that of the blackbody radiation limit (6.15 W m−2), is achieved under a forward bias of only 1 K. Then it surges with elevated temperature gradients (Fig. 2a) since more phonon modes are excited at higher temperatures. This intensive heat transfer is mainly attributed to the contribution of evanescent waves mediated by the cavity phonon polariton modes thermally excited between the bi-planar paring of hBN/PDMS/SiO2 and hBN/SiO2. Two main sharp peaks centered at ∼1.6 × 1014 rad s−1 and ∼2.8 × 1014 rad s−1 (Fig. 2b), corresponding to the energies of HPhPs within two RBs of the hBN layer, dominate the spectral heat flux, together with two other small peaks mediated by the optical phonon energies of SiO2 (Fig. S2 in the ESI). When the temperature bias is reversed and increases from −1 K to −200 K, the heat flux Q considerably decreases at first due to the enlarged gap size and then gradually increases due to the increased temperature gradients (Fig. 2a). The reason for the former behavior is that the SP-mediated evanescent modes decay exponentially in the vacuum gap, leading to a considerable decrease in the number of available surface modes with increasing gap size. More specifically, only SP modes with the decay length l approximately equal to or larger than the gap size (ld) can be tunneled across the gap,22 thus contributing to the Q enhancement. This naturally happens when the forward gap size is 10 nm. In contrast, nearly no SP modes can be tunneled when the reverse gap size enlarges and exceeds the decay length (l < d), resulting in a considerable decrease in Q. This trend has been well elucidated by the profile of the reverse spectral heat flux (Fig. 2b), which closely resembles that of the forward one but has an amplitude being more than 103 lower. In this way, a strong asymmetry in heat fluxes is created between the forward and reverse scenarios, thus achieving high TRFs. For this hBN-based NFRTR, a maximal TRF of ∼900 is obtained under |ΔT| = 200 K. As schematically illustrated in Fig. 2c and d, these high TRFs originate from the contrasted spatial distributions of the SP-mediated evanescent fields between the forward and reverse scenarios, rather than their spectral distributions, thus providing a rectification mechanism that differs from the previous studies.


image file: d2nr04350e-f2.tif
Fig. 2 Thermal rectification performance of NFRTRs with hBN- and graphene/hBN-based parings, together with the underlying mechanisms. (a) Radiative heat flux (solid lines with markers) and TRFs (dashed lines) as a function of temperature bias ΔT. Note that the TRF characterizes the relative enhancement of Q between the forward (+ΔT) and reverse (−ΔT) scenarios under the same absolute temperature difference (|ΔT|). Thus TRFs are plotted as a function of |ΔT|. (b) Spectral heat flux under both forward and reverse biases of 200 K, corresponding to forward and reverse gap sizes of 10 nm and 1210 nm. (c and d) Schematic illustration of the underlying mechanisms for the proposed NFRTR designs under (c) forward and (d) reverse bias conditions.

The performance of NFRTRs can be further improved by symmetrically stacking a graphene monolayer on the radiative layer of the original heterostructure. The heat flux for this graphene/hBN-based NFRTR is about 25.98 MW m−2 under a forward bias of 200 K, which is about 13.25 times higher than that of the hBN-based design. This significant enhancement is attributed to the effective coupling of SPPs in graphene and SPhPs in both hBN and SiO2, which forms the hybridized polaritonic modes enabling a more broadband spectral heat flux (the red solid line in Fig. 2b). The underlying mechanism can be further elucidated using the PTP contour maps in Fig. 3. Notably, to better illustrate the roles played by the SPs, Fig. 3a–d show the PTP mediated by the SP modes only, i.e., ξp(ω,β) is calculated over k0 < β < kc. Two distinct features can be observed. First, the PTP bands in Fig. 3b appear both broader and brighter than their counterparts in Fig. 3a, indicating the heat transfer enhancement mediated by the effective polaritonic coupling in the graphene/hBN-based design. Second, the PTP bands under the reverse bias significantly decay in large wavevectors β, and become invisible for β > 8 × 107 m−1 (Fig. 3c and d). The reason lies in that the decay length l of SP modes is approximately equal to the inverse of the momentum (wavevector β), i.e., l = 1/β,22 indicating that SP modes with larger β have smaller l and thus lead to faster decays. Besides, the cutoff wavevector kc, i.e., the upper limit of β for the contributing modes, is inversely proportional to the gap size kc ≅ 1/d.27,28,42 The reverse gap size is 1210 nm in these cases, leading to a greatly reduced kc and also the contributing mode numbers compared with those in the forward cases. This also explains the trend of its reverse heat flux which gradually decreases and then saturates as ΔT changes from −1 K to −200 K (Fig. 2a). More hybridized polaritonic modes with large β in the graphene/hBN-based design decay with increasing d compared to the case of the hBN-based design (Fig. 3), thus leading to a faster decrease of Q. This difference eventually results in a much higher contrast of Q for the graphene–hBN design, e.g. with a TRF of ∼6936 achieved under a 200 K temperature bias, which is 7.7 times that of the counterpart of the hBN-based one (Fig. 2a). These TRFs can be further optimized by tailoring the chemical potential of the monolayer graphene and the hBN thickness, as shown in Fig. S4 in the ESI.


image file: d2nr04350e-f3.tif
Fig. 3 Contour maps displaying the PTP of p-polarized SP modes thermally excited between distinct pairings: (a) hBN/PDMS/SiO2-hBN/SiO2 and (b) graphene/hBN/PDMS/SiO2-graphene/hBN/SiO2 under a forward temperature bias of 200 K and a gap size of 10 nm. (c and d) Corresponding PTP contour maps under the same reverse bias with a corresponding gap size of 1210 nm. The red dashed lines illustrate the vacuum light line.

3.2 Unlimited potentials due to high tunability and design flexibility

To further demonstrate the rectification potential of the proposed scheme, we computed the Q for the above-mentioned two pairings with d ranging from 10 nm to 1 μm under both ΔT of 20 K and 100 K. Near-field heat transfer occurs through the evanescent coupling of thermally excited surface resonances, thus making Q scale inversely with the gap size d.23 Here, a power-law model Q = γdn (n < 0) is used to determine their quantitative relationship, where the exponent n is a geometry-dependent factor and γ is the fitting coefficient.23

As shown in Fig. 4a and c, fitting curves indicate a near-perfect d−2 scaling relationship of Q for all cases, which is consistent with previous studies for similar parallel-plane configurations.23,28,43–45 This dependence signifies the strong contribution of surface modes (i.e., SPPs in graphene, SPhPs in hBN and SiO2, and their hybridized modes) in the near-field radiative heat transfer. Detailed fitting parameters can be found in Fig. S6 of the ESI. This relationship also provides an intuitive understanding of the sharp gap-size dependence of the radiative heat flux, as well as an optimistic upper bound for the gap-size variation-induced thermal rectification. As indicated by the different γ values, the two pairings exhibit different d-dependences of Q. The NFRTR based on the graphene/hBN heterostructure exhibits a much stronger dependence than the hBN-based one, due to the effective coupling of SPPs and SPhPs in the combination of polar dielectrics with IR plasmonic materials. This combination finally results in higher TRFs, particularly with larger gap size variations. Record-high TRFs of ∼9597 and ∼9980 have been obtained for the hBN- and graphene/hBN-based designs under df = 10 nm, dr = 1000 nm, and ΔT = 20 K, respectively. Furthermore, these TRFs remain nearly unchanged for both designs when ΔT increases from 20 K to 100 K (Fig. 4b and d), even though the amplitude of Q increases significantly (Fig. 4).


image file: d2nr04350e-f4.tif
Fig. 4 Gap-size dependence of near-field radiative heat flux for hBN- and graphene/hBN-based pairings under constant temperature biases of (a) 20 K and (c) 100 K, together with the corresponding power-law fitting curves. TRFs as a function of gap size variations for both designs under the same temperature biases of (b) 20 K and (d) 100 K.

TRFs achieved for NFRTRs in this work have been compared with those in previous publications, as listed in Table S1 (ESI). An improvement of 1–4 orders of magnitude has been achieved, indicating the tremendous potential of the proposed scheme for designing highly-efficient NFRTRs. Here we would like to emphasize that the key to achieving ultra-high TRFs in our proposed design scheme for NFRTRs is to obtain a strong gap-size dependence of the radiative heat flux. Specifically, measures that enable the further enhancement of the forward heat flux but the rapid decay of the reverse heat flux with increasing gap size would help to improve the TRFs. Hence, we anticipate that other IR polaritonic materials such as α-MoO3 and BP,46,47 and elaborate nanostructures such as multilayers and meta-surfaces,48,49 particularly those featuring hyperbolic properties with broadband responses and large transverse wavevectors, can serve as alternative radiative layers to achieve extra gains. Besides, both the active layer and the substrate can be appropriately selected and matched with the radiative layers for further optimization. When it comes to the experimental implementation of this NFRTR concept, challenges have to be tackled such as maintaining the surface parallelism alignment under dynamic variations of the gap size at the nanoscale, avoiding undesired deformation of the heterostructures induced by temperature changes, and optimizing the active layer (e.g., the CTE and thickness) to compromise between the available gap size variations and the compactness requirement of related devices.

It should be mentioned that the corresponding TRFs for both designs decrease more than 20 times as the forward gap size increases from 10 nm to 50 nm, with the maximum TRFs decreasing to ∼400 (Fig. 4b and d), due to the rapid decay of the SP modes with increasing gap size. However, these TRFs are still much higher than those achieved in most of the previously reported NFRTRs (Table S1 in the ESI). More importantly, the proposed scheme offers higher design flexibility and is much simpler to implement in experiments. The design flexibility of this scheme, as well as the achievable superior performance, has also been demonstrated by replacing the radiative layer with traditional polar dielectrics such as SiO2 and SiC, as discussed in section S5 in the ESI.

4. Conclusions

In summary, we have proposed a novel scheme to design NFRTRs by stacking a pair of polaritonic layers separately on a thermally-expanding layer and a rigid substrate. In this scheme, the symmetrical polaritonic layers provide stable radiative channels for heat flux, while the thermally-expanding layer can modulate the gap size with flipped temperature biases to allow high contrasts in heat flux. This scheme relies on the spatial distributions of SP-mediated evanescent fields, which decay exponentially with increasing gap size. The hBN-based paring has been demonstrated to yield a record-high TRF of ∼104 due to the thermally excited phonon polaritons, which can be further boosted by the polaritonic hybridizations in the graphene/hBN-based heterostructures. This study paves the way to introduce more radiative material candidates for NFRTRs with high efficiency, thus opening new avenues for their long-awaited experimental realization.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors greatly acknowledge the financial support from the National Natural Science Foundation of China (Grant No. 51925203), the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDB36000000), and the Special Research Assistant Program of the Chinese Academy of Sciences.

References

  1. N. Li, J. Ren, L. Wang, G. Zhang, P. Hänggi and B. Li, Rev. Mod. Phys., 2012, 84, 1045–1066 CrossRef.
  2. M. J. Martínez-Pérez, A. Fornieri and F. Giazotto, Nat. Nanotechnol., 2015, 10, 303–307 CrossRef PubMed.
  3. H. Wang, S. Hu, K. Takahashi, X. Zhang, H. Takamatsu and J. Chen, Nat. Commun., 2017, 8, 1–8 CrossRef PubMed.
  4. M. Kasprzak, M. Sledzinska, K. Zaleski, I. Iatsunskyi, F. Alzina, S. Volz, C. M. Sotomayor Torres and B. Graczykowski, Nano Energy, 2020, 78, 105261 CrossRef CAS.
  5. C. W. Chang, D. Okawa, A. Majumdar and A. Zettl, Science, 2006, 314, 1121–1124 CrossRef CAS PubMed.
  6. J. Zhao, D. Wei, A. Gao, H. Dong, Y. Bao, Y. Jiang and D. Liu, Appl. Therm. Eng., 2020, 176, 115410 CrossRef.
  7. P. Ben-Abdallah and S. A. Biehs, Phys. Rev. Lett., 2014, 112, 1–5 CrossRef PubMed.
  8. A. Ghanekar, J. Ji and Y. Zheng, Appl. Phys. Lett., 2016, 109, 123106 CrossRef.
  9. Y. Li, W. Li, T. Han, X. Zheng, J. Li, B. Li, S. Fan and C. W. Qiu, Nat. Rev. Mater., 2021, 6, 488–507 CrossRef CAS.
  10. C. R. Otey, W. T. Lau and S. Fan, Phys. Rev. Lett., 2010, 104, 1–4 CrossRef PubMed.
  11. K. Joulain, Y. Ezzahri, J. Drevillon, B. Rousseau and D. De Sousa Meneses, Opt. Express, 2015, 23, A1388–A1397 CrossRef CAS PubMed.
  12. G. Xu, J. Sun, H. Mao and T. Pan, J. Quant. Spectrosc. Radiat. Transfer, 2018, 220, 140–147 CrossRef CAS.
  13. D. Feng, S. K. Yee and Z. M. Zhang, Appl. Phys. Lett., 2021, 119, 181111 CrossRef CAS.
  14. Y. Yang, S. Basu and L. Wang, Appl. Phys. Lett., 2013, 103, 163101 CrossRef.
  15. A. Fiorino, D. Thompson, L. Zhu, R. Mittapally, S. A. Biehs, O. Bezencenet, N. El-Bondry, S. Bansropun, P. Ben-Abdallah, E. Meyhofer and P. Reddy, ACS Nano, 2018, 12, 5174–5179 CrossRef PubMed.
  16. Y. Liu, Y. Tian, F. Chen, A. Caratenuto, X. Liu, M. Antezza and Y. Zheng, Appl. Phys. Lett., 2021, 119, 123101 CrossRef CAS.
  17. Q. Li, H. He, Q. Chen and B. Song, Phys. Rev. Appl., 2021, 16, 014069 CrossRef CAS.
  18. B. Song, D. Thompson, A. Fiorino, Y. Ganjeh, P. Reddy and E. Meyhofer, Nat. Nanotechnol., 2016, 11, 509–514 CrossRef CAS PubMed.
  19. L. Tang, J. Desutter and M. Francoeur, ACS Photonics, 2020, 7, 1304–1311 CrossRef CAS.
  20. I. Latella, P. Ben-Abdallah and M. Nikbakht, Phys. Rev. B, 2021, 104, 1–8 CrossRef.
  21. K. Kim, B. Song, V. Fernández-Hurtado, W. Lee, W. Jeong, L. Cui, D. Thompson, J. Feist, M. T. H. Reid, F. J. García-Vidal, J. C. Cuevas, E. Meyhofer and P. Reddy, Nature, 2015, 528, 387–391 CrossRef CAS PubMed.
  22. M. Francoeur, Nat. Nanotechnol., 2015, 10, 206–208 CrossRef CAS PubMed.
  23. R. St-Gelais, L. Zhu, S. Fan and M. Lipson, Nat. Nanotechnol., 2016, 11, 515–519 CrossRef CAS PubMed.
  24. M. R. Sergei, A. K. Yurii and I. T. Valeryan, Principles of Statistical Radiophysics, Springer, New York, 1989 Search PubMed.
  25. S. Dai, Z. Fei, Q. Ma, A. S. Rodin, M. Wagner, A. S. McLeod, M. K. Liu, W. Gannett, W. Regan, K. Watanabe, T. Taniguchi, M. Thiemens, G. Dominguez, A. H. Castro Neto, A. Zettl, F. Keilmann, P. Jarillo-Herrero, M. M. Fogler and D. N. Basov, Science, 2014, 343, 1125–1129 CrossRef CAS PubMed.
  26. B. Song, Y. Ganjeh, S. Sadat, D. Thompson, A. Fiorino, V. Fernández-Hurtado, J. Feist, F. J. Garcia-Vidal, J. C. Cuevas, P. Reddy and E. Meyhofer, Nat. Nanotechnol., 2015, 10, 253–258 CrossRef CAS PubMed.
  27. P. Ben-Abdallah and K. Joulain, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 82, 1–4 CrossRef.
  28. S. A. Biehs, E. Rousseau and J. J. Greffet, Phys. Rev. Lett., 2010, 105, 3–6 CrossRef PubMed.
  29. A. Kausar, Polym.-Plast. Technol. Mater., 2020, 59, 1148–1166 CAS.
  30. Y. Hu, J. Liu, L. Chang, L. Yang, A. Xu, K. Qi, P. Lu, G. Wu, W. Chen and Y. Wu, Adv. Funct. Mater., 2017, 27, 1704388 CrossRef.
  31. Y. Yang, Y. Liu and Y. Shen, Adv. Funct. Mater., 2020, 30, 1–7 Search PubMed.
  32. T. I. Lee, M. S. Kim and T. S. Kim, Polym. Test., 2016, 51, 181–189 CrossRef CAS.
  33. Y. H. Lin, S. W. Kang and T. Y. Wu, Appl. Therm. Eng., 2009, 29, 573–580 CrossRef CAS.
  34. N. Bowden, W. T. S. Huck, K. E. Paul and G. M. Whitesides, Appl. Phys. Lett., 1999, 75, 2557–2559 CrossRef CAS.
  35. B. Yang, D. Pan, X. Guo, H. Hu and Q. Dai, Int. J. Therm. Sci., 2022, 176, 107493 CrossRef CAS.
  36. A. Vakil and N. Engheta, Science, 2011, 332, 1291–1294 CrossRef CAS PubMed.
  37. P. Li, I. Dolado, F. J. Alfaro-Mozaz, F. Casanova, L. E. Hueso, S. Liu, J. H. Edgar, A. Y. Nikitin, S. Vélez and R. Hillenbrand, Science, 2018, 359, 892–896 CrossRef CAS PubMed.
  38. Z. Yuan, R. Chen, P. Li, A. Y. Nikitin, R. Hillenbrand and X. Zhang, ACS Photonics, 2020, 7, 2610–2617 CrossRef CAS.
  39. E. D. Palik, Handbook of Optical Constants of Solids, Academic P., San Diego, USA, 1998 Search PubMed.
  40. V. Chiloyan, J. Garg, K. Esfarjani and G. Chen, Nat. Commun., 2015, 6, 6755 CrossRef CAS PubMed.
  41. K. Kloppstech, N. Könne, S. A. Biehs, A. W. Rodriguez, L. Worbes, D. Hellmann and A. Kittel, Nat. Commun., 2017, 8, 1–6 CrossRef PubMed.
  42. S. A. Biehs, M. Tschikin and P. Ben-Abdallah, Phys. Rev. Lett., 2012, 109, 1–5 CrossRef PubMed.
  43. J. B. Pendry, J. Phys.: Condens. Matter, 1999, 11, 6621–6633 CrossRef CAS.
  44. J.-P. Mulet, K. Joulain, R. Carminati and J.-J. Greffet, Microscale Thermophys. Eng., 2002, 6, 209–222 CrossRef.
  45. M. Ghashami, H. Geng, T. Kim, N. Iacopino, S. K. Cho and K. Park, Phys. Rev. Lett., 2018, 120, 175901 CrossRef CAS PubMed.
  46. W. Ma, P. Alonso-González, S. Li, A. Y. Nikitin, J. Yuan, J. Martín-Sánchez, J. Taboada-Gutiérrez, I. Amenabar, P. Li, S. Vélez, C. Tollan, Z. Dai, Y. Zhang, S. Sriram, K. Kalantar-Zadeh, S. T. Lee, R. Hillenbrand and Q. Bao, Nature, 2018, 562, 557–562 CrossRef CAS PubMed.
  47. D. Correas-Serrano, A. Alù and J. S. Gomez-Diaz, Phys. Rev. B, 2017, 96, 1–9 CrossRef.
  48. B. Zhao, B. Guizal, Z. M. Zhang, S. Fan and M. Antezza, Phys. Rev. B, 2017, 95, 1–9 Search PubMed.
  49. V. Fernández-Hurtado, F. J. García-Vidal, S. Fan and J. C. Cuevas, Phys. Rev. Lett., 2017, 118, 1–6 CrossRef PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2nr04350e

This journal is © The Royal Society of Chemistry 2022