Abstract
Coupling of surface plasmon polaritons (SPPs) supported by graphene and hyperbolic phonon polaritons (HPPs) supported by hyperbolic materials (HMs) could effectively promote photon tunneling, and hence the radiative heat transfer. In this work, we investigate the polariton hybridization phenomena on near-field radiative heat transfer (NFRHT) in multilayer heterostructures, which consist of periodic graphene/α-MoO3 cells. Numerical results show that increasing the graphene/α-MoO3 cells can effectively enhance the NFRHT when the vacuum gap is less than 50 nm, but suppresses the enhanced performance with larger gap distance. This depends on the coupling of SPPs and HPPs in the periodic structure, which is analyzed by the energy transmission coefficients distributed in the wavevector space. The influence of the thickness of the α-MoO3 film and the chemical potential of graphene on the NFRHT is investigated. The findings in this work may guide designing high-performance near-field energy transfer and conversion devices based on coupling polaritons.
1 Introduction
Near-field radiative heat transfer (NFRHT) has been demonstrated to exceed the blackbody limit by several orders of magnitude [1–13]. When the gap distance of two objects is less than characteristic thermal wavelength, coupling effect of the evanescent waves will provide an efficient channel for photon transport, making the near-field radiative heat flux (NFRHF) greatly enhanced [14–28]. Such huge radiative heat flux has promising applications in noncontact refrigeration [29], thermal rectifications [30, 31], and thermophotovoltaics [32–36].
Previous studies of the NFRHT based on hyperbolic materials (HMs) show that the NFRHF can be promoted in a wide frequency range owing to the excitation of hyperbolic phonon polaritons (HPPs) [37–43]. HMs are a kind of materials with opposite signs of the permittivity components and exhibit superior hyperbolic dispersion properties [44–49]. Specifically, the natural HMs have remarkable advantages because of manifest hyperbolic dispersion without designing into nanostructures. α-MoO3 is a natural biaxial hyperbolic crystal with three wide bands [50]. Liu et al. have proposed a near-field radiative modulator, which demonstrated that effective modulation of NFRHT can be achieved by mechanical rotation of a hyperbolic material [51]. Wu et al. have exhibited that the NFRHT between two bulk α-MoO3 is much larger than that between two hBN [52]. However, the above studies mainly focus on the NFRHT between single materials and are less involved in composite structures based on HMs.
Graphene is an excellent two-dimensional material that can provide lots of resonances to facilitate photon tunneling [53]. It has been demonstrated that surface plasmon polaritons (SPPs) supported by graphene could couple HPPs in HMs to impact NFRHT [54–57]. In fact, the NFRHT between single-layer HMs covered by graphene has been thoroughly studied [58, 59]. Particularly, it is demonstrated that the single-cell structure is limited in both enhancement and modulation of NFRHT. Recent studies on the multilayer graphene/hBN heterostructures show that the coupling of SPPs in graphene and HPPs in hBN has the potential to further enhance NFRHT [60, 61]. Compared with hBN, α-MoO3 has broader hyperbolic bandwidth and longer polariton lifetime [62, 63]. Such a wide hyperbolic band allows HPPs to be excited in a wider frequency range, greatly promoting the NFRHT. Moreover, owing to the strong directional of the HPPs excited in α-MoO3, the NFRHT exhibits highly anisotropic characteristics, which makes it possible to achieve the modulation of the NFRHT. However, the effect of polariton hybridization phenomena on NFRHT in periodic graphene/α-MoO3 heterostructures has rarely been investigated, and the underlying physical mechanism needs further discussion.
Here, we take periodic graphene/α-MoO3 cells as a platform, and discuss the polariton hybridization phenomena on NFRHT. The characteristics of NFRHT were analyzed and compared between multilayer structures for different units. The coupling of the SPPs and HPPs explains the underlying physical mechanism. The energy transmission coefficients in the wavevector space are analyzed. Finally, the influence of the thickness of the α-MoO3 film and the chemical potential of graphene on NFRHT is investigated.
2 Theory and method
Figure 1 shows the schematic of NFRHT between two periodic multilayer structures that consists of α-MoO3 film covered by graphene. The gap distance between the two objects is d, and h is the thickness of each α-MoO3 film. The temperature of emitter is T1 = 300 K, and temperature of receiver is T2 = 0 K. The number of cells is N. This work considers four different structures: single-cell, double-cell, five-cell, and ten-cell. The NFRHT in this work is along the [010] crystal direction of α-MoO3.

Schematic of the NFRHT between two periodic multilayer heterostructures. The number of cells is N and the cell configuration consists of α-MoO3 film covered by graphene.
The permittivity of α-MoO3 is given by:
Detailed parameters can be found in Ref. [64]. Figure 2 shows real permittivity components of α-MoO3, and the shaded area represents three different hyperbolic bands. Graphene is an excellent two-dimensional material [65, 66]. In this work, to simplify the computational model, we ignore the effect of the background [67]. Thus, the conductivity of graphene is modelled by [58]:

The real permittivity components of α-MoO3 vary in angular frequency. Hyperbolic bands are shaded in different colors.
Detailed parameters can be found in Ref. [68]. In the calculation, the graphene is modeled as a layer of thickness Δ = 0.3 nm with an effective permittivity
The NFRHF can be described as [70]:
where
where R and T represent Fresnel’s reflection/transmission coefficients for p or s polarization, which can be expressed as:
where D is a Fabry–Perot-like denominator matrix and is obtained by
3 Results and discussion
To start, we calculate the NFRHT between periodic multilayer graphene/α-MoO3 heterostructures for different cells, which is illustrated in Figure 3(a). The heat flux drops sharply when d increases due to the evanescent effects. When d < 50 nm, Q increases with larger number of graphene/α-MoO3 cells in configurations. The ten-cell configuration shows the largest radiative heat flux compared to that of the other three configurations (single-cell, double-cell, and five-cell). Particularly, when the vacuum gap is 10 nm, the heat flux of ten-cell configuration reaches 2.49 × 106 W/m2, which is 1.76-fold higher than that of the single-cell structure. This result can be confirmed by the variation of spectral heat flux with angular frequency in Figure 3(b). Note that the ten-cell configuration obviously exhibits the largest area between spectral heat flux and angular frequency. Moreover, to study the coupling effects of α-MoO3 and graphene, the spectral heat flux of graphene without α-MoO3 films (marked as graphene in the legend), and bare α-MoO3 films without graphene (marked as α-MoO3 in the legend) are shown in Figure 3(b). The spectral heat flux of the graphene-cell configuration covers a broadband, while the spectral heat flux of the α-MoO3-cell configuration is only covered from 1.03 × 1014 to 1.89 × 1014 rad/s. In addition, the spectral heat flux of the bare α-MoO3 film is much higher than that of the graphene sheet. The single-cell configuration represents the graphene-covered α-MoO3 film. The spectral heat flux of single-cell covers a broader frequency band over the Reststrahlen bands due to the broadband supported by graphene plasmons. The value of spectral heat flux of single-cell can beat that of the graphene sheet but is still smaller than that of the bare α-MoO3 within the Reststrahlen bands. The spectral heat flux in the Reststrahlen band of α-MoO3 films is suppressed by covering graphene. When more layers of graphene/α-MoO3 heterostructures are added to the configurations, the spectral heat flux will be enhanced by the strong hybrid/coupled multiple graphene SPPs and HPPs of α-MoO3 films, which directly explains the enhancement with the increase of the number of graphene-α-MoO3 cells in configurations.

NFRHF varies in gap distance for different configurations. (a) Total heat flux varies in vacuum gap from 10 nm to 1000 nm for four configurations. (b) Spectral heat flux varies in angular frequency for six configurations at d = 10 nm. (c) Spectral heat flux varies in angular frequency for six configurations at d = 100 nm. The α-MoO3 in the legend denotes the α-MoO3-cell configuration. The graphene in the legend denotes the graphene-cell configuration.
However, when vacuum gap is larger than 50 nm, the heat flux of single-cell configuration shows the largest thermal radiation of NFRHT compared to the other three configurations. The variation of spectral heat flux with angular frequency can confirm the phenomenon in Figure 3(c), which shows the change phenomenon opposite to that at narrow gaps. Spectral heat flux gradually decreases with increase of graphene-α-MoO3 cells. This can be explained by the rapid attenuation of the coupling of SPPs and HPPs of the multilayer structure in the wavevector space due to the increased spacing [13, 60, 71, 72]. Since the thickness of each α-MoO3 layer is much smaller than the vacuum spacing, it makes the coupling of the total hybridization mode weaken or disappear. In addition, the optical/material loss of the multilayer structure has a greater negative influence on the polaritons mode than that of the single-layer structure, which ultimately makes the multilayer hybridization mode have no contribution in the larger transverse wavevector. Therefore, the NFRHT between single-layer structures is superior to that of multilayer structures.
Moreover, we calculated energy transmission coefficient distribution for four configurations of multilayer structure at angular frequency ω = 0.75 × 1014 rad/s, which is demonstrated in Figure 4. When the angular frequency is ω = 0.75 × 1014 rad/s, the permittivity of α-MoO3 is ε x = 6.1193 + j0.0066, ε y = 21.2119 + j0.1837, ε z = 2.6852 + j0.0003. Because the conditions of ε x > 0, ε y > 0, ε z > 0, the HPPs in α-MoO3 cannot be excited at this angular frequency. In particular, when the angular frequency is outside the hyperbolic band of α-MoO3, the main contribution of NFRHT between multilayer structures is the SPPs. The dispersion relation of single-cell structure can be described by [73]:
where

Energy transmission coefficient varies in dimensionless wavevector at 0.75 × 1014 rad/s and vacuum gap d = 10 nm for different configurations, (a) single-cell, (b) double-cell, (c) five-cell, and (d) ten-cell. The green solid line represents the dispersion relation.
There are several sharp peaks in the spectral heat flux curves at the 10 nm vacuum gap. The energy transmission coefficients are further analyzed at different angular frequencies for the underlying physics of the four configurations. At ω = 0.97 × 1014 rad/s outside the Reststrahlen band of α-MoO3, the permittivity of α-MoO3 is ε x = 6.6731 + j0.0135, ε y = 73.8813 + j4.3890, ε z = 2.7317 + j0.0005, respectively. Since three real components of the permittivity of α-MoO3 are positive, there are no HPPs excited in α-MoO3 shown in Figure 5(a). For the suspended graphene sheets (Figure 5(b)), two concentric rings of coupled SPPs occur at ω = 0.97 × 1014 rad/s. The green solid line represents the dispersion relation, which can be described by [74]:
where

Energy transmission coefficient varies in dimensionless wavevector at 0.97 × 1014 rad/s and vacuum gap d = 10 nm for different configurations, (a) α-MoO3-cell configuration (α-MoO3 films without graphene), (b) graphene-cell configuration (single-cell without α-MoO3 films), (c) single-cell, (d) double-cell, (e) five-cell, and (f) ten-cell. The green solid line represents the dispersion relation.
When ω = 1.16 × 1014 rad/s, permittivity of α-MoO3 is ε x = 7.7096 + j0.0311, ε y = −21.9670 + j0.8186, ε z = 2.8018 + j0.0009, respectively. According to eight cases in Ref. [58], volume-confined hyperbolic polaritons (VHPs) are excited in α-MoO3 on the upper and lower sides (Figure 6(a)) onto the k x –k y plane due to the signs of three permittivity components Re[ε x ] > 0, Re[ε y ] < 0, Re[ε z ] > 0. The boundary lines can be represented by:

Energy transmission coefficient varies in dimensionless wavevector at 1.16 × 1014 rad/s and vacuum gap d = 10 nm for different configurations, (a) α-MoO3-cell configuration (α-MoO3 films without graphene), (b) graphene-cell configuration (single-cell without α-MoO3 films), (c) single-cell, (d) double-cell, (e) five-cell, and (f) ten-cell. The white dashed lines represent k y = ±0.59k x . The green line represents the dispersion relation. Note that the purpose of using the dashed line in (c) is to avoid shading the energy transfer coefficient.
Therefore, the asymptotes are k y = ±0.59k x , denoted as green dashed lines shown in Figure 6(a). Figure 6(b) shows SPPs excited in graphene-cell configuration at ω = 1.16 × 1014 rad/s. Particularly, discrete VHPs of a concave shape is supported by α-MoO3 (Figure 6(a)) and excited in a wide region of wavevector, while SPPs supported by graphene (Figure 6(b)) are excited in a small region of wavevector. It is clear that the dispersion curves match well with the energy transmission coefficients in wavevector space. In the case of single-cell configuration, HPPs couple with SPPs to form coupling polaritons with a weakened concave shape combined with a ring shape shown in Figure 6(c). These coupling polaritons are excited in a smaller region of wavevector than that of HPPs in Figure 6(a). As a result, spectral heat flux decreases at 1.16 × 1014 rad/s when α-MoO3 film is covered by graphene. In addition, more cell configurations could support more continuous coupling polaritons shown in Figure 6(d)–(f). Consequently, the ten-cell configuration exhibits the highest radiative heat flux than other three configurations at 1.16 × 1014 rad/s.
When ω = 1.5 × 1014 rad/s, the permittivity of α-MoO3 is ε x = 31.6522 + j2.2487, ε y = −1.4059 + j0.0625, ε z = 3.1613 + j0.0043, respectively. VHPs are excited in α-MoO3 on upper and lower sides shown in Figure 7(a), and the asymptotes are k y = ±4.74k x , denoted as green-dashed lines. VHPs supported by α-MoO3 shown in Figure 7(a) could couple with SPPs supported by graphene shown in Figure 7(b) to form the coupling polaritons shown in Figure 7(c). Compared to the VHPs in Figure 7(a), the coupling polaritons are in a smaller region on upper and lower sides and in wide region on left and right sides in wavevector space, which helps to enhance the spectral heat flux at 1.5 × 1014 rad/s. As the number of graphene-α-MoO3 cells increases, the coupling polaritons barely change on left and right sides, and gradually enhance on upper and lower sides. Consequently, ten-cell configuration exhibits the highest radiative heat flux than that of the other three configurations at 1.5 × 1014 rad/s. Furthermore, SPPs are highly enhanced in large areas, although VHPs are excited in limited azimuthal angles. Hence, the coupled polaritons make the spectral heat flux of the single-cell configuration much higher than that of the α-MoO3 cell configuration and the graphene cell configuration.

Energy transmission coefficient varies in dimensionless wavevector at 1.5 × 1014 rad/s and vacuum gap d = 10 nm for different configurations, (a) α-MoO3-cell configuration (α-MoO3 films without graphene), (b) graphene-cell configuration (single-cell without α-MoO3 films), (c) single-cell, (d) double-cell, (e) five-cell, and (f) ten-cell. The white dashed lines represent k y = ±4.74k x . The green solid line represents the dispersion relation.
At ω = 1.85 × 1014 rad/s located at the edge of the Reststrahlen band in α-MoO3, the permittivity of α-MoO3 is ε x = 0.2547 + j0.0506, ε y = 1.8660 + j0.0196, ε z = −2.3527 + j0.2051, respectively. As demonstrated in Figure 8(a), VHPs are excited at all azimuthal angles within wide wavevectors, since the sign of ε z is negative, and the signs of both ε x and ε y are positive. However, SPPs supported by graphene (Figure 8(b)) are excited in a small region of wavevectors. The dispersion curves in Figure 8(a) and (b) are calculated from Eq. (7). The coupling polaritons are excited in small wavevectors demonstrated in Figure 8(c). As the number of cell increases, the wavevector range of coupling polaritons excitation is further compressed. Consequently, the single-cell configuration shows the highest radiative heat flux than the other three configurations at 1.85 × 1014 rad/s.

Energy transmission coefficient varies in dimensionless wavevector at 1.85 × 1014 rad/s and vacuum gap d = 10 nm for different configurations, (a) α-MoO3-cell configuration (α-MoO3 films without graphene), (b) graphene-cell configuration (single-cell without α-MoO3 films), (c) single-cell, (d) double-cell, (e) five-cell, and (f) ten-cell. The green solid line represents the dispersion relation.
Moreover, we investigated the effect of the thickness of α-MoO3 on the NFRHT between multilayer graphene/α-MoO3 heterostructures, as shown in Figure 9. At the case of μ = 0.2 eV and d = 10 nm, the total heat flux increases significantly with the number of cells, especially when the α-MoO3 layer is very thin. When the thickness of α-MoO3 is 10 nm, the heat flux between ten-cell structures is 23.57 × 105 W/m2, which is 87.81 % larger than that of single-cell structure (12.55 × 105 W/m2). For the single-cell structure, the heat flux monotonically increases with the thickness of α-MoO3 film. However, the heat flux between the multi-cell structures increases and then decreases with the thickness of α-MoO3 film, which agrees well with that in multilayer graphene-hBN heterostructures in Ref. [60]. In addition, the effect of the number of cells on the heat flux decreases as the thickness of the α-MoO3 film increases. When the thickness of the α-MoO3 film is large enough (>1000 nm), the heat flux hardly varies with the number of cells.

Total heat flux versus α-MoO3 thickness for four different configurations. The vacuum gap is 10 nm.
The spectral heat flux between multilayer graphene/α-MoO3 heterostructures in four configurations for different thicknesses of α-MoO3 film is calculated in Figure 10. In the case of h = 10 nm, the spectral heat flux changes drastically with the number of cells. When the frequency is outside the hyperbolic band of α-MoO3, the graphene-supported SPPs are the main reason for the variation of the spectral heat flux. Since the distance between graphene is small, increasing the number of units is beneficial to enhance the coupling of SPPs between different graphene, further promoting NFRHT. When the frequency is within the hyperbolic band of α-MoO3, the main reason for the spectral heat flux variation is the coupling of graphene-supported SPPs with α-MoO3-supported HPPs. When the number of cells increases, the coupled polaritons are further promoted, significantly enhancing the spectral heat flux. Notably, when the frequency is in the Band III hyperbolic region of α-MoO3, increasing the number of cells suppresses the NFRHT. The case of h = 23 nm is similar to that of h = 10 nm, differing only at high frequencies (2.5 × 1014 rad/s – 3 × 1014 rad/s), that is, the spectral heat flux is almost independent of the number of cells in that frequency range. When the thickness of α-MoO3 is 100 nm, the change in spectral heat flux occurs mainly within the hyperbolic band of α-MoO3. Increasing the number of cells promotes the coupling of SPPs and HPPs, and therefore the spectral heat flux is enhanced. The spectral heat flux varies little with the number of cells outside the hyperbolic band of α-MoO3, which is due to the large thickness of the α-MoO3 films, which weakens the coupling of SPPs between different graphene.

Spectral heat flux versus angular frequency for four different configurations at different thicknesses.
Finally, the effect of chemical potential of graphene on the NFRHT between multilayer graphene/α-MoO3 heterostructures is discussed, as shown in Figure 11. Note d = 10 nm and h = 20 nm. For the case of μ = 0.1 eV, Figure 11(a) shows that as the number of cells increases, the spectral heat flux increases significantly, especially in the hyperbolic band of α-MoO3 and in the low-frequency region (<1 × 1014 rad/s). For the ten-cell structure, the total heat flux is 47.65 × 105 W/m2, which is 33.4 % more than that of the single-cell structure (35.72 × 105 W/m2). When the chemical potential of graphene is μ = 0.3 eV, the NFRHT is significantly suppressed. We find the number of cells affects the spectral heat flux in the high-frequency region (>2 × 1014 rad/s). The total heat flux is 16.90 × 105 W/m2 for the ten-cell structure, which is 115 % larger than that of the single-cell structure (7.86 × 105 W/m2), which exhibits more dramatic variations of the spectral heat flux.

Spectral heat flux versus angular frequency for four configurations at different chemical potentials of graphene.
4 Conclusions
We theoretically investigate the effect of polariton hybridization phenomena on NFRHT, using periodic graphene/α-MoO3 cells as a platform. The radiative heat transfer characteristics of four different kinds of periodic cell structures are compared and analyzed. When the thickness of each α-MoO3 is fixed, our calculation shows that when the vacuum gap is less than 50 nm, adding graphene/α-MoO3 cells is effective in enhancing the NFRHT. However, when the vacuum gap is large, adding cells would suppress the NFRHT between multilayer structures with fixed thickness of each layer. The underlying physics can be explained by the coupling of SPPs and HPPs, which is equivalently illustrated by the energy transmission coefficients distribution in wavevector space. Moreover, we discuss the influence of the thickness of α-MoO3 on the NFRHT. The heat flux monotonously increases with increasing film thickness for single-cell structures, while it has an optimal value for other multi-cell structures. Finally, we investigate the effect of the chemical potential of graphene on the NFRHT. The findings of this work may help to design tunable near-field thermal radiation systems.
Funding source: National Natural Science Foundation of China
Award Identifier / Grant number: 52106099
Funding source: Science and Technology Innovation Development Project of Yantai
Award Identifier / Grant number: 2022GCCRC158
Funding source: Natural Science Foundation of Shandong Province
Award Identifier / Grant number: ZR2022YQ57
Funding source: Taishan Scholars Program
Funding source: China Postdoctoral Science Foundation
Award Identifier / Grant number: 2022M710122
-
Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: The authors acknowledge the support of the National Natural Science Foundation of China (52106099), the Natural Science Foundation of Shandong Province (ZR2022YQ57), the Taishan Scholars Program, the Science and Technology Innovation Development Project of Yantai (2022GCCRC158), and the China Postdoctoral Science Foundation (2022M710122).
-
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
-
Data availability: Data will be made available on request.
References
[1] Z. Zhang, Nano/Microscale Heat Transfer, Switzerland, Springer Nature, 2020.10.1007/978-3-030-45039-7Suche in Google Scholar
[2] X. Liu, L. Wang, and Z. Zhang, “Near-field thermal radiation: recent progress and outlook,” Nanoscale Microscale Thermophys. Eng., vol. 19, pp. 98–126, 2015. https://doi.org/10.1080/15567265.2015.1027836.Suche in Google Scholar
[3] X. Liu, R. Zhang, and Z. Zhang, “Near-field radiative heat transfer with doped-silicon nanostructured metamaterials,” Int. J. Heat Mass Transfer, vol. 73, pp. 389–398, 2014. https://doi.org/10.1016/j.ijheatmasstransfer.2014.02.021.Suche in Google Scholar
[4] R. St-Gelais, B. Guha, L. Zhu, S. Fan, and M. Lipson, “Demonstration of strong near-field radiative heat transfer between integrated nanostructures,” Nano Lett., vol. 14, pp. 6971–6975, 2014. https://doi.org/10.1021/nl503236k.Suche in Google Scholar PubMed
[5] K. Shi, Z. Chen, Y. Xing, et al.., “Near-field radiative heat transfer modulation with an ultrahigh dynamic range through mode Mismatching,” Nano Lett., vol. 22, pp. 7753–7760, 2022. https://doi.org/10.1021/acs.nanolett.2c01286.Suche in Google Scholar PubMed PubMed Central
[6] Y. Sun, Y. Hu, K. Shi, J. Zhang, D. Feng, and X. Wu, “Negative differential thermal conductance between Weyl semimetals nanoparticles through vacuum,” Phys. Scr., vol. 97, p. 095506, 2022. https://doi.org/10.1088/1402-4896/ac8843.Suche in Google Scholar
[7] M. Luo, J. Zhao, L. Liu, and M. Antezza, “Radiative heat transfer and radiative thermal energy for two-dimensional nanoparticle ensembles,” Phys. Rev. B, vol. 102, p. 024203, 2020. https://doi.org/10.1103/physrevb.102.024203.Suche in Google Scholar
[8] M. Luo, J. Zhao, L. Liu, B. Guizal, and M. Antezza, “Many-body effective thermal conductivity in phase-change nanoparticle chains due to near-field radiative heat transfer,” Int. J. Heat Mass Transfer, vol. 166, p. 120793, 2021. https://doi.org/10.1016/j.ijheatmasstransfer.2020.120793.Suche in Google Scholar
[9] K. Shi, F. Bao, N. He, and S. He, “Near-field heat transfer between graphene-Si grating heterostructures with multiple magnetic-polaritons coupling,” Int. J. Heat Mass Transfer, vol. 134, pp. 1119–1126, 2019. https://doi.org/10.1016/j.ijheatmasstransfer.2019.01.037.Suche in Google Scholar
[10] Y. Hu, Y. Sun, Z. Zheng, J. Song, K. Shi, and X. Wu, “Rotation-induced significant modulation of near-field radiative heat transfer between hyperbolic nanoparticles,” Int. J. Heat Mass Transfer, vol. 189, p. 122666, 2022. https://doi.org/10.1016/j.ijheatmasstransfer.2022.122666.Suche in Google Scholar
[11] P. Ben-Abdallah and S. Biehs, “Near-field thermal transistor,” Phys. Rev. Lett., vol. 112, p. 044301, 2014. https://doi.org/10.1103/physrevlett.112.044301.Suche in Google Scholar PubMed
[12] M. Lim, J. Song, S. Lee, and B. Lee, “Tailoring near-field thermal radiation between metallo-dielectric multilayers using coupled surface plasmon polaritons,” Nat. Commun., vol. 9, pp. 1–9, 2018. https://doi.org/10.1038/s41467-018-06795-w.Suche in Google Scholar PubMed PubMed Central
[13] K. Shi, Z. Chen, X. Xu, J. Evans, and S. He, “Optimized colossal near‐field thermal radiation enabled by manipulating coupled plasmon polariton geometry,” Adv. Mater., vol. 33, p. 2106097, 2021. https://doi.org/10.1002/adma.202106097.Suche in Google Scholar PubMed PubMed Central
[14] K. Shi, Y. Sun, Z. Chen, et al.., “Colossal enhancement of near-field thermal radiation across hundreds of nanometers between millimeter-scale plates through surface plasmon and phonon polaritons coupling,” Nano Lett., vol. 19, pp. 8082–8088, 2019. https://doi.org/10.1021/acs.nanolett.9b03269.Suche in Google Scholar PubMed
[15] J. Dong, J. Zhao, and L. Liu, “Long-distance near-field energy transport via propagating surface waves,” Phys. Rev. B, vol. 97, p. 075422, 2018. https://doi.org/10.1103/physrevb.97.075422.Suche in Google Scholar
[16] J. Dong, W. Zhang, and L. Liu, “Nonreciprocal thermal radiation of nanoparticles via spin-directional coupling with reciprocal surface modes,” Appl. Phys. Lett., vol. 119, p. 021104, 2021. https://doi.org/10.1063/5.0057446.Suche in Google Scholar
[17] M. He, H. Qi, Y. Ren, Y. Zhao, and M. Antezza, “Graphene-based thermal repeater,” Appl. Phys. Lett., vol. 115, p. 263101, 2019. https://doi.org/10.1063/1.5132995.Suche in Google Scholar
[18] A. Ott and S. Biehs, “Thermal rectification and spin-spin coupling of nonreciprocal localized and surface modes,” Phys. Rev. B, vol. 101, p. 155428, 2020. https://doi.org/10.1103/physrevb.101.155428.Suche in Google Scholar
[19] A. Ott, R. Messina, P. Ben-Abdallah, and S. Biehs, “Radiative thermal diode driven by nonreciprocal surface waves,” Appl. Phys. Lett., vol. 114, p. 163105, 2019. https://doi.org/10.1063/1.5093626.Suche in Google Scholar
[20] R. St-Gelais, L. Zhu, S. Fan, and M. Lipson, “Near-field radiative heat transfer between parallel structures in the deep subwavelength regime,” Nat. Nanotechnol., vol. 11, pp. 515–519, 2016. https://doi.org/10.1038/nnano.2016.20.Suche in Google Scholar PubMed
[21] J. Song and Q. Cheng, “Near-field heat transfer between graphene and anisotropic magneto-dielectric hyperbolic metamaterials,” Phys. Rev. B, vol. 94, p. 125419, 2016. https://doi.org/10.1103/physrevb.94.125419.Suche in Google Scholar
[22] C. Zhou, X. Wu, Y. Zhang, and H. Yi, “Amplification and modulation effect of elliptical surface polaritons on a thermal diode,” Int. J. Heat Mass Transfer, vol. 180, p. 121794, 2021. https://doi.org/10.1016/j.ijheatmasstransfer.2021.121794.Suche in Google Scholar
[23] H. Wu, Y. Huang, L. Cui, and K. Zhu, “Active magneto-optical control of near-field radiative heat transfer between graphene sheets,” Phys Rev Appl, vol. 11, p. 054020, 2019. https://doi.org/10.1103/physrevapplied.11.054020.Suche in Google Scholar
[24] T. Kralik, P. Hanzelka, M. Zobac, V. Musilova, T. Fort, and M. Horak, “Strong near-field enhancement of radiative heat transfer between metallic surfaces,” Phys. Rev. Lett., vol. 109, p. 224302, 2012. https://doi.org/10.1103/physrevlett.109.224302.Suche in Google Scholar
[25] Y. Zhang, H. Yi, and H. Tan, “Near-field radiative heat transfer between black phosphorus sheets via anisotropic surface plasmon polaritons,” ACS Photonics, vol. 5, pp. 3739–3747, 2018. https://doi.org/10.1021/acsphotonics.8b00776.Suche in Google Scholar
[26] K. Shi, F. Bao, and S. He, “Spectral control of near-field thermal radiation with periodic cross resonance metasurfaces,” IEEE J. Quantum Electron., vol. 54, p. 7000107, 2018. https://doi.org/10.1109/jqe.2018.2791639.Suche in Google Scholar
[27] K. Shi, R. Liao, G. Cao, F. Bao, and S. He, “Enhancing thermal radiation by graphene-assisted hBN/SiO2 hybrid structures at the nanoscale,” Opt. Express, vol. 126, pp. A591–A600, 2018. https://doi.org/10.1364/oe.26.00a591.Suche in Google Scholar PubMed
[28] X. Liu, J. Shen, and Y. Xuan, “Near-field thermal radiation of nanopatterned black phosphorene mediated by topological transitions of phosphorene plasmons,” Nanoscale Microscale Thermophys. Eng., vol. 23, pp. 188–199, 2019. https://doi.org/10.1080/15567265.2019.1578310.Suche in Google Scholar
[29] K. Chen, P. Santhanam, S. Sandhu, L. Zhu, and S. Fan, “Heat-flux control and solid-state cooling by regulating chemical potential of photons in near-field electromagnetic heat transfer,” Phys. Rev. B, vol. 91, p. 134301, 2015. https://doi.org/10.1103/physrevb.91.134301.Suche in Google Scholar
[30] C. Otey, W. Lau, and S. Fan, “Thermal rectification through vacuum,” Phys. Rev. Lett., vol. 104, p. 154301, 2010. https://doi.org/10.1103/physrevlett.104.154301.Suche in Google Scholar PubMed
[31] L. Zhu, C. Otey, and S. Fan, “Ultrahigh-contrast and large-bandwidth thermal rectification in near-field electromagnetic thermal transfer between nanoparticles,” Phys. Rev. B, vol. 88, p. 184301, 2013. https://doi.org/10.1103/physrevb.88.184301.Suche in Google Scholar
[32] C. Lucchesi, D. Cakiroglu, J. Perez, et al.., “Near-field thermophotovoltaic conversion with high electrical power density and cell efficiency above 14%,” Nano Lett., vol. 21, pp. 4524–4529, 2021. https://doi.org/10.1021/acs.nanolett.0c04847.Suche in Google Scholar PubMed
[33] T. Inoue, T. Koyama, D. Kang, K. Ikeda, T. Asano, and S. Noda, “One-chip near-field thermophotovoltaic device integrating a thin-film thermal emitter and photovoltaic cell,” Nano Lett., vol. 19, pp. 3948–3952, 2019. https://doi.org/10.1021/acs.nanolett.9b01234.Suche in Google Scholar PubMed
[34] A. Fiorino, L. Zhu, D. Thompson, R. Mittapally, P. Reddy, and E. Meyhofer, “Nanogap near-field thermophotovoltaics,” Nat. Nanotechnol., vol. 13, pp. 806–811, 2018. https://doi.org/10.1038/s41565-018-0172-5.Suche in Google Scholar PubMed
[35] T. Inoue, K. Ikeda, B. Song, et al.., “Integrated near-field thermophotovoltaic device overcoming blackbody limit,” ACS Photonics, vol. 8, pp. 2466–2472, 2021. https://doi.org/10.1021/acsphotonics.1c00698.Suche in Google Scholar
[36] R. Wang, J. Lu, X. Wu, J. Peng, and J. Jiang, “Tuning topological transitions in twisted thermophotovoltaic systems,” arXiv:2205.07666, 2022.10.1103/PhysRevApplied.19.044050Suche in Google Scholar
[37] Y. Guo, C. Cortes, S. Molesky, and Z. Jacob, “Broadband super-Planckian thermal emission from hyperbolic metamaterials,” Appl. Phys. Lett., vol. 101, p. 131106, 2012. https://doi.org/10.1063/1.4754616.Suche in Google Scholar
[38] Y. Guo and Z. Jacob, “Thermal hyperbolic metamaterials,” Opt. Express, vol. 21, pp. 15014–15019, 2013. https://doi.org/10.1364/oe.21.015014.Suche in Google Scholar
[39] S. Biehs, M. Tschikin, R. Messina, and P. Ben-Abdallah, “Super-Planckian near-field thermal emission with phonon-polaritonic hyperbolic metamaterials,” Appl. Phys. Lett., vol. 102, p. 131106, 2013. https://doi.org/10.1063/1.4800233.Suche in Google Scholar
[40] X. Liu, R. Zhang, and Z. Zhang, “Near-field thermal radiation between hyperbolic metamaterials: graphite and carbon nanotubes,” Appl. Phys. Lett., vol. 103, p. 213102, 2013. https://doi.org/10.1063/1.4832057.Suche in Google Scholar
[41] C. Simovski, S. Maslovski, I. Nefedov, and S. Tretyakov, “Optimization of radiative heat transfer in hyperbolic metamaterials for thermophotovoltaic applications,” Opt. Express, vol. 21, pp. 14988–15013, 2013. https://doi.org/10.1364/oe.21.014988.Suche in Google Scholar
[42] S. Lang, M. Tschikin, S. Biehs, A. Petrov, and M. Eich, “Large penetration depth of near-field heat flux in hyperbolic media,” Appl. Phys. Lett., vol. 104, p. 121903, 2014. https://doi.org/10.1063/1.4869490.Suche in Google Scholar
[43] O. Miller, S. Johnson, and A. Rodriguez, “Effectiveness of thin films in lieu of hyperbolic metamaterials in the near field,” Phys. Rev. Lett., vol. 112, p. 157402, 2014. https://doi.org/10.1103/physrevlett.112.157402.Suche in Google Scholar
[44] P. Shekhar, J. Atkinson, and Z. Jacob, “Hyperbolic metamaterials: fundamentals and applications,” Nano Convergence, vol. 1, pp. 1–17, 2014. https://doi.org/10.1186/s40580-014-0014-6.Suche in Google Scholar PubMed PubMed Central
[45] A. Poddubny, I. Iorsh, P. Belov, and Y. Kivshar, “Hyperbolic metamaterials,” Nat. Photonics, vol. 7, pp. 948–957, 2013. https://doi.org/10.1038/nphoton.2013.243.Suche in Google Scholar
[46] Y. Guo, W. Newman, C. L. Cortes, and Z. Jacob, “Applications of hyperbolic metamaterial substrates,” Adv. Optoelectron., vol. 2012, p. 452502, 2012. https://doi.org/10.1155/2012/452502.Suche in Google Scholar
[47] W. Huang, T. Folland, F. Sun, et al.., “In-plane hyperbolic polariton tuner in terahertz and long-wave infrared regimes,” arXiv:2206.10433, 2022.10.21203/rs.3.rs-1735989/v1Suche in Google Scholar
[48] M. Ye, B. Qiang, S. Zhu, et al.., “Nano‐optical engineering of anisotropic phonon resonances in a hyperbolic α‐MoO3 metamaterial,” Adv. Opt. Mater., vol. 10, p. 2102096, 2022. https://doi.org/10.1002/adom.202102096.Suche in Google Scholar
[49] O. Takayama and A. V. Lavrinenko, “Optics with hyperbolic materials,” J. Opt. Soc. Am. B, vol. 36, pp. 38–48, 2019. https://doi.org/10.1364/josab.36.000f38.Suche in Google Scholar
[50] X. Wu and C. Fu, “Near-field radiative modulator based on dissimilar hyperbolic materials with in-plane anisotropy,” Int. J. Heat Mass Transfer, vol. 168, p. 120908, 2021. https://doi.org/10.1016/j.ijheatmasstransfer.2021.120908.Suche in Google Scholar
[51] X. Liu and Y. Xuan, “Super-Planckian thermal radiation enabled by hyperbolic surface phonon polaritons,” Sci. China Technol. Sci., vol. 59, pp. 1680–1686, 2016. https://doi.org/10.1007/s11431-016-0480-9.Suche in Google Scholar
[52] X. Wu, C. Fu, and Z. Zhang, “Near-field radiative heat transfer between two α-MoO3 biaxial crystals,” J. Heat Transfer, vol. 142, p. 072802, 2020.10.1115/1.4046968Suche in Google Scholar
[53] A. Grigorenko, M. Polini, and K. Novoselov, “Graphene plasmonics,” Nat. Photonics, vol. 6, pp. 749–758, 2012. https://doi.org/10.1038/nphoton.2012.262.Suche in Google Scholar
[54] R. Messina, J. Hugonin, J. Greffet, et al.., “Tuning the electromagnetic local density of states in graphene-covered systems via strong coupling with graphene plasmons,” Phys. Rev. B, vol. 87, p. 085421, 2013. https://doi.org/10.1103/physrevb.87.085421.Suche in Google Scholar
[55] A. Wang, Z. Zheng, and Y. Xuan, “Near-field radiative thermal control with graphene covered on different materials,” J. Quant. Spectrosc. Radiat. Transf., vol. 180, pp. 117–125, 2016. https://doi.org/10.1016/j.jqsrt.2016.04.018.Suche in Google Scholar
[56] H. Hajian, I. Rukhlenko, V. Erçağlar, G. Hanson, and E. Ozbay, “Epsilon-near-zero enhancement of near-field radiative heat transfer in BP/hBN and BP/α-MoO3 parallel-plate structures,” Appl. Phys. Lett., vol. 120, p. 112204, 2022. https://doi.org/10.1063/5.0083817.Suche in Google Scholar
[57] C. Zhou, Y. Zhang, Z. Torbatian, D. Novko, M. Antezza, and H. Yi, “Photon tunneling reconstitution in black phosphorus/hBN heterostructure,” Phys. Rev. Mater., vol. 6, p. 075201, 2022. https://doi.org/10.1103/physrevmaterials.6.075201.Suche in Google Scholar
[58] X. Wu and R. Liu, “Near-field radiative heat transfer between graphene covered biaxial hyperbolic materials,” ES Energy Environ., vol. 10, pp. 66–72, 2020. https://doi.org/10.30919/esee8c939.Suche in Google Scholar
[59] B. Zhao and Z. Zhang, “Enhanced photon tunneling by surface plasmon–phonon polaritons in Graphene/hBN heterostructures,” J. Heat Transfer, vol. 139, p. 022701, 2017. https://doi.org/10.1115/1.4034793.Suche in Google Scholar
[60] K. Shi, F. Bao, and S. He, “Enhanced near-field thermal radiation based on multilayer graphene-hBN heterostructures,” ACS Photonics, vol. 4, pp. 971–978, 2017. https://doi.org/10.1021/acsphotonics.7b00037.Suche in Google Scholar
[61] B. Zhao, B. Guizal, Z. Zhang, S. Fan, and M. Antezza, “Near-field heat transfer between graphene/hBN multilayers,” Phys. Rev. B, vol. 95, p. 245437, 2017. https://doi.org/10.1103/physrevb.95.245437.Suche in Google Scholar
[62] H. Liu, Y. Hu, Q. Ai, M. Xie, and X. Wu, “Spontaneous emission modulation in biaxial hyperbolic van der Waals material,” J. Appl. Phys., vol. 132, p. 175105, 2022. https://doi.org/10.1063/5.0120203.Suche in Google Scholar
[63] T. Folland and J. Caldwell, “Precise control of infrared polarization using crystal vibrations,” Nature, vol. 562, pp. 499–501, 2018. https://doi.org/10.1038/d41586-018-07087-5.Suche in Google Scholar PubMed
[64] H. Liu, Q. Ai, M. Ma, Z. Wang, and M. Xie, “Prediction of spectral absorption of anisotropic α-MoO3 nanostructure using deep neural networks,” Int. J. Therm. Sci., vol. 177, p. 107587, 2022. https://doi.org/10.1016/j.ijthermalsci.2022.107587.Suche in Google Scholar
[65] E. Hwang and S. Sarma, “Dielectric function, screening, and plasmons in two-dimensional graphene,” Phys. Rev. B, vol. 75, p. 205418, 2007. https://doi.org/10.1103/physrevb.75.205418.Suche in Google Scholar
[66] S. Badalyan and F. Peeters, “Enhancement of Coulomb drag in double-layer graphene structures by plasmons and dielectric background inhomogeneity,” Phys. Rev. B, vol. 86, p. 121405, 2012. https://doi.org/10.1103/physrevb.86.121405.Suche in Google Scholar
[67] H. Hu, N. Chen, H. Teng, et al.., “Doping-driven topological polaritons in graphene/α-MoO3 heterostructures,” Nat. Nanotechnol., vol. 17, pp. 940–946, 2022. https://doi.org/10.1038/s41565-022-01185-2.Suche in Google Scholar PubMed PubMed Central
[68] M. Jablan, H. Bulja, and M. Soljacic, “Plasmonics in graphene at infrared frequencies,” Phys. Rev. B, vol. 80, p. 245435, 2009. https://doi.org/10.1103/physrevb.80.245435.Suche in Google Scholar
[69] A. Vakil and N. Engheta, “Transformation optics using graphene,” Science, vol. 332, pp. 1291–1295, 2011. https://doi.org/10.1126/science.1202691.Suche in Google Scholar PubMed
[70] S. Biehs, P. Ben-Abdallah, F. Rosa, K. Joulain, and J. Greffet, “Nanoscale heat flux between nanoporous materials,” Opt. Express, vol. 19, pp. A1088–A1103, 2011. https://doi.org/10.1364/oe.19.0a1088.Suche in Google Scholar
[71] Y. Zhang, C. Wang, H. Yi, and H. Tan, “Multiple surface plasmon polaritons mediated near-field radiative heat transfer between graphene/vacuum multilayer,” J. Quant. Spectrosc. Radiat. Transf., vol. 221, pp. 138–146, 2018. https://doi.org/10.1016/j.jqsrt.2018.09.029.Suche in Google Scholar
[72] H. Iizuka and S. Fan, “Significant enhancement of near-field electromagnetic heat transfer in a multilayer structure through multiple surface-states coupling,” Phys. Rev. Lett., vol. 120, p. 063901, 2018. https://doi.org/10.1103/physrevlett.120.063901.Suche in Google Scholar PubMed
[73] A. Bapat, S. Dixit, Y. Gupta, T. Low, and A. Kumar, “Gate tunable light–matter interaction in natural biaxial hyperbolic van der Waals heterostructures,” Nanophotonics, vol. 11, pp. 2329–2340, 2022. https://doi.org/10.1515/nanoph-2022-0034.Suche in Google Scholar
[74] G. A. Pérez, K. V. Voronin, V. S. Volkov, P. A. González, and A. Y. Nikitin, “Analytical approximations for the dispersion of electromagnetic modes in slabs of biaxial crystals,” Phys. Rev. B, vol. 100, p. 235408, 2019. https://doi.org/10.1103/physrevb.100.235408.Suche in Google Scholar
© 2023 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Artikel in diesem Heft
- Frontmatter
- Research Articles
- Reaction nanoscopy of ion emission from sub-wavelength propanediol droplets
- Polariton hybridization phenomena on near-field radiative heat transfer in periodic graphene/α-MoO3 cells
- Experimental observations of communication in blackout, topological waveguiding and Dirac zero-index property in plasma sheath
- Strongly coupled Raman scattering enhancement revealed by scattering-type scanning near-field optical microscopy
- Terahertz nanospectroscopy of plasmon polaritons for the evaluation of doping in quantum devices
- Spin–orbit interactions in plasmonic crystals probed by site-selective cathodoluminescence spectroscopy
- Implementing of infrared camouflage with thermal management based on inverse design and hierarchical metamaterial
- Stable single-mode operation of distributed feedback quantum cascade laser under high current via a grating reflector
- Wide-angle deep ultraviolet antireflective multilayers via discrete-to-continuous optimization
- Active 3D positioning and imaging modulated by single fringe projection with compact metasurface device
- Enhanced cutoff energies for direct and rescattered strong-field photoelectron emission of plasmonic nanoparticles
Artikel in diesem Heft
- Frontmatter
- Research Articles
- Reaction nanoscopy of ion emission from sub-wavelength propanediol droplets
- Polariton hybridization phenomena on near-field radiative heat transfer in periodic graphene/α-MoO3 cells
- Experimental observations of communication in blackout, topological waveguiding and Dirac zero-index property in plasma sheath
- Strongly coupled Raman scattering enhancement revealed by scattering-type scanning near-field optical microscopy
- Terahertz nanospectroscopy of plasmon polaritons for the evaluation of doping in quantum devices
- Spin–orbit interactions in plasmonic crystals probed by site-selective cathodoluminescence spectroscopy
- Implementing of infrared camouflage with thermal management based on inverse design and hierarchical metamaterial
- Stable single-mode operation of distributed feedback quantum cascade laser under high current via a grating reflector
- Wide-angle deep ultraviolet antireflective multilayers via discrete-to-continuous optimization
- Active 3D positioning and imaging modulated by single fringe projection with compact metasurface device
- Enhanced cutoff energies for direct and rescattered strong-field photoelectron emission of plasmonic nanoparticles