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

Dirac semimetal-assisted near-field radiative thermal rectifier and thermostat based on phase transition of vanadium dioxide

Open Access Open Access

Abstract

The near-field thermal radiation has broad application prospects in micro-nano-scale thermal management technology. In this paper, we report the Dirac semimetal-assisted (AlCuFe quasicrystal) near-field radiative thermal rectifier (DSTR) and thermostat (DST), respectively. The DSTR is made of a Dirac semimetal-covered vanadium dioxide (VO2) plate and silicon dioxide (SiO2) plate separated by a vacuum gap. The left and right sides of DST are consisted of the SiO2 covered with Dirac semimetal, and the intermediate plate is the VO2. The strong coupling of the surface electromagnetic modes between the Dirac semimetal, SiO2, and insulating VO2 leads to enhance near-field radiative transfer. In the DSTR, the net radiative heat flux of VO2 in the insulating state is much larger than that in metallic state. When the vacuum gap distance d=100 nm, Fermi level EF=0.20 eV, and film thickness t=12 nm, the global rectification factor of DSTR is 3.5, which is 50% higher than that of structure without Dirac semimetal. In the DST, the equilibrium temperature of the VO2 can be controlled accurately to achieve the switching between the metallic and insulating state of VO2. When the vacuum gap distance d=60 nm, intermediate plate thickness δ=30 nm, and film thickness t=2 nm, with the modulation of Fermi level between 0.05-0.15 eV, the equilibrium temperature of VO2 can be controlled between 325-371 K. In brief, when the crystalline state of VO2 changes between the insulating and metallic state with temperature, the active regulation of near-field thermal radiation can be realized in both two-body and three-body parallel plate structure. This work will pave a way to further improve performance of near-field radiative thermal management and modulation.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The heat transfer quantity can exceed several orders of magnitude over the black body limit when the feature size of the radiation heat transfer system is smaller than the peak wavelength determined by Wien's displacement law [1]. In the near-field thermal radiation (NFTR) the evanescent wave replaces the propagating wave as the key to the heat transfer process. This significant enhancement of radiation heat transfer is expected to be used in new energy conversion technologies, and thermal management of nano components, such as electroluminescent cooling [2,3], subwavelength imaging [4,5], thermophotovoltaics [68], to name a few. Recently, NFTR is mainly focused on active regulation of heat transfer quantity.

In order to achieve active regulation, the radiative thermal switch and transistor, which are composed of the black phosphorus or graphene covered with silicon carbide, SiO2, and VO2, have become research hotspots in the field of NFTR [912]. The study of NFTR began with a simple two parallel plates structure. The radiative heat transfer quantity between two parallel plates is calculated accurately by Polder and van Hove for the first time [13]. Since then, a rapid progress has been seen in the field of NFTR, no matter in theoretical or experimental studies.

The research on active control of the NFTR can be roughly divided into two-body and multi-body structures. The research on NFTR of two-body structures is earlier and deeper. Francoeur et al. [14] deduced the expression of near-field radiative heat flux between two finite thickness plates. Biehs et al. [15] used the scattering matrix method to study the NFTR between the two parallel plates composed of multilayer hyperbolic metamaterials. Various types of surface electromagnetic modes can be excited under near-field mechanisms by utilizing materials such as metals [16], polar dielectric materials [17,18], hyperbolic materials [19,20], and two-dimensional materials [21,22]. As for these materials, active control of NFTR can be achieved through reasonable structural design. Yang et al. [23] investigated the effect of magnetic polaritons (MPs) inside the silicon carbide (SiC) grating microstructures on the NFTR. Zhang et al. [24] studied the NFTR between two suspended monolayer black phosphorus (BP) sheets, and investigated the effect of the number of layers, electron density of BP and the mechanical rotation on the NFTR. In addition, a magnetic field is a common method for achieving active thermal regulation and applied to the NFTR system. Moncada-Villa et al. [25] demonstrated the NFTR of doped semiconductors medium can be severely affected by the application of a static magnetic field. Similarly, Song et al. [26] regulated near-field thermal radiation between two multilayer hyperbolic metamaterial plates by utilizing an external magnetic field successfully. As a phase transition material, the vanadium dioxide (VO2) is introduced in NFTR to broaden the application range due to its unique properties. Zwol et al. [27] researched the NFTR between VO2 thin films. The results indicated that the near-field radiative heat flux can be regulated broadly. Yang et al. [28] proposed the concept of thermal rectifier based on the characteristic of VO2. In order to achieve active regulation and enhancement of the heat transfer quantity, the graphene is introduced in NFTR because it is possible to achieve electromagnetic modulation from mid-infrared to terahertz wavelengths [29]. Ilic et al. [30] first studied the role of SPPs of graphene in the NFTR by using the structure of two parallel graphene films. Messina et al. [31] achieved enhancement of the NFTR between polar materials by covering graphene films on polar material substrates. Zhang and Chang et al. [32,33] analyzed the graphene-assisted NFTR configurations by calculating both the near-field dispersion relations and Rabi frequencies. He et al. [34] realized the magnetoplasmonic manipulation of NFTR by adding an external magnetic field on the twisted graphene gratings. Zheng et al. [35] reported a graphene-assisted near-field radiative thermal rectifier and achieved a larger rectification factor under the same conditions compared to the VO2-SiO2 thermal rectifier.

In practical applications, the simple two-body structure cannot meet the demand of the heat transfer of complex equipment obviously. Therefore, the NFTR of the multi-body structures has attracted more and more attentions. Ben Abdallah et al. [36] studied the NFTR between simple multi-body system composed of small interacting body. Choubdar et al. [37] demonstrated the heat flux can be switched effectively by changing the relative direction of ellipsoidal nanoparticles, especially the intermediate nanoparticles. Zheng et al. [38] realized flexible control of the heat flux by changing the thickness and position of the suspended film between two parallel plates. Ott et al. [39] realized the thermal rectification of NFTR between two particles by using the nonreciprocal surface waves excited by the magneto-optical material substrate. Xu et al. [40] built an effective multilayer model to approach the NFTR between random rough surfaces of SiC. Chen and Dong et al. [41,42] investigated the NFTR between two or more SiC/Au nanoparticles inside a cavity configuration composed of two semi-infinite SiC plates. Latella et al. [43] studied the near-field heat engine of a three-body structure and demonstrated that the energy conversion process driven by three-body photon tunneling imparts superior thermodynamic performance to the three-body structure heat engine compared to the two-body structure. When phase transition materials (VO2) and two-dimensional materials (graphene) are applied in NFTR, the multi-body structure can achieve more functions. Song et al. [44] studied the NFTR effect of a three-body structure composed of graphene/SiC core-shell nanoparticles. Zhang et al. [45] analyzed the effect of the separation distances, number of layers and chemical potentials on NFTR in graphene/vacuum multilayers. He et al. [46] studied the steady-state temperature distribution of multilayer graphene systems and observed a unique multi-body enhancement phenomenon. Ben Abdallah et al. [47] designed a three-body structure based on the VO2 and SiO2 plates, which realized the transistor-like effect in NFTR. The thermal transistors have three key elements: the source, the drain, and the gate, which is similar to the electronic counterparts. The effective modulation of heat transfer from the source to the drain can be achieved by manipulating the temperature of the gate. Recently, Li et al. [48] reported an experimental demonstration of thermal transistor effect and an amplification parameter of heat flux over 20 is reached in the experiment. Such structures not only enable heat flux modulation and amplification but also facilitate the design of various functional devices, such as near-field thermal storage devices, thermal photon logic operators, and near-field thermostats [49]. For instance, Ito et al. [50] designed a near-field thermal memory and studied the radiative heat transfer hysteresis between a VO2 film and a fused silica substrate under steady-state conditions. Ben Abdallah et al. [51] introduced the concept of thermal radiation logic gates based on phase-change materials, and discussed the implementation of NOT, OR, and AND gates operating in the near-field regime in detail. He et al. [52] coated the graphene on the simple three-body parallel plate structure, and a near-field thermostat was designed.

In the existing research, graphene is one of the key materials for achieving active regulation of NFTR. However, graphene, as an ultra-thin two-dimensional material, lacks the physical space that could be manipulated and controlled. Moreover, graphene is vulnerable to interference from the external dielectric environment, and the optical absorbance of single-layer graphene is weak [53]. In order to overcome the defects of graphene, researchers pay attention to a new topological quantum material called Dirac semimetal. This kind of material has similar energy band structure and linear dispersion relationship with graphene, and the dielectric function can also be adjusted by regulating Fermi level. Meanwhile, the structure of the topological quantum material is stable and easy to prepare [54]. Uchida et al. [55] introduced a precise growth technique for Dirac semimetal Cd3As2 thin films. The research indicated the Cd3As2 thin films with a thickness between 12 and 100 nm can be prepared using precise growth technique. Moreover, mechanical stripping and selenization technique can be employed to produce semi-metallic films at the atomic level [56,57]. Guo et al. [58] realized the active regulation of the Fermi level of Cd3As2 material successfully through the doping of Mn elements. In addition, similar to the graphene, the Fermi level of Dirac semimetal can be controlled by using a grid voltage control method [59,60]. The Fermi level was closer to the Dirac point by increasing the doping of impurities and applying a higher bias voltage. Dirac semimetal, as a substitute material of graphene, is applied in the field of optics widely. Kotov et al. [61] gave the calculation method of conductivity and dielectric function of Dirac semimetal through theoretical derivation. The research results proved that Dirac semimetal can support SPPs which is similar to graphene. Hu et al. [62] utilized Dirac semimetal and VO2 to realize the dynamic adjustment of resonance frequency and absorption intensity of optical absorption peak. At present, research on the Dirac semimetal in the NFTR fields is limited. Xu et al. [63] coated Dirac semimetal films on the SiO2 plates, and studied the variation of energy transmission coefficient with the thickness of Dirac semimetal, Fermi energy level and degeneracy factor. However, to the best of author’s knowledge, there is no investigation about modulating functions about Dirac semimetal in the two-body and three-body systems.

In order to further explore the performance of Dirac semimetal in the active regulation of NFTR, this paper utilized the Dirac semimetal, VO2, and SiO2 to design the two-body thermal rectifier and three-body thermostat, respectively. The paper is structured as follows. In section 2, the dielectric function and other parameters of materials is given, and calculation method of two-body and three-body parallel plate structure is introduced. In section 3, the NFTR of two-body and three-body structures is analyzed in detailed. Meanwhile, the influence of different parameters on the energy transmission coefficient, net radiative heat flux, and equilibrium temperature is studied, and the interaction forms of various electromagnetic surface modes are explored. Finally, the conclusion is given in section 4.

2. Model introduction

2.1 Heat transfer structure

The configuration of the near-field radiative thermal rectifier is showed in Fig. 1. The DSTR is consisted of a VO2 plate covered with Dirac semimetal and a SiO2 plate, in which the plates are semi-infinite. As shown in Fig. 1, the two plates are separated by a vacuum gap d and the thickness of Dirac semimetal film is t. VO2 behaves like an insulator and exhibits anisotropy, when the temperature of VO2 is lower than phase change temperature 341 K. On the contrary, VO2 behaves like metal and exhibits isotropy.

 figure: Fig. 1.

Fig. 1. Heat transfer model of two-body structure thermal rectifier. Media 1 to 4 represent VO2 plate, Dirac semimetal film, vacuum gap and SiO2 plate, respectively. The temperature of the VO2 and SiO2 plate is set as 300 K and 353 K, and the scenario is defined as forward heat transfer. The heat transfer quantity of NFTR is expressed as ΦF. The temperature of VO2 and SiO2 plates is set as 300 K and 353 K, and the scenario is defined as forward heat transfer. The heat transfer quantity of NFTR is expressed as ΦR.

Download Full Size | PDF

On the basis of the two-body structure thermal rectifier and inspired by the idea of thermal transistors [49], a near-field thermostat based on SiO2-VO2-SiO2 three-body structure covered with Dirac semimetal film is designed, as shown in Fig. 2. The structure is composed of three parallel plates. The left is a semi-infinite SiO2 plate covered with a Dirac semimetal film with the thickness t1, and the temperature is fixed at T1. The middle is a VO2 plate with the thickness δ. The right is a semi-infinite SiO2 plate covered with a Dirac semimetal film with the thickness t2, and the temperature is fixed at T3 (T1 > T3). The middle plate is separated from the left and right plates by a vacuum gap d.

 figure: Fig. 2.

Fig. 2. Heat transfer model of three-body thermostat. Media 1-3 represent the SiO2 plate on the left (400 K), VO2 plate, and SiO2 plate on the right (300 K), and the two Dirac semimetal films are represented by 4 and 5. The heat flux released by the high temperature plate 1 is represented as Φ1, and the heat flux absorbed by low temperature plate 3 is represented as Φ3.

Download Full Size | PDF

2.2 Dielectric function calculation model

Determining the physical properties of materials is the premise for accurate calculation of near-field radiation heat transfer, and the basis for guiding near-field thermal radiation regulation. The dielectric function of Dirac semimetal can be written as [61,63]

$${\varepsilon _\textrm{D}} = {\varepsilon _b} - \frac{{g\alpha c}}{{6\pi {v_F}}}\left[ {{{\left( {\frac{{2{E_F}}}{{\hbar \Omega }}} \right)}^2} + \frac{1}{3}{{\left( {\frac{{2\pi {k_B}T}}{{\hbar \Omega }}} \right)}^2} + i\pi G\left( {\frac{{\hbar \Omega }}{2}} \right) - 8\int_0^{{E_c}} {\frac{{G(\xi )- G\left( {\frac{{\hbar \Omega }}{2}} \right)}}{{{{({\hbar \Omega } )}^2} - 4{\xi^2}}}\xi d\xi } } \right]$$
where, εb represents the background dielectric function, g denotes the degenerate factor of 3D Dirac point; α refers to the fine structure constant; c indicates the speed of light; vF corresponds to the Fermi velocity; EF means the Fermi level; kB is the Boltzmann constant; T is the temperature of Dirac semimetal; $\hbar$ is the Planck’s constant divided by 2π; Ω=ω+−1, where ${\tau ^{ - 1}} = {{ev_F^2} / {\mu {E_F}}}$ is the relaxation time and µ indicates the carrier mobility; Fermi Dirac distribution function G(ξ)=sinh(ξ/kBT)/[cosh(ξ/kBT)+cosh(EF/kBT)]; Ec means the cut-off energy. In present study, the AlCuFe quasicrystal, which is a typical Dirac semimetal material, is employed to analyze the performance of the DSTR and DST. The degeneracy factor, carrier mobility, and cut-off energy of the AlCuFe quasicrystal are 40, 6.42m2/($\textrm{V} \cdot \textrm{s}$), and 3EF, respectively [63,64].

The crystal structure of VO2 will be transformed near 341 K, and the dielectric function will change accordingly. VO2 is in the insulating state when the temperature is lower than 341 K, and the dielectric function shows anisotropic characteristics. Assuming that the optical axis is the z-axis and perpendicular to the surface, the dielectric function of VO2 can be expressed as a tensor [28]

$$\overline{\overline \varepsilon } = \left[ {\begin{array}{{ccc}} {{\varepsilon_ \bot }}&0&0\\ 0&{{\varepsilon_ \bot }}&0\\ 0&0&{{\varepsilon_\parallel }} \end{array}} \right]$$
where the ordinary dielectric function ε and extraordinary dielectric function ε can be calculated by Lorentz model, as following [65]
$$\varepsilon (\omega = {\varepsilon _\infty } + \sum\limits_{j = 1}^N {\left[ {\frac{{({{S_j}\omega_j^2} )}}{{({\omega_j^2 - i{\gamma_j}\omega - {\omega^2}} )}}} \right]}$$
where ω indicates the angular frequency; ε denotes the high frequency constant; Sj means the phonon intensity; ωj represents the phonon frequency; γj refers to the damping coefficient; the subscript j represents the jth optical phonon frequency.

VO2 is in the metallic state when the temperature is higher than 341 K, and the dielectric function can be described by Drude model as [65]

$$\varepsilon (\omega ={-} \frac{{{\varepsilon _\infty }\omega _p^2}}{{{\omega ^2} + i{\omega _c}\omega }}$$
where ωp is the plasma frequency, ωc is the collision frequency. The relative dielectric function of SiO2 adopts Palik's data [66].

2.3 Calculation model of near-field thermal radiation

For the DSTR, the net radiative heat flux can be written as [13,67]

$$\Phi = \int_0^\infty {\frac{{d\omega }}{{2\pi }}\varphi (\omega )}$$
where φ(ω) is the spectral radiative heat flux, which can be expressed as [35,67]
$$\varphi (\omega ) = \frac{{\hbar \omega {n_{12}}}}{{4{\pi ^2}}}\int_0^\infty {\xi (\omega ,\kappa )\kappa d\kappa }$$
where n12(ω)=n1(ω)-n2(ω) is the difference for the average energy of the harmonic oscillators at different temperatures, in which ni(ω) = [exp(ħω/kBTi)-1]-1; $\kappa$ denotes the component of the wave vector, which is parallel to the interface; ξ(ω,κ) refers to the energy transmission coefficient. Since the dielectric function of VO2 in the insulating state is anisotropic, the contribution of the propagating wave (κ<ω/c) to the NFTR can be expressed as [28,35,68]
$${\xi _{\textrm{prop}}}(\omega ,\kappa ) = \frac{{({1 - {{|{R_{321}^s} |}^2}} )({1 - {{|{r_{34}^s} |}^2}} )}}{{{{|{1 - R_{321}^sr_{34}^s{e^{2ik_{z,3}^sd}}} |}^2}}} + \frac{{({1 - {{|{R_{321}^p} |}^2}} )({1 - {{|{r_{34}^p} |}^2}} )}}{{{{|{1 - R_{321}^pr_{34}^p{e^{2ik_{z,3}^pd}}} |}^2}}}$$

The contribution of the evanescent wave (κ>ω/c) to the NFTR is expressed as [28,35,68]

$${\xi _{\textrm{evan}}}(\omega ,\kappa ) = \frac{{4{\mathop{\rm Im}\nolimits} ({R_{321}^s} ){\mathop{\rm Im}\nolimits} ({r_{34}^s} ){e^{ - 2{\mathop{\rm Im}\nolimits} ({k_{z,3}^s} )d}}}}{{{{|{1 - R_{321}^sr_{34}^s{e^{2ik_{z,3}^sd}}} |}^2}}} + \frac{{4{\mathop{\rm Im}\nolimits} ({R_{321}^p} ){\mathop{\rm Im}\nolimits} ({r_{34}^p} ){e^{ - 2{\mathop{\rm Im}\nolimits} ({k_{z,3}^p} )d}}}}{{{{|{1 - R_{321}^pr_{34}^p{e^{2ik_{z,3}^pd}}} |}^2}}}$$
where d refers to the distance of vacuum gap between two plates; ${\mathop{\rm Im}\nolimits}$ indicates to the imaginary part; Rs 321 and Rp 321 indicate the reflection coefficients under s and p polarizations when the electromagnetic wave enters into VO2 plate covered by Dirac semimetal film from vacuum; rs 34 and rp 34 represent the reflection coefficients under s and p polarizations when the electromagnetic wave enters into the SiO2 plate from vacuum.

Rs 321 and Rp 321 can be expressed respectively as [63]

$$R_{321}^s = \frac{{\left( {1 - \frac{{k_{z,1}^s}}{{k_{z,3}^s}}} \right)\cos k_{z,2}^st + i\left( {\frac{{k_{z,2}^s}}{{k_{z,3}^s}} - \frac{{k_{z,1}^s}}{{k_{z,2}^s}}} \right)\sin k_{z,2}^st}}{{\left( {1 + \frac{{k_{z,1}^s}}{{k_{z,3}^s}}} \right)\cos k_{z,2}^st - i\left( {\frac{{k_{z,2}^s}}{{k_{z,3}^s}} + \frac{{k_{z,1}^s}}{{k_{z,2}^s}}} \right)\sin k_{z,2}^st}}$$
$$R_{321}^p = \frac{{\left( {1 - \frac{{k_{z,1}^p}}{{{\varepsilon_1}k_{z,3}^p}}} \right)\cos k_{z,2}^pt + i\left( {\frac{{k_{z,2}^p}}{{{\varepsilon_2}k_{z,3}^p}} - \frac{{{\varepsilon_2}k_{z,1}^p}}{{{\varepsilon_1}k_{z,2}^p}}} \right)\sin k_{z,2}^pt}}{{\left( {1 + \frac{{k_{z,1}^p}}{{{\varepsilon_1}k_{z,3}^p}}} \right)\cos k_{z,2}^pt - i\left( {\frac{{k_{z,2}^p}}{{{\varepsilon_2}k_{z,3}^p}} + \frac{{{\varepsilon_2}k_{z,1}^p}}{{{\varepsilon_1}k_{z,2}^p}}} \right)\sin k_{z,2}^pt}}$$
rs 34 and rp 34 can be expressed respectively as [35,68]
$$r_{34}^s = \frac{{k_{z,3}^s - k_{z,4}^s}}{{k_{z,3}^s + k_{z,4}^s}}$$
$$r_{34}^p = \frac{{{\varepsilon _4}k_{z,3}^p - {\varepsilon _3}k_{z,4}^p}}{{{\varepsilon _4}k_{z,3}^p + {\varepsilon _3}k_{z,4}^p}}$$
where εi is the dielectric function of medium i; ks z,i and kp z,i are the vertical component of the wave vector in medium i under s and p polarizations, ks z,i and kp z,i can be expressed as [35,68]
$$k_{z,i}^s = \sqrt {{\varepsilon _{ \bot ,i}}{{({{\omega / c}} )}^2} - {\kappa ^2}}$$
$$k_{z,i}^p = \sqrt {{\varepsilon _{ \bot ,i}}{{({{\omega / c}} )}^2} - ({{{{\varepsilon_{ \bot ,i}}{\kappa^2}} / {{\varepsilon_{{\parallel} ,i}}}}} )}$$
where ε⊥,2 and ε‖,2 are the components of dielectric function perpendicular to the optical axis and parallel to the optical axis respectively. Equations (13) and (14) can be employed to compute the vertical component of the wave vector when VO2 is in the metallic state.

For the DST, the net radiative heat flux absorbed by plate 3 can be written as [13,69,70]

$${\Phi _3} = \int_0^\infty {\frac{{d\omega }}{{2\pi }}{\varphi _3}(\omega )}$$
where φ3(ω) is the spectral radiative heat flux, which can be written as [13,71,72]
$${\varphi _3}(\omega )\textrm{ = }\hbar \omega \sum\limits_{j = s,p} {\int {\frac{{{d^2}\kappa }}{{{{({2\pi } )}^2}}}[{{n_{12}}(\omega )\xi_j^{1 - 2}({\omega ,\kappa } )+ {n_{23}}(\omega )\xi_j^{2 - 3}({\omega ,\kappa } )} ]} }$$
where j = s, p represents s and p polarizations, respectively; nij(ω)=ni(ω)-nj(ω)refers to the difference for the average energy of the harmonic oscillators at different temperatures, where ni(ω) = [exp(ħω/kBTi)-1]-1; ξ1-2 j(ω,κ) and ξ2-3 j(ω,κ) denote the energy transmission coefficients between plates 1 and 2 and between plates 2 and 3 in the case of j polarization, respectively.

ξ1-2 j(ω,κ) and ξ2-3 j(ω,κ) can be expressed as [52,71]

$$\xi _j^{1 - 2}({\omega ,\kappa } )\textrm{ = }\left\{ {\begin{array}{{c}} {\frac{{{{|{\tau_2^j} |}^2}({1 - {{|{\rho_1^j} |}^2}} )({1 - {{|{\rho_3^j} |}^2}} )}}{{{{|{1 - \rho_{12}^j\rho_3^j{e^{2i{k_{z,0}}d}}} |}^2}{{|{1 - \rho_1^j\rho_2^j{e^{2ik_{z,0}^dd}}} |}^2}}},\textrm{ }\kappa < \frac{\omega }{c}}\\ {\frac{{4{{|{\tau_2^j} |}^2}{\mathop{\rm Im}\nolimits} ({\rho_1^j} ){\mathop{\rm Im}\nolimits} ({\rho_3^j} ){e^{ - 4{\mathop{\rm Im}\nolimits} ({{k_{z,0}}} )d}}}}{{{{|{1 - \rho_{12}^j\rho_3^j{e^{2i{k_{z,0}}d}}} |}^2}{{|{1 - \rho_1^j\rho_2^j{e^{2ik_{z,0}^dd}}} |}^2}}},\textrm{ }\kappa > \frac{\omega }{c}} \end{array}} \right.$$
$$\xi _j^{2 - 3}({\omega ,\kappa } )\textrm{ = }\left\{ {\begin{array}{{c}} {\frac{{({1 - {{|{\rho_{12}^j} |}^2}} )({1 - {{|{\rho_3^j} |}^2}} )}}{{{{|{1 - \rho_{12}^j\rho_3^j{e^{2i{k_{z,0}}d}}} |}^2}}},\textrm{ }\kappa < \frac{\omega }{c}}\\ {\frac{{4{\mathop{\rm Im}\nolimits} ({\rho_{12}^j} ){\mathop{\rm Im}\nolimits} ({\rho_3^j} ){e^{ - 2{\mathop{\rm Im}\nolimits} ({{k_{z,0}}} )d}}}}{{{{|{1 - \rho_{12}^j\rho_3^j{e^{2i{k_{z,0}}d}}} |}^2}}},\textrm{ }\kappa > \frac{\omega }{c}} \end{array}} \right.$$
where τj 2 is the transmission coefficient of the plate 2 under j (j = s, p) polarization; ρj i is the reflection coefficient of the plate i (i = 1, 2, 3) under j polarization; ρj 12 is the whole reflection coefficient composed of plate 1 and plate 2 under j polarization.

τj 2, ρj i, and ρj 12 can be expressed as [46,69]

$$\tau _2^j = \frac{{t_2^j\overline {t_2^j} {e^{ik_{z,2}^j{\delta _2}}}}}{{1 - {{({r_2^j} )}^2}{e^{ik_{z,2}^j{\delta _2}}}}}\quad \quad \rho _i^j = r_i^j\frac{{1 - {e^{ik_{z,i}^j{\delta _2}}}}}{{1 - {{({r_i^j} )}^2}{e^{ik_{z,i}^j{\delta _2}}}}}\quad \quad \rho _{12}^j = \rho _2^j + \frac{{\rho _1^j{{({\tau_2^j} )}^2}{e^{2i{k_{z,0}}d}}}}{{1 - \rho _1^j\rho _2^j{e^{2i{k_{z,0}}d}}}}$$
where δi is the thickness of plate i; rj i is the vacuum-medium Fresnel reflection coefficient of plate i in the case of j polarization; tj 2 is the vacuum-medium transmission coefficient of plate 2 in the case of j polarization; $\overline {t_2^j}$ is the medium-vacuum transmission coefficient of plate 2 in the case of j polarization.

For SiO2 plates 1 and 3 covered by the Dirac semimetal film (marked as D) with thickness t, rj i(i = 1,3) can be expressed as [46,63,69]

$$r_i^s = \frac{{\left( {1 - \frac{{{k_{z,i}}}}{{{k_{z,0}}}}} \right)\cos {k_{z,D}}t + i\left( {\frac{{{k_{z,D}}}}{{{k_{z,0}}}} - \frac{{{k_{z,i}}}}{{{k_{z,D}}}}} \right)\sin {k_{z,D}}t}}{{\left( {1 + \frac{{{k_{z,i}}}}{{{k_{z,0}}}}} \right)\cos {k_{z,D}}t - i\left( {\frac{{{k_{z,D}}}}{{{k_{z,0}}}} + \frac{{{k_{z,i}}}}{{{k_{z,D}}}}} \right)\sin {k_{z,D}}t}}$$
$$r_i^p = \frac{{\left( {1 - \frac{{{k_{z,i}}}}{{{\varepsilon_i}{k_{z,0}}}}} \right)\cos {k_{z,D}}t + i\left( {\frac{{{k_{z,D}}}}{{{\varepsilon_D}{k_{z,0}}}} - \frac{{{\varepsilon_D}{k_{z,i}}}}{{{\varepsilon_i}{k_{z,D}}}}} \right)\sin {k_{z,D}}t}}{{\left( {1 + \frac{{{k_{z,i}}}}{{{\varepsilon_i}{k_{z,0}}}}} \right)\cos {k_{z,D}}t - i\left( {\frac{{{k_{z,D}}}}{{{\varepsilon_D}{k_{z,0}}}} + \frac{{{\varepsilon_D}{k_{z,i}}}}{{{\varepsilon_i}{k_{z,D}}}}} \right)\sin {k_{z,D}}t}}$$
where kz,α is the vertical component of the wave vector in the medium α ($\alpha = 0,i,D$), in which 0, i, and D represent vacuum, SiO2 and Dirac semimetal; εα is the relative dielectric function of the material α.

For the intermediate plate 2, rj 2 can be expressed as [35,46,69]

$$r_2^s = \frac{{{k_{z,0}} - k_{z,2}^s}}{{{k_{z,0}} + k_{z,2}^s}},\quad \quad r_2^p = \frac{{{\varepsilon _{ \bot ,2}}{k_{z,0}} - k_{z,2}^p}}{{{\varepsilon _{ \bot ,2}}{k_{z,0}} + k_{z,2}^p}}$$
where ks z,2 and kp z,2 are the vertical component of the wave vector in the VO2 under s and p polarizations, respectively; ε⊥,2 is the component of dielectric function perpendicular to the optical axis in the VO2.

tj 2 and $\overline {t_2^j}$ can be expressed as [46,52,69]

$$t_2^s = \frac{{2{k_{z,0}}}}{{{k_{z,0}} + k_{z,2}^s}},\textrm{ }\quad \textrm{ }t_2^p = \frac{{2\sqrt {{\varepsilon _{ \bot ,2}}} {k_{z,0}}}}{{{\varepsilon _{ \bot ,2}}{k_{z,0}} + k_{z,2}^p}}$$
$$\overline {t_2^s} = \frac{{2k_{z,2}^s}}{{{k_{z,0}} + k_{z,2}^s}},\textrm{ }\quad \textrm{ }\overline {t_2^p} = \frac{{2\sqrt {{\varepsilon _{ \bot ,2}}} k_{z,2}^p}}{{{\varepsilon _{ \bot ,2}}{k_{z,0}} + k_{z,2}^p}}$$

Similarly, the spectral radiative heat flux released by plate 1 with temperature T1 can be written as [52,71,72]

$${\varphi _1}(\omega )= \hbar \omega \sum\limits_{j = s,p} {\int {\frac{{{d^2}\kappa }}{{{{({2\pi } )}^2}}}[{{n_{23}}(\omega )\xi_j^{2 - 3,\ast }({\omega ,\kappa } )+ {n_{12}}(\omega )\xi_j^{1 - 2,\ast }({\omega ,\kappa } )} ]} }$$

The calculation of energy transmission coefficient only needs to exchange the subscript 1 and 3 in Eqs. (17) and (18).

3. Results and discussion

3.1 Properties of dielectric functions

The dielectric function of the material is the basis of calculation of the energy transmission coefficient and net radiative heat flux [24]. The real and imaginary parts of the relative dielectric function of Dirac semimetal, VO2 and SiO2 are shown in Figs. 35, respectively.

 figure: Fig. 3.

Fig. 3. The relative dielectric function of Dirac semimetal (a) real part; (b) imaginary part

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. The relative dielectric function of VO2 (a) real part; (b) imaginary part

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. The relative dielectric function of SiO2 (a) real part; (b) imaginary part

Download Full Size | PDF

Most of the electromagnetic surface modes are closely related to the dielectric function of the material in NFTR. Dirac semimetal supports surface plasmon polaritons (SPPs), and the electromagnetic mode can be effectively regulated by adjusting Fermi level. Figure 3 shows the relative dielectric functions of Dirac semimetal with different Fermi levels. The effect of SPPs on NFTR will be enhanced with the increase of the relative dielectric function. SiO2 supports surface phonon polarizations (SPhPs). When the real part of the dielectric function is less than 0 (Re(ε) < 0), SPhPs of the vacuum-SiO2 interface can be excited under the p-polarization. VO2 supports SPhPs, type I and type II surface hyperbolic modes (HMsI and HMsII). When the real part of the ordinary dielectric function and extraordinary dielectric function are both less than 0 (Re(ε) < 0 and Re(ε) < 0), SPhPs can be excited under p-polarization. HMsI can only be excited when Re(ε) > 0 and Re(ε) < 0, however, HMsII can only be excited when Re(ε) < 0 and Re(ε) > 0.

3.2 Discussion on two-body structure near-field thermal rectifier

Two-body parallel plate structure is one of the focuses in the NFTR. In the two-body structure, the variation trend of quantity of the heat transfer with various parameters can be observed directly, and the interaction of various electromagnetic surface modes can be explored effectively. Therefore, the NFTR between DSTR is explored at first. The forward and reverse heat transfer between two-body structure is investigated because VO2 shows different properties in insulating and metallic state.

The forward and reverse energy transmission coefficients of DSTR under different combinations of the vacuum gap distance d, Fermi level EF, and film thickness t are shown in Fig. 6. As shown in Fig. 6, the energy transmission coefficient decreases with the increase of the vacuum gap distance d. The energy transmission coefficient of forward heat transfer move towards the high frequently and low wave vector as the Fermi level EF increases. Meanwhile, the energy transmission coefficient of reverse heat transfer is also enhanced slightly. With the increase of film thickness t, the energy transmission coefficient is shifted to direction of low wave vector furtherly (see Fig. 6(d)).

 figure: Fig. 6.

Fig. 6. Energy transmission coefficients of forward and reverse heat transfer of two-body structure under different parameter combinations. (a) d = 10 nm, EF =0.15 eV, t = 5 nm; (b) d = 20 nm, EF =0.15 eV, t = 5 nm; (c) d = 10 nm, EF =0.18 eV, t = 5 nm; (d) d = 10 nm, EF =0.15 eV, t = 10 nm. In Fig, the left side represents the energy transmission coefficient of forward heat transfer, and the right side represents the energy transmission coefficient of reverse heat transfer.

Download Full Size | PDF

With the vacuum gap distance d = 50 nm, Fermi level EF = 0.25 eV, and film thickness t = 4 nm, meanwhile, the structure without Dirac semimetal film under the same vacuum gap, the forward and reverse spectral radiative heat flux are shown in Fig. 7. The rectification effect of the two-body structure can be significantly enhanced by utilizing the Dirac semimetal film (see Fig. 7). This enhancement is achieved by an additional heat transfer near ω=2.14 × 1014 rad/s during forward heat transfer, while the reverse heat transfer is almost consistent with the non-covered structure. The forward net radiative heat flux of the DSTR, which is obtained by integral, is 2.22 × 104 W/m2. The forward net radiative heat flux of the structure without Dirac semimetal film is 1.47 × 104 W/m2. The forward net radiative heat flux is increased by 50% as the Dirac semimetal film is applied. On the contrary, the reverse net radiative heat flux of the DSTR is slightly smaller than that of the structure without Dirac semimetal film.

 figure: Fig. 7.

Fig. 7. Variation trends of the spectral radiative heat flux with different conditions. The black solid line and black dashed line represent the forward heat transfer scenario of the film-covered structure (d = 50 nm, EF = 0.25 eV, and t = 4 nm) and non-covered structure, respectively, while the red solid line and red dashed line illustrates the reverse heat transfer scenario of the film-covered structure and non-covered structure, respectively.

Download Full Size | PDF

In order to illustrate rectification effect of DSTR, the global rectification factor is defined as M, which can be expressed as [35]

$$M = \frac{{{\varphi _\textrm{F}}}}{{{\varphi _\textrm{R}}}}$$
where φF refers to the net radiative heat flux of forward heat transfer scenario; φR denotes the net radiative heat flux of reverse heat transfer scenario.

To offer an intuitive explanation of the rectification effect variation caused by the coverage of the Dirac semimetal film, the global rectification factor of the DSTR is shown in Fig. 8(a), where the Dirac semimetal film changes between 0-20 nm and Fermi level varies from 0.05-0.25 eV. In this part, the vacuum gap is set as 100 nm because it is attainable in existing near-field thermal radiation experimental research. In Fig. 8(a), the black contour line represents the global rectification factor of the structure without Dirac semimetal film, and the value is 2.45. It is evident that the global rectification factor of the DSTR is larger than that of the structure without Dirac semimetal film in the vast majority of cases (see Fig. 8(a)). However, as the film thickness and Fermi level increase, the surface electromagnetic mode is suppressed, resulting in a deterioration of the rectification effect of film-covered structure. The numerical experiments demonstrate that the rectification effect of the DSTR can be enhanced by adjusting the Fermi level and thickness of the Dirac semimetal film (Meanwhile, the rectification effect of the DSTR is larger than that of non-film structure). It offers valuable insights for optimizing the Dirac semimetal film thickness and the Fermi level in experimental studies.

 figure: Fig. 8.

Fig. 8. (a) Contour nephogram of the global rectification factor of DSTR as a function of film thickness and Fermi level. The black contour line represents the global rectification factor of non-coverd structure; (b) Contour nephogram of the global rectification factor of DSTR as a function of vacuum gap distance and Fermi level.

Download Full Size | PDF

In practical applications, the active regulation of the film thickness is challenging, but the vacuum gap is controllable within a certain range. Meanwhile, the Fermi level can be controlled using grid voltage. The influence of the vacuum gap distance d and Fermi level EF on the global rectification factor is investigated, where the thickness of the Dirac semimetal film is 10 nm. Figure 8(b) shows the global rectification factor with vacuum gap distance between 50-200 nm and Fermi level between 0.05-0.25 eV. As shown in Fig. 8(b), the maximum global rectification factor is 4.4, when the Dirac semimetal film is existed. For the non-covered structure, as the vacuum gap varies between 50 nm and 200 nm, the global rectification factor decreases from 2.76 to 2.04. The optical Fermi level with different vacuum gap and film thickness can be determined by numerical experiment to help in selecting the best experimental parameter settings.

The variation of global rectification factor with temperature is studied, in which the value of the film thickness, vacuum gap distance, and Fermi level are 10 nm, 100 nm, and 0.18 eV. The reason is that the combination of parameters has good rectification effect and can be realized by existing experimental technology. To ensure that the VO2 maintains in the insulating state under the forward heat transfer scenario and maintains in the metallic state under the reverse heat transfer scenario, the temperature of the low-temperature plate is fixed at 300 K while the temperature range of the high-temperature plate is varied from 350 K to 450 K. Therefore, the temperature difference varies from 50 K to 150 K. Figure 9 shows the variation of global rectification factor of the DSTR with temperature difference. As a comparison, the global rectification factor of the structure without Dirac semimetal film is also shown in Fig. 9. It can be observed that when the temperature difference increases, the decreasing trend of the global rectification factor in film-covered structure is smoother that in the non-covered structure.

 figure: Fig. 9.

Fig. 9. The variation of global rectification factor with temperature difference under specific combination of parameters. The black solid line represents the global rectification factor of film-covered structure, while the red line represents the global rectification factor of non-covered structure.

Download Full Size | PDF

3.3 Discussion on three-body structure near-field thermostat

Based on the research of two-body parallel plate structure, the middle plate is introduced to form a three-body parallel plate structure to expand the application value of NFTR furtherly (see Fig. 2). Based on the principle of the near-field thermostat [52], the temperature of the intermediate plate can be controlled by changing the parameters of the Dirac semimetal film. Furthermore, an independent near-field radiative thermal transistor-like effect without external heat flux can be realized by combining with the phase-change characteristics of VO2.

The energy transmission coefficients of the three-body parallel plate structure are shown in Fig. 10, in which the Fermi level and thickness of the Dirac semimetal films on both sides are equal. In Fig. 10, the energy transmission coefficient from the intermediate plate to the low-temperature plate is significantly higher than that from the high-temperature plate to the intermediate plate whether in the insulating or metallic state. The primary reason for this phenomenon is that although the Fermi level and thickness of Dirac semimetal films are equal, the two films are at different temperatures. It leads to that the differences exist in the dielectric function. The symmetry of the energy transmission coefficient is affected. Meanwhile, compared to the insulating state, the distribution of the energy transmission coefficient of the metallic state is more concentrated, but the value is smaller (see Fig. 10). The phenomenon is consistent with the results in the two-body structure. This fact can be predicted that when VO2 is in the metallic state, the net radiative heat flux is also less than that of the insulating state greatly in three-body structure.

 figure: Fig. 10.

Fig. 10. Energy transmission coefficient of three-body structure when EF1 = EF2 = 0.15 eV and t1 = t2 = 5 nm. (a) VO2 in the insulating state; (b) VO2 in the metallic state. In each figure, the left side represents the energy transmission coefficient between plate 1 and plate 2, and the right side represents the energy transmission coefficient between plate 2 and plate 3.

Download Full Size | PDF

The thermal equilibrium of the three-body structure is investigated to explain the working mechanism furtherly. When the net radiative heat flux Φ1 released by plate 1 is equal to the net radiative heat flux Φ3 absorbed by plate 3, the heat transfer reaches the steady state and plate 2 reaches the equilibrium temperature T2,eq. At this time, the net radiative heat flux of the three-body structure is defined as the equilibrium heat flux Φeq. To illustrate the effect of the Dirac semimetal film on heat transfer in DST, a contour nephogram of the equilibrium temperature T2,eq and equilibrium heat flux Φeq is calculated when the Fermi levels and film thickness range from 0.05-0.15 eV and 2-20 nm, and the thickness of the intermediate plate and vacuum gap are 30 nm and 60 nm as shown in Fig. 11. In addition, a contour line representing the equilibrium heat flux of non-film covered structure is existed in Fig. 11(b) to demonstrate the role of the Dirac semimetal film directly. It can be clearly observed that the Dirac semimetal film can enhance the heat transfer quantity of the DST in most cases. In Fig. 11(b), the maximum equilibrium heat flux is 2.62 × 104 W/m2, which is nearly double that in the structure without Dirac semimetal film (1.57 × 104 W/m2).

 figure: Fig. 11.

Fig. 11. Contour nephograms of equilibrium temperature T2,eq and equilibrium heat flux Φeq varying with Fermi level EF and film thickness t at intermediate plate initial temperature T2 = 400 K. (a) equilibrium temperature T2,eq; (b) equilibrium heat flux Φeq. The black solid line in (b) represents the equilibrium heat flux of non-film covered structure.

Download Full Size | PDF

The principle of the near-field thermostat effect is introduced in the part. Meanwhile, the realization approach of thermal transistor-like effect is discussed in DST. Figure 12 shows the equilibrium temperature T2,eq and equilibrium heat flux Φeq change with the Fermi level, when the initial temperature of the intermediate plate is 300 K and thickness of the Dirac semimetal film is 2 nm. As shown in Fig. 12(a), the equilibrium temperature of the intermediate plate can be effectively controlled within a range of 325-371 K, as the Fermi level ranges from 0.05 eV to 0.15 eV. Notably, this temperature includes the phase transition temperature of VO2 (the phase transition temperature of VO2 is 341 K). This fact illustrates the DST can realize the thermostat effect. The Fig. 12(b) reveals the significant difference in the equilibrium heat flux at different Fermi level. The difference is mainly caused by the phase transition properties of VO2. As shown in Fig. 12(b), the structure demonstrates a remarkable thermal transistor-like effect, where the maximum heat flux near 5.5 × 104 W/m2 and the minimum heat flux is around 1.9 × 104 W/m2. This substantial difference of 3 times highlights the outstanding thermal modulation capabilities of the DST.

 figure: Fig. 12.

Fig. 12. Contour nephograms of equilibrium temperature T2,eq and equilibrium heat flux Φeq varying with Fermi level EF at intermediate plate initial temperature T2 = 300 K. (a) equilibrium temperature T2,eq; (b) equilibrium heat flux Φeq. In Fig. 10, black solid line in (a) represents the phase transition temperature contour of VO2, which corresponds to the black solid line in (b).

Download Full Size | PDF

When the initial temperature of the intermediate plate changes from 300 K to 400 K, the equilibrium temperatures T2,eq and equilibrium heat flux Φeq of the intermediate plate under different Fermi level are shown in Fig. 13(a) and (b), respectively. Combined with the Fig. 12, the interesting phenomenon that the initial temperature of the intermediate plate will affect the final equilibrium temperature can be found in this study. In order to explain the phenomenon and explore the regulation mechanism of equilibrium heat flux, the heat flux released by plate 1 and absorbed by plate 3 with different Fermi level are investigated in detail. The change process of the intermediate plate temperature is also analyzed. The thickness of Dirac semimetal films on both sides is set as 2 nm, and Fermi level of Dirac semimetal film covered on the low-temperature SiO2 plate is 0.15 eV. The Fermi level of Dirac semimetal film covered on the high-temperature SiO2 plate is set as 0.05 eV, 0.08 eV and 0.12 eV, respectively. As the intermediate plate temperature changes between 300 K and 400 K, the heat flux released by plate 1 and absorbed by plate 3 are shown in Fig. 14. Figure 14 also depicts the results of non-covered structure for comparison purposes.

 figure: Fig. 13.

Fig. 13. Contour nephograms of equilibrium temperature T2,eq and equilibrium heat flux Φeq varying with Fermi level EF at intermediate plate initial temperature T2 = 400 K. (a) equilibrium temperature T2,eq; (b) equilibrium heat flux Φeq. Auxiliary lines are the same as those in Fig. 12.

Download Full Size | PDF

 figure: Fig. 14.

Fig. 14. Heat flux released by plate 1 and absorbed by plate 3 under different conditions. The solid line represents the heat flux released by plate 1, and the dotted line represents the heat flux absorbed by plate 3. The black line combination corresponds to case 1, the blue line combination corresponds to case 2, and the red line combination corresponds to case 3.

Download Full Size | PDF

In Fig. 14, the four intersection points might be the equilibrium temperature T2,eq of the intermediate plate 2 under the thermal equilibrium state of the system. As shown in Fig. 14, when the Fermi level of Dirac semimetal covered on high-temperature SiO2 plate is 0.05 eV and 0.12 eV, the equilibrium temperature of the intermediate plate is fixed at 325 K or 352 K (points 1 and 4). The equilibrium temperature is located in the region of insulating and metallic state of VO2, respectively. For the structure without Dirac semimetal film, the result is similar to the case of the Fermi level of 0.05 eV. An interesting phenomenon that two different intersection points are existed in absorption and release heat flux in the whole temperature range from 300 K to 400 K occurs, when the Fermi level of Dirac semimetal covered on high-temperature SiO2 plate is 0.07 eV. The corresponding temperature of intersection point 2 is located in the insulating state region of VO2, and the corresponding temperature of intersection point 3 is located in the metallic state region. When the initial temperature T2 of the intermediate plate is lower than the phase change temperature 341 K, the equilibrium temperature T2,eq is equal to 333 K (point 2). The equilibrium temperature T2,eq is equal to 346 K (point 3) if the initial temperature T2 is higher than the phase change temperature 341 K. This is the reason that the equilibrium temperature of the intermediate plate is affected by the initial temperature.

In order to explore the role of Dirac semimetal in the three-body structure near-field thermostat. The coupling effect between the SPPs supported by Dirac semimetal, SPhPs supported by SiO2, and SPhPs, HMsI, and HMsII supported by VO2 is investigated in detail. The p-polarized energy transmission coefficients (${\xi_{2p}^{1-}}+{\xi_{3p}^{2-}}/2$ of the three-body structure are calculated for different Fermi levels of Dirac semimetal. It can be seen obviously from Fig. 15(a) that a new mode of energy transmission coefficient appears in the low-frequency region due to the addition of Dirac semimetal. The new mode of energy transmission coefficient is caused by the SPPs supported by Dirac semi metal. Meanwhile, the energy transmission coefficient increases as with the increasement of the Fermi level significantly. As shown in Fig. 15(b), the Fermi level of Dirac semimetal films has no obvious effect on the energy transmission coefficient of the three-body structure. This is due to the fact that when the intermediate plate VO2 is in the metallic state, VO2 cannot support any electromagnetic surface mode. In addition, the SPhPs supported by SiO2, the SPPs supported by Dirac semimetal, and the interaction between them are suppressed greatly.

 figure: Fig. 15.

Fig. 15. The p-polarized energy transmission coefficients (${\xi_{2p}^{1-}}+{\xi_{3p}^{2-}}/2$ of the 5 nm film-covered three-body structure under different Fermi levels of Dirac semimetal. (a) intermediate plate VO2 is in insulating state; (b) intermediate plate VO2 is in metallic state. The four pictures from left to right show the situation without Dirac semimetal film coverage and when the Fermi levels of Dirac semimetal film are 0.05 eV, 0.10 eV and 0.15 eV respectively.

Download Full Size | PDF

4. Conclusions

This paper focuses on the active regulation of NFTR. We report the Dirac semimetal-assisted DSTR and DST, respectively. The DSTR is made of a Dirac semimetal-covered vanadium dioxide VO2 plate and SiO2 plate separated by a vacuum gap. The left and right sides of DST are consisted of the SiO2 covered with Dirac semimetal, and the intermediate plate is the VO2. The following conclusions are obtained:

  • 1. In the NFTR, the Dirac semimetal supports the SPPs, which can interact with SPhPs supported by SiO2, and HMsI and HMsII supported by VO2.
  • 2. In the two-body structure, the net radiative heat flux of VO2 in the insulating state is significantly higher than that of VO2 plate in the metallic state. The thermal rectification can be implemented. The NFTR of the two-body structure can be adjusted by changing the Fermi level EF, thickness of the Dirac semimetal film t, and vacuum gap distance d.
  • 3. In the three-body structure, the equilibrium temperature of the intermediate plate can be adjusted to make it slightly higher and lower than the phase transition temperature based on the principle of a near field thermostat. The equilibrium heat flux will step with the switching of VO2 between the metallic and insulating state, so as to realize the near-field thermal transistor-like effect.

Funding

National Natural Science Foundation of China (52106125); Science and Technology Program of Hunan Province (2021RC4006).

Disclosures

The authors declare no conflicts of interest

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. S. A. Biehs, O. Huth, and F. Rueting, “Near-field radiative heat transfer for structured surfaces,” Phys. Rev. B 78(8), 085414 (2008). [CrossRef]  

2. Y. Chen, Z. Zheng, and Y. Xuan, “Effective modulation of the near-field heat flux with radiative thermal switch based on electrochromic effects of tungsten trioxide,” J. Quant. Spectrosc. Radiat. Transfer 218, 171–177 (2018). [CrossRef]  

3. K. F. Chen, P. Santhanam, and S. H. Fan, “Near-field enhanced negative luminescent refrigeration,” Phys. Rev. Appl. 6(2), 024014 (2016). [CrossRef]  

4. Y. De Wilde, F. Formanek, R. Carminati, B. Gralak, P.-A. Lemoine, K. Joulain, J.-P. Mulet, Y. Chen, and J.-J. Greffet, “Thermal radiation scanning tunnelling microscopy,” Nature 444(7120), 740–743 (2006). [CrossRef]  

5. J. Chen, B. X. Wang, and C. Y. Zhao, “Scattering-type multi-probe scanning thermal microscope based on near-field thermal radiation,” Int. J. Heat Mass Transfer 181(12), 121869 (2021). [CrossRef]  

6. T. Inoue, T. Suzuki, K. Ikeda, and S. Noda, “Near-field thermophotovotaic devices with surrounding non-contact reflectors for efficient photon recycling,” Opt. Express 29(7), 11133–11143 (2021). [CrossRef]  

7. A. Narayanaswamy and G. Chen, “Surface modes for near field thermophotovoltaics,” Appl. Phys. Lett. 82(20), 3544–3546 (2003). [CrossRef]  

8. O. Ilic, M. Jablan, J. D. Joannopoulos, I. Celanovic, and M. Soljacic, “Overcoming the black body limit in plasmonic and graphene near-field thermophotovoltaic systems,” Opt. Express 20(S3), A366–A384 (2012). [CrossRef]  

9. M. J. He, H. Qi, Y. T. Ren, Y. J. Zhao, Y. Zhang, J. D. Shen, and M. Antezza, “Radiative thermal switch driven by anisotropic black phosphorus plasmons,” Opt. Express 28(18), 26922–26934 (2020). [CrossRef]  

10. V. B. Svetovoy, P. J. van Zwol, and J. Chevrier, “Plasmon enhanced near-field radiative heat transfer for graphene covered dielectrics,” Phys. Rev. B 85(15), 155418 (2012). [CrossRef]  

11. M. Lim, S. S. Lee, and B. J. Lee, “Near-field thermal radiation between graphene-covered doped silicon plates,” Opt. Express 21(19), 22173–22185 (2013). [CrossRef]  

12. X. L. Liu, L. Wang, and Z. M. Zhang, “Near-field thermal radiation: recent progress and outlook,” Nanoscale Microscale Thermophys. Eng. 19(2), 98–126 (2015). [CrossRef]  

13. D. Polder and M. van Hove, “Theory of radiative heat transfer between closely spaced bodies,” Phys. Rev. B 4(10), 3303–3314 (1971). [CrossRef]  

14. M. Francoeur, M. P. Menguc, and R. Vaillon, “Spectral tuning of near-field radiative heat flux between two thin silicon carbide films,” J. Phys. D: Appl. Phys. 43(7), 075501 (2010). [CrossRef]  

15. S. A. Biehs, M. Tschikin, R. Messina, and P. Ben-Abdallah, “Super-Planckian near-field thermal emission with phonon-polaritonic hyperbolic metamaterials,” Appl. Phys. Lett. 102(13), 131106 (2013). [CrossRef]  

16. A. Nemilentsau and S. V. Rotkin, “Optimization of nanotube thermal interconnects for near-field radiative heat transport,” Appl. Phys. Lett. 101(6), 063115 (2012). [CrossRef]  

17. J. P. Mulet, K. Joulain, R. Carminati, and J. J. Greffet, “Enhanced radiative heat transfer at nanometric distances,” Microscale Thermophys. Eng. 6(3), 209–222 (2002). [CrossRef]  

18. E. Tervo, Z. M. Zhang, and B. Cola, “Collective near-field thermal emission from polaritonic nanoparticle arrays,” Phys. Rev. Materials 1(1), 015201 (2017). [CrossRef]  

19. Y. Guo, C. L. Cortes, S. Molesky, and Z. Jacob, “Broadband super-Planckian thermal emission from hyperbolic metamaterials,” Appl. Phys. Lett. 101(13), 131106 (2012). [CrossRef]  

20. X. L. Liu and Y. M. Xuan, “Super-Planckian thermal radiation enabled by hyperbolic surface phonon polaritons,” Sci. China Technol. Sci. 59(11), 1680–1686 (2016). [CrossRef]  

21. J. Yang, W. Du, Y. S. Su, Y. Fu, S. X. Gong, S. L. He, and Y. G. Ma, “Observing of the super-Planckian near-field thermal radiation between graphene sheets,” Nat. Commun. 9(1), 4033 (2018). [CrossRef]  

22. J. D. Shen, S. Guo, X. L. Liu, B. A. Liu, W. T. Wu, and H. He, “Super-Planckian thermal radiation enabled by coupled quasi-elliptic 2D black phosphorus plasmons,” Appl. Therm. Eng. 144, 403–410 (2018). [CrossRef]  

23. Y. Yang, P. Sabbaghi, and L. P. Wang, “Effect of magnetic polaritons in SiC deep gratings on near-field radiative transfer,” Int. J. Heat Mass Transfer 108, 851–859 (2017). [CrossRef]  

24. Y. Zhang, H. L. Yi, and H. P. Tan, “Near-Field Radiative Heat Transfer between Black Phosphorus Sheets via Anisotropic Surface Plasmon Polaritons,” ACS Photonics 5(9), 3739–3747 (2018). [CrossRef]  

25. E. Moncada-Villa, V. Fernandez-Hurtado, F. J. Garcia-Vidal, A. Garcia-Martin, and J. C. Cuevas, “Magnetic field control of near-field radiative heat transfer and the realization of highly tunable hyperbolic thermal emitters,” Phys. Rev. B 92(12), 125418 (2015). [CrossRef]  

26. J. L. Song, Q. Cheng, L. Lu, B. W. Li, K. Zhou, B. Zhang, Z. X. Luo, and X. P. Zhou, “Magnetically Tunable Near-Field Radiative Heat Transfer in Hyperbolic Metamaterials,” Phys. Rev. Applied 13(2), 024054 (2020). [CrossRef]  

27. P. J. van Zwol, K. Joulain, P. Ben-Abdallah, and J. Chevrier, “Phonon polaritons enhance near-field thermal transfer across the phase transition of VO2,” Phys. Rev. B 84(16), 161413 (2011). [CrossRef]  

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

29. A. Wang, Z. Zheng, and Y. Xuan, “Near-field radiative thermal control with graphene covered on different materials,” J. Quant. Spectrosc. Radiat. Transfer 180, 117–125 (2016). [CrossRef]  

30. O. Ilic, M. Jablan, J. D. Joannopoulos, I. Celanovic, H. Buljan, and M. Soljacic, “Near-field thermal radiation transfer controlled by plasmons in graphene,” Phys. Rev. B 85(15), 155422 (2012). [CrossRef]  

31. R. Messina, M. Tschikin, S.-A. Biehs, and P. Ben-Abdallah, “Fluctuation-electrodynamic theory and dynamics of heat transfer in systems of multiple dipoles,” Phys. Rev. B 88(10), 104307 (2013). [CrossRef]  

32. W. B. Zhang, B. X. Wang, and C. Y. Zhao, “Active control and enhancement of near-field heat transfer between dissimilar materials by strong coupling effects,” Int. J. Heat Mass Transfer 188(12), 122588 (2022). [CrossRef]  

33. J. Y. Chang, Y. Yang, and L. P. Wang, “Enhanced energy transfer by near-field coupling of a nanostructured metamaterial with a graphene-covered plate,” J. Quant. Spectrosc. Radiat. Transfer 184, 58–67 (2016). [CrossRef]  

34. M. J. He, H. Qi, Y. T. Ren, Y. J. Zhao, and M. Antezza, “Magnetoplasmonic manipulation of nanoscale thermal radiation using twisted graphene gratings,” Int. J. Heat Mass Transfer 150(11), 119305 (2020). [CrossRef]  

35. Z. H. Zheng, X. L. Liu, A. Wang, and Y. Xuan, “Graphene-assisted near-field radiative thermal rectifier based on phase transition of vanadium dioxide (VO2),” Int. J. Heat Mass Transfer 109, 63–72 (2017). [CrossRef]  

36. P. Ben-Abdallah, S. A. Biehs, and K. Joulain, “Many-body radiative heat transfer theory,” Phys. Rev. Lett. 107(11), 114301 (2011). [CrossRef]  

37. O. R. Choubdar and M. Nikbakht, “Radiative heat transfer between nanoparticles: Shape dependence and three-body effect,” J. Appl. Phys. 120(14), 144303 (2016). [CrossRef]  

38. Z. H. Zheng and Y. M. Xuan, “Enhancement or suppression of the near-field radiative heat transfer between two materials,” Nanoscale Microscale Thermophys. Eng. 15(4), 237–251 (2011). [CrossRef]  

39. A. Ott, R. Messina, P. Ben-Abdallah, and S.-A. Biehs, “Radiative thermal diode driven by nonreciprocal surface waves,” Appl. Phys. Lett. 114(16), 163105 (2019). [CrossRef]  

40. D. Y. Xu, A. Bilal, J. M. Zhao, L. H. Liu, and Z. M. Zhang, “Near-field radiative heat transfer between rough surfaces modeled using effective media with gradient distribution of dielectric function,” Int. J. Heat Mass Tran 142(11), 118432 (2019). [CrossRef]  

41. J. Chen, B. X. Wang, and C. Y. Zhao, “Near-field radiative heat transport between nanoparticles inside a cavity configuration,” Int. J. Heat Mass Tran 196(12), 123213 (2022). [CrossRef]  

42. J. Dong, J. M. Zhao, and L. H. Liu, “Near-field radiative heat transfer between clusters of dielectric nanoparticles,” J. Quant. Spectrosc. Radiat. Transfer 197, 114–122 (2017). [CrossRef]  

43. I. Latella, A. Perez-Madrid, J. M. Rubi, S. A. Biehs, and P. Ben-Abdallah, “Heat Engine Driven by Photon Tunneling in Many-Body Systems,” Phys. Rev. Appl. 4(1), 011001 (2015). [CrossRef]  

44. J. L. Son, Q. Cheng, Z. X. Luo, X. P. Zhou, and Z. M. Zhang, “Modulation and splitting of three-body radiative heat flux via graphene/SiC core-shell nanoparticles,” Int. J. Heat Mass Tran 140, 80–87 (2019). [CrossRef]  

45. Y. Zhang, C. H. Wang, H. L. Yi, and H. P. Tan, “Multiple surface plasmon polaritons mediated near-field radiative heat transfer between graphene/vacuum multilayers,” J. Quant. Spectrosc. Radiat. Transfer 221, 138–146 (2018). [CrossRef]  

46. M. J. He, H. Qi, Y. F. Wang, Y. T. Ren, W. H. Cai, and L. M. Ruan, “Near-field radiative heat transfer in multilayered graphene system considering equilibrium temperature distribution,” Opt. Express 27(16), A953–A966 (2019). [CrossRef]  

47. P. Ben-Abdallah and S. A. Biehs, “Near-Field Thermal Transistor,” Phys. Rev. Lett. 112(4), 044301 (2014). [CrossRef]  

48. Y. X. Li, Y. D. Dang, S. Zhang, X. R. Li, Y. Jin, P. Ben-Abdallah, J. B. Xu, and Y. G. Ma, “Radiative Thermal Transistor,” Phys. Rev. Appl. 20(2), 024061 (2023). [CrossRef]  

49. I. Latella, O. Marconot, J. Sylvestre, L. G. Frechette, and P. Ben-Abdallah, “Dynamical Response of a Radiative Thermal Transistor Based on Suspended Insulator-Metal-Transition Membranes,” Phys. Rev. Appl. 11(2), 024004 (2019). [CrossRef]  

50. K. Ito, K. Nishikawa, and H. Iizuka, “Multilevel radiative thermal memory realized by the hysteretic metal-insulator transition of vanadium dioxide,” Appl. Phys. Lett. 108(5), 053507 (2016). [CrossRef]  

51. P. Ben-Abdallah and S. A. Biehs, “Towards Boolean operations with thermal photons,” Phys. Rev. B 94(24), 241401 (2016). [CrossRef]  

52. M. J. He, H. Qi, Y. Li, Y. T. Ren, W. H. Cai, and L. M. Ruan, “Graphene-mediated near field thermostat based on three-body photon tunneling,” Int. J. Heat Mass Tran 137, 12–19 (2019). [CrossRef]  

53. V. Berry, “Impermeability of graphene and its applications,” Carbon 62, 1–10 (2013). [CrossRef]  

54. Y. F. Wu, L. Zhang, C. Z. Li, Z. S. Zhang, S. Liu, Z. M. Liao, and D. Yu, “Dirac semimetal heterostructures: 3D Cd3As2 on 2D graphene,” Adv. Mater. 30(34), 1707547 (2018). [CrossRef]  

55. M. Uchida, Y. Nakazawa, S. Nishihaya, K. Akiba, M. Kriener, Y. Kozuka, A. Miyake, Y. Taguchi, M. Tokunaga, N. Nagaosa, Y. Tokura, and M. Kawasaki, “Quantum Hall states observed in thin films of Dirac semimetal Cd3As2,” Nat. Commun. 8(1), 2274 (2017). [CrossRef]  

56. B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein, R. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A. McGuire, D. H. Cobden, W. Yao, D. Xiao, P. Jarillo-Herrero, and X. D. Xu, “Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit,” Nature 546(7657), 270–273 (2017). [CrossRef]  

57. L. H. Zeng, D. Wu, S. H. Lin, C. Xie, H. Y. Yuan, W. Lu, S. P. Lau, Y. Chai, L. B. Luo, Z. J. Li, and Y. H. Tsang, “Controlled synthesis of 2D palladium diselenide for sensitive photodetector applications,” Adv. Funct. Mater. 29(1), 1806878 (2019). [CrossRef]  

58. J. Guo, X. G. Zhao, N. K. Sun, X. F. Xiao, W. Liu, and Z. D. Zhang, “Tunable quantum Shubnikov-de Hass oscillations in antiferromagnetic topological semimetal Mn-doped Cd3As2,” J. Mater. Sci. Technol. 76, 247–253 (2021). [CrossRef]  

59. K. Z. Shi, Z. Y. Chen, X. N. Xu, J. Evans, and S. L. He, “Optimized colossal near-field thermal radiation enabled by manipulating coupled plasmon polariton geometry,” Adv. Mater. 33(52), 2106097 (2021). [CrossRef]  

60. K. Z. Shi, Z. Y. Chen, Y. X. Xing, J. X. Yang, X. A. Xu, J. S. Evans, and S. L. He, “Near-field radiative heat transfer modulation with an ultrahigh dynamic range through mode mismatching,” Nano Lett. 22(19), 7753–7760 (2022). [CrossRef]  

61. O. V. Kotov and Y. E. Lozovik, “Dielectric response and novel electromagnetic modes in three-dimensional Dirac semimetal films,” Phys. Rev. B 93(23), 235417 (2016). [CrossRef]  

62. B. Hu, M. Huang, P. Li, and J. Yang, “Dynamically dual-tunable dual-band to four-band metamaterial absorbers based on bulk Dirac semimetal and vanadium dioxide,” J. Opt. Soc. Amer. A 39(3), 383–391 (2022). [CrossRef]  

63. G. D. Xu, J. Sun, H. M. Mao, Z. L. Cao, and X. Y. Ma, “Modulating Near-Field Radiative Heat Transfer through Thin Dirac Semimetal Films,” Nanoscale and Microscale Thermophys. Eng. 25(2), 101–115 (2021). [CrossRef]  

64. H. Y. Meng, X. J. Shang, X. X. Xue, K. Z. Tang, S. X. Xia, X. Zhai, Z. R. Liu, J. H. Chen, H. J. Li, and L. L. Wang, “Bidirectional and dynamically tunable THz absorber with Dirac semimetal,” Opt. Express 27(21), 31062–31074 (2019). [CrossRef]  

65. A. S. Barker, H. W. Verleur, and H. J. Guggenheim, “Infrared optical properties of vanadium dioxide above and below transition temperature,” Phys. Rev. Lett. 17(26), 1286–1289 (1966). [CrossRef]  

66. E. D. Palik, Handbook of optical constants of solids, handbook of optical constants of solids (Academic, 1985).

67. S. A. Biehs, E. Rousseau, and J. J. Greffet, “Mesoscopic description of radiative heat transfer at the nanoscale,” Phys. Rev. Lett. 105(23), 234301 (2010). [CrossRef]  

68. S. A. Biehs, “Thermal heat radiation, near-field energy density and near-field radiative heat transfer of coated materials,” Eur. Phys. J. B 58(4), 423–431 (2007). [CrossRef]  

69. I. Latella, P. Ben-Abdallah, S. A. Biehs, M. Antezza, and R. Messina, “Radiative heat transfer and nonequilibrium Casimir-Lifshitz force in many-body systems with planar geometry,” Phys. Rev. B 95(20), 205404 (2017). [CrossRef]  

70. I. Latella, S. A. Biehs, R. Messina, A. W. Rodriguez, and P. Ben-Abdallah, “Ballistic near-field heat transport in dense many-body systems,” Phys. Rev. B 97(3), 035423 (2018). [CrossRef]  

71. R. Messina, M. Antezza, and P. Ben-Abdallah, “Three-body amplification of photon heat tunneling,” Phys. Rev. Lett. 109(24), 244302 (2012). [CrossRef]  

72. R. Messina and M. Antezza, “Three-body radiative heat transfer and Casimir-Lifshitz force out of thermal equilibrium for arbitrary bodies,” Phys. Rev. A 89(5), 052104 (2014). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

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

Alert me when this article is cited.


Figures (15)

Fig. 1.
Fig. 1. Heat transfer model of two-body structure thermal rectifier. Media 1 to 4 represent VO2 plate, Dirac semimetal film, vacuum gap and SiO2 plate, respectively. The temperature of the VO2 and SiO2 plate is set as 300 K and 353 K, and the scenario is defined as forward heat transfer. The heat transfer quantity of NFTR is expressed as ΦF. The temperature of VO2 and SiO2 plates is set as 300 K and 353 K, and the scenario is defined as forward heat transfer. The heat transfer quantity of NFTR is expressed as ΦR.
Fig. 2.
Fig. 2. Heat transfer model of three-body thermostat. Media 1-3 represent the SiO2 plate on the left (400 K), VO2 plate, and SiO2 plate on the right (300 K), and the two Dirac semimetal films are represented by 4 and 5. The heat flux released by the high temperature plate 1 is represented as Φ1, and the heat flux absorbed by low temperature plate 3 is represented as Φ3.
Fig. 3.
Fig. 3. The relative dielectric function of Dirac semimetal (a) real part; (b) imaginary part
Fig. 4.
Fig. 4. The relative dielectric function of VO2 (a) real part; (b) imaginary part
Fig. 5.
Fig. 5. The relative dielectric function of SiO2 (a) real part; (b) imaginary part
Fig. 6.
Fig. 6. Energy transmission coefficients of forward and reverse heat transfer of two-body structure under different parameter combinations. (a) d = 10 nm, EF =0.15 eV, t = 5 nm; (b) d = 20 nm, EF =0.15 eV, t = 5 nm; (c) d = 10 nm, EF =0.18 eV, t = 5 nm; (d) d = 10 nm, EF =0.15 eV, t = 10 nm. In Fig, the left side represents the energy transmission coefficient of forward heat transfer, and the right side represents the energy transmission coefficient of reverse heat transfer.
Fig. 7.
Fig. 7. Variation trends of the spectral radiative heat flux with different conditions. The black solid line and black dashed line represent the forward heat transfer scenario of the film-covered structure (d = 50 nm, EF = 0.25 eV, and t = 4 nm) and non-covered structure, respectively, while the red solid line and red dashed line illustrates the reverse heat transfer scenario of the film-covered structure and non-covered structure, respectively.
Fig. 8.
Fig. 8. (a) Contour nephogram of the global rectification factor of DSTR as a function of film thickness and Fermi level. The black contour line represents the global rectification factor of non-coverd structure; (b) Contour nephogram of the global rectification factor of DSTR as a function of vacuum gap distance and Fermi level.
Fig. 9.
Fig. 9. The variation of global rectification factor with temperature difference under specific combination of parameters. The black solid line represents the global rectification factor of film-covered structure, while the red line represents the global rectification factor of non-covered structure.
Fig. 10.
Fig. 10. Energy transmission coefficient of three-body structure when EF1 = EF2 = 0.15 eV and t1 = t2 = 5 nm. (a) VO2 in the insulating state; (b) VO2 in the metallic state. In each figure, the left side represents the energy transmission coefficient between plate 1 and plate 2, and the right side represents the energy transmission coefficient between plate 2 and plate 3.
Fig. 11.
Fig. 11. Contour nephograms of equilibrium temperature T2,eq and equilibrium heat flux Φeq varying with Fermi level EF and film thickness t at intermediate plate initial temperature T2 = 400 K. (a) equilibrium temperature T2,eq; (b) equilibrium heat flux Φeq. The black solid line in (b) represents the equilibrium heat flux of non-film covered structure.
Fig. 12.
Fig. 12. Contour nephograms of equilibrium temperature T2,eq and equilibrium heat flux Φeq varying with Fermi level EF at intermediate plate initial temperature T2 = 300 K. (a) equilibrium temperature T2,eq; (b) equilibrium heat flux Φeq. In Fig. 10, black solid line in (a) represents the phase transition temperature contour of VO2, which corresponds to the black solid line in (b).
Fig. 13.
Fig. 13. Contour nephograms of equilibrium temperature T2,eq and equilibrium heat flux Φeq varying with Fermi level EF at intermediate plate initial temperature T2 = 400 K. (a) equilibrium temperature T2,eq; (b) equilibrium heat flux Φeq. Auxiliary lines are the same as those in Fig. 12.
Fig. 14.
Fig. 14. Heat flux released by plate 1 and absorbed by plate 3 under different conditions. The solid line represents the heat flux released by plate 1, and the dotted line represents the heat flux absorbed by plate 3. The black line combination corresponds to case 1, the blue line combination corresponds to case 2, and the red line combination corresponds to case 3.
Fig. 15.
Fig. 15. The p-polarized energy transmission coefficients (${\xi_{2p}^{1-}}+{\xi_{3p}^{2-}}/2$ of the 5 nm film-covered three-body structure under different Fermi levels of Dirac semimetal. (a) intermediate plate VO2 is in insulating state; (b) intermediate plate VO2 is in metallic state. The four pictures from left to right show the situation without Dirac semimetal film coverage and when the Fermi levels of Dirac semimetal film are 0.05 eV, 0.10 eV and 0.15 eV respectively.

Equations (26)

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

ε D = ε b g α c 6 π v F [ ( 2 E F Ω ) 2 + 1 3 ( 2 π k B T Ω ) 2 + i π G ( Ω 2 ) 8 0 E c G ( ξ ) G ( Ω 2 ) ( Ω ) 2 4 ξ 2 ξ d ξ ]
ε ¯ ¯ = [ ε 0 0 0 ε 0 0 0 ε ]
ε ( ω = ε + j = 1 N [ ( S j ω j 2 ) ( ω j 2 i γ j ω ω 2 ) ]
ε ( ω = ε ω p 2 ω 2 + i ω c ω
Φ = 0 d ω 2 π φ ( ω )
φ ( ω ) = ω n 12 4 π 2 0 ξ ( ω , κ ) κ d κ
ξ prop ( ω , κ ) = ( 1 | R 321 s | 2 ) ( 1 | r 34 s | 2 ) | 1 R 321 s r 34 s e 2 i k z , 3 s d | 2 + ( 1 | R 321 p | 2 ) ( 1 | r 34 p | 2 ) | 1 R 321 p r 34 p e 2 i k z , 3 p d | 2
ξ evan ( ω , κ ) = 4 Im ( R 321 s ) Im ( r 34 s ) e 2 Im ( k z , 3 s ) d | 1 R 321 s r 34 s e 2 i k z , 3 s d | 2 + 4 Im ( R 321 p ) Im ( r 34 p ) e 2 Im ( k z , 3 p ) d | 1 R 321 p r 34 p e 2 i k z , 3 p d | 2
R 321 s = ( 1 k z , 1 s k z , 3 s ) cos k z , 2 s t + i ( k z , 2 s k z , 3 s k z , 1 s k z , 2 s ) sin k z , 2 s t ( 1 + k z , 1 s k z , 3 s ) cos k z , 2 s t i ( k z , 2 s k z , 3 s + k z , 1 s k z , 2 s ) sin k z , 2 s t
R 321 p = ( 1 k z , 1 p ε 1 k z , 3 p ) cos k z , 2 p t + i ( k z , 2 p ε 2 k z , 3 p ε 2 k z , 1 p ε 1 k z , 2 p ) sin k z , 2 p t ( 1 + k z , 1 p ε 1 k z , 3 p ) cos k z , 2 p t i ( k z , 2 p ε 2 k z , 3 p + ε 2 k z , 1 p ε 1 k z , 2 p ) sin k z , 2 p t
r 34 s = k z , 3 s k z , 4 s k z , 3 s + k z , 4 s
r 34 p = ε 4 k z , 3 p ε 3 k z , 4 p ε 4 k z , 3 p + ε 3 k z , 4 p
k z , i s = ε , i ( ω / c ) 2 κ 2
k z , i p = ε , i ( ω / c ) 2 ( ε , i κ 2 / ε , i )
Φ 3 = 0 d ω 2 π φ 3 ( ω )
φ 3 ( ω )  =  ω j = s , p d 2 κ ( 2 π ) 2 [ n 12 ( ω ) ξ j 1 2 ( ω , κ ) + n 23 ( ω ) ξ j 2 3 ( ω , κ ) ]
ξ j 1 2 ( ω , κ )  =  { | τ 2 j | 2 ( 1 | ρ 1 j | 2 ) ( 1 | ρ 3 j | 2 ) | 1 ρ 12 j ρ 3 j e 2 i k z , 0 d | 2 | 1 ρ 1 j ρ 2 j e 2 i k z , 0 d d | 2 ,   κ < ω c 4 | τ 2 j | 2 Im ( ρ 1 j ) Im ( ρ 3 j ) e 4 Im ( k z , 0 ) d | 1 ρ 12 j ρ 3 j e 2 i k z , 0 d | 2 | 1 ρ 1 j ρ 2 j e 2 i k z , 0 d d | 2 ,   κ > ω c
ξ j 2 3 ( ω , κ )  =  { ( 1 | ρ 12 j | 2 ) ( 1 | ρ 3 j | 2 ) | 1 ρ 12 j ρ 3 j e 2 i k z , 0 d | 2 ,   κ < ω c 4 Im ( ρ 12 j ) Im ( ρ 3 j ) e 2 Im ( k z , 0 ) d | 1 ρ 12 j ρ 3 j e 2 i k z , 0 d | 2 ,   κ > ω c
τ 2 j = t 2 j t 2 j ¯ e i k z , 2 j δ 2 1 ( r 2 j ) 2 e i k z , 2 j δ 2 ρ i j = r i j 1 e i k z , i j δ 2 1 ( r i j ) 2 e i k z , i j δ 2 ρ 12 j = ρ 2 j + ρ 1 j ( τ 2 j ) 2 e 2 i k z , 0 d 1 ρ 1 j ρ 2 j e 2 i k z , 0 d
r i s = ( 1 k z , i k z , 0 ) cos k z , D t + i ( k z , D k z , 0 k z , i k z , D ) sin k z , D t ( 1 + k z , i k z , 0 ) cos k z , D t i ( k z , D k z , 0 + k z , i k z , D ) sin k z , D t
r i p = ( 1 k z , i ε i k z , 0 ) cos k z , D t + i ( k z , D ε D k z , 0 ε D k z , i ε i k z , D ) sin k z , D t ( 1 + k z , i ε i k z , 0 ) cos k z , D t i ( k z , D ε D k z , 0 + ε D k z , i ε i k z , D ) sin k z , D t
r 2 s = k z , 0 k z , 2 s k z , 0 + k z , 2 s , r 2 p = ε , 2 k z , 0 k z , 2 p ε , 2 k z , 0 + k z , 2 p
t 2 s = 2 k z , 0 k z , 0 + k z , 2 s ,     t 2 p = 2 ε , 2 k z , 0 ε , 2 k z , 0 + k z , 2 p
t 2 s ¯ = 2 k z , 2 s k z , 0 + k z , 2 s ,     t 2 p ¯ = 2 ε , 2 k z , 2 p ε , 2 k z , 0 + k z , 2 p
φ 1 ( ω ) = ω j = s , p d 2 κ ( 2 π ) 2 [ n 23 ( ω ) ξ j 2 3 , ( ω , κ ) + n 12 ( ω ) ξ j 1 2 , ( ω , κ ) ]
M = φ F φ R
Select as filters


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