Abstract
Controlling and manipulating radiative heat transfer remains a pivotal challenge in both scientific inquiry and technological advancement, traditionally tackled through the precise geometric design of metastructures. However, geometrical optimization cannot break the inherent shackles of local modes within individual meta-atoms, which hinders sustained progress in radiative heat transfer. Here, we propose a comprehensive strategy based on interatomic displacement to achieve superior heat transfer performance while obviating the need for increasingly complex structural designs. This meta-atomic displacement strategy enables a shift from quasi-isolated localized resonances to extended nonlocal resonant modes induced by strong interactions among neighboring meta-atoms, resulting in a radiative heat conductance that surpasses other previously reported geometrical structures. Furthermore, this meta-atomic displacement strategy can be seamlessly applied to various metastructures, offering significant implications for advancing thermal science and next-generation energy devices.
1 Introduction
Radiative heat transfer (RHT) is ubiquitous in nature, spanning from gigantic galaxies to microscopic atomic structures [1], [2], [3]. Effective manipulation of RHT is vital for mitigating diverse challenges such as global climate change [4], [5] and the overheating of electronics [6]. In this context, the question regarding the fundamental limits of RHT is attracting a lot of attentions. Since then, researchers investigated RHT in a variety of systems with objects of different shapes and materials, in pursuit of optimal radiative strategies [7], [8]. Among them, thermophotonic metastructurals are considered the most promising strategy [9]. Conventionally, the thermophotonic metastructures focus on the structural design of individual meta-atoms, trying to continuously optimize the local response of single meta-atom to pursue higher radiation performance [10], [11], [12], as schematically shown in Figure 1. For instance, Fernández-Hurtado et al. achieved much greater room-temperature radiative heat conductance than any unstructured material to date by constructing Si-based metastructures featuring two-dimensional periodic arrays of holes [13]. Motivated by the extraordinary effects, increasingly intricate micro- and nanostructures have sprung up, expanding the family of thermophotonic metastructures [14], [15], [16].
![Figure 1:
Structural and energetic properties of interatom-displacement-driven thermophotonic metasurface. (a) Schematics of RHT between two metastructures separated by a vacuum gap d, which have temperatures T and T + ΔT, respectively. (b) Conceptual diagram to design thermophotonic metastructures, illustrating the differences between the traditional and proposed approaches. The traditional approaches focus on the independent electromagnetic behavior of meta-atoms. In contrast, the proposed approaches rely on periodic displacement between meta-atoms to introduce additional interaction effects, stimulating stronger collective electromagnetic modes. (c) The heat transfer coefficient H for the regular grating, the more complex meta-atoms (cavity surface-plasmon polaritons [CSPPs] [13]), and proposed metastructures with strengthened interatom displacement, at different gaps.](/document/doi/10.1515/nanoph-2024-0729/asset/graphic/j_nanoph-2024-0729_fig_001.jpg)
Structural and energetic properties of interatom-displacement-driven thermophotonic metasurface. (a) Schematics of RHT between two metastructures separated by a vacuum gap d, which have temperatures T and T + ΔT, respectively. (b) Conceptual diagram to design thermophotonic metastructures, illustrating the differences between the traditional and proposed approaches. The traditional approaches focus on the independent electromagnetic behavior of meta-atoms. In contrast, the proposed approaches rely on periodic displacement between meta-atoms to introduce additional interaction effects, stimulating stronger collective electromagnetic modes. (c) The heat transfer coefficient H for the regular grating, the more complex meta-atoms (cavity surface-plasmon polaritons [CSPPs] [13]), and proposed metastructures with strengthened interatom displacement, at different gaps.
Nevertheless, the complexity of these meta-atoms is both a blessing and a curse. As the most straightforward approach, enhancing radiative heat transfer by persistently refining more intricate metacellular architectures is undoubtedly feasible [17]. Unfortunately, constrained by the degrees of freedom of the local modes, the independent response of each meta-atom cannot achieve a sustainable improvement with the increase in structural complexity [18], [19], [20]. The potential for further enhancing radiative heat transfer performance would diminish as the complexity of the meta-atom further increases. With the advancement in electronics and energy technologies, there is an increasing need for heat transfer performance [21], [22]. However, the current study of thermophotonic metastructures is mostly limited to the investigation of the local response of a single meta-atom, which cannot support further development of radiative heat transfer. Therefore, overcoming the current bottleneck in developing thermophotonic metastructures and finding a new general strategy to improve radiative heat transfer remains a formidable challenge.
In response to this challenge, we resort to displacement between meta-atoms to achieve an extraordinary thermal response, marking the first demonstration of the effects of meta-atomic displacement in RHT, as schematically depicted in Figure 1b. Utilizing rigorous coupled wave analysis, it is demonstrated that introducing inter-element displacement into traditional metastructures could markedly amplify radiative energy transfer, while surpassing conventional approaches that rely solely on optimizing meta-atom configurations. This also shows that the inter-element displacement effects are not just carriers of thermal information (as previously reported [23], [24]), but can help better manipulate thermal energy transfer. We then develop a nonlocal effective medium approach to predict non-trivial fingerprint of this thermo-metastructures, and demonstrate rigorously that the underlying physical mechanism responsible for this remarkable behavior is the existence of nonlocal electromagnetic response mode enabled by meta-atomic displacement. Moreover, we further demonstrate this inter-element displacement effect allows us to achieve a much higher radiative thermal conductivity than other metastructure to date, almost a factor of two higher than the metastructures with the previously reported maximum.
2 Extraordinary energy feature in interatom displacement
To illustrate our general strategy, first we concentrate on an instance of two mirrored metastructures formed by 2D alternating arrays on a semi-infinite planar substrate (see Figure 1b). Two mirrored metastructures are separated by a vacuum gap d. A conventional subwavelength grating comprises alternating strips of the core with a width w and the cladding groove with a width G, arrayed with a subwavelength period P = G + w along the direction perpendicular to the strips (x-axis). Referring to Figure 1b, in the proposed metastructures, the nanostrips are periodically partitioned into rectangular nanoblocks with a pitch L along the y-axis. The rectangular nanoblocks are then periodically dislocated by a distance Δ/2 in the x-direction. This dislocation introduces meta-atomic displacement of configuration assignment into the regular grating, which in turn enhances the interactions between the meta-atoms. The thickness of nonlocal metasurface is fixed at 200 nm. For the simplicity of analysis, the dielectric function of the substrate is set to 1. The filling ratio can be defined as f = w/P. Since the width of each strip and the dimension of the period along the x direction are the same, the structural displacement does not affect the filling ratio of the strips in the metastructures. The nanostrips are constructed from silicon (Si) with a doping concentration of 1020cm−3.
Theoretically, we combine fluctuational electrodynamics (FED) [25], [26] and rigorous coupled wave analysis (RCWA) [27], [28], [29] to reveal an radiative thermal effect of this metastructures. Our main goals focus on the analysis of the heat transfer coefficient (HTC), i.e. the radiative thermal conductance per unit area, at room temperature (300 K). In the framework of FED, the HTC between two arbitrary periodic metastructures can be expressed as follows [30]
where, h(ω) is the spectral heat transfer coefficient.
Let us start the discussion of the results by illustrating the main finding of our work. Figure 1c describes the room-temperature HTC versus the gap size for three metastructures with P = 50 nm and f = 0.4. This result is compared with the HTC for the regular grating and more complex meta-atoms [cavity surface-plasmon polaritons (CSPPs) structure] [13] with the same P and f. It is worth noting that when the Si plate is patterned as a regular grating, its HTC in the deep near-field regions (d < 50 nm) is already very significant, several times larger than the corresponding result for Si plates, as has been confirmed in many studies [10], [16]. In order to achieve a further breakthrough in radiative heat transfer, the structure of CSPPs meta-atom has been proposed [13]. As shown in Figure 1c, the HTC in both deep near-field regions is significantly improved when the meta-atoms structure is converted from the conventional grating to this CSPPs metacells.
Intriguingly, a more pronounced increase in radiative heat transfer is observed upon introducing a meta-atomic displacement, as seen in Figure 1c. This can be attributed to the system geometry, i.e., this interatom displacement introduces additional interactions into the system and optimizes the collective response behavior, thereby improving the radiative heat transfer performance. Taking d = 20 nm as an example, the HTC enhancement due to the meta-atomic displacement effect is 400 % of the increase amplitude from conventional idea of designing the metacell as CSPPs structure. However, it should be noted that there is a limit to this enhancement, and the enhancement resulting from this meta-atomic displacement disappears when the spacing is too large (see Section III of the SM [33] for details).
As shown in Figure 2a, the radiative heat transfer has a pronounced sensitivity to meta-atomic displacement. An increase in the radiative heat flux of the metastructures is observed when a small meta-atomic displacement (Δ/w = 0.2) is introduced. Upon reaching a degree of meta-atomic displacement of 0.9 Δ/w, the radiative heat transfer attains its maximum, exceeding that of a regular grating by 40 % and that of a CSPPs meta-atom by 28 %. Nevertheless, further reinforcement of the meta-atomic displacement cannot provide a sustained enhancement of the RHT. As the degree of misalignment increases above 0.9 Δ/w, or the two Si blocks are completely separated (δ = w), it can be observed that radiative heat transfer experiences a significant decline. This nonlinear enhancement indicates the complex wave response mechanisms in our thermo-metastructures, similar to optical [34], [35], acoustic [36], and other metastructures with strong interatom interactions.

Heat transfer enhancement induced by the interatom displacement effect. (a) The heat transfer coefficient H versus the meta-atomic displacement degree Δ. (b) The spectral heat transfer coefficient h(ω) as a function of the frequency ω. The different lines correspond to different meta-atomic displacement degree Δ.
To clarify the excellent properties of this interatom displacement metastructure, we present the spectral heat transfer coefficient in Figure 2b. It is noticed that the meta-atomic displacement does not affect the polarisation distribution of the system, which is still dominated by TM waves at this point, and a detailed analysis can be found in SM [33]. This spectral heat transfer coefficient indicates the energy levels carried by thermal photons of different frequencies. It can be seen that meta-atomic displacement broaden the spectral bandwidth while significantly intensifying the spectral peaks. As demonstrated in Figure 2b, increasing Δ/w from 0 to 0.9 results in a 28 % heightening in the spectral heat transfer coefficient (from 2.5 to 3.2 pW m−2 rad−1 s K−1). The meta-atomic displacements cannot result in a significant shift of the spectral peak. As Δ/w increases from 0 to 1, the frequency of spectral peak remains within the range of 0.1–0.13 eV/ℏ. The results reveal crucial significance that it demonstrates this interatom-dislocated metastructure can play crucial role in thermophotovoltaics [37] and electroluminescent refrigeration [6]. This feature enhances the power of the mentioned apparatus while maintaining optimal efficiency.
3 Local-nonlocal transition of thermophotons mode
The thermophotons tunnelling probability indicates the tunneling probability of thermal photons between the emitter and receiver. k0 = ω/c being the wavenumber in vacuum. We also depict the thermophotons tunnelling properties of the regular grating for comparison in Figure 3a. It can be observed that the thermophotons tunnelling of the system exhibits a clear hyperbolic character, which is also consistent with previous studies [38], [39], [40]. In addition to the exact RCWA theory, we also employ the local effective medium theory (EMT) to facilitate the analysis. This approach treats the nanostructures as equivalent homogeneous biaxial plates, offering a computationally simple and rapid solution [14]. The effective dielectric function [ɛxx,emt, ɛyy,emt, ɛzz,emt] of regular grating can be expressed in SM [33]. The local EMT theory can accurately predict the thermophotons tunnelling properties, thereby indicating the localized nature of the mode of this metasurface.

The thermophotonic tunnelling probability of regular grating (without meta-atomic displacement) for (a) exact solution (RCWA) and (b) EMT solution. The thermophotonic tunnelling probability of proposed thermo-metastructures (with meta-atomic displacement) for (c) exact solution (RCWA) and (d) EMT solution. The nonlocal effective model is shown in Eq. (2) (see also Ref. SM for the parameters). Electric field profiles (|E z |) of meta-atom with displacements of (e) Δ/w = 0 and (f) Δ/w = 0.9. The frequency is fixed at 0.11 eV/ℏ. The evolution of electric field distributions|FFT(E z )| of metastructure in response to different meta-atomic displacements in momentum space: (g) Δ/w = 0 and (h) Δ/w = 0.9. The field is excited by a dipole polarized along z placed 20 nm above the metastructure.
It is noteworthy that a notable alteration of the thermophotons properties of the metasurface can be obtained by adding meta-atomic displacement (from hyperbolic to dumbbell-like) (see Figure 3c). It demonstrates that a larger Δ suppresses the wavevector region of stronger thermophoton tunnelling (ξ > 0.8). However, a more significant meta-atomic displacement expands the bright band of weak thermophoton tunnelling (ξ < 0.5) into a wider wavevector region, which effectively counteracts the recession of strong thermophoton tunnelling while intensifies the spectral heat flux of the metastructure. Moreover, conventional local EMT theory fails to predict the nonlocal behavior of this metastructure. To address this limitation, we propose a nonlocal EMT model to precisely characterize the thermophotons behavior in metastructures. Given the absence of straightforward analytical expressions for nonlocal corrections associated with meta-atomic displacement, we incorporate these nonlocal corrections into the EMT model using a Taylor series expansion as a reference. This approach significantly reduces the prediction error of the electromagnetic response, particularly in the context of introducing meta-atomic displacement to metamaterials [41].
where, the subscript i represents the direction x and y. The mentioned nonlocal corrections can be applied to the imaginary and real parts of the equivalent permittivity along different directions (see SM [33] for parameter details). It can be observed that incorporating nonlocal corrections significantly improves the agreement between the EMT-predicted thermophoton tunnelling coefficients and the exact solution, as seen in Figure 3d. The strong nonlocal behavior of metastructures can be attributed to the phenomenon of continuous reconstruction of the electric field distribution, thereby facilitating a transition from quasi-isolated localized resonances to extended nonlocal modes (see Figure 3e and f). This extended nonlocal resonant modes between the meta-atoms induced by strong interunit interactions further contribute to a significant change in the behaviour of surface polariton (see Figure 3g and h).
4 Interatom displacement in thermophotonic metastructures
The presented strategy for enhancing radiative heat transfer by using periodic displacement between meta-atoms is not confined to a certain metasurface with the mentioned units. Instead, it is a general approach that can be employed for various thermophotonic metastructures, such as rectangular nanowires [10], circular nanorods [14], square nanorods [40], elliptical nanorods, square cavity [13], and many others, as shown in Figure 4 (the structural parameters can be seen as SM [33]). Silicon serves as the matrix material for these structures. These meta-atoms were previously regarded as artificial structures with exceptional radiative heat transfer performance. The square cavity structure, in particular, was previously considered to be a metastructures with room-temperature radiative thermal conductivity that can be much greater than any unstructured material [13]. Figure 4 indicates that when the distribution of meta-atoms is rearranged (i.e., a meta-atomic displacement is applied to the conventional distribution between the meta-atoms) to increase the interactions correlation between different elements, it leads to an overall increase in radiative heat transfer. Note that the reorganisation process preserves the original filling ratio f and the f is the optimum for the different structures. The HTC of cavity structure in Figure 4 is higher than that in Figure 1 due to the difference in f between the two. Surprisingly, the heat transfer coefficient at room temperature can reach up to 1.7 times that of previously reported structures at the highest level after interatom interactions enhancement of the square cavity structure. The strategy is equally effective for radiative enhancement at different temperatures, as can be seen in the SM [33].
![Figure 4:
Dependence of the HTC on the introduction of meta-atomic displacement effect between meta-atoms for various designs (Si-based), which are labelled (I) to (VI). All square units have a 50 nm period. Furthermore, filling ratios remain consistent before and after the interactions effects. These designs and materials are regarded as promising candidates with the potential to exhibit high heat transfer performance [10], [13], [14], [40], [42], [43], [44], [45], [46], [47].](/document/doi/10.1515/nanoph-2024-0729/asset/graphic/j_nanoph-2024-0729_fig_004.jpg)
Dependence of the HTC on the introduction of meta-atomic displacement effect between meta-atoms for various designs (Si-based), which are labelled (I) to (VI). All square units have a 50 nm period. Furthermore, filling ratios remain consistent before and after the interactions effects. These designs and materials are regarded as promising candidates with the potential to exhibit high heat transfer performance [10], [13], [14], [40], [42], [43], [44], [45], [46], [47].
Remarkably, our findings reveal that the introduction of meta-atomic displacement effects enable silicontraditionally regarded as a material with modest radiative propertiesto surpass high-performance materials such as Ga2O3 in radiative heat transfer. Notably, this photonics strategy is not limited to silicon but can also be extended to materials like Ga2O3, MoO3, and others, offering a versatile approach to significantly enhance their radiative heat transfer capabilities. Moreover, the primary focus of this work is on the formulation and validation of the concept of enhanced radiative heat transfer with displacement. Consequently, the global optimisation of the arrangement of structural unit is not involved in the aforementioned calculations. However, it is anticipated that the optimisation of the displacement of the structural units by certain global optimisation methodology will lead to further enhancement of the heat transfer performance [17], [48], [49]. Although it is possible to enhance the radiative heat transfer in the thermal metasurface with this method, there is still a considerable gap between the current HTC and the ideal HTC limit [50], [51], [52]. Let’s take the optimal bulk plasmonic material in Ref. [52] as an example, which at a vacuum gap of 20 nm is still close to three times that of the highest HTC in Figure 4.
5 Conclusions
We have proposed a conceptual framework to achieve unprecedented radiative heat transfer by exploiting interatomic displacement effects. This approach facilitates interactions among meta-atoms by introducing meta-atomic displacement that reconfigure structural periodicity, thereby enabling a transition from quasi-isolated localized resonances to extended nonlocal modes. Remarkably, the results reveal that this displacement-driven strategy can significantly amplify radiative heat transfer, yielding radiative heat conductances that surpass those of other proposed structures. These observed thermal responses suggest that radiative heat transfer can be effectively manipulated through introducing meta-atomic displacement effects into the distribution of meta-atoms, eliminating the need for increasingly complex metastructure designs. We contend that this approach offers a definitive pathway for advancing research in radiative heat transfer, paving the way for innovative applications in thermal management [53], thermophotovoltaics [37], photonic cooling [54], [55], thermocomputation [56], and near-field imaging [57].
Funding source: Ministry of Education – Singapore
Award Identifier / Grant number: A-8000107-01-00
Funding source: National Research Foundation, Singapore (NRF) under NRF’s Medium Sized Centre: Singapore Hybrid-Integrated NextGeneration Electronics (SHINE) Centre funding programme
Funding source: The Research Grants Council (RGC) of the Hong Kong Special Administrative Region
Funding source: National Natural Science Foundation of China
Award Identifier / Grant number: 523B2060
Award Identifier / Grant number: U22A20210
Funding source: Science and Technology Project of Jiangsu Province
Award Identifier / Grant number: BZ2022056
Funding source: The ANR/RGC Joint Research Scheme sponsored by the French National Research Agency
Award Identifier / Grant number: A-HKUST604/20
Funding source: Postdoctoral Fellowship Program of CPSF
Award Identifier / Grant number: GZB20240951
Acknowledgments
HLY acknowledges the support from the National Natural Science Foundation of China (Grant No. U22A20210). CWQ acknowledges the financial support by the Ministry of Education, Republic of Singapore (Grant No.: A-8002978-00-00), the National Research Foundation, Singapore (NRF) under NRFs Medium Sized Centre: Singapore Hybrid-Integrated Next-Generation – Electronics (SHINE) Centre funding programme, and the Science and Technology Project of Jiangsu Province (Grant No. BZ2022056). CLZ acknowledges the support from the National Natural Science Foundation of China (Grant No. 523B2060) and the Postdoctoral Fellowship Program of CPSF (Grant No. GZB20240951). MA acknowledges the ANR/RGC Joint Research Scheme sponsored by the French National Research Agency (Grant No. A-HKUST604/20) and the Research Grants Council (RGC) of the Hong Kong Special Administrative Region and the National Research Foundation.
-
Research funding: This work acknowledges the support from the National Natural Science Foundation of China (Grant No.U22A20210, Grant No.523B2060), the Postdoctoral Fellowship Program of CPSF (Grant No.GZB20240951), the Fundamental Research Funds for the Central Universities (Grant No.HIT.DZIJ.2023098), the ANR/RGC Joint Research Scheme sponsored by the French National Research Agency (Grant No.A-HKUST604/20), and the Ministry of Education, Singapore (Grant No. A-8002978-00-00). This work was also partially supported by the Research Grants Council (RGC) of the Hong Kong Special Administrative Region and the National Research Foundation, Singapore (NRF) under NRF’s Medium Sized Centre: Singapore Hybrid-Integrated Next-Generation μElectronics (SHINE) Centre funding program.
-
Author contributions: Investigation – CLZ and YH; coding, simulations, data analysis, and visualization – CLZ; writing – CLZ (original draft) and YH; review – MA and YZ; editing CLZ and SY; conceptualization – CLZ and CWQ; supervision – CLZ and CWQ. All authors have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results and approved the final version of the manuscript.
-
Conflict of interest: Authors state no conflicts of interest.
-
Informed consent: Informed consent was obtained from all individuals included in this study.
-
Ethical approval: The conducted research is not related to either human or animals use.
-
Data availability: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.
References
[1] Z. M. Zhang, Nano/Microscale Heat Transfer, Switzerland, Springer, 2020.10.1007/978-3-030-45039-7Search in Google Scholar
[2] S.-A. Biehs, R. Messina, P. S. Venkataram, A. W. Rodriguez, J. C. Cuevas, and P. Ben-Abdallah, “Near-field radiative heat transfer in many-body systems,” Rev. Mod. Phys., vol. 93, no. 2, p. 025009, 2021, https://doi.org/10.1103/revmodphys.93.025009.Search in Google Scholar
[3] H. P. Zhou, et al.., “Interface engineering of highly efficient perovskite solar cells,” Science, vol. 345, no. 6196, pp. 542–546, 2014, https://doi.org/10.1126/science.1254050.Search in Google Scholar PubMed
[4] Y. C. Peng, et al.., “Nanoporous polyethylene microfibres for large-scale radiative cooling fabric,” Nat. Sustain., vol. 1, no. 2, pp. 105–112, 2018, https://doi.org/10.1038/s41893-018-0023-2.Search in Google Scholar
[5] D. J. Fan, T. Burger, S. McSherry, B. Lee, A. Lenert, and S. R. Forrest, “Near-perfect photon utilization in an air-bridge thermophotovoltaic cell,” Nature, vol. 586, no. 7828, pp. 237–241, 2020, https://doi.org/10.1038/s41586-020-2717-7.Search in Google Scholar PubMed
[6] L. X. Zhu, A. Fiorino, D. Thompson, R. Mittapally, E. Meyhofer, and P. Reddy, “Near-field photonic cooling through control of the chemical potential of photons,” Nature, vol. 566, no. 7743, pp. 239–244, 2019, https://doi.org/10.1038/s41586-019-0918-8.Search in Google Scholar PubMed
[7] P. S. Venkataram, S. Molesky, W. L. Jin, and A. W. Rodriguez, “Fundamental limits to radiative heat transfer: The limited role of nanostructuring in the near-field,” Phys. Rev. Letts., vol. 124, no. 1, p. 013904, 2020, https://doi.org/10.1103/physrevlett.124.013904.Search in Google Scholar PubMed
[8] L. Zhang, F. Monticone, and O. D. Miller, “All electromagnetic scattering bodies are matrix-valued oscillators,” Nat. Commun., vol. 14, no. 1, p. 7724, 2023, https://doi.org/10.1038/s41467-023-43221-2.Search in Google Scholar PubMed PubMed Central
[9] Y. Li, et al.., “Transforming heat transfer with thermal metamaterials and devices,” Nat. Rev. Mater., vol. 6, no. 6, pp. 488–507, 2021, https://doi.org/10.1038/s41578-021-00283-2.Search in Google Scholar
[10] A. W. Rodriguez, et al.., “Frequency-selective near-field radiative heat transfer between photonic crystal slabs: A computational approach for arbitrary geometries and materials,” Phys. Rev. Lett., vol. 107, no. 11, p. 114302, 2011, https://doi.org/10.1103/physrevlett.107.114302.Search in Google Scholar PubMed
[11] O. D. Miller, S. G. Johnson, and A. W. Rodriguez, “Shape-independent limits to near-field radiative heat transfer,” Phys. Rev. Lett., vol. 115, no. 20, p. 204302, 2015, https://doi.org/10.1103/physrevlett.115.204302.Search in Google Scholar PubMed
[12] L. Tang, L. M. Correa, M. Francoeur, and C. Dames, “Corner- and edge-mode enhancement of near-field radiative heat transfer,” Nature, vol. 629, no. 8010, pp. 67–73, 2024. https://doi.org/10.1038/s41586-024-07279-2.Search in Google Scholar PubMed
[13] V. Fernández-Hurtado, F. J. García-Vidal, S. H. Fan, and J. C. Cuevas, “Enhancing near-field radiative heat transfer with si-based metasurfaces,” Phys. Rev. Letts., vol. 118, no. 20, p. 203901, 2017, https://doi.org/10.1103/physrevlett.118.203901.Search in Google Scholar
[14] S.-A. Biehs, M. Tschikin, and P. Ben-Abdallah, “Hyperbolic metamaterials as an analog of a blackbody in the near field,” Phys. Rev. Lett., vol. 109, no. 10, p. 104301, 2012, https://doi.org/10.1103/physrevlett.109.104301.Search in Google Scholar
[15] A. W. Rodriguez, M. T. H. Reid, J. D. J. Varela, F. Capasso, F. Capasso, and S. G. Johnson, “Anomalous near-field heat transfer between a cylinder and a perforated surface,” Phys. Rev. Lett., vol. 110, no. 1, p. 014301, 2013, https://doi.org/10.1103/physrevlett.110.014301.Search in Google Scholar PubMed
[16] X. L. Liu and Z. M. Zhang, “Near-field thermal radiation between metasurfaces,” ACS Photonics, vol. 2, no. 9, pp. 1320–1326, 2015, https://doi.org/10.1021/acsphotonics.5b00298.Search in Google Scholar
[17] W. L. Jin, S. Molesky, Z. Lin, and A. W. Rodriguez, “Material scaling and frequency-selective enhancement of near-field radiative heat transfer for lossy metals in two dimensions via inverse design,” Phys. Rev. B, vol. 99, no. 4, p. 041403, 2019, https://doi.org/10.1103/physrevb.99.041403.Search in Google Scholar
[18] K. Shastri and F. Monticone, “Nonlocal flat optics,” Nat. Photonics, vol. 17, no. 1, pp. 36–47, 2023, https://doi.org/10.1038/s41566-022-01098-5.Search in Google Scholar
[19] K. Wang, et al.., “Quantum metasurface for multiphoton interference and state reconstruction,” Science, vol. 361, no. 6407, pp. 1104–1107, 2018, https://doi.org/10.1126/science.aat8196.Search in Google Scholar PubMed
[20] J. H. Song, J. van de Groep, S. J. Kim, and M. L. Brongersma, “Non-local metasurfaces for spectrally decoupled wavefront manipulation and eye tracking,” Nat. Nanotechnol., vol. 16, no. 11, pp. 1224–1230, 2021, https://doi.org/10.1038/s41565-021-00967-4.Search in Google Scholar PubMed
[21] C. L. Zhou, et al.., “Unconventional thermophotonic charge density wave,” Phys. Rev. Lett., vol. 133, no. 6, p. 066902, 2024, https://doi.org/10.1103/physrevlett.133.066902.Search in Google Scholar PubMed
[22] G. Q. Xu, Y. H. Yang, X. Zhou, H. S. Chen, A. Alú, and C. W. Qiu, “Diffusive topological transport in spatiotemporal thermal lattices,” Nat. Phys., vol. 18, no. 4, pp. 450–456, 2022, https://doi.org/10.1038/s41567-021-01493-9.Search in Google Scholar
[23] A. C. Overvig, S. A. Mann, and A. Alú, “Thermal metasurfaces: Complete emission control by combining local and nonlocal light-matter interactions,” Phys. Rev. X, vol. 11, no. 2, p. 021050, 2021, https://doi.org/10.1103/physrevx.11.021050.Search in Google Scholar
[24] J. R. Nolen, A. C. Overvig, M. Overvig, and A. Alú, “Local control of polarization and geometric phase in thermal metasurfaces,” Nat. Nanotechnol., vol. 19, no. 11, pp. 1627–1634, 2024, https://doi.org/10.1038/s41565-024-01763-6.Search in Google Scholar PubMed
[25] B. Song, D. Thompson, A. Fiorino, Y. Ganjeh, P. Reddy, and E. Meyhofer, “Radiative heat conductances between dielectric and metallic parallel plates with nanoscale gaps,” Nat. Nanotechnol., vol. 11, no. 6, pp. 509–514, 2016, https://doi.org/10.1038/nnano.2016.17.Search in Google Scholar PubMed
[26] S. Shen, A. Narayanaswamy, and G. Chen, “Surface phonon polaritons mediated energy transfer between nanoscale gaps,” Nano. Letts., vol. 9, no. 8, pp. 2909–2913, 2009, https://doi.org/10.1021/nl901208v.Search in Google Scholar PubMed
[27] J. Lussange, R. Guerout, F. S. S. Rosa, J.-J. Greffet, A. Lambrecht, and S. Reynaud, “Radiative heat transfer between two dielectric nanogratings in the scattering approach,” Phys. Rev. B, vol. 86, no. 8, p. 085432, 2012, https://doi.org/10.1103/physrevb.86.085432.Search in Google Scholar
[28] J. Dai, S. A. Dyakov, and M. Yan, “Radiative heat transfer between two dielectric-filled metal gratings,” Phys. Rev. B, vol. 93, no. 15, p. 155403, 2016, https://doi.org/10.1103/physrevb.93.155403.Search in Google Scholar
[29] Y. Yang and L. P. Wang, “Spectrally enhancing near-field radiative transfer between metallic gratings by exciting magnetic polaritons in nanometric vacuum gaps,” Phys. Rev. Lett., vol. 117, no. 4, p. 044301, 2016, https://doi.org/10.1103/physrevlett.117.044301.Search in Google Scholar PubMed
[30] R. Messina and M. Antezza, “Scattering-matrix approach to casimir-lifshitz force and heat transfer out of thermal equilibrium between arbitrary bodies,” Phys. Rev. A, vol. 84, no. 4, p. 042102, 2011, https://doi.org/10.1103/physreva.84.042102.Search in Google Scholar
[31] P. Ben-Abdallah and S. A. Biehs, “Near-field thermal transistor,” Phys. Rev. Letts., vol. 112, no. 4, p. 044301, 2014, https://doi.org/10.1103/physrevlett.112.044301.Search in Google Scholar PubMed
[32] A. I. Volokitin and B. N. J. Persson, “Near-field radiative heat transfer and noncontact friction,” Rev. Mod. Phys., vol. 79, no. 4, pp. 1291–1329, 2007, https://doi.org/10.1103/revmodphys.79.1291.Search in Google Scholar
[33] Supplementary material at [URL], which additionally includes Refs.[58]–[63], for further details: S1. The computational details for the radiative heat transfer and the nonlocal effective medium theory; S2. Impact of temperature on the meta-atomic displacement effect; S3. The radiative heat transfer of designing thermophotonic metastructures at different structure parameters; S4. Thermal photons evolution in designing thermophotonic metastructures with different structure parameters; S5. The derivation and result of electromagnetic property with different structure parameters.Search in Google Scholar
[34] A. Overvig, S. A. Mann, and A. Alú, “Spatio-temporal coupled mode theory for nonlocal metasurfaces,” Light. Sci. Appl., vol. 13, no. 1, p. 28, 2024, https://doi.org/10.1038/s41377-023-01350-9.Search in Google Scholar PubMed PubMed Central
[35] H. Kwon, D. Sounas, A. Cordaro, A. Polman, and A. Alú, “Nonlocal metasurfaces for optical signal processing,” Phys. Rev. Letts., vol. 121, no. 17, p. 173004, 2018, https://doi.org/10.1103/physrevlett.121.173004.Search in Google Scholar
[36] Y. F. Zhu, A. Merkel, K. Donda, S. W. Fan, L. Y. Cao, and B. Assouar, “Nonlocal acoustic metasurface for ultrabroadband sound absorption,” Phys. Rev. B, vol. 103, no. 6, p. 064102, 2021, https://doi.org/10.1103/physrevb.103.064102.Search in Google Scholar
[37] A. Fiorino, L. X. Zhu, D. Thompson, R. Mittapally, P. Reddy, and E. Meyhofer, “Nanogap near-field thermophotovoltaics,” Nat. Nanotechnol., vol. 13, no. 9, pp. 806–811, 2018, https://doi.org/10.1038/s41565-018-0172-5.Search in Google Scholar PubMed
[38] X. L. Liu and Z. M. Zhang, “Giant enhancement of nanoscale thermal radiation based on hyperbolic graphene plasmons,” Appl. Phys. Lett., vol. 107, no. 14, p. 143114, 2015, https://doi.org/10.1063/1.4932958.Search in Google Scholar
[39] N. S. alakhova, I. M. Fradkin, S. A. Dyakov, and N. A. Gippius, “Fourier modal method for moire lattices,” Phys. Rev. B, vol. 104, no. 8, p. 085424, 2021, https://doi.org/10.1103/physrevb.104.085424.Search in Google Scholar
[40] Y. Z. Hu, H. G. Li, Y. G. Zhu, and Y. Yang, “Enhanced near-field radiative heat transport between graphene metasurfaces with symmetric nanopatterns,” Phys. Rev. Appl., vol. 14, no. 4, p. 044054, 2020, https://doi.org/10.1103/physrevapplied.14.044054.Search in Google Scholar
[41] M. Coppolaro, G. Castaldi, and V. Galdi, “Anomalous light transport induced by deeply subwavelength quasiperiodicity in multilayered dielectric metamaterials,” Phys. Rev. B, vol. 102, no. 7, p. 75107, 2020, https://doi.org/10.1103/physrevb.102.075107.Search in Google Scholar
[42] C. L. Zhou, G. M. Tang, Y. Zhang, M. Antezza, and H. L. Yi, “Radiative heat transfer in a low-symmetry bravais crystal,” Phys. Rev. B, vol. 106, no. 15, p. 155404, 2022, https://doi.org/10.1103/physrevb.106.155404.Search in Google Scholar
[43] X. H. Wu, C. J. Fu, and Z. M. Zhang, “Near-field radiative heat transfer between two α-moo3 biaxial crystals,” J. Heat. Trans.-T ASME, vol. 142, no. 7, p. 072802, 2020, https://doi.org/10.1115/1.4046968.Search in Google Scholar
[44] L. Lu, et al.., “Enhanced near-field radiative heat transfer between graphene/hBN systems,” Small, vol. 18, no. 19, p. 2108032, 2022, https://doi.org/10.1002/smll.202108032.Search in Google Scholar PubMed
[45] M. Ghashami, H. Y. Geng, T. Kim, N. Iacopino, S. K. Cho, and K. Park, “Precision measurement of phonon-polaritonic near-field energy transfer between macroscale planar structures under large thermal gradients,” Phys. Rev. Letts., vol. 120, no. 17, p. 175901, 2018, https://doi.org/10.1103/physrevlett.120.175901.Search in Google Scholar
[46] R. St-Gelais, L. X. Zhu, S. H. Fan, and M. Lipson, “Near-field radiative heat transfer between parallel structures in the deep subwavelength regime,” Nat. Nanotechnol., vol. 11, no. 6, pp. 515–519, 2016, https://doi.org/10.1038/nnano.2016.20.Search in Google Scholar PubMed
[47] J. DeSutter, L. Tang, and M. Francoeur, “A near-field radiative heat transfer device,” Nat. Nanotechnol., vol. 14, no. 8, pp. 751–755, 2019, https://doi.org/10.1038/s41565-019-0483-1.Search in Google Scholar PubMed
[48] W. L. Jin, R. Messina, and A. W. Rodriguez, “Overcoming limits to near-field radiative heat transfer in uniform planar media through multilayer optimization,” Opt. Express, vol. 25, no. 13, pp. 14746–14759, 2017, https://doi.org/10.1364/oe.25.014746.Search in Google Scholar PubMed
[49] J. J. Garcia-Esteban, J. Bravo-Abad, and J. C. Cuevas, “Deep learning for the modeling and inverse design of radiative heat transfer,” Phys. Rev. Appl., vol. 16, no. 6, p. 064006, 2021, https://doi.org/10.1103/physrevapplied.16.064006.Search in Google Scholar
[50] O. D. Miller, et al.., “Limits to the optical response of graphene and two-dimensional materials,” Nano Letts., vol. 17, no. 9, pp. 5408–5415, 2017, https://doi.org/10.1021/acs.nanolett.7b02007.Search in Google Scholar PubMed
[51] S. Basu and Z. M. Zhang, “Maximum energy transfer in near-field thermal radiation at nanometer distances,” J. Appl. Phys., vol. 105, no. 9, p. 093535, 2009, https://doi.org/10.1063/1.3125453.Search in Google Scholar
[52] L. Zhang and O. D. Miller, “Optimal materials for maximum large-area near-field radiative heat transfer,” ACS Photonics, vol. 7, no. 11, pp. 3116–3129, 2020, https://doi.org/10.1021/acsphotonics.0c01176.Search in Google Scholar
[53] C. R. Otey, W. T. Lau, and S. H. Fan, “Thermal rectification through vacuum,” Phys. Rev. Letts., vol. 104, no. 15, p. 154301, 2010, https://doi.org/10.1103/physrevlett.104.154301.Search in Google Scholar PubMed
[54] K. F. Chen, P. Santhanam, S. Sandhu, L. X. Zhu, and S. H. 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, no. 13, p. 134301, 2015, https://doi.org/10.1103/physrevb.91.134301.Search in Google Scholar
[55] G. M. Tang, L. Zhang, Y. Zhang, J. Chen, and C. T. Chan, “Near-field energy transfer between graphene and magneto-optic media,” Phys. Rev. Letts., vol. 127, no. 24, p. 247401, 2021, https://doi.org/10.1103/physrevlett.127.247401.Search in Google Scholar
[56] V. Kubytskyi, S.-A. Biehs, and P. Ben-Abdallah, “Radiative bistability and thermal memory,” Phys. Rev. Letts., vol. 113, no. 7, p. 074301, 2014, https://doi.org/10.1103/physrevlett.113.074301.Search in Google Scholar
[57] S. Kawata, Y. Inouye, and P. Verma, “Plasmonics for near-field nano-imaging and superlensing,” Nat. Photonics, vol. 3, no. 7, pp. 388–394, 2009, https://doi.org/10.1038/nphoton.2009.111.Search in Google Scholar
[58] J. L. Wise and D. M. Basko, “Near field versus far field in radiative heat transfer between two-dimensional metals,” Phys. Rev. B, vol. 103, no. 16, p. 165423, 2021, https://doi.org/10.1103/physrevb.103.165423.Search in Google Scholar
[59] K. F. Chen, B. Zhao, and S. H. Fan, “Mesh a free electromagnetic solver for far-field and near-field radiative heat transfer for layered periodic structures,” Comput. Phys. Commun., vol. 231, pp. 163–172, 2018, https://doi.org/10.1016/j.cpc.2018.04.032.Search in Google Scholar
[60] J. Dai, S. A. Dyakov, S. I. Bozhevolnyi, and M. Yan, “Near-field radiative heat transfer between metasurfaces a full-wave study based on two-dimensional grooved metal plates,” Phys. Rev. B, vol. 94, no. 12, p. 125431, 2016, https://doi.org/10.1103/physrevb.94.125431.Search in Google Scholar
[61] S. Basu, B. J. Lee, and Z. M. Zhang, “Near-field radiation calculated with an improved dielectric function model for doped silicon,” J. Heat Trans.-T ASME, vol. 132, no. 2, p. 023302, 2010, https://doi.org/10.1115/1.4000179.Search in Google Scholar
[62] X. L. Liu, B. Zhao, and Z. M. Zhang, “Enhanced near-field thermal radiation and reduced casimir stiction between doped-Si gratings,” Phys. Rev. A, vol. 91, no. 6, p. 062510, 2015, https://doi.org/10.1103/physreva.91.062510.Search in Google Scholar
[63] X. H. Wu, C. J. Fu, and Z. M. Zhang, “Influence of hBN orientation on the near-field radiative heat transfer between graphene/hBN heterostructures,” J. Photon. Energy, vol. 9, no. 3, p. 032702, 2019, https://doi.org/10.1117/1.jpe.9.032702.Search in Google Scholar
Supplementary Material
This article contains supplementary material (https://doi.org/10.1515/nanoph-2024-0729).
© 2025 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.