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

Tunable bifunctional terahertz metamaterial device based on Dirac semimetals and vanadium dioxide

Open Access Open Access

Abstract

A tunable bifunctional terahertz (THz) metamaterial device based on Dirac semimetal films (DSFs) and VO2 is presented. The insulator-to-metal phase transition of VO2 enables bifunctional asymmetric transmission and dual-directional absorption to be switched in the THz range. When VO2 serves as a dielectric, tunable broadband asymmetric transmission of linearly polarized THz waves can be achieved. When VO2 is in a metallic state, the proposed device acts as a tunable dual-directional absorber with perfect absorption in both illumination directions. In each case, the response can be tuned by varying the Fermi energy of the DSFs. This offers a new pathway for the development of tunable multifunctional THz metamaterial devices.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Terahertz (THz) radiation occurs between microwave and infrared frequencies [1]. It shows great promise and is likely to become the working frequency of future wireless networks, such as sixth-generation (6G) systems [2]. Few natural materials interact with THz radiation, so there is considerable interest in functional devices that operate in the so-called ‘THz gap’ [3]. Metamaterials are artificially engineered sub-wavelength electromagnetic materials with extraordinary physical properties [4], which makes them a desirable solution to this problem [5]. The first THz metamaterial absorber was proposed and fabricated by Tao et al. in 2008 [6]. Since then, there has been extensive research into different THz metamaterial absorbers including single-, dual-, multi-, and broadband [710]. Asymmetric transmission of linearly polarized waves, which can be regarded as linear conversion dichroism, has been largely realized in THz metamaterials [1113] following their proposal and experimental demonstration by Menzel et al. [14]. However, the practical applications of these devices are limited by the fact that the operating frequency and working intensity cannot be adjusted after fabrication.

Recently, tunable metamaterial functional devices that use two-dimensional materials (such as graphene, black phosphorus, and vanadium dioxide (VO2), and so on) have attracted much interest [1524]. Among them, VO2 is a phase-change material that undergoes a reversible insulator-to-metal transition in response to electrical, thermal, and optical stimuli [2527]. Its conductivity can change by four or five orders of magnitude during this transition [28], which makes it a promising candidate for tuning THz metamaterial functional devices. With the aim of producing tunable THz metamaterial functional devices, Lv et al. [29] produced a strong asymmetric transmission in a THz metamaterial that consisted of two 90° twisted E-shaped resonators embedded with VO2. M. Liu et al. [30] demonstrated a temperature-controlled asymmetric transition with linearly polarized THz waves by exploiting the VO2 phase transition. Song et al. [31] investigated a broadband polarization-independent VO2-based THz metamaterial absorber that had dynamically tunable absorption. H. Liu et al. [32] proposed a broadband tunable THz metamaterial absorber that included VO2 and could achieve between 5% and 100% absorption.

Dirac semimetal films (DSFs) are a three-dimensional analog of graphene and they are of significant interest to researchers due to their ultrahigh carrier mobility [33], robustness against environmental defects or excess conductive bulk states [34], and dynamically adjustable permittivity function (The in situ electron doping method can be used to adjust the Fermi energy.) [3536]. Dai et al. [37] achieved broadband tunable asymmetric transmission for linearly polarized THz waves in a Dirac semimetal-based metamaterial consisting of two double-T resonators. Meng et al. [38] reported a bidirectional and tunable THz absorber based on bulk Dirac semimetals. However, these THz metamaterial functional devices have a single function which is fixed once the device has been fabricated. Therefore, it is desirable to integrate two or more diversified functionalities into a single device. Some studies have reported THz metamaterial devices capable of multifunctional performance [3943]; however, there are no devices that can achieve both functions with dual-directional absorption and asymmetric transmission.

This study proposes a theoretical switchable bifunctional metamaterial device based on VO2 and DSFs in the THz region. The proposed device can utilize the insulator-to-metal phase transition of VO2 to switch from a dual-directional absorber, which obtains perfect absorption regardless of whether incident light propagates in a backward or forward direction, to another function with broadband asymmetric transmission of linearly polarized waves. In both cases, the response of the proposed device shows blue-shift by increasing the Fermi energy of the DSFs. Furthermore, when the device acts as absorber, the state can switch between a reflector (with reflectance of 55%) and an absorber (with absorptance of 96.2%) by dynamically varying the conductivity of the VO2. This research can be used to inspire the design of tunable multifunctional devices in the THz range and also can be utilized in other frequency bands by scaling down the structure.

2. Design and theoretical analysis

Figure 1 illustrates a unit cell belonging to the proposed bifunctional THz metamaterial device, which consists of five layers. The second and third dielectric layers consist of a tp = 10 µm-thick polyimide spacer separated by a VO2 layer. Two DSFs-based square split-ring resonators are patterned on opposite sides of two dielectric layers. The DSFs and VO2 layers are td = 0.7 and tv = 0.5 µm thick, respectively. Polyimide can be considered as a lossy dielectric with a relative permittivity of 3.5 + 0.00945i. The DSFs layers contain two simple structures in the form of a square split ring resonator. To introduce symmetry broken along the propagation direction, the bottom DSFs layer structure is the mirror image in the x direction after the top DSFs layer structure is rotated 90°. These square split rings have side L = 40 µm, gap g = 6 µm, and width w = 6 µm. In this work, all the calculations were performed using the finite difference frequency domain method in CST Microwave Studio. In simulations, the unit cell boundary conditions were employed along the x- and y-directions with periodicity P = 55 µm and the open boundary condition was along the z direction.

 figure: Fig. 1.

Fig. 1. Schematic of the DSFs- and VO2-based switchable bifunctional THz metamaterial device: (a) top, (b) bottom, (c) side, and (d) perspective views of the structure with P = 55 µm, L = 40 µm, g = 6 µm, w = 6 µm, td = 0.7 µm, tp=10 µm, and tv = 0.5 µm..

Download Full Size | PDF

At the long-wavelength limit, the complex surface conductivity of the DSFs can be calculated using the Kubo formalism from random-phase approximation theory [44]. The interband and intraband conductivity of the DSFs can be defined by [45]

$${\sigma _{{\mathop{\rm {int}}} \textrm{er}}} = \frac{{i{e^2}g\omega }}{{3{\pi ^2}\hbar {v_F}}}\left[ { - \frac{{\pi i}}{2}\frac{{G({\hbar \omega /2} )}}{4} + \int_0^\infty {\left( {\frac{{G(E )- G({\hbar \omega /2} )}}{{{\hbar^2}{\omega^2} - 4{E^2}}}} \right)} EdE} \right]$$
$${\sigma _{{\mathop{\rm {int}}} ra}} = \frac{{i{e^2}}}{\hbar }\frac{{g{k_F}}}{{6{\pi ^2}\Omega }}\left( {1 + \frac{{{\pi^2}}}{3}{{\left( {\frac{T}{{{E_F}}}} \right)}^2}} \right)$$
here, $G(E )\textrm{ = }n({ - E} )- n(E )= \frac{{\sinh ({{E \mathord{\left/ {\vphantom {E T}} \right.} T}} )}}{{\cosh ({{{{E_F}} \mathord{\left/ {\vphantom {{{E_F}} T}} \right.} T}} )+ \cosh ({{E \mathord{\left/ {\vphantom {E T}} \right.} T}} )}}$ where n(E) is the Fermi distribution function, $E\textrm{ = }\hbar \omega /2$. In addition, $\omega$ is the angular frequency, e is the charge of an electron, $\hbar$ is the reduced Planck constant, and ${E_F}$ is the Fermi energy. Furthermore, $\Omega = \hbar \omega /{E_F}\textrm{ + }i\hbar {\tau ^{ - 1}}/{E_F}$, ${v_F} \approx {10^6}m/s$ is the Fermi velocity, ${k_F} = {E_F}/({\hbar {v_F}} )$ is the Fermi momentum, and the degeneracy factor is g = 40. Thus, the dynamic conductivity of the DSFs is $\sigma \textrm{ = }{\sigma _{{\mathop{\rm {int}}} \textrm{er}}}\textrm{ + }{\sigma _{{\mathop{\rm {int}}} \textrm{ra}}}$. Considering the interband electronic transitions and using the two-band model, the permittivity of the DSFs can be expressed as [46]
$$\varepsilon \textrm{ = }{\varepsilon _\textrm{b}} + i\frac{\sigma }{{\omega {\varepsilon _0}}}$$
where ɛ0 is the permittivity of a vacuum and the effective background dielectric is ${\varepsilon _\textrm{b}}$ = 1 for g = 40 (AlCuFe quasi-crystals [47]).

In the THz range, the relative permittivity of VO2 is described by the Drude model ${\varepsilon _{\textrm{v}{\textrm{o}_2}}}(\omega )\textrm{ = }{\varepsilon _\infty } - \frac{{{{({{\omega_\textrm{p}}({{\sigma_{\textrm{v}{\textrm{o}_2}}}} )} )}^2}}}{{{\omega ^2}\textrm{ + }i\gamma \omega }}$ where the permittivity at an infinite frequency is ${\varepsilon _\infty }\textrm{ = }12$ and the damping frequency is $\gamma \textrm{ = }5.75 \times {10^{13}}{\textrm{s}^{ - 1}}$[48]. Moreover, ${\omega _\textrm{p}}({{\sigma_{\textrm{v}{\textrm{o}_2}}}} )$ is the conductivity-dependent plasma frequency at ${\sigma _{\textrm{v}{\textrm{o}_2}}}$ which can be described as ${\omega _\textrm{p}}^2({{\sigma_{\textrm{v}{\textrm{o}_2}}}} )\textrm{ = }\frac{{{\sigma _{\textrm{v}{\textrm{o}_2}}}}}{{{\sigma _0}}}{\omega _\textrm{p}}^2({{\sigma_0}} )$ (${\sigma _0}\textrm{ = }3 \times {10^5}\textrm{S} \cdot {\textrm{m}^{ - 1}}$ and ${\omega _\textrm{p}}({{\sigma_0}} )\textrm{ = }1.4 \times {10^{15}}\textrm{rad} \cdot {\textrm{s}^{ - 1}}$). It was assumed that VO2 transitioned from an insulator to a full metal state when the conductivity increased from 10 to 100 000 S/m. The VO2 phase-transition process can be realized using thermal, electrical, or optical stimuli [27].

VO2 transitions from an insulating dielectric state to a metallic state. The proposed bifunctional THz metamaterial device uses this behavior to switch from a functional device with tunable dual-directional absorption to tunable broadband asymmetric transmission. When VO2 is in the metallic state, the proposed device acts as a tunable dual-directional absorber because of the highly reflective characteristics in the middle of the VO2 layer. However, when VO2 serves as a dielectric, the device achieves broadband asymmetric transmission of linearly polarized waves which can be dynamically tuned by varying the Fermi energy of the DSFs. The caption of Fig. 1 shows the respective values of the parameters used throughout in all the calculations, and the initial Fermi energy was 0.15 eV. The tunable asymmetric transmission and dual-directional absorption were achieved under a fixed VO2 conductivity of 10 S/m and 100000 S/m, respectively. When the proposed device acts as an absorber, the absorptance can be calculated as A=1-|S11|2-|S21|2 where the reflection |S11| and transmission |S21| parameters are obtained from the simulation. For the asymmetric transmission performance, the T matrix, commonly called the Jones matrix, is utilized to better understand the effect of asymmetric transmission for linearly polarized waves [14]. The T matrix links the complex amplitudes of the incident and transmitted fields and can be expressed as

$$\left( {\begin{array}{c} {{T_x}}\\ {{T_y}} \end{array}} \right) = \left( {\begin{array}{cc} {{T_{xx}}}&{{T_{xy}}}\\ {{T_{yx}}}&{{T_{yy}}} \end{array}} \right)\left( {\begin{array}{c} {{I_x}}\\ {{I_y}} \end{array}} \right) = T_{lin}^f\left( {\begin{array}{c} {{I_x}}\\ {{I_y}} \end{array}} \right)$$
Here the superscript f and subscript $lin$ indicate forward propagation and linear polarization; Ix (Iy) and Tx (Ty) represent the complex amplitudes of the incident and transmitted waves; and Txx (Tyy) and Txy (Tyx) represent the co-polarization and cross-polarization transmission coefficients, respectively. According to the reciprocity theorem, the transmission matrix $T_{lin}^b$ for backward propagation can be written as
$$T_{lin}^b = \left( {\begin{array}{cc} {{T_{xx}}}&{ - {T_{yx}}}\\ { - {T_{xy}}}&{{T_{yy}}} \end{array}} \right)$$
where the superscript b indicates propagation in the backward direction. In addition, for circular polarization, the right- and left-hand circularly polarized waves can be denoted by + and −, respectively. The T matrix for circularly polarized wave can be calculated from the transmission coefficient of linearly polarized wave as
$$T_{circ}^f = \left( {\begin{array}{cc} {{T_{ +{+} }}}&{{T_{ +{-} }}}\\ {{T_{ -{+} }}}&{{T_{ -{-} }}} \end{array}} \right) = \frac{1}{2}\left( {\begin{array}{cc} {{T_{xx}} + {T_{yy}} + i({T_{xy}} - {T_{yx}})}&{{T_{xx}} - {T_{yy}} - i({T_{xy}} + {T_{yx}})}\\ {{T_{xx}} - {T_{yy}} + i({T_{xy}} + {T_{yx}})}&{{T_{xx}} + {T_{yy}} - i({T_{xy}} - {T_{yx}})} \end{array}} \right)$$

The degree of the asymmetric transmission effect can be characterized by the asymmetric transmission parameter which is defined as the total difference between the transmission propagating in two opposite directions. Thus, the asymmetric transmission parameter for linearly and circularly polarized waves can be expressed as [14]

$$\Delta _{lin}^{(x)} = {|{{T_{yx}}} |^2} - {|{{T_{xy}}} |^2} ={-} \Delta _{lin}^{(y)}$$
$$\Delta _{circ}^{( + )} = {|{{T_{ -{+} }}} |^2} - {|{{T_{ +{-} }}} |^2} ={-} \Delta _{circ}^{( - )}$$

According to these equations, the asymmetric transmission effect for linearly and circularly polarized can be realized when the off-diagonal components of the T matrix are not identical. Thus, the asymmetric transmission effect for linearly and circularly polarized waves should satisfy the conditions |Tyx|≠|Txy| and |Txx|≠|Tyy|, respectively.

3. Results and discussion

The response of the proposed device was first demonstrated for the case where VO2 was in the insulating dielectric state with conductivity of 10 S/m. Figures 2(a) and (b) present the four transmission matrix elements of a linearly polarized wave propagating in the backward (−z) and forward (+z) directions when the Fermi energy of the DSFs was 0.15 eV. Under each propagation direction, the co-polarized transmission coefficients |Txx| and |Tyy| remained equal across the frequency ranges, while the cross-polarized transmission coefficients |Txy| and |Tyx| were different. Figure 2(a) shows that |Tyx| reached two transmission peaks at 1.5044 and 1.85 THz with maxima of 0.81 and 0.725. In contrast, |Txy| remained below 0.24 in this frequency range. A comparison of Figs. 2(a) and (b) shows that the cross-polarized transmission coefficients were exchanged with each other when the wave propagated forward direction. Thus, these phenomena are consistent with the conditions required for the asymmetric transmission effect which results in asymmetric transmission of linearly polarized waves.

 figure: Fig. 2.

Fig. 2. Transmission coefficients of linearly polarized waves in (a) backward (-z) and (b) forward (+z) directions as well as the (c) PCR and (d) total transmission for x- and y-polarized waves propagating in the backward (−z) direction when the Fermi energy was 0.15 eV.

Download Full Size | PDF

To provide more information about this asymmetric transmission, the polarization conversion ratio (PCR) and total transmissions of x- and y-polarized waves propagating backward direction were introduced, as shown in Figs. 2(c) and (d). The equations PCRx=|Tyx|2/(|Txx|2 +|Tyx|2) and PCRy=|Txy|2/(|Tyy|2 +|Txy|2) were used to illustrate the performance of the x- and y-linear PCRx and PCRy shown in Fig. 2(c). PCRx was greater than 0.75 between 1.44 and 2.3 THz, and it exhibited two peaks of 0.857 and 0.96 at 1.53 and 2.138 THz, respectively. In comparison, PCRy remained lower across this frequency range. The total transmissions of the x- and y-polarized waves were defined as |Tx|2=|Txx|2+|Tyx|2 and |Ty|2=|Tyy|2+|Txy|2 to demonstrate the intensity of the transmitted wave for an incident wave propagating backwards. Figure 2(d) shows that the x-polarized wave achieved a maximum transmission of 0.785 and it maintained a high transmitted intensity across a relatively wide bandwidth. In contrast, the total transmission of the y-polarized wave was small. These results suggest that the x-polarized wave achieved relatively high polarization conversion efficiency and that it transmitted well with backward propagation, which demonstrates the asymmetric transmission effect.

The asymmetric transmission parameters for linearly and circularly polarized waves based on Eqs. (7) and (8) are illustrated in Fig. 3(a). The value of the asymmetric transmission parameter $\Delta _{lin}^x$ reached a maximum of 0.6 and was approximately opposite to $\Delta _{lin}^y$. Therefore, under the functional with asymmetric transmission, the proposed device allowed x- to y-polarization conversions and blocked y- to x-polarization conversions when a linearly polarized wave propagated backwards. In addition, the asymmetric transmission parameter for circularly polarized waves was approximately zero, which confirms that there was no asymmetric transmission effect for circularly polarized waves. Furthermore, Fig. 3(b) shows the four transmission matrix elements for circularly polarized waves propagating backwards. The two elements of the co-polarization transmission coefficients were quite different (|T++|≠|T|), while the cross-polarization transmission coefficients were equal (|T+-|=|T-+|≠0) across the entire frequency range. Therefore, the asymmetric transmission effect for circularly polarized waves cannot be realized in the THz range. Moreover, the optical activity which characterizes the rotation of the polarization plane for a linearly polarized wave propagating through a chiral medium can be defined by the polarization azimuth:$\theta = 0.5 \cdot {\tan^{ - 1}}[{2|{{T_{xx}}} ||{{T_{yx}}} |\cos \varphi /({{|{{T_{xx}}} |}^2} - {{|{{T_{yx}}} |}^2})} ]$, where the phase difference is $\varphi = {\varphi _{xx}} - {\varphi _{yx}} = \arg ({{T_{xx}}} )- \arg ({{T_{yx}}} )$. The polarization state of the transmitted wave resulting from an x-polarized incident wave propagating backwards is described by the polarization ellipse at the low frequency transmission peak for the cross-polarized transmission coefficients |Tyx|, as shown in Fig. 3(c). At 1.5044 THz, the polarization azimuth was rotated by −83.5511°, which illustrates x- to y-polarization conversion.

 figure: Fig. 3.

Fig. 3. (a) Asymmetric transmission parameter for linearly and circularly polarized incident waves and (b) transmission coefficients for circularly polarized waves propagating in the backward direction. (c) Polarization ellipse for the transmitted wave at 1.5044 THz with an incident x-polarized wave propagating in the backward direction (EF = 0.15 eV).

Download Full Size | PDF

To better understand the asymmetric transmission effect for linearly polarized waves, the electric field distributions for x-polarized incident waves propagating backwards and forwards are shown in Fig. 4. When the x-polarized waves propagate backwards, the direction of the electric field has a 90° polarization rotation, which indicates that the incident wave is transformed into a y-polarized output wave at 1.5044 THz. In contrast, the direction of the electric field of the output wave does not change when the incident wave propagates forward, as shown in Fig. 4(b). Thus, x-polarized incident wave cannot be converted into y-polarized output wave and transmission occurs at a low level. As described above, the proposed device selectively allows x- to y-polarization conversions for linearly polarized waves propagating backwards.

 figure: Fig. 4.

Fig. 4. Calculated electric field distributions at 1.5044 THz for x-polarized wave incident along the (a) backward and (b) forward directions (EF = 0.15 eV).

Download Full Size | PDF

Based on the adjustable conductivity of DSFs, tunable asymmetric transmission can be achieved by changing the Fermi energy. Figures 5(a) and (b) show the co-polarized transmission coefficients, cross-polarized transmission coefficients, and transmitted polarization state with various Fermi energies for linearly polarized waves propagating backwards. Figure 5(a) shows that the four transmission coefficients are blueshifted as the Fermi energy increases. The corresponding polarization ellipses for the low frequency transmission peaks of the cross-polarized transmission coefficients |Tyx| are shown in Fig. 5(b). The polarization azimuths were rotated by −80.6309°, −85.6865°, and −83.5511° at Fermi energies EF= 0.07, 0.11, and 0.15 eV, which are close to 90° for x-polarized waves propagating backwards. Therefore, the asymmetric transmission effect can be tuned by varying the Fermi energy in the THz frequency regime.

 figure: Fig. 5.

Fig. 5. (a) Transmission coefficients of linearly polarized waves and (b) polarization ellipses of the transmitted waves at corresponding frequencies for x-polarized waves propagating backwards with Fermi energies of 0.07–0.15 eV (frequencies indicated on the right hand side).

Download Full Size | PDF

In addition, the asymmetric transmission effect for linearly polarized wave can be controlled by transforming the state of the VO2. Figure 6 shows the VO2-conductivity-dependence of the total transmission and asymmetric transmission parameters for x-polarized waves propagating backwards. As the metallic properties of VO2 increase, the rising conductivity leads to a continuous decrease in transmission, as shown in Fig. 6(a). When the VO2 is in the insulator state (10 S/m), $\triangle _{lin}^x$ reaches a maximum of 0.6 at a frequency of 1.504 THz and then remains relatively high across the bandwidth, see Fig. 6(b). As the VO2 transitions to a metallic state, the asymmetric transmission parameter continuously decreases because the transmittance of the device gradually decreases. Thus, asymmetric transmission occurs when VO2 is in the insulator state and it may be switched off by increasing its conductivity.

 figure: Fig. 6.

Fig. 6. (a) Total transmission and (b) asymmetric transmission parameters for x-polarized waves propagating backwards at various VO2 conductivities (EF = 0.15 eV).

Download Full Size | PDF

Next, the response of the bifunctional THz metamaterial device was investigated for VO2 in the metallic state with conductivity of 100000 S/m. In this case, the metallic state of VO2 prevents transmission. Therefore, a DSFs-Polyimide-VO2 structure in the upper half of the device and a VO2-Polyimide-DSFs structure in the lower half of the device can be regarded as three-layer traditional metamaterial absorbers. Because these structures are integrated into a single device, the proposed device can achieve dual-directional absorption when incident light illuminates from backward or forward directions, which improves the absorption efficiency. Figure 7 presents the absorptance spectra under x- and y-polarized waves incident along the backward and forward directions. When a linearly polarized wave illuminates from the backward direction, the absorptance spectra show high absorptance of 98.2% at 1.97 THz and 96.2% at 1.298 THz for x- and y-polarized waves, respectively. Accordingly, when a linearly polarized wave illuminates in the forward direction, the absorptance spectra show 98.2% absorptance at 1.97 THz and 96.2% at 1.298 THz are obtained for y- and x-polarized waves, respectively. Thus, perfect absorption can be achieved regardless of whether incident light illuminates in the backward or forward directions. Furthermore, a comparison of Figs. 7(a) and (b) shows that the absorptance spectra for x- and y-polarized waves are interchanged when the illumination direction is reversed. Hence, the response of the proposed device may be mainly analyzed for linearly polarized waves illuminating from the backward direction alone in the following.

 figure: Fig. 7.

Fig. 7. Calculated absorptance spectra for x- and y-polarized waves incident along the (a) backward and (b) forward directions when the DSFs Fermi energy is assumed to be 0.15 eV.

Download Full Size | PDF

To gain insight into the physical mechanism of absorption, the electric field distributions of the top and bottom DSFs layers and the middle VO2 layer for x- and y-polarized waves incident from both directions are shown in Fig. 8. Figures 8(a) and (d) show that opposite induced charges accumulate on the upper and lower sides, or the left and right sides, of the square split ring resonator depending on the direction of the electric field. This forms the electrical dipole resonance. A reverse charge distribution can be realized in the VO2 layer due to a coupling effect with the electrical dipole resonance. In addition, multiple electric dipole resonances and an opposing charge distribution between the two layers are formed, as shown in Figs. 8(b) and (c). Therefore, electromagnetic energy is consumed which results in perfect absorption at the corresponding absorption peaks.

 figure: Fig. 8.

Fig. 8. Calculated electric field distributions in the top and bottom DSFs layers and the middle VO2 layer for (a) y-polarized and (b) x-polarized waves incident along the backward direction, and (c) y-polarized and (d) x-polarized waves incident along the forward direction at their corresponding absorption peaks (the frequencies are indicated below the respective distributions).

Download Full Size | PDF

Figure 9 depicts the absorptance spectra for x- and y-polarized waves propagating backwards under different Fermi energies. It is obvious that the absorptance is blueshifted as the Fermi energy of the DSFs increases from 0.03 to 0.15 eV. In addition, as the Fermi energy of the DSFs increases, the absorptance for both polarizations increases because the relative impedance of the proposed device gradually approaches that of free-space. Therefore, when the proposed device acts as an absorber, the dual-directional absorption response frequency can be tuned by varying the Fermi energy.

 figure: Fig. 9.

Fig. 9. Calculated absorptance spectra for (a) y-polarized and (b) x-polarized waves propagating backwards under different Fermi energies.

Download Full Size | PDF

Based on the insulator-to-metal transition property of VO2, the reflectance, transmittance, and absorptance spectra for y-polarized waves at various VO2 conductivities are shown in Figs. 10(a)–(c). Clearly, the reflectance, transmittance, and absorptance of the device can be dynamically controlled by changing the conductivity of the VO2. When it is in the insulator state, the reflectance is highest and the corresponding absorptance is lowest at 1.298 THz. In this case, the device may be regarded to have the function of absorption in the off state and reflectance of 55%. As the conductivity of the VO2 increases, the absorptance can be adjusted from 10% to 96.2% at the frequency of absorption peak. When the VO2 is in the fully metallic state, the proposed device, which acts as dual-directional absorber, is in the on state and achieves 96.2% absorption. Thus, the state of the device can be switched between a reflector (with reflectance of 55%) and an absorber (with absorptance of 96.2%) by adjusting the conductivity of the VO2. Impedance matching theory was used to gain insight into the physics of the proposed device [49]. S-parameter retrieval gives the relative impedance as

$$Z ={\pm} \sqrt {\frac{{{{({1 + {S_{11}}} )}^2} - S_{21}^2}}{{{{({1 - {S_{11}}} )}^2} - S_{21}^2}}} $$
here, Z = Z1/Z0, where Z1 represents the equivalent surface impedance of the device and Z0 ≈ 377 Ω denotes the free space impedance. When Z = 1, the relative impedance of the device matches that of free space. This shows that perfect absorption can be achieved for this device. Figures 10(d) and (e) show the real and imaginary parts of the relative impedance when VO2 is in metallic state (100000 S/m) and insulator state (10 S/m) under a y-polarized wave propagating backwards. Figure 10(d) shows that the real part is close to 1 while the imaginary part approaches zero at the frequency of the absorption peak (1.298 THz), which means the impedance matching performance is perfect at this time. However, when VO2 is in the insulator state, the relative impedance of the device does not match that of free space, which results in poor absorption.

 figure: Fig. 10.

Fig. 10. Calculated (a) reflectance, (b) transmittance, and (c) absorptance spectra for y-polarized waves at various VO2 conductivities. Real and imaginary parts of the relative impedance when VO2 is in the (d) metallic and (e) insulator states under y-polarized waves propagating backwards (EF = 0.15 eV).

Download Full Size | PDF

Moreover, the robustness of the proposed bifunctional device with perfect absorption under oblique incident waves was investigated. Figures 11(a) and (b) depict the absorptance spectra for y- and x-polarized illumination as a function of the incidence angle and frequency for incidence angles between 0 and 80°. The peak absorption remains greater than 80% up to 65° for y-polarized waves and exhibits slight blueshift. The absorption spectra sustain high absorbance over a wide range of incidence angles up to 80°, as shown in Fig. 11(b), and a slight blueshift is observed for incident angles greater than 70°. In general, these results show that good performance can be achieved for both polarizations over a wide range of oblique incident angles, which is beneficial in practical applications.

 figure: Fig. 11.

Fig. 11. Calculated color maps for absorptance as a function of incidence angle and frequency with incidence angles between 0° and 80° for (a) y-polarized and (b) x-polarized waves propagating backwards (EF = 0.15 eV).

Download Full Size | PDF

Finally, we provide the comparison of the proposed device with others at THz frequencies. In the introduction, we pointed out that some works reported multifunctional THz metamaterial devices. The Ref. [39] firstly designed a type of VO2-based metasurfaces which can switch from a broadband absorber to a reflecting surface, and then achieved polarization-insensitive beam-steering and polarization-splitting based on the meta-atom design. The Ref. [40] presented a device integrated with perfect absorption and polarization conversion, in which a square loop resonator is embedded in VO2. The Ref. [41] proposed a bifunctional switchable metamaterial device with broadband absorption and polarization conversion. The Ref. [42] achieved the bifunctionality of perfect absorption and high transmission by utilized the design of metallic ring/silica/VO2/SiO2/subwavelength metallic mesh/silicon structure. The Ref. [43] reported a gold cross/silica/VO2/gold split ring resonator/SiO2 structure with dual functionalities of electromagnetically induced transparency and absorption. As we know that there is no relevant work has investigated bifunctional THz metamaterial device for asymmetric transmission and dual-directional absorption, Table 1 compares the proposed device with previously reported single-functional THz metamaterial devices with related functions.

Tables Icon

Table 1. Comparisons between single-function THz metamaterial devices.

4. Conclusions

In conclusion, a DSFs- and VO2-based switchable bifunctional THz metamaterial device was investigated. When the VO2 serves as a dielectric, tunable asymmetric transmission of linearly polarized waves was achieved across a wide range of THz frequencies. In this case, the asymmetric transmission experienced blueshift as the Fermi energy of the DSFs increased. This function could be turned on and off by controlling the state of the VO2. When the VO2 was in the metallic state, the proposed device acted as a tunable dual-directional absorber. It achieved 96.2% absorptance at 1.298 THz (98.2% absorptance at 1.97 THz) and 98.2% absorptance at 1.97 THz (96.2% absorptance at 1.298 THz) for y- and x-polarized waves when a linearly polarized waves illuminated from the backward (forward) direction. The absorptance could be tuned from 10% to 96.2% at the frequency of the absorption peak by varying the conductivity of the VO2 and blueshift occurred as the Fermi energy of the DSFs increased. Thus, the device can switch between a reflector (with reflectance of 55%) and an absorber (with absorptance of 96.2%). Moreover, the proposed device under the absorption function has excellent absorption properties over a wide range of incidence angles. Hence, it is promising for applications in the THz range such as optical isolation, switching, modulation, and sensing.

Funding

National Natural Science Foundation of China (61875106, 61775123); Key Research and Development Program of Shandong Province (2019GGX104039, 2019GGX104053); Shandong University of Science and Technology Top-Notch Talent Project for Young Teachers (BJRC20160505); Shandong Graduate Student Tutor Guidance Ability Promotion Program Project (SDYY17030); National Key Research and Development Program of China (2017YFA0701000).

Disclosures

The authors declare no conflicts of interest.

References

1. M. Tonouchi, “Cutting-edge terahertz technology,” Nat. Photonics 1(2), 97–105 (2007). [CrossRef]  

2. I. F. Akyildiz, J. M. Jornet, and C. Han, “Terahertz band: Next frontier for wireless communications,” Phys. Commun. 12, 16–32 (2014). [CrossRef]  

3. G. P. Williams, “Filling the THz gap—high power sources and applications,” Rep. Prog. Phys. 69(2), 301–326 (2006). [CrossRef]  

4. V. M. Shalaev, “Optical negative-index metamaterials,” Nat. Photonics 1(1), 41–48 (2007). [CrossRef]  

5. H.-T. Chen, J. F. O’Hara, A. K. Azad, and A. J. Taylor, “Manipulation of terahertz radiation using metamaterials,” Laser Photonics Rev. 5(4), 513–533 (2011). [CrossRef]  

6. H. Tao, N. I. Landy, C. M. Bingham, X. Zhang, R. D. Averitt, and W. J. Padilla, “A metamaterial absorber for the terahertz regime: design, fabrication and characterization,” Opt. Express 16(10), 7181–7188 (2008). [CrossRef]  

7. C. M. Watts, X. Liu, and W. J. Padilla, “Metamaterial electromagnetic wave absorbers,” Adv. Mater. 24(23), OP98–OP120 (2012). [CrossRef]  

8. S. Daniel and P. Bawuah, “Swastika-shaped microslots as a dual-band metamaterial absorber in the terahertz range,” OSA Continuum 2(1), 216–224 (2019). [CrossRef]  

9. B. X. Wang, X. Zhai, G. Z. Wang, W. Q. Huang, and L. L. Wang, “Design of a four-band and polarization-insensitive terahertz metamaterial absorber,” IEEE Photonics J. 7(1), 1–8 (2015). [CrossRef]  

10. J. Zhu, Z. Ma, W. Sun, F. Ding, Q. He, L. Zhou, and Y. Ma, “Ultra-broadband terahertz metamaterial absorber,” Appl. Phys. Lett. 105(2), 021102 (2014). [CrossRef]  

11. M. Kang, J. Chen, H. X. Cui, Y. Li, and H. T. Wang, “Asymmetric transmission for linearly polarized electromagnetic radiation,” Opt. Express 19(9), 8347–8356 (2011). [CrossRef]  

12. C. Huang, Y. Feng, J. Zhao, and T. Jiang, “Asymmetric electromagnetic wave transmission of linear polarization via polarization conversion through chiral metamaterial structures,” Phys. Rev. B 85(19), 195131 (2012). [CrossRef]  

13. D. J. Liu, Z. Y. Xiao, X. L. Ma, K. K. Xu, J. Y. Tang, and Z. H. Wang, “Broadband asymmetric transmission and polarization conversion of a linearly polarized wave based on chiral metamaterial in terahertz region,” Wave Motion 66, 1–9 (2016). [CrossRef]  

14. C. Menzel, C. Helgert, C. Rockstuhl, E.-B. Kley, A. Tünnermann, T. Pertsch, and F. Lederer, “Asymmetric transmission of linearly polarized light at optical metamaterials,” Phys. Rev. Lett. 104(25), 253902 (2010). [CrossRef]  

15. L. Qi, C. Liu, and S. M. A. Shah, “A broad dual-band switchable graphene-based terahertz metamaterial absorber,” Carbon 153, 179–188 (2019). [CrossRef]  

16. J. Wang and Y. Jiang, “Infrared absorber based on sandwiched two-dimensional black phosphorus metamaterials,” Opt. Express 25(5), 5206–5216 (2017). [CrossRef]  

17. Z. Zhang, J. Yang, Y. Han, X. He, J. Zhang, J. Huang, D. Chen, S. Xu, and W. Xie, “Actively tunable terahertz electromagnetically induced transparency analogue based on vanadium-oxide-assisted metamaterials,” Appl. Phys. A 126(3), 199 (2020). [CrossRef]  

18. A. Chen and Z. Song, “Wideband polarization-insensitive dielectric switch for mid-infrared waves realized by phase change material Ge3Sb2Te6,” EPL 126(2), 27004 (2019). [CrossRef]  

19. M. Wei, Z. Song, Y. Deng, Y. Liu, and Q. Chen, “Large-angle mid-infrared absorption switch enabled by polarization-independent GST metasurfaces,” Mater. Lett. 236, 350–353 (2019). [CrossRef]  

20. M. Zhang and Z. Song, “Terahertz bifunctional absorber based on a graphene-spacer-vanadium dioxide-spacer-metal configuration,” Opt. Express 28(8), 11780–11788 (2020). [CrossRef]  

21. S. X. Xia, X. Zhai, L. L. Wang, and S. C. Wen, “Plasmonically induced transparency in double-layered graphene nanoribbons,” Photonics Res. 6(7), 692–702 (2018). [CrossRef]  

22. S. X. Xia, X. Zhai, L. L. Wang, and S. C. Wen, “Polarization-independent plasmonic absorption in stacked anisotropic 2D material nanostructures,” Opt. Lett. 45(1), 93–96 (2020). [CrossRef]  

23. S. X. Xia, X. Zhai, Y. Huang, J. Q. Liu, L. L. Wang, and S. C. Wen, “Multi-band perfect plasmonic absorptions using rectangular graphene gratings,” Opt. Lett. 42(15), 3052–3055 (2017). [CrossRef]  

24. X. Luo, X. Zhai, L. Wang, and Q. Lin, “Enhanced dual-band absorption of molybdenum disulfide using a plasmonic perfect absorber,” Opt. Express 26(9), 11658–11666 (2018). [CrossRef]  

25. P. U. Jepsen, B. M. Fischer, A. Thoman, H. Helm, J. Y. Suh, R. Lopez, and R. F. Haglund, “Metal-insulator phase transition in a VO2 thin film observed with terahertz spectroscopy,” Phys. Rev. B 74(20), 205103 (2006). [CrossRef]  

26. Q. Y. Wen, H. W. Zhang, Q. H. Yang, Y. S. Xie, K. Chen, and Y. L. Liu, “Terahertz metamaterials with VO2 cut-wires for thermal tunability,” Appl. Phys. Lett. 97(2), 021111 (2010). [CrossRef]  

27. C. Zhang, G. Zhou, J. Wu, Y. Tang, Q. Wen, S. Li, J. Han, B. Jin, J. Chen, and P. Wu, “Active control of terahertz waves using vanadium-dioxide-embedded metamaterials,” Phys. Rev. Appl. 11(5), 054016 (2019). [CrossRef]  

28. H. Liu, J. Lu, and X. R. Wang, “Metamaterials based on the phase transition of VO2,” Nanotechnology 29(2), 024002 (2018). [CrossRef]  

29. T. T. Lv, Y. X. Li, H. F. Ma, Z. Zhu, Z. P. Li, C. Y. Guan, J. H. Shi, H. Zhang, and T. J. Cui, “Hybrid metamaterial switching for manipulating chirality based on VO2 phase transition,” Sci. Rep. 6(1), 23186 (2016). [CrossRef]  

30. M. Liu, Q. Xu, X. Chen, E. Plum, H. Li, X. Zhang, C. Zhang, C. Zou, J. Han, and W. Zhang, “Temperature-Controlled asymmetric transmission of electromagnetic waves,” Sci. Rep. 9(1), 4097 (2019). [CrossRef]  

31. Z. Song, K. Wang, J. Li, and Q. H. Liu, “Broadband tunable terahertz absorber based on vanadium dioxide metamaterials,” Opt. Express 26(6), 7148–7154 (2018). [CrossRef]  

32. H. Liu, Z. H. Wang, L. Li, Y. X. Fan, and Z. Y. Tao, “Vanadium dioxide-assisted broadband tunable terahertz metamaterial absorber,” Sci. Rep. 9(1), 5751 (2019). [CrossRef]  

33. T. Liang, Q. Gibson, M. N. Ali, M. Liu, R. J. Cava, and N. P. Ong, “Ultrahigh mobility and giant magnetoresistance in the Dirac semimetal Cd3As2,” Nat. Mater. 14(3), 280–284 (2015). [CrossRef]  

34. M. Neupane, S.-Y. Xu, R. Sankar, N. Alidoust, G. Bian, C. Liu, I. Belopolski, T.-R. Chang, H.-T. Jeng, H. Lin, A. Bansil, F. Chou, and M. Z. Hasan, “Observation of a three-dimensional topological Dirac semimetal phase in high-mobility Cd3As2,” Nat. Commun. 5(1), 3786 (2014). [CrossRef]  

35. Z. K. Liu, J. Jiang, B. Zhou, Z. J. Wang, Y. Zhang, H. M. Weng, D. Prabhakaran, S.-K. Mo, H. Peng, P. Dudin, T. Kim, M. Hoesch, Z. Fang, X. Dai, Z. X. Shen, D. L. Feng, Z. Hussain, and Y. L. Chen, “A stable three-dimensional topological Dirac semimetal Cd3As2,” Nat. Mater. 13(7), 677–681 (2014). [CrossRef]  

36. Z. K. Liu, B. Zhou, Y. Zhang, Z. J. Wang, H. M. Weng, D. Prabhakaran, S.-K. Mo, Z. X. Shen, Z. Fang, X. Dai, Z. Hussain, and Y. L. Chen, “Discovery of a three-dimensional topological Dirac semimetal, Na3Bi,” Science 343(6173), 864–867 (2014). [CrossRef]  

37. L. Dai, Y. Zhang, J. F. O’Hara, and H. Zhang, “Controllable broadband asymmetric transmission of terahertz wave based on Dirac semimetals,” Opt. Express 27(24), 35784–35796 (2019). [CrossRef]  

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

39. X. Li, S. Tang, F. Ding, S. Zhong, Y. Yang, T. Jiang, and J. Zhou, “Switchable multifunctional terahertz metasurfaces employing vanadium dioxide,” Sci. Rep. 9(1), 5454 (2019). [CrossRef]  

40. H. He, X. Shang, L. Xu, J. Zhao, W. Cai, J. Wang, C. Zhao, and L. Wang, “Thermally switchable bifunctional plasmonic metasurface for perfect absorption and polarization conversion based on VO2,” Opt. Express 28(4), 4563–4570 (2020). [CrossRef]  

41. Z. Song and J. Zhang, “Achieving broadband absorption and polarization conversion with a vanadium dioxide metasurface in the same terahertz frequencies,” Opt. Express 28(8), 12487–12497 (2020). [CrossRef]  

42. L. Chen and Z. Song, “Simultaneous realizations of absorber and transparent conducting metal in a single metamaterial,” Opt. Express 28(5), 6565–6571 (2020). [CrossRef]  

43. Z. Song, A. Chen, J. Zhang, and J. Wang, “Integrated metamaterial with functionalities of absorption and electromagnetically induced transparency,” Opt. Express 27(18), 25196–25204 (2019). [CrossRef]  

44. G. D. Liu, X. Zhai, H. Y. Meng, Q. Lin, Y. Huang, C. J. Zhao, and L. L. Wang, “Dirac semimetals based tunable narrowband absorber at terahertz frequencies,” Opt. Express 26(9), 11471–11480 (2018). [CrossRef]  

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

46. H. Chen, H. Zhang, M. Liu, Y. Zhao, X. Guo, and Y. Zhang, “Realization of tunable plasmon-induced transparency by bright-bright mode coupling in Dirac semimetals,” Opt. Mater. Express 7(9), 3397–3407 (2017). [CrossRef]  

47. T. Timusk, J. P. Carbotte, C. C. Homes, D. N. Basov, and S. G. Sharapov, “Three-dimensional Dirac fermions in quasicrystals as seen via optical conductivity,” Phys. Rev. B 87(23), 235121 (2013). [CrossRef]  

48. Z. Song, Y. Deng, Y. Zhou, and Z. Liu, “Terahertz toroidal metamaterial with tunable properties,” Opt. Express 27(4), 5792–5797 (2019). [CrossRef]  

49. D. R. Smith, D. C. Vier, T. Koschny, and C. M. Soukoulis, “Electromagnetic parameter retrieval from inhomogeneous metamaterials,” Phys. Rev. E 71(3), 036617 (2005). [CrossRef]  

50. J. X. Zhao, J. L. Song, T. Y. Xu, T. X. Yang, and J. H. Zhou, “Controllable linear asymmetric transmission and perfect polarization conversion in a terahertz hybrid metal-graphene metasurface,” Opt. Express 27(7), 9773–9781 (2019). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (11)

Fig. 1.
Fig. 1. Schematic of the DSFs- and VO2-based switchable bifunctional THz metamaterial device: (a) top, (b) bottom, (c) side, and (d) perspective views of the structure with P = 55 µm, L = 40 µm, g = 6 µm, w = 6 µm, td = 0.7 µm, tp=10 µm, and tv = 0.5 µm..
Fig. 2.
Fig. 2. Transmission coefficients of linearly polarized waves in (a) backward (-z) and (b) forward (+z) directions as well as the (c) PCR and (d) total transmission for x- and y-polarized waves propagating in the backward (−z) direction when the Fermi energy was 0.15 eV.
Fig. 3.
Fig. 3. (a) Asymmetric transmission parameter for linearly and circularly polarized incident waves and (b) transmission coefficients for circularly polarized waves propagating in the backward direction. (c) Polarization ellipse for the transmitted wave at 1.5044 THz with an incident x-polarized wave propagating in the backward direction (EF = 0.15 eV).
Fig. 4.
Fig. 4. Calculated electric field distributions at 1.5044 THz for x-polarized wave incident along the (a) backward and (b) forward directions (EF = 0.15 eV).
Fig. 5.
Fig. 5. (a) Transmission coefficients of linearly polarized waves and (b) polarization ellipses of the transmitted waves at corresponding frequencies for x-polarized waves propagating backwards with Fermi energies of 0.07–0.15 eV (frequencies indicated on the right hand side).
Fig. 6.
Fig. 6. (a) Total transmission and (b) asymmetric transmission parameters for x-polarized waves propagating backwards at various VO2 conductivities (EF = 0.15 eV).
Fig. 7.
Fig. 7. Calculated absorptance spectra for x- and y-polarized waves incident along the (a) backward and (b) forward directions when the DSFs Fermi energy is assumed to be 0.15 eV.
Fig. 8.
Fig. 8. Calculated electric field distributions in the top and bottom DSFs layers and the middle VO2 layer for (a) y-polarized and (b) x-polarized waves incident along the backward direction, and (c) y-polarized and (d) x-polarized waves incident along the forward direction at their corresponding absorption peaks (the frequencies are indicated below the respective distributions).
Fig. 9.
Fig. 9. Calculated absorptance spectra for (a) y-polarized and (b) x-polarized waves propagating backwards under different Fermi energies.
Fig. 10.
Fig. 10. Calculated (a) reflectance, (b) transmittance, and (c) absorptance spectra for y-polarized waves at various VO2 conductivities. Real and imaginary parts of the relative impedance when VO2 is in the (d) metallic and (e) insulator states under y-polarized waves propagating backwards (EF = 0.15 eV).
Fig. 11.
Fig. 11. Calculated color maps for absorptance as a function of incidence angle and frequency with incidence angles between 0° and 80° for (a) y-polarized and (b) x-polarized waves propagating backwards (EF = 0.15 eV).

Tables (1)

Tables Icon

Table 1. Comparisons between single-function THz metamaterial devices.

Equations (9)

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

σ i n t er = i e 2 g ω 3 π 2 v F [ π i 2 G ( ω / 2 ) 4 + 0 ( G ( E ) G ( ω / 2 ) 2 ω 2 4 E 2 ) E d E ]
σ i n t r a = i e 2 g k F 6 π 2 Ω ( 1 + π 2 3 ( T E F ) 2 )
ε  =  ε b + i σ ω ε 0
( T x T y ) = ( T x x T x y T y x T y y ) ( I x I y ) = T l i n f ( I x I y )
T l i n b = ( T x x T y x T x y T y y )
T c i r c f = ( T + + T + T + T ) = 1 2 ( T x x + T y y + i ( T x y T y x ) T x x T y y i ( T x y + T y x ) T x x T y y + i ( T x y + T y x ) T x x + T y y i ( T x y T y x ) )
Δ l i n ( x ) = | T y x | 2 | T x y | 2 = Δ l i n ( y )
Δ c i r c ( + ) = | T + | 2 | T + | 2 = Δ c i r c ( )
Z = ± ( 1 + S 11 ) 2 S 21 2 ( 1 S 11 ) 2 S 21 2
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.