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

Multi-band terahertz metamaterial absorber composed of concentric square patch and ring resonator array

Open Access Open Access

Abstract

An equivalent circuit model (ECM) to analyze a single-layered graphene multi-band metamaterial absorber was developed. This absorber consists of concentric square patch and ring resonator arrays and operates in the terahertz (THz) region. To validate our analysis based on the ECM, we also conducted numerical simulations using the finite element method (FEM) within CST software. Additionally, we have explained the absorption behavior of the metamaterial using the coupled mode theory (CMT). This absorber design, with its single-layer structure, tunability, and triple absorption bands, offers promise for applications in THz devices and systems. Notably, it achieves an average absorption of 99% for three bands and the absorption reaches 100% in the frequency range of 4 to 6.5 THz. The correlation of ECM and CMT analyses with the FEM simulations validate the accuracy and the effectiveness of these simplified approaches in comprehending the resonant characteristics of the metamaterial absorber.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Materials that can absorb light are fundamental elements in optical engineering, and they are utilized in various fields where wave absorption is crucial. These domains encompass wireless communication, sensing, and imaging. Over time, significant efforts have been dedicated to developing versatile tools in the terahertz (THz) range. This is primarily because of their possibilities in creating optical devices such as modulators, sensors, and absorbers [13]. The graphene-based resonator arrays play an indispensable role in constructing absorbers that achieve high performance [48].

Numerous studies have used arrays of graphene resonators within a bias scheme. This scheme emphasizes the significance of the Fermi energy level in controlling the device behavior across a range of frequencies [48]. Extensive research has been undertaken to realize various graphene resonator-based absorbers operating in the THz range [4,6,7,9,10].

Typically, comprehensive simulations like Finite Element Methods (FEM) [911] are employed to model such devices, which demand substantial computational power and memory resources. However, the adoption of the Equivalent Circuit Model (ECM) and the concept of equivalent conductivity (EC) [10,1215] to represent graphene configurations offers a simplified mathematical framework model these metamaterials. In our earlier paper [10], the ECM is based on EC and transfer matrix, and the geometry is contained of two finite parallel graphene ribbons per unit cell. In our earlier paper [15], the ECM is based on EC and admittance-based transmission line theory, and the geometry is contained of a double-sided comb resonator array. In our earlier paper [16], the ECM is based on ABCD matrix and S-parameters, and the geometry is contained of U-shaped resonator array. Additionally, the Coupled Mode Theory (CMT) approach has been applied to metamaterials featuring multiple resonators per unit cell, referred to as supercells [17,18].

Among the published configurations, the graphene square patch arrays [1921] have already well-established ECM representations. In addition to them, different kind of THz absorber designs have been introduced, including wideband, narrowband, and multi-band absorbers, all based on graphene patterns and some including theoretical approaches of ECM or CMT methodologies [2224].

In this paper, we present a tunable metamaterial absorber operating in the terahertz (THz) region, utilizing a graphene-based resonator array. This absorber comprises an array of concentric square patch and ring resonators, arranged to exhibit multi-band absorption characteristics. Recently, similar kind of absorber design was reported but it was based on aluminum resonator array with graphene ribbons with more complicated geometry acting as a THz dual-band filter with transmission over 80%. That paper is also lacking any theoretical approach to model its characteristics. They used graphene ribbons to make the transmission spectrum of the filter tunable [25]. To enhance the functional characteristics of metamaterial, we replaced a metal resonator array with graphene and modelled its behavior exploiting ECM. The ECM is based on EC, impedance-based transmission line theory, and variational method (eigenvalue and eigenfunction problem). The impedance of the overall pattern is modelled by EC. The impedance of inner square patch array is modelled by variational method [20]. The impedance of the outer square patch is then obtained which is the combination of EC (overall pattern) and variational method (inner square patch array). Since the metamaterial design is made of supercells, the CMT approach is proposed as well. The designed metamaterial was evaluated as a multi-band absorber, but it can also be used in other applications like filtering and sensing in THz devices and systems.

2. Materials and methods

Fig. 1 displays the metamaterial absorber design from different angles: periodic, unit cell, front, and side views. The layers making the metamaterial from the top to the bottom are respectively ion gel, graphene, Teflon, and gold (Fig. 1(d)). A layer of ion gel with a permittivity value of 2.0164 and a thickness of 150 nm is positioned atop the metamaterial. This ion gel layer serves the purpose of biasing the underlying graphene resonator layer [26].

 figure: Fig. 1.

Fig. 1. (a) Periodic, (b) unit cell, and (c) front, and (d) side views of the designed graphene-based THz metamaterial absorber composed of concentric square patch and ring resonator array. The layers from the top to the bottom are respectively ion gel, graphene, Teflon, and gold. A layer of ion gel is used for biasing the resonator array and a gold layer is used beneath the metamaterial to avoid the transmission of the electromagnetic waves.

Download Full Size | PDF

The metamaterial consists of a lower gold layer characterized by high conductivity (4.56 × 107 S/m) and a relatively large thickness (0.5 µm) relative to the wavelengths in the THz range. This layer acts as a reflective mirror against THz radiation. On this gold layer, a dielectric layer made of Teflon is placed, upon which graphene patterns are arranged. In this setup, graphene is represented as an ultra-thin layer with a thickness of 0.335 nm and is placed over the Teflon layer with a thickness of 8 µm, also known as polytetrafluoroethylene (PTFE). Teflon possesses a low permittivity value (ɛd = 2.1) and an exceptionally low loss tangent (tanδ = 0.0003), making it a suitable substrate for THz absorbers. The design also features a graphene-patterned layer consisting of periodic arrays of concentric square patch and ring resonators, placed on the Teflon dielectric layer.

To fabricate the metamaterial, the following steps can be followed: Initially, deposit an 8 µm thick Teflon dielectric with gold coating on one side using chemical vapor deposition (CVD) [27]. Next, generate the graphene patterns through a standard lithography process on the Teflon dielectric [28]. The ion gel dielectric is then put on top of the graphene resonator array via thermal evaporation [29].

The optimized dimensional values are listed in Table 1. Numerical simulations were performed based on the finite element method (FEM) in the frequency domain solver of CST microwave studio 2018. Periodic boundary conditions were applied in the x and y directions, while an absorbing boundary condition was implemented in the z direction. The tetrahedral mesh was utilized for meshing the metamaterial [10,15,30]. The metamaterial was optimized using the genetic algorithm optimization technique within the CST software [10,15,31]. The unit cell dimensions were chosen as Px = Py = 16 µm, which is smaller than λmin = 46.15 µm, the wavelength corresponding to fmax = 6.5 THz, the upper limit of the simulated frequency range. This sizing choice aimed to prevent the propagation of higher-order Floquet modes [32,33].

Tables Icon

Table 1. Structural dimensions and their optimized values for the metamaterial absorber of Fig. 1

The overall thickness of the metamaterial, comprising ion gel, graphene, Teflon, and gold layers, was 8.65 µm, approximately 0.187 ×λmin. This indicated that the metamaterial’s thickness remained relatively thin within the simulated frequency range.

The relative permittivity of graphene is modeled by [10,15]:

$${\varepsilon _g} = 1 - j\frac{{{\sigma _g}}}{{\omega {\varepsilon _0}\Delta }}$$
where σg represents the surface conductivity of graphene, ω stands for angular frequency, ε0 is the permittivity of a vacuum, and Δ corresponds to the thickness of graphene (assumed as 0.335 nm). The value of σg encompasses inter- and intra-band electron transition contributions as determined by the Kubo formula [34]:
$${\sigma _g}(\omega )= {\sigma _{{\mathop{\rm int}} er}}(\omega )+ {\sigma _{{\mathop{\rm int}} ra}}(\omega )$$
$${\sigma _{{\mathop{\rm int}} er}}(\omega )= \frac{{{e^2}}}{{4\hbar }}\left[ {H\left( {\frac{\omega }{2}} \right) - \frac{{4j\omega }}{\pi }\int_0^\infty {\frac{{H(\xi )- H\left( {\frac{\omega }{2}} \right)}}{{{\omega^2} - 4{\xi^2}}}d\xi } } \right]$$
$${\sigma _{{\mathop{\rm int}} ra}}(\omega )= \frac{{2{k_B}{e^2}T}}{{\pi {\hbar ^2}}}\ln \left[ {2\cosh \left( {\frac{{{E_f}}}{{2{k_B}T}}} \right)} \right]\frac{j}{{j{\tau ^{ - 1}} - \omega }}$$
$$H(\xi )= \frac{{\sinh \left( {\frac{{\hbar \xi }}{{{k_B}T}}} \right)}}{{\cosh \left( {\frac{{{E_f}}}{{{k_B}T}}} \right) + \cosh \left( {\frac{{\hbar \xi }}{{{k_B}T}}} \right)}}$$

Here, ${\hbar}$ represents the reduced Planck constant, kB which equals 1.38 × 10−23 J/K is Boltzmann's constant, e denoting the electron charge is 1.6 × 10−19 C, T stands for temperature (300 K), and ξ is used as the variable of integration. The parameter τ signifies the relaxation time, obtained by [35]:

$$\tau = \frac{{\mu {E_f}}}{{ev_f^2}}$$
where vf = 106 m/s is the Fermi velocity of graphene and µ denotes the carrier mobility of graphene.

The propagation constant of the electromagnetic wave on graphene-vacuum configuration can be calculated by [16]:

$$\beta = {k_0}\sqrt {1 - {{\left( {\frac{2}{{{\eta_0}\mathrm{\sigma }}}} \right)}^2}} $$
where β, k0, and η0 are respectively the propagation constant of electromagnetic wave on graphene-vacuum configuration, the wave vector of incident light wave, and the vacuum impedance. External bias voltage is applied between the graphene resonator array and the bottom gold layer to change the Ef . The relationship between Ef and bias voltage (V > 0) is approximately considered as [16]:
$${E_f} = \hbar {v_F}\sqrt {\pi {\varepsilon _0}{\varepsilon _d}\frac{V}{{ed}}} $$

The metamaterial is illuminated by TE mode. The ECMs of the graphene resonator layer and the whole metamaterial (shown in Fig. 1) are given in Fig. 2. The inner patch resonator is modeled by Z1 impedance. As the metamaterial is illuminated by TE mode, the incident electric field which is along the y-axis is normal to the upper and lower gaps between the resonators. So, these gaps are modeled by the c capacitance in the ECM. The whole impedance of the inner square patch resonator and the distances between the inner patch and outer ring is modeled by Z3. The impedance of the outer ring resonator is modeled by Z2. The impedance of the whole graphene resonator layer is modeled by Zg. The whole metamaterial is modeled by the transmission lines. The input impedances for each section of the metamaterial absorber are shown in Fig. 2(b). The thickness of the graphene layer is much smaller than the minimum wavelength in the considered simulated wavelength range, so the graphene resonator layer is modeled as a point load [36,37].

 figure: Fig. 2.

Fig. 2. Equivalent circuit model (ECM) of (a) the graphene-based resonator array and (b) the whole metamaterial absorber.

Download Full Size | PDF

In order to calculate the absorption by use of the ECM approach, the formulas are calculated by a MATLAB code. The parameters inside the formula are considered with their SI unit in the code.

The equivalent conductivity of the graphene resonator layer σg is calculated by the Fresnel equation [13]:

$${\sigma _g} = \frac{{\sec ({{\theta_{in}}} )- \sqrt {{\varepsilon _d}} \sec ({{\theta_{out}}} )- r\left( {\sec ({{\theta_{in}}} )+ \sqrt {{\varepsilon_d}} \sec ({{\theta_{out}}} )} \right)}}{{{Z_0}({1 + r} )}}$$
where θin stands for the angle of the incident illuminated wave, εd represents the relative permittivity of the dielectric substrate (Teflon), θout represents the angle of the transmitted wave, r shows the reflection of the graphene resonator array obtained in CST. To calculate r and minimize the effect of the Teflon substrate on r since graphene should always be located on a substrate for further investigation, the graphene resonator array is located on the Teflon dielectric half-space. The obtained r from CST is substituted in the MATLAB code for Eq. 6. and Z0 corresponds to the impedance of vacuum (377 Ω). The relation between θin and θout is:
$$\sin ({{\theta_{out}}} )= \sqrt {\frac{1}{{{\varepsilon _d}}}} \sin ({{\theta_{in}}} )$$

Conductivity is equal to admittance. Impedance is the inverse of admittance, so we have:

$${Z_g} = \frac{1}{{{Y_g}}} = \frac{1}{{{\sigma _g}}}$$
$${Z_g} = \frac{{{Z_0}({1 + r} )}}{{\sec ({{\theta_{in}}} )- \sqrt {{\varepsilon _d}} \sec ({{\theta_{out}}} )- r\left( {\sec ({{\theta_{in}}} )+ \sqrt {{\varepsilon_d}} \sec ({{\theta_{out}}} )} \right)}}$$

Extraction of the fundamental mode of the graphene-based inner square patch array could be done by a variational approach. The equivalent impedance of the inner square patch array is calculated by [20]:

$${Z_1} = {P_x}{P_y}\left( {\sigma_g^{ - 1} + \frac{q}{{j\omega {\varepsilon_{eff}}}}} \right)\frac{{{k_1}}}{{s_1^2}}$$
where Px and Py are the unit cell dimensions in the x and y directions, εeff, q, k1, and s1 are respectively the effective permittivity, the eigenvalues, and the coefficients of the array of the inner square patches calculated by the proposed trial eigenfunctions and the geometrical constant PxPy.

εeff are calculated by:

$${\varepsilon _{eff}} = \frac{{{\varepsilon _0}({{\varepsilon_{ig}} + {\varepsilon_d}} )}}{2}$$
k1 and s1 are calculated by [20]:
$${k_1} = {\int\limits_S {|{{\xi_x}({x,y} )} |} ^2}dS$$
$${s_1} = \int\limits_S {{\xi _x}({x,y} )dS}$$

In the context of a periodic arrangement of square graphene patches, the interaction between these patches leads to a modification in the eigenvalues of an individual graphene patch. By disregarding the influence of the disturbance on the eigenfunctions, the eigenvalues for the ensemble of graphene square patches can be expressed as [20]:

$$q = \lambda + \frac{1}{{\int\limits_S {{{|{\xi ({x,y} )} |}^2}dS} }} \times \sum\limits_{({p,q} )\ne `0} {\int\limits_S {\int\limits_{S^{\prime}} {\frac{{{{\nabla ^{\prime}}_T}.\xi ({x^{\prime},y^{\prime}} ){\nabla _T}.\xi ({x,y} )}}{{4\pi \sqrt {{{({x - x^{\prime} - p{p_1}} )}^2} + {{({y - y^{\prime} - q{p_1}} )}^2}} }}} } } dS^{\prime}dS$$
where q is the shifted eigenvalue. The third trial eigenfunction is reported to be the most accurate and complicated one and it can be obtained as [20]:
$$\lambda = 0.262\frac{\pi }{l}$$

In this method, suitable trial eigenfunctions must adhere to the boundary conditions. Considering the polarization of the incident wave, the current vector is presumed to align with the x-direction. The normal component of the surface current on the edges of graphene should be zero. As a result, the boundary condition that needs to be met is as follows [20]:

$${\xi _x}({x,y} )= {\left. { - \frac{{\partial \psi ({x,y} )}}{{\partial x}}} \right|_{x ={\pm} {\raise0.7ex\hbox{$l$} \!\mathord{/ {\vphantom {l 2}}}\!\lower0.7ex\hbox{$2$}}}} = 0$$

By employing the technique of variable separation, we express the eigenfunction of the fundamental mode in two dimensions as a multiplication of two distinct one-dimensional functions, as follows [20]:

$$\psi ({x,y} )= \Psi (x ).\Phi (y )$$

The approximates of three trial eigenvalues and eigenfunctions for the fundamental mode of the graphene-based inner square patch array satisfying Eq. 17 for the ψ (x, y) are extracted and given in Table 1 of [20]. The second and third eigenfunctions are as follows [20]:

$${\Psi _2}(x )= 1.2\left( {\frac{l}{2}} \right)\left[ {\frac{1}{2}\left[ {\left( {\frac{{2x}}{l}} \right)\sqrt {1 - {{\left( {\frac{{2x}}{l}} \right)}^2}} + {{\sin }^{ - 1}}\left( {\frac{{2x}}{l}} \right)} \right]} \right] - 0.106\left( {\frac{l}{2}} \right)\left[ { - \frac{{2x}}{l}{{\left( {1 - {{\left( {\frac{{2x}}{l}} \right)}^2}} \right)}^{\frac{3}{2}}}} \right]$$
$${\Phi _3}(y )= \cosh \left( {\alpha \frac{{2y}}{l}} \right)$$

The fitting parameter α ≈ 0.67 is based on the best fit to the current distribution obtained from full-wave simulation results in [20].

Since the metamaterial is illuminated by TE wave, the incident electric field is along the y-axis and normal to the upper and lower gaps between the inner and outer resonators. These gaps are modeled by capacitances. The capacitances are calculated by:

$$c = {\varepsilon _{eff}}\frac{l}{g}$$

The impedance of the capacitances is obtained by:

$${Z_c} = \frac{{2g}}{{j\omega {\varepsilon _0}({{\varepsilon_{ig}} + {\varepsilon_d}} )l}}$$

By having Z1 and c, Z3 (shown in Fig. 2(a)) can be obtained by:

$${Z_3} = {Z_1} + 2{Z_c}$$

So:

$${Z_3} = {P_x}{P_y}\left( {\sigma_g^{ - 1} + \frac{q}{{j\omega {\varepsilon_{eff}}}}} \right)\frac{{{k_1}}}{{s_1^2}} + \frac{{4g}}{{j\omega {\varepsilon _0}({{\varepsilon_{ig}} + {\varepsilon_d}} )l}}$$

Then, following the circuit of Fig. 2(a) which is containing of three parallel impedances, Zg is obtained by:

$${Z_g} = \frac{{{Z_2}.{Z_3}}}{{{Z_2} + {Z_3}}}$$

By substituting Eqs. 9 and 2323 in Eq. 24, the impedance of the outer square ring array Z2 can be calculated by:

$${Z_2} = \frac{{{a_1}}}{{{a_2}}}$$
$${a_1} = {Z_0}({1 + r} )[{l{P_x}{P_y}{k_1}({j\omega {\varepsilon_{eff}} + q{\sigma_g}} )+ 2g{\sigma_g}s_1^2} ]$$
$$\begin{array}{l} {a_2} = [{l{P_x}{P_y}{k_1}({j\omega {\varepsilon_{eff}} + q{\sigma_g}} )+ 2g{\sigma_g}s_1^2} ].\left[ \begin{array}{l} \cos ({{\theta_{in}}} )- \sqrt {{\varepsilon_d}} \cos ({{\theta_{out}}} )- \\ r\left( {\cos ({{\theta_{in}}} )+ \sqrt {{\varepsilon_d}} \cos ({{\theta_{out}}} )} \right) \end{array} \right] - \\ {Z_0}({1 + r} )j\omega {\varepsilon _{eff}}l{\sigma _g}s_1^2 \end{array}$$

The gold layer impedance Zgold is zero as it acts as a perfect reflector in the considered simulation ranges [15]. Zin1 in Fig. 2(b) is calculated by [36]:

$${Z_{in1}} = {Z_d}\frac{{{Z_{gold}} + j{Z_d}\tan ({{\beta_d}d} )}}{{{Z_d} + j{Z_{gold}}\tan ({{\beta_d}d} )}}$$
where Zd and βd are respectively the impedance of the Teflon dielectric substrate and the propagation constant of the THz electromagnetic waves in the Teflon substrate. Since Zgold is zero, Eq. 26 simplifies to:
$${Z_{in1}} = j{Z_d}\tan ({{\beta_d}d} )$$

As shown in Fig. 2(b), Zin2 is made of Zin1 and Zg which are parallel. So:

$${Z_{in2}} = \frac{{j{Z_d}\tan ({{\beta_d}d} ){Z_0}({1 + r} )}}{{{Z_0}({1 + r} )+ j{Z_d}\tan ({{\beta_d}d} )\left[ {\cos ({{\theta_{in}}} )- \sqrt {{\varepsilon_d}} \cos ({{\theta_{out}}} )- r\left( \begin{array}{l} \cos ({{\theta_{in}}} )\\ + \sqrt {{\varepsilon_d}} \cos ({{\theta_{out}}} )\end{array} \right)} \right]}}$$

The input impedance of the whole metamaterial absorber Zin can be obtained by [36]:

$${Z_{in}} = {Z_{ig}}\frac{{{Z_{in2}} + j{Z_{ig}}\tan ({{\beta_{ig}}{d_{ig}}} )}}{{{Z_{ig}} + j{Z_{in2}}\tan ({{\beta_{ig}}{d_{ig}}} )}}$$
where Zig and βig are respectively the impedance of the ion gel and the propagation constant of the THz electromagnetic waves in the ion gel. The impedance of ion gel/Teflon Zig/d can be obtained by [36]:
$${Z_{ig/d}} = \frac{{{Z_0}}}{{\sqrt {{\varepsilon _{ig/d}}} }}\cos ({{\theta_{in/d}}} )$$
where θin and θd are respectively the angle of the incident illuminated wave and the electrical length of the substrate. θd is calculated by:
$${\theta _d} = \frac{{d\omega \sqrt {{\varepsilon _d}} }}{c}$$
βig/d are calculated by:
$${\beta _{ig/d}} = \frac{\omega }{c}\sqrt {{\varepsilon _{ig/d}}}$$

The scattering parameter S11 is calculated by:

$${S_{11}} = \frac{{{Z_{in}} - {Z_0}}}{{{Z_{in}} + {Z_0}}}$$

The absorption is calculated by:

$$A = 1 - {|{{S_{11}}} |^2}$$

The summarized procedure of the ECM approach is as follows:

Step 1: reflection of the graphene resonator array layer r in CST

Step 2: Zg obtained by Eq. 9

Step 3: Z1 obtained by Eq. 10

Step 4: c obtained by Eq. 20

Step 5: Z3 obtained by Eqs. 22 and 23

Step 6: Z2 obtained by Eq. 25

Moreover, the absorption characteristics of metamaterial absorbers can be elucidated using the coupled-mode theory (CMT). CMT falls under the category of parametric theoretical models, providing a framework to unveil the underlying physical principles governing the interaction between artificial atoms of the super cells (unit cells containing more than one resonator) of the metamaterials. According to the principles of CMT, the absorption can be determined by the expression outlined as [17,18]:

$$A = \sum\limits_{i = 1}^n {\left( {\frac{{4{\gamma_i}{\delta_i}}}{{{{({\omega - {\omega_i}} )}^2} + {{({{\gamma_i} + {\delta_i}} )}^2}}}} \right)}$$

Since our designed metamaterial has three absorption resonances, n which shows the number of the resonances of the metamaterial is equal to 3. ${\omega_{i}}$ represents the resonance frequency, γi and δi denote the time rate of the amplitude change and the dissipative losses in the guided resonance of the metamaterial, respectively.

3. Results and discussion

Absorption spectrum of the designed metamaterial of Fig. 1 obtained by numerical simulation in CST software is given in Fig. 3(a). As its shown, the metamaterial has three resonances in the considered frequency range of 4-6.5 THz. The average of the absorption peaks is 99%. Moreover, the electric field distributions of the metamaterial absorber in the three resonance frequencies are given in Fig. 3(b-d). As it is clear from the field distributions, the whole pattern combination of the inner and outer resonators creates the resonances. Some parts of the pattern have more influence and some parts of the pattern have less influence in formation of the resonances. The inner square patch array or the outer square ring array cannot produce the resonances itself.

 figure: Fig. 3.

Fig. 3. (a) Absorption spectrum of the metamaterial absorber of Fig. 1. Electric field distributions of the metamaterial absorber of Fig. 1 in (b) 4.66 THz, (c) 5.70 THz, and (d) 6.22 THz.

Download Full Size | PDF

Fabrication of perfect graphene patterns is challenging, and the presence of defects in graphene leads to a reduction in its mobility, which subsequently affects its relaxation time [16,38]. According to Eq. 3, change of τ will change µ while the other parameters in Eq. 3 kept unchanged. Considering this condition, to assess the performance of the absorber under varying relaxation time of graphene, the absorption spectra of the device were obtained for three relaxation time values: τ = 1.5, 1.6, and 1.7 ps, all while keeping Ef unchanged and equal to 0.7 eV. Consequently, based on Eq. 3, we have different values for µ as well. The results are presented in Fig. 4. Remarkably, the change of the τ (consequently µ) has no influence on the resonance frequencies of the absorption. However, an increase in τ (consequently µ) corresponds to increased absorption peaks within the absorption spectra. This phenomenon is attributed to the fact that an elongated τ results in an augmented contribution of carriers to plasma oscillation. Therefore, the absorption of the metamaterial increases due to the increase of τ (consequently µ) [16,39,40].

 figure: Fig. 4.

Fig. 4. Absorption spectra of the absorber of Fig. 1 for three different values of τ (consequently µ) while keeping Ef unchanged and equal to 0.7 eV,

Download Full Size | PDF

The spectra of the real part of the β (Eq. 4), of the graphene layer for three different values of Ef are calculated and plotted in Fig. 5(a). As it is shown, the real part of the β decreases as the Ef increases. Thus the resonance frequencies of the absorption peaks increase with the increase of Ef, leading to a blueshift effect [41,42]. By increasing Ef, the imaginary part of the β (Eq. 4) for the graphene layer (Fig. 5(b)) which shows the losses of the electromagnetic waves in graphene decrease. Thus, the absorption increases with the Ef increase. Absorption spectra of the metamaterial absorber of Fig. 1 for three different values of Ef are given in Fig. 5(c). External bias voltage which is applied between the graphene resonator array and the bottom gold reflector should be changed for manipulating the Ef. The relationship between Ef and the external bias voltage (V > 0) is given in Eq. 5 [16,39]. Overall, the three absorption resonances show high average absorption for this absorber in the Ef of 0.7-0.8 eV. However, based on Eq. 3, the change of Ef will change τ. As shown in Fig. 4, the change of τ will slightly affect the absorption spectrum which can be neglected.

 figure: Fig. 5.

Fig. 5. (a) The real part of β (Eq. (4)) of the graphene layer, (b) the imaginary part of β (Eq. (4)) of the graphene layer, and (c) Absorption spectra of the absorber of Fig. 1 for three different values of Ef.

Download Full Size | PDF

We have simulated the metamaterial absorber of Fig. 1 for different values of the Teflon substrate thickness d with the goal of achieving the highest average absorption. Absorption spectra of the absorber for three different values of d are given in Fig. 6(a). As it is shown, the highest average absorption of the absorber reaches 99% when d is equal to 8 µm. So 8 µm is the optimized value of the substrate thickness.

 figure: Fig. 6.

Fig. 6. Absorption spectra of the absorber of Fig. 1 for three different values of (a) d (thickness of Teflon substrate), (b) g and l, and (c) g and L.

Download Full Size | PDF

The resonance frequency depends on the gap, g. g cannot change itself and its change will also affect the other parameters. There are two ways for studying the change of g: 1-Change of g and l while other parameters are kept unchanged. 2-Change of g and L while other parameters are kept unchanged. The results are shown in Figs. 6(b) and 6(c), respectively. Change of g and l, or g and L will affect the resonance frequencies and also the amplitudes of the absorptions. As depicted in Fig. 6(b), the highest average absorption is obtained for g = 1 and l = 9 µm. As shown in Fig. 6(c), the highest average absorption is obtained for g = 1 and L = 11 µm.

The real and imaginary parts of the impedance of the inner square patch resonator array (Eq. 10 and Z1 in Fig. 2(a)) are given in Fig. 7. The real part of Z1 is almost constant and positive with a small value showing almost small constant loss and the resistive nature of the inner square patch resonator array. The imaginary part contains both positive and negative sections showing the inductive and capacitive natures of the inner square patch resonator array.

 figure: Fig. 7.

Fig. 7. The real and imaginary parts of the impedance of the inner square patch resonator array (Z1 in Fig. 2(a)).

Download Full Size | PDF

Gap is defined as the distance between the inner square patch and outer square ring. The incident electric field is along the y axis, and it is perpendicular to the gap in the up and lower sides. These distances are modeled by capacitors in Fig. 2(a). Since the incident electric field is parallel to the gap in the right and left sides, these distances cannot be modeled by capacitors. The gap capacitance is shown by c in Fig. 2(a). c is obtained by Eq. (20) and it is equal to 164 pF.

To obtain the reflection spectrum of the graphene resonator layer (r at Eq. (9)), the graphene resonator array is located on a half-space Teflon substrate. The configuration is simulated, and the reflection of the resonator array is obtained. The result is given in Fig. 8.

 figure: Fig. 8.

Fig. 8. Reflection spectrum of the graphene resonator array layer when it is located on a half-space Teflon dielectric substrate to minimize the substrate effects on the reflection spectrum of the graphene layer.

Download Full Size | PDF

The obtained reflection in Fig. 8 is exploited in Eq. (9) to calculate the real and imaginary parts of the impedance of the whole graphene resonator array Zg. The results are shown in Fig. 9. The real part is positive showing the resistive nature and the imaginary part has both positive and negative sections showing the inductive and capacitive natures of the resonator array. As a consequence, from Figs. 7(a) and 9(a), loss is mainly produced by the outer square ring resonator array in this metamaterial absorber.

 figure: Fig. 9.

Fig. 9. The real and imaginary parts of the impedance of the whole graphene resonator array layer (Zg in Fig. 2(a)).

Download Full Size | PDF

The real and imaginary parts of the impedance of the outer square ring array are given in Fig. 10. As it is clear from Fig. 10(a), the outer ring array produces a great loss for the metamaterial. The real part of 2Z2 is positive and the imaginary part contains both positive and negative sections.

 figure: Fig. 10.

Fig. 10. The real and imaginary parts of the half impedance of the outer square ring resonator array (2Z2 in Fig. 2(a)).

Download Full Size | PDF

Absorption spectra of the metamaterial absorber of Fig. 1 obtained by CST, ECM, and CMT are given in Fig. 11. The obtained spectra are in good agreement. The parameters of the CMT approach in Eq. (35) and their values are given in Table 2.

 figure: Fig. 11.

Fig. 11. Absorption spectrum of the metamaterial absorber of Fig. 1 obtained by CST simulation, equivalent circuit model (ECM), and coupled mode theory (CMT) approaches.

Download Full Size | PDF

Tables Icon

Table 2. Parameters of the CMT approach in Eq. (35) and their values

Our designed multi-band graphene-based THz metamaterial absorber is compared with some previously published ones in Table 3. As can be seen, the overall features of our proposed metamaterial are improved compared to each of Refs. [4346].

Tables Icon

Table 3. Comparison of the multi-band graphene-based THz metamaterial absorbers

4. Conclusion

In this paper, a tunable multi-band metamaterial absorber is constructed from graphene-based concentric square patch and ring resonator array in the terahertz (THz) region. Equivalent circuit modeling (ECM) approach by MATLAB code is presented for the absorber. Moreover, the absorption behavior of the metamaterial is studied through coupled mode theory (CMT) since it is made of super cells. The simulation outcomes, obtained via the finite element method (FEM) in CST Microwave Studio Software, exhibit consistency with the ECM and the CMT results. The proposed metamaterial absorber is notable for being tunable, single-layered, featuring two resonators within each unit cell, showing three absorption bands with absorption reaching 100%, and demonstrating an average absorption of 99% for three bands. Furthermore, the suggested device holds promise for potential applications in future controllable THz devices and systems.

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. P. Yu, X. Chen, Z. Yi, et al., “A numerical research of wideband solar absorber based on refractory metal from visible to near infrared,” Opt. Mater. 97, 109400 (2019). [CrossRef]  

2. C. Liang, Y. Zhang, Z. Yi, et al., “A broadband and polarization-independent metamaterial perfect absorber with monolayer Cr and Ti elliptical disks array,” Results Phys. 15, 102635 (2019). [CrossRef]  

3. M. Li, C. Liang, Y. Zhang, et al., “Terahertz wideband perfect absorber based on open loop with cross nested structure,” Results Phys. 15, 102603 (2019). [CrossRef]  

4. M.-R. Nickpay, M. Danaie, and A. Shahzadi, “Graphene-based metamaterial absorber for refractive index sensing applications in terahertz band,” Diamond Relat. Mater. 130, 109539 (2022). [CrossRef]  

5. S. K. Gupta and P. K. Basu, “Tunability in graphene based metamaterial absorber structures in mid-infrared region,” IEEE Photonics J. 14(3), 1–5 (2022). [CrossRef]  

6. K. V. Babu and G. N. J. Sree, “Design and circuit analysis approach of graphene-based compact metamaterial-absorber for terahertz range applications,” Opt. Quantum Electron. 55(9), 769 (2023). [CrossRef]  

7. M.-R. Nickpay, M. Danaie, and A. Shahzadi, “Graphene-based tunable quad-band fan-shaped split-ring metamaterial absorber and refractive index sensor for THz spectrum,” Micro Nanostruct. 173, 207473 (2023). [CrossRef]  

8. S. Rangasamy, A. M. Khansadurai, G. Venugopal, et al., “Graphene-based O-shaped metamaterial absorber design with broad response for solar energy absorption,” Opt. Quantum Electron. 55(1), 90 (2023). [CrossRef]  

9. T. Chen, W. Jiang, and X. Yin, “Dual-band ultrasensitive terahertz sensor based on tunable graphene metamaterial absorber,” Superlattices Microstruct. 154, 106898 (2021). [CrossRef]  

10. S. Asgari and T. Fabritius, “Terahertz graphene-based multi-functional anisotropic metamaterial and its equivalent circuit model,” Sci. Rep. 13(1), 3433 (2023). [CrossRef]  

11. J. Bai, W. Shen, J. Shi, et al., “A non-volatile tunable terahertz metamaterial absorber using graphene floating gate,” Micromachines 12(3), 333 (2021). [CrossRef]  

12. M. Rahmanzadeh, H. Rajabalipanah, and A. Abdolali, “Analytical investigation of ultrabroadband plasma–graphene radar absorbing structures,” IEEE Trans. Plasma Sci. 45(6), 945–954 (2017). [CrossRef]  

13. M. Rahmanzadeh, H. Rajabalipanah, and A. Abdolali, “Multilayer graphene-based metasurfaces: robust design method for extremely broadband, wide-angle, and polarization-insensitive terahertz absorbers,” Appl. Opt. 57(4), 959–968 (2018). [CrossRef]  

14. M. Rahmanzadeh, A. Abdolali, A. Khavasi, et al., “Adopting image theorem for rigorous analysis of a perfect electric conductor–backed array of graphene ribbons,” J. Opt. Soc. Am. B 35(8), 1836–1844 (2018). [CrossRef]  

15. S. Asgari and T. Fabritius, “Numerical simulation and equivalent circuit model of multi-band terahertz absorber composed of double-sided graphene comb resonator array,” IEEE Access 11, 36052–36063 (2023). [CrossRef]  

16. S. Asgari and T. Fabritius, “Equivalent circuit model of graphene chiral multi-band metadevice absorber composed of U-shaped resonator array,” Opt. Express 28(26), 39850–39867 (2020). [CrossRef]  

17. H. Jiang, Y. Wang, Z. Cui, et al., “Vanadium dioxide-based terahertz metamaterial devices switchable between transmission and absorption,” Micromachines 13(5), 715 (2022). [CrossRef]  

18. H. A. Haus, Waves and Fields in Optoelectronics (Prentice Hall, 1983).

19. Y. R. Padooru, A. B. Yakovlev, C. S. Kaipa, et al., “Dual capacitive-inductive nature of periodic graphene patches: Transmission characteristics at low-terahertz frequencies,” Phys. Rev. B 87(11), 115401 (2013). [CrossRef]  

20. S. B. Parizi, M. R. Tavakol, and A. Khavasi, “Deriving surface impedance for 2-D arrays of graphene patches using a variational method,” IEEE J. Quantum Electron. 53(1), 1–6 (2017). [CrossRef]  

21. S. K. Ghosh, S. Das, and S. Bhattacharyya, “Graphene-based metasurface for tunable absorption and transmission characteristics in the near mid-infrared region,” IEEE Trans. Antennas Propag. 70(6), 4600–4612 (2022). [CrossRef]  

22. I. Rezaei, A. A. Mohammad Khani, S. Biabanifard, et al., “A reconfigurable narrow and wide band multi bias graphene based THz absorber,” Opt. Laser Technol. 151, 107996 (2022). [CrossRef]  

23. M. D. Bahloli, A. Bordbar, R. Basiri, et al., “A tunable multi-band absorber based on graphene metasurface in terahertz band,” Opt. Quantum Electron. 54(11), 708 (2022). [CrossRef]  

24. R. Cheng, Y. Zhou, X. Wu, et al., “Mechanically tunable multi-band terahertz absorber based on overlapping graphene nanoribbon arrays,” Results Phys. 52, 106817 (2023). [CrossRef]  

25. P.-J. Wu, W.-C. Tsai, and C.-S. Yang, “Electrically tunable graphene-based multi-band terahertz metamaterial filters,” Opt. Express 31(1), 469–478 (2023). [CrossRef]  

26. N. Matthaiakakis, X. Yan, H. Mizuta, et al., “Tuneable strong optical absorption in a graphene-insulator-metal hybrid plasmonic device,” Sci. Rep. 7(1), 7303 (2017). [CrossRef]  

27. J. Kang, D. Shin, S. Bae, et al., “Graphene transfer: key for applications,” Nanoscale 4(18), 5527–5537 (2012). [CrossRef]  

28. M. D. Astorino, R. Fastampa, F. Frezza, et al., “Polarization-maintaining reflection-mode THz timedomain spectroscopy of a polyimide based ultra-thin narrow-band metamaterial absorber,” Sci. Rep. 8(1), 1985 (2018). [CrossRef]  

29. Z. Yi, C. Liang, X. Chen, et al., “Dual-band plasmonic perfect absorber based on graphene metamaterials for refractive index sensing application,” Micromachines 10(7), 443 (2019). [CrossRef]  

30. R. Sarkar, D. Ghindani, K. M. Devi, et al., “Independently tunable electromagnetically induced transparency effect and dispersion in a multi-band terahertz metamaterial,” Sci. Rep. 9(1), 18068 (2019). [CrossRef]  

31. J. McCall, “Genetic algorithms for modelling and optimisation,” J. Comput. Appl. Maths. 184(1), 205–222 (2005). [CrossRef]  

32. M. Rahmanzadeh, A. Khavasi, and B. Rejaei, “Analytical method for the diffraction of an electromagnetic wave by subwavelength graphene ribbons,” J. Opt. Soc. Am. B 38(3), 953–960 (2021). [CrossRef]  

33. M. Rahmanzadeh, A. Khavasi, and B. Rejaei, “Analytical method for diffraction analysis and design of perfect-electric-conductor backed graphene ribbon metagratings,” Opt. Express 29(18), 28935–28952 (2021). [CrossRef]  

34. G. W. Hanson, “Dyadic Green’s functions and guided surface waves for a surface conductivity model of graphene,” J. Appl. Phys. 103(6), 064302 (2008). [CrossRef]  

35. Z. Wu, B. Xu, M. Yan, et al., “Tunable terahertz perfect absorber with a graphene-based double split-ring structure,” Opt. Mater. Express 11(1), 73–79 (2021). [CrossRef]  

36. D.K. Cheng, Field and Wave Electromagnetics (Addison-Wesley, 1989).

37. A. Andryieuski and A. Lavrinenko, “Graphene metamaterials based tunable terahertz absorber: effective surface conductivity approach,” Opt. Express 21(7), 9144–9155 (2013). [CrossRef]  

38. Y. Zhao, Y. Huo, B. Man, et al., “Grating-assisted surface plasmon resonance for enhancement of optical harmonic generation in graphene,” Plasmonics 14(6), 1911–1918 (2019). [CrossRef]  

39. Z. Yi, J. Chen, C. Cen, et al., “Tunable graphene-based plasmonic perfect metamaterial absorber in the THz region,” Micromachines 10(3), 194 (2019). [CrossRef]  

40. H. Li, J. Niu, and G. Wang, “Dual-band, polarization-insensitive metamaterial perfect absorber based on monolayer graphene in the mid-infrared range,” Results Phys. 13, 102313 (2019). [CrossRef]  

41. S. Asgari, N. Granpayeh, and T. Fabritius, “Controllable terahertz cross-shaped three-dimensional graphene intrinsically chiral metastructure and its biosensing application,” Opt. Commun. 474, 126080 (2020). [CrossRef]  

42. H. Li, L. L. Wang, B. Sun, et al., “Tunable mid-infrared plasmonic band-pass filter based on a single graphene sheet with cavities,” J. Appl. Phys. 116(22), 224505 (2014). [CrossRef]  

43. J. Wu, “A polarization insensitive dual-band tunable graphene absorber at the THz frequency,” Phys. Lett. A 384(35), 126890 (2020). [CrossRef]  

44. M.-R. Nickpay, M. Danaie, and A. Shahzadi, “Design of a graphene-based multi-band metamaterial perfect absorber in THz frequency region for refractive index sensing,” Phys. E (Amsterdam, Neth.) 138, 115114 (2022). [CrossRef]  

45. W. Chen, C. Li, D. Wang, et al., “A dual ultra-broadband switchable high-performance terahertz absorber based on hybrid graphene and vanadium dioxide,” Phys. Chem. Chem. Phys. 25(30), 20414–20421 (2023). [CrossRef]  

46. O. Samy, M. Belmoubarik, T. Otsuji, et al., “A voltage-tuned terahertz absorber based on MoS2/graphene nanoribbon structure,” Nanomaterials 13(11), 1716 (2023). [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 (11)

Fig. 1.
Fig. 1. (a) Periodic, (b) unit cell, and (c) front, and (d) side views of the designed graphene-based THz metamaterial absorber composed of concentric square patch and ring resonator array. The layers from the top to the bottom are respectively ion gel, graphene, Teflon, and gold. A layer of ion gel is used for biasing the resonator array and a gold layer is used beneath the metamaterial to avoid the transmission of the electromagnetic waves.
Fig. 2.
Fig. 2. Equivalent circuit model (ECM) of (a) the graphene-based resonator array and (b) the whole metamaterial absorber.
Fig. 3.
Fig. 3. (a) Absorption spectrum of the metamaterial absorber of Fig. 1. Electric field distributions of the metamaterial absorber of Fig. 1 in (b) 4.66 THz, (c) 5.70 THz, and (d) 6.22 THz.
Fig. 4.
Fig. 4. Absorption spectra of the absorber of Fig. 1 for three different values of τ (consequently µ) while keeping Ef unchanged and equal to 0.7 eV,
Fig. 5.
Fig. 5. (a) The real part of β (Eq. (4)) of the graphene layer, (b) the imaginary part of β (Eq. (4)) of the graphene layer, and (c) Absorption spectra of the absorber of Fig. 1 for three different values of Ef.
Fig. 6.
Fig. 6. Absorption spectra of the absorber of Fig. 1 for three different values of (a) d (thickness of Teflon substrate), (b) g and l, and (c) g and L.
Fig. 7.
Fig. 7. The real and imaginary parts of the impedance of the inner square patch resonator array (Z1 in Fig. 2(a)).
Fig. 8.
Fig. 8. Reflection spectrum of the graphene resonator array layer when it is located on a half-space Teflon dielectric substrate to minimize the substrate effects on the reflection spectrum of the graphene layer.
Fig. 9.
Fig. 9. The real and imaginary parts of the impedance of the whole graphene resonator array layer (Zg in Fig. 2(a)).
Fig. 10.
Fig. 10. The real and imaginary parts of the half impedance of the outer square ring resonator array (2Z2 in Fig. 2(a)).
Fig. 11.
Fig. 11. Absorption spectrum of the metamaterial absorber of Fig. 1 obtained by CST simulation, equivalent circuit model (ECM), and coupled mode theory (CMT) approaches.

Tables (3)

Tables Icon

Table 1. Structural dimensions and their optimized values for the metamaterial absorber of Fig. 1

Tables Icon

Table 2. Parameters of the CMT approach in Eq. (35) and their values

Tables Icon

Table 3. Comparison of the multi-band graphene-based THz metamaterial absorbers

Equations (40)

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

ε g = 1 j σ g ω ε 0 Δ
σ g ( ω ) = σ int e r ( ω ) + σ int r a ( ω )
σ int e r ( ω ) = e 2 4 [ H ( ω 2 ) 4 j ω π 0 H ( ξ ) H ( ω 2 ) ω 2 4 ξ 2 d ξ ]
σ int r a ( ω ) = 2 k B e 2 T π 2 ln [ 2 cosh ( E f 2 k B T ) ] j j τ 1 ω
H ( ξ ) = sinh ( ξ k B T ) cosh ( E f k B T ) + cosh ( ξ k B T )
τ = μ E f e v f 2
β = k 0 1 ( 2 η 0 σ ) 2
E f = v F π ε 0 ε d V e d
σ g = sec ( θ i n ) ε d sec ( θ o u t ) r ( sec ( θ i n ) + ε d sec ( θ o u t ) ) Z 0 ( 1 + r )
sin ( θ o u t ) = 1 ε d sin ( θ i n )
Z g = 1 Y g = 1 σ g
Z g = Z 0 ( 1 + r ) sec ( θ i n ) ε d sec ( θ o u t ) r ( sec ( θ i n ) + ε d sec ( θ o u t ) )
Z 1 = P x P y ( σ g 1 + q j ω ε e f f ) k 1 s 1 2
ε e f f = ε 0 ( ε i g + ε d ) 2
k 1 = S | ξ x ( x , y ) | 2 d S
s 1 = S ξ x ( x , y ) d S
q = λ + 1 S | ξ ( x , y ) | 2 d S × ( p , q ) 0 S S T . ξ ( x , y ) T . ξ ( x , y ) 4 π ( x x p p 1 ) 2 + ( y y q p 1 ) 2 d S d S
λ = 0.262 π l
ξ x ( x , y ) = ψ ( x , y ) x | x = ± l / l 2 2 = 0
ψ ( x , y ) = Ψ ( x ) . Φ ( y )
Ψ 2 ( x ) = 1.2 ( l 2 ) [ 1 2 [ ( 2 x l ) 1 ( 2 x l ) 2 + sin 1 ( 2 x l ) ] ] 0.106 ( l 2 ) [ 2 x l ( 1 ( 2 x l ) 2 ) 3 2 ]
Φ 3 ( y ) = cosh ( α 2 y l )
c = ε e f f l g
Z c = 2 g j ω ε 0 ( ε i g + ε d ) l
Z 3 = Z 1 + 2 Z c
Z 3 = P x P y ( σ g 1 + q j ω ε e f f ) k 1 s 1 2 + 4 g j ω ε 0 ( ε i g + ε d ) l
Z g = Z 2 . Z 3 Z 2 + Z 3
Z 2 = a 1 a 2
a 1 = Z 0 ( 1 + r ) [ l P x P y k 1 ( j ω ε e f f + q σ g ) + 2 g σ g s 1 2 ]
a 2 = [ l P x P y k 1 ( j ω ε e f f + q σ g ) + 2 g σ g s 1 2 ] . [ cos ( θ i n ) ε d cos ( θ o u t ) r ( cos ( θ i n ) + ε d cos ( θ o u t ) ) ] Z 0 ( 1 + r ) j ω ε e f f l σ g s 1 2
Z i n 1 = Z d Z g o l d + j Z d tan ( β d d ) Z d + j Z g o l d tan ( β d d )
Z i n 1 = j Z d tan ( β d d )
Z i n 2 = j Z d tan ( β d d ) Z 0 ( 1 + r ) Z 0 ( 1 + r ) + j Z d tan ( β d d ) [ cos ( θ i n ) ε d cos ( θ o u t ) r ( cos ( θ i n ) + ε d cos ( θ o u t ) ) ]
Z i n = Z i g Z i n 2 + j Z i g tan ( β i g d i g ) Z i g + j Z i n 2 tan ( β i g d i g )
Z i g / d = Z 0 ε i g / d cos ( θ i n / d )
θ d = d ω ε d c
β i g / d = ω c ε i g / d
S 11 = Z i n Z 0 Z i n + Z 0
A = 1 | S 11 | 2
A = i = 1 n ( 4 γ i δ i ( ω ω i ) 2 + ( γ i + δ i ) 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.