Next Article in Journal
Partially Coherent Cylindrical Vector Sources
Previous Article in Journal
Effect of Contact Angle on Friction Properties of Superhydrophobic Nickel Surface
Previous Article in Special Issue
Single-Layered Phase-Change Metasurfaces Achieving Polarization- and Crystallinity-Dependent Wavefront Manipulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Sensitivity Sensor Based on Diametrical Graphene Strip Plasma-Induced Transparency

1
School of Electronic Engineering and Automation, Guilin University of Electronic Technology, Guilin 541004, China
2
Guangxi Key Laboratory of Automatic Detecting Technology and Instruments, Guilin 541004, China
3
Shanxi Key Laboratory of Advanced Semiconductor Optoelectronic Devices and Integrated Systems, Jincheng 048000, China
4
Jincheng Research Institute of Opto-Mechatronics Industry, Jincheng 048000, China
5
Department of Computer Science and Engineering, Texas A&M University, College Station, TX 77843-3112, USA
*
Author to whom correspondence should be addressed.
Photonics 2023, 10(7), 830; https://doi.org/10.3390/photonics10070830
Submission received: 26 June 2023 / Revised: 12 July 2023 / Accepted: 14 July 2023 / Published: 17 July 2023
(This article belongs to the Special Issue Terahertz Metamaterials and Device Applications)

Abstract

:
In this paper, two parallel graphene strip structures are adopted to achieve tunable plasma-induced transparency (PIT) sensors in the terahertz band. Both graphene bands act as bright modes, and a PIT window appears due to the weak hybridization between them. A Lorentzian oscillation coupling model is fitted to the simulation results of the proposed structure by the finite-difference time-domain (FDTD) method and is in good agreement with the simulation results. The performance of the PIT system can be controlled by tuning the geometrical parameters of the structure. In addition, the resonant frequency of the PIT window can be dynamically adjusted by changing the chemical potential and carrier mobility of the graphene strips. When the chemical potential of graphene increases from 0.2 eV to 1 eV, the amplitude modulation depth of the PIT window (2.832 THz, 3.684 THz, and 4.386 THz) can reach 92.39%, 96.14%, and 90.4%, respectively. Furthermore, due to its dispersion characteristics, the realized PIT window has a sensitive response to the surrounding medium, and the sensitivity can be as high as 1.25 THz/RIU. This PIT effect-based graphene microstructure has important implications for the future design of terahertz modulators, optical switches, and ultrasensitive sensors.

1. Introduction

Electronically induced transparency (EIT) is a quantum interference effect that occurs in three-energy atomic systems and produces sharp windows of transparency within a broad absorption spectrum [1,2]. However, the research and practical applications of EIT in atomic systems are greatly limited by specific experimental conditions such as low temperature, coherent pumping, and high intensity [3]. To solve this problem, many researchers have endeavored to model the classical EIT effect in a new way. Surface plasmon polaritons (SPPs) supported at the metal–dielectric interface have the ability to confine electromagnetic waves in the subwavelength range and overcome the classical diffraction limit [4,5]. Therefore, plasma-induced transparency (PIT), an optical effect similar to EIT, has attracted much attention due to its remarkable advantages and wide practical applications [6]. In general, there are two schemes to achieve PIT: bright–dark mode coupling (direct coupling) [7,8,9,10,11] and bright–bright mode coupling (indirect coupling) [12,13,14,15,16]. Unlike the first approach, which is based on the phase extinction interference between the bright and dark modes, the second approach stems from the detuning between the bright and light modes, which has attracted the attention of an increasing number of researchers. In recent years, structures formed by conventional metallic materials such as Al, Cu, Ag, Au, etc. with dielectric layer materials are also capable of forming PIT window phenomena [17,18,19]. However, these aforementioned structures composed of metallic materials have huge propagation losses, making it difficult to control the dielectric constants of the metals, which will result in low modulation ranges. Changing the geometrical parameters of the conventional metallic material pattern layers to achieve dynamic control of the PIT window is essential, but the possibility of large-scale remanufacturing is still limited by the complex structures and processes, making it difficult to use widely. Moreover, it is inconvenient to adjust the geometric parameters of the metallic material pattern layer once the device is fabricated. Meanwhile, this method has the obvious drawback that the PIT peak is shifted as the structure is adjusted, which hinders the practical application effect of bright–bright mode coupled PIT in filters and modulators [20,21,22]. To address this problem, graphene, an emerging two-dimensional material composed of honeycomb carbon atoms, has been proposed to design tunable PIT devices because its plasmonic response can be actively tuned by a controlled graphene chemical potential [23,24,25,26]. Due to their superior optical properties, including flexible tunability, tight field confinement, and low propagation loss [27,28,29], various graphene-based PIT devices, such as graphene–metal hybrid structures [30,31], multilayer graphene structures [32,33,34], and graphene-based substructures [35,36], have been widely studied to tune the PIT effect. Compared to the complex structure of patterned graphene rings with multiple PIT window effects [37,38,39], our proposed structure is not only simple, but also obtains a higher sensitivity. Therefore, it has a greater potential to be widely used in practical applications. Furthermore, it is possible to analyze the effect of different graphene nanoribbon parameters (e.g., length, width, intermediate distance, position, etc.) [40,41].
In this paper, we design a 2D graphene metamaterial structure to achieve the PIT effect at terahertz frequencies. The structure consists of two parallel graphene strips. Compared with the above structures, it is very simple and easy to fabricate this metamaterial structure. By analyzing the surface electric field distribution of the transmission peak, weak hybridization between two bright modes with frequency detuning produces a PIT optical response in the terahertz region. The FDTD simulation results of the proposed structure are in high agreement with the theoretical calculations of the Lorentzian oscillation coupling model. Meanwhile, the effects of structural parameters on the PIT, such as the length and width of each of the two graphene strips, the distance between the two strips, and the carrier mobility of graphene, were investigated. In addition, the maximum modulation depth of the PIT window can reach 96.14% by changing the chemical potential of graphene. Furthermore, the proposed graphene microstructure also has an extremely high sensitivity of 1.25 THz/RIU to the variation of the surrounding medium. The sensing performance is greatly improved compared to the recently reported metamaterial sensors based on the PIT effect in the terahertz region.

2. Metamaterial Structure and FDTD Simulation Model

Figure 1a shows a schematic view of the proposed graphene metamaterial structure. The polarization angle θ is the angle between the electric field polarization direction and the x-axis. From the figure, it can be observed that the graphene microstructure consists of two parallel graphene strips and a Si substrate array cell periodically. The length L, thickness, and index of the Si substrate in Figure 1b are 4 μm, 0.15 μm, and 3.42, respectively. The length and width of the long graphene strip (strip1) are L1 and W1, respectively. The length and width of the short graphene strip (strip2) are L2 and W2, respectively. The distance between two parallel graphene strips is S. In the final fabrication of graphene metamaterials, we can lay an ionic gel layer on top of the graphene pattern layer by mechanical spin coating and deposit gold grid contacts above the ionic gel layer and below the substrate. Since the ultra-thin ionic gel layer has little effect on the spectrum as an electrode medium, it can be neglected in the simulation. When we change the gate voltage, the chemical potential of graphene can be changed by the ionic gel layer, and, finally, the dynamic regulation of graphene conductivity is achieved [39]. In this paper, a full-wave numerical simulation of the transmission spectrum of this system is performed using the FDTD method. In order to balance the simulation time and accuracy, a suitable grid is used during the simulation calculations. Periodic boundary conditions are applied in the x and y directions, and perfectly matched layers are placed along the incident plane of light in the z direction.
According to the Kubo model of conductivity of monolayer graphene, the optical properties of graphene depend on the angular frequency ω , the scattering rate Γ , the chemical potential μ c , and the temperature T . Its conductivity consists of two components, intraband electron–photon scattering and direct interband electron leap, as follows [42]:
σ ( ω , Γ , μ c , T ) = σ i n t e r ( ω , Γ , μ c , T ) + σ i n t r a ( ω , Γ , μ c , T )
σ i n t e r = i e 2 4 π l n 2 | μ c | ( ω + i 2 Γ ) 2 | μ c | + ( ω + i 2 Γ )
  σ i n t r a = i e 2 k B T π 2 ( ω + i 2 Γ ) [ μ c k B T + 2 I n ( 1 + e μ c k B T ) ]
where e   is the electron charge, ℏ is the approximate Planck constant, and k B is the Boltzmann constant. The scattering rate is related to the electron relaxation time ( τ ) through 2 Γ =   τ 1 . In the terahertz range, the direct interband electron jumps can be neglected due to | μ c | ω / 2 , so the conductivity of monolayer graphene is mainly generated by intraband electron–photon scattering. In our proposed system, the temperature T is set to 300 K, which leads to k B T < < μ c . Thus, the conductivity of monolayer graphene can be similar to the metallic Drude expressions [43,44], as follows:
σ g = i e 2 μ c π 2 ( ω + i 2 Γ )
To further analyze the properties of graphene, its propagation constant needs to be calculated. The propagation constant β and the effective refractive index n e f f of the conducting mode can be expressed as follows [45,46]:
β = k 0 ϵ 1 ( 2 ε 1 η 0 σ g ) 2
n e f f = β k 0
Here, k 0 is the wave number in free space, η 0 is the intrinsic impedance, and ε 1 is the relative dielectric constant of dielectric silicon. Therefore, we can calculate the real and imaginary parts of the effective refractive index with frequency from Equation (6), as shown in Figure 2.

3. Results and Discussion

To investigate the mechanism of inducing transparent windows in the proposed metamaterial structure, we numerically calculated the transmission spectra along the x-polarization direction. We discuss the simulation results for the structure of the cell in Figure 3. Apparently, the two graphene strips in Figure 3a produce a single PIT peak window plotted with red lines at 4.68 THz. In order to understand the formation process of the PIT, we need to understand the contribution of each part. In particular, the black line in Figure 3a represents the bright mode contributed by the short graphene strip (strip2), and the blue line represents the bright mode contributed by the long graphene strip (strip1). Due to the mode interactions in the proposed structure, PIT valleys associated with the two dips are obtained at 3.63 THz and 5.61 THz, with a transmission of 0.88 at 4.68 THz.
The z-component electric field distributions of the two transmission dips (3.63 and 5.61 THz) and the transmission peak (4.68 THz) are shown in Figure 3b–d, which further contribute to the understanding of the PIT effect obtained through bright–bright mode coupling. As can be seen in Figure 3b, corresponding to the dip at 3.63 THz, only strip1 is strongly excited by the incident light due to the interference effect between strip1 and strip2, but the field intensity around strip2 is very weak. In contrast, in Figure 3c, the field distribution of strip1 and strip2 bands at 5.61 THz is completely opposite to that in Figure 3b. When the 4.68 THz frequency is applied in Figure 3d, the two graphene strips are excited simultaneously due to resonance detuning. At the same time, the phases of two strips at the same end are opposite.
To better illustrate the formation process of the PIT transparent window, we use the Lorentz oscillatory coupling model to fit the parameters to the FDTD simulation results. In the coupling model, the incident plane wave is denoted by E ˜ ( ω ) e i ω t , the bright mode 1 resonator by M ˜ 1 ( ω ) e i ω t , and the bright mode 2 resonator by M ˜ 2 ( ω ) e i ω t . By definition, the Lorentz oscillatory coupling model under bright-mode coupling can be expressed as [47,48]:
[ ω ω 1 + i γ 1 k ˜ k ˜ ω ω 2 + i γ 2 ] [ M ˜ 1 M ˜ 2 ] = [ g 1 E ˜ g 2 E ˜ ]
where ω 1 , ω 2 , γ 1 , γ 2 represent the resonant frequencies and damping factors of bright mode 1 and bright mode 2, respectively; k ˜ represents the coupling coefficient between the two bright modes; g 1 represents the coupling strength between bright mode 1 and the electromagnetic field; and g 2 represents the coupling strength between bright mode 2 and the electromagnetic field.
Since the energy dissipation of the metamaterial structure is mainly determined by the imaginary part of the magnetic permeability, the transmittance of the graphene metamaterial structure can be simply expressed as [44]:
T ( ω ) = 1 | M ˜ 1 E ˜ | 2 = 1 | ( g 1 ( ω ω 2 + i γ 2 ) g 2 k ˜ ) E ˜ ( ω ω 1 + i γ 1 ) ( ω ω 2 + i γ 2 ) k ˜ 2 | 2
By using the Lorentz oscillatory coupling model in Equation (8), we have numerically fitted the FDTD simulation transmission curves for the case of cell structure size Lx = Ly = 4 μm, L1 = 3.5 μm, W1 = 1.3 μm, L2 = 2.1 μm, W2 = 1 μm, S = 0.4 μm in Figure 1b, as shown in Figure 4.
The fitted parameters are as follows: ω 1 = 5.39 THz, ω 2 = 3.67 THz, g 1 = 0.714 THz, g 2 = 6.44 THz, k ˜ = 0.0941 THz, γ 1 = 0.396 THz, and γ 2 = 0.389 THz. As can be seen from Figure 4, the Lorentzian coupling model curves are in good agreement with the simulation results. Therefore, the reliability and accuracy of the PIT curve obtained from the FDTD simulation are further verified.
The proposed graphene metamaterial structure has a very typical PIT phenomenon. When the graphene chemical potential increases from 0.2 eV to 1.0 eV, the PIT transparent window not only has a significant blue shift but also satisfies the conclusion f α 0 c μ c 2 π 2 h L , where f is the resonant frequency of the microstructure, α 0 the structural constant, and L the length of the graphene cut line [49]. Meanwhile, the surface conductivity of the graphene metamaterial depends on its chemical potential, so the position of the PIT window can be effectively tuned by applying a bias voltage. The PIT effect will become obvious as the chemical potential increases. The amplitude modulation of the PIT window can be achieved by controlling the chemical potential of graphene. To evaluate the modulation performance of the metamaterial structure, a PIT window with μ c = 0.6 eV was chosen as the study object. Three typical resonant frequencies of the PIT window were f d i p 1 = 2.832 THz, f p e a k   = 3.684 THz, and f d i p 2   = 4.386 THz. Figure 5b illustrates the variation of transmission peak and transmission dip with the chemical potential of graphene. The amplitude modulation performance at the three frequencies is analyzed by introducing the modulation depth M D , defined as M D = | T m a x T m i n | / T m a x , where T m a x and T m i n denote the maximum and minimum transmission amplitudes at the resonant frequencies, respectively. The modulation depth at the three resonant frequencies of f d i p 1   , f p e a k   , and f d i p 2 can reach 92.39%, 96.14%, and 90.4%, respectively. On the other hand, the high and low transmission amplitude values can be set to 1 and 0 due to the large variation in transmission amplitude at the three typical frequencies. The optical switch can be set to “on” or “off” by simply adjusting the chemical potential.
Then, we analyzed the effect of different carrier mobilities of graphene on the PIT window. It is well known that for graphene, higher carrier mobility can reduce the transmission loss and achieve higher transmittance in the transmission spectrum, as shown in Figure 6a. In addition, with the increase in carrier mobility, the plasma-induced transparent window has a red-shift phenomenon, as shown in Figure 6b.
Moreover, as shown in Figure 7a,b, the effect of length L1 on the PIT window remains unchanged when L2, W2 and W1. Clearly, length L1 determines the coupling strength between the two graphene strips and the resonant frequency of bright mode 1. When the length of strip1 is around 3.5 μm, the induced electric field strength of the two graphene strips is almost equal, and frequency detuning can be easily achieved, which can lead to a more pronounced PIT window. When the difference between length L1 of strip1 and 3.5 μm is large, the difference between the induced electric field intensity of the two graphene strips increases, which leads to the deformation of the PIT window. Similarly, as shown in Figure 7c,d, the effect of length L2 on the PIT window remains unchanged when L1, W1, and W2 of strip2. The results of the analysis are similar to the effect of length L1 on the PIT window. Length L2 determines the coupling strength between the two graphene strips and the resonant frequency of bright mode 2. The analysis shows that the resonant frequencies of the corresponding bright modes will be red-shifted as the lengths of strip1 and strip2 increase, respectively, while the peak value of the transmission spectrum of the corresponding PIT will increase and the dip value will decrease. Similarly, as shown in Figure 7e–h, we analyzed the effect of the widths W1 and W2 on the PIT window, respectively. The simulation results show that the PIT windows formed by weak hybridization of the two graphene strips are relatively stable when only the width W1 is changed. When only the width W2 is changed, the resonant frequency of the corresponding bright mode will be blue-shifted as the width W2 increases, and the peak value of the transmission spectrum of the corresponding PIT will increase and the dip value will decrease.
Next, we consider the effect of the distance S between two parallel graphene strips on the PIT window. Figure 8 shows the transmission spectral response of the proposed structure when the geometrical parameter S takes different values. From Figure 8a,b, it can be seen that there is no PIT effect when S = 0 μm. When S gradually increases, the PIT effect appears. The resonant frequencies of the bright modes corresponding to strip 1 and strip 2 will be blue-shifted, while the peak and dip2 values of the transmission spectrum of the corresponding PIT will increase and the dip1 value will decrease. Combining Figure 8a,b, it can be seen that the PIT phenomenon changes only slightly when the distance S > 0.6 μm. Therefore, parameter S > 0.6 μm has a minor effect on PIT Interestingly, the Lorentzian oscillation coupling model fitting parameters are shown in Table 1. Results show that the resonant frequencies ω 1 , ω 2 , γ 1 , γ 2 , and k ˜ in the theoretical model converge to stable values at S > 0.6 μm. Since the FDTD simulation calculations fit well with the Lorentzian oscillation coupling model, the convergence of the fitted parameter values indicates that a better transparency window is attributed to weaker hybridization [13] and that the Lorentzian oscillation coupling model can quantitatively describe the hybridization through the fitted parameters.
We also analyzed the sensitivity of the microstructure and calculated the variation of the transmission spectrum. As shown in Figure 9a, when the refractive index of the surrounding medium is increased from 1 to 1.5 with 0.1 intervals, the PIT window will show a significant redshift, and the position of the transmission peak moves from 4.712 THz to 4.085 THz. According to Figure 9b, the linear fit is consistent with the expectation between the frequency shift of the transmission peak and the refractive index. By fitting the curve, the sensitivity of the graphene metamaterial can be as high as 1.253 THz/RIU. Table 2 summarizes the performance of the graphene PIT system and compares it with previous studies that exploited the PIT effect [50,51,52,53]. It is noteworthy that the sensitivity values of the proposed structure are much higher than the results reported in recent studies. The results indicate that the proposed graphene structure is able to effectively sense changes in the refractive index of the environment. It has a reference value for the application of high-precision sensors in the terahertz region.
Figure 10a,b show the transmission spectra at different polarization angles varying from 0° to 90° with a step of 20°. When the incident light polarization angle is below 60°, we can see that the resonant frequencies do not change significantly, while the dip values of the transmission spectra of the corresponding PIT increase and the dip value changes significantly. The PIT window gradually disappears when the incident light polarization angle is greater than 60°. These analyses suggest that the proposed PIT graphene metamaterial structure is sensitive to the polarization angle because it is not highly symmetric. In Figure 10c,d, the incident angle varies with a step of 10° from 0° to 60°, and the spectrum changes slightly only when the incident angle approaches 60°. In summary, the sensor has a certain polarization angle and wide incidence angle insensitivity characteristics that can be applied to a complex test environment with large interference.

4. Conclusions

In summary, this work presents a numerical and theoretical study of a PIT system based on bright–bright mode-coupled graphene metamaterials. Both graphene bands act as bright modes, and a PIT window appears due to the weak hybridization between them. The analysis of the chemical potential and carrier mobility of graphene as well as the geometrical parameters of the structure shows that these parameters can modulate the resonant frequency and resonant intensity of the PIT window. Meanwhile, the Lorentzian oscillation coupling mode is fitted to the simulation results of the proposed structure by the finite difference in time domain (FDTD) method, and the simulation results are in good agreement with the Lorentzian coupling model curves. It is noteworthy that when the chemical potential of graphene is changed, the PIT window is blue-shifted, and the modulation depths of transmission peaks and transmission valleys reach 96.14%, 92.39%, and 90.4%, respectively. The damping factors γ 1 and γ 2 reach 0.36 THz and 0.42 THz, respectively. They are stabilized when the distance between two parallel graphene strips is greater than 0.6 μm. Furthermore, due to its dispersive properties, the realized PIT window has a sensitive response to the surrounding medium up to 1.25 THz/RIU. This PIT effect-based graphene structure has important implications for the future design of terahertz modulators, optical switches, and ultrasensitive sensors.

Author Contributions

Conceptualization and model, A.Z.; numerical simulation, A.Z. and P.B.; writing—original draft preparation, A.Z. and P.B.; investigation and formal analysis C.H.; writing—review and editing, A.Z., P.B. and L.C.; project administration and resources, R.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work is partially funded by the National Natural Science Foundation of China (62161008), the Guangxi Natural Science Foundation Joint Funding Project (2018GXNSFAA138115), Guangxi Key Laboratory of Automatic Detecting Technology and Instruments (YQ22110), the Open Project Program of Shanxi Key Laboratory of Advanced Semiconductor Optoelectronic Devices and Integrated Systems (No. 2023SZKF04, No. 2023SZKF10), Shanxi Province Science and Technology Major Program (No. 202201030201009).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data supporting the findings of this study are available within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fleischhauer, M.; Imamoglu, A.; Marangos, J.P. Electromagnetically induced transparency: Optics in coherent media. Rev. Mod. Phys. 2005, 77, 633–673. [Google Scholar] [CrossRef] [Green Version]
  2. Taubert, R.; Hentschel, M.; Giessen, H. Plasmonic analog of electromagnetically induced absorpti-on: Simulations, experiments, and coupled oscillator analysis. J. Opt. Soc. Am. B 2013, 30, 3123. [Google Scholar] [CrossRef] [Green Version]
  3. Papasimakis, N.; Fu, Y.H.; Fedotov, V.A.; Prosvirnin, S.L.; Tsai, D.P.; Zheludev, N.I. Metamaterial with polarization and direction insensitive resonant transmission response mimicking electromagnetically induced transparency. Appl. Phys. Lett. 2009, 94, 211902. [Google Scholar] [CrossRef] [Green Version]
  4. Shamshirband, S.; Malvandi, A.; Karimipour, A.; Goodarzi, M.; Afrand, M.; Petković, D.; Dahari, M.; Mahmoodian, N. Performance investigation of micro- and nano-sized particle erosion in a 90° elbow using an ANFIS model. Powder Technol. 2015, 284, 336–343. [Google Scholar] [CrossRef]
  5. Noh, S.M.C.; Shamshirband, S.; Petković, D.; Penny, R.; Zakaria, R. Adaptive neuro-fuzzy appraisal of plasmonic studies on morphology of deposited silver thin films having different thicknesses. Plasmonics 2014, 9, 1189–1196. [Google Scholar] [CrossRef]
  6. Zakaria, R.; Noh, S.M.C.; Petković, D.; Shamshirband, S.; Penny, R. RETRACTED: Investigation of plasmonic studies on morphology of deposited silver thin films having different thicknesses by soft computing methodologies—A comparative study. Phys. E Low-Dimens. Syst. Nanostructures 2014, 63, 317–323. [Google Scholar] [CrossRef]
  7. Zhang, S.; Genov, D.A.; Wang, Y.; Liu, M.; Zhang, X. Plasmon-Induced transparency in metamaterials. Phys. Rev. Lett. 2008, 101, 047401. [Google Scholar] [CrossRef] [Green Version]
  8. Liu, N.; Weiss, T.; Mesch, M.; Langguth, L.; Eigenthaler, U.; Hirscher, M.; Sönnichsen, C.; Giessen, H. Planar metamaterial analogue of electromagnetically induced transparency for plasmonic sensing. Nano Lett. 2009, 10, 1103–1107. [Google Scholar] [CrossRef]
  9. Yang, Y.; Kravchenko, I.I.; Briggs, D.P.; Valentine, J. All-dielectric metasurface analogue of electromagnetically induced transparency. Nat. Commun. 2014, 5, 5753. [Google Scholar] [CrossRef] [Green Version]
  10. Han, S.; Singh, R.; Cong, L.; Yang, H. Engineering the fano resonance and electromagnetically induced transparency in near-field coupled bright and dark metamaterial. J. Phys. D Appl. Phys. 2014, 48, 035104. [Google Scholar] [CrossRef]
  11. Wei, Z.; Li, X.; Zhong, N.; Tan, X.; Zhang, X.; Liu, H.; Meng, H.; Liang, R. Analogue Electromagne-tically Induced Transparency Based on Low-loss Metamaterial and its Application in Nanosensor and Slow-light Device. Plasmonics 2016, 12, 641–647. [Google Scholar] [CrossRef]
  12. Fu, G.-L.; Zhai, X.; Li, H.-J.; Xia, S.-X.; Wang, L.-L. Tunable plasmon-induced transparency based on bright-bright mode coupling between two parallel graphene nanostrips. Plasmonics 2016, 11, 1597–1602. [Google Scholar] [CrossRef]
  13. Jin, X.-R.; Park, J.; Zheng, H.; Lee, S.; Lee, Y.; Rhee, J.Y.; Kim, K.W.; Cheong, H.S.; Jang, W.H. Highly-dispersive transparency at optical frequencies in planar metamaterials based on two-bright-mode coupling. Opt. Express 2011, 19, 21652. [Google Scholar] [CrossRef] [PubMed]
  14. Ding, G.-W.; Liu, S.-B.; Zhang, H.-F.; Kong, X.-K.; Li, H.-M.; Li, B.-X.; Liu, S.-Y.; Li, H. Tunable electromagnetically induced transparency at terahertz frequencies in coupled graphene metamaterial. Chin. Phys. B 2015, 24, 118103. [Google Scholar] [CrossRef]
  15. Hu, X.; Yuan, S.; Armghan, A.; Liu, Y.; Jiao, Z.; Lv, H.; Zeng, C.; Huang, Y.; Huang, Q.; Wang, Y.; et al. Plasmon induced transparency and absorption in bright–bright mode coupling metamaterials: A radiating two-oscillator model analysis. J. Phys. D Appl. Phys. 2016, 50, 025301. [Google Scholar] [CrossRef]
  16. Zhang, H.; Cao, Y.; Liu, Y.; Li, Y.; Zhang, Y. A novel graphene metamaterial design for tunable terahertz plasmon induced transparency by two bright mode coupling. Opt. Commun. 2017, 391, 9–15. [Google Scholar] [CrossRef]
  17. Chen, M.; Singh, L.; Xu, N.; Singh, R.; Zhang, W.; Xie, L. Terahertz sensing of highly absorptive water-methanol mixtures with multiple resonances in metamaterials. Opt. Express 2017, 25, 14089. [Google Scholar] [CrossRef] [PubMed]
  18. Pan, W.; Yan, Y.; Ma, Y.; Shen, D. A terahertz metamaterial based on electromagnetically induced transparency effect and its sensing performance. Opt. Commun. 2019, 431, 115–119. [Google Scholar] [CrossRef]
  19. Fang, Z.; Pan, C.; Xue, Y.; Wu, B.; Wu, E. Polarization control of plasmon-induced transparency in metamaterials with reversibly convertible bright and dark modes. Appl. Opt. 2021, 60, 10689. [Google Scholar] [CrossRef]
  20. Wang, X.; Chen, C.; Gao, P.; Dai, Y.; Zhao, J.; Lu, X.; Wan, Y.; Zhao, S.; Liu, H. Slow-Light and sensing performance analysis based on plasmon-induced transparency in terahertz graphene metasurface. IEEE Sens. J. 2023, 23, 4794–4801. [Google Scholar] [CrossRef]
  21. Xiao, G.; Zhou, S.; Yang, H.; Lin, Z.; Li, H.; Liu, X.; Chen, Z.; Sun, T.; Wangyang, P.; Li, J. Dynamically tunable and multifunctional polarization beam splitters based on graphene metasurfaces. Nanomaterials 2022, 12, 3022. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, Z.; Xu, Y.; Chen, F.; Cheng, S.; Yi, Z.; Xiao, G.; Li, Y.; Jiang, J.; Zhou, X.; Chen, Z. Tunable multi-narrowband perfect absorber based on graphene and black phosphorus metamaterial. Optik 2022, 270, 169932. [Google Scholar] [CrossRef]
  23. Hong, Q.; Xiong, F.; Xu, W.; Zhu, Z.; Liu, K.; Yuan, X.; Zhang, J.; Qin, S. Towards high performance hybrid two-dimensional material plasmonic devices: Strong and highly anisotropic plasmonic resonances in nanostructured graphene-black phosphorus bilayer. Opt. Express 2018, 26, 22528. [Google Scholar] [CrossRef] [PubMed]
  24. Nong, J.; Tang, L.; Lan, G.; Luo, P.; Guo, C.; Yi, J.; Wei, W. Wideband tunable perfect absorption of graphene plasmons via attenuated total reflection in Otto prism configuration. Nanophotonics 2020, 9, 645–655. [Google Scholar] [CrossRef] [Green Version]
  25. Nong, J.; Tang, L.; Lan, G.; Luo, P.; Li, Z.; Huang, D.; Yi, J.; Shi, H.; Wei, W. Enhanced graphene plasmonic mode energy for highly sensitive molecular fingerprint retrieval. Laser Photonics Rev. 2020, 15, 2000300. [Google Scholar] [CrossRef]
  26. Zhou, X.; Xu, Y.; Li, Y.; Cheng, S.; Yi, Z.; Xiao, G.; Wang, Z.; Chen, Z. Multi-frequency switch and excellent slow light based on tunable triple plasmon-induced transparency in bilayer graphene metamaterial. Commun. Theor. Phys. 2022, 74, 115501. [Google Scholar] [CrossRef]
  27. Xia, S.; Zhai, X.; Wang, L.; Wen, S. Plasmonically induced transparency in in-plane isotropic and anisotropic 2D materials. Opt. Express 2020, 28, 7980. [Google Scholar] [CrossRef]
  28. Nong, J.; Tang, L.; Lan, G.; Luo, P.; Li, Z.; Huang, D.; Shen, J.; Wei, W. Combined visible plasmons of ag nanoparticles and infrared plasmons of graphene nanoribbons for high-performance surface-enhanced raman and infrared spectroscopies. Small 2020, 17, 2004640. [Google Scholar] [CrossRef]
  29. Nong, J.; Wei, W.; Lan, G.; Luo, P.; Guo, C.; Yi, J.; Tang, L. Resolved infrared spectroscopy of aqueous molecules employing tunable graphene plasmons in an otto prism. Anal. Chem. 2020, 92, 15370–15378. [Google Scholar] [CrossRef]
  30. Yu, A.; Guo, X.; Zhu, Y.; Balakin, A.V.; Shkurinov, A.P. Metal-graphene hybridized plasmon induced transparency in the terahertz frequencies. Opt. Express 2019, 27, 34731. [Google Scholar] [CrossRef]
  31. Dong, Z.; Sun, C.; Si, J.; Deng, X. Tunable polarization-independent plasmonically induced transparency based on metal-graphene metasurface. Opt. Express 2017, 25, 12251. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, T.; Liu, Q.; Dan, Y.; Yu, S.; Han, X.; Dai, J.; Xu, K. Machine learning and evolutionary algorithm studies of graphene metamaterials for optimized plasmon-induced transparency. Opt. Express 2020, 28, 18899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Zhou, F.; Wang, Y.; Zhang, X.; Wang, J.; Liu, Z.; Luo, X.; Zhang, Z.; Gao, E. Dynamically adjustable plasmon-induced transparency and switching application based on bilayer graphene metamaterials. J. Phys. D Appl. Phys. 2020, 54, 054002. [Google Scholar] [CrossRef]
  34. Feng, Y.; Li, Z.; Zhao, Q.; Chen, P.; Wang, J. Evaluation of Fano resonance and phase analysis of plasma induced transparency in photonic nanostructure based on equivalent circuit analysis. J. Opt. 2022, 24, 035001. [Google Scholar] [CrossRef]
  35. Xiao, B.; Tong, S.; Fyffe, A.; Shi, Z. Tunable electromagnetically induced transparency based on graphene metamaterials. Opt. Express 2020, 28, 4048. [Google Scholar] [CrossRef]
  36. Gao, E.; Li, H.; Liu, Z.; Xiong, C.; Liu, C.; Ruan, B.; Li, M.; Zhang, B. Terahertz multifunction switch and optical storage based on triple plasmon-induced transparency on a single-layer patterned graphene metasurface. Opt. Express 2020, 28, 40013. [Google Scholar] [CrossRef]
  37. Xu, H.; Li, M.; Chen, Z.; He, L.; Dong, Y.; Li, X.; Wang, X.; Nie, G.; He, Z.; Zeng, B. Optical tunable multifunctional applications based on graphene metasurface in terahertz. Phys. Scripta. 2023, 98, 045511. [Google Scholar] [CrossRef]
  38. Zhou, X.; Xu, Y.; Li, Y.; Cheng, S.; Yi, Z.; Xiao, G.; Wang, Z.; Chen, Z. High-sensitive refractive index sensing and excellent slow light based on tunable triple plasmon-induced transparency in monolayer graphene based metamaterial. Commun. Theor. Phys. 2022, 75, 015501. [Google Scholar] [CrossRef]
  39. Ge, J.; You, C.; Feng, H.; Li, X.; Wang, M.; Dong, L.; Veronis, G.; Yun, M. Tunable dual plasmon-induced transparency based on a monolayer graphene metamaterial and its terahertz sensing performance. Opt. Express 2020, 28, 31781. [Google Scholar] [CrossRef]
  40. Yang, G.; Liu, Z.; Zhou, F.; Zhuo, S.; Qin, Y.; Luo, X.; Ji, C.; Xie, Y.; Yang, R. Effect of symmetry breaking on multi-plasmon-induced transparency based on single-layer graphene metamaterials with strips and rings. J. Opt. Soc. Am. A 2023, 40, 841. [Google Scholar] [CrossRef]
  41. Gu, X.; Zhang, H.-F.; Li, M.-Y.; Chen, J.-Y.; He, Y. Theoretical analysis of tunable double plasmon induced transparency in three-ellipse-shaped resonator coupled waveguide. Acta Phys. Sin. 2022, 71, 247301. [Google Scholar] [CrossRef]
  42. Zhu, A.; Bu, P.; Hu, C.; Niu, J.; Mahapatra, R. High extinction ratio 4×2 encoder based on electro-optical grphene plasma structure. Photonics 2023, 10, 216. [Google Scholar] [CrossRef]
  43. Xu, H.; Zhao, M.; Zheng, M.; Xiong, C.; Zhang, B.; Peng, Y.; Li, H. Dual plasmon-induced transparency and slow light effect in monolayer graphene structure with rectangular defects. J. Phys. D Appl. Phys. 2018, 52, 025104. [Google Scholar] [CrossRef]
  44. Zhao, M.; Xu, H.; Xiong, C.; Zhang, B.; Liu, C.; Xie, W.; Li, H. Tunable slow light effect based on dual plasmon induced transparency in terahertz planar patterned graphene structure. Results Phys. 2019, 15, 102796. [Google Scholar] [CrossRef]
  45. Gan, C.H.; Chu, H.S.; Li, E.P. Synthesis of highly confined surface plasmon modes with doped graphene sheets in the midinfrared and terahertz frequencies. Phys. Rev. B 2012, 85, 125431. [Google Scholar] [CrossRef] [Green Version]
  46. Vakil, A.; Engheta, N. Transformation optics using graphene. Science 2011, 332, 1291–1294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Chen, M.; Xiao, Z.; Lv, F.; Cui, Z.; Xu, Q. Dynamically tunable electromagnetically induced transparency-like effect in terahertz metamaterial based on graphene cross structures. IEEE J. Sel. Top. Quantum Electron. 2022, 28, 1–9. [Google Scholar] [CrossRef]
  48. Tang, B.; Jia, Z.; Huang, L.; Su, J.; Jiang, C. Polarization-Controlled dynamically tunable electromagnetically induced transparency-like effect based on graphene metasurfaces. IEEE J. Sel. Top. Quantum Electron. 2021, 27, 1–6. [Google Scholar] [CrossRef]
  49. Ding, J.; Arigong, B.; Ren, H.; Zhou, M.; Shao, J.; Lu, M.; Chai, Y.; Lin, Y.; Zhang, H. Tuneable complementary metamaterial structures based on graphene for single and multiple transparency windows. Sci. Rep. 2014, 4, 6128. [Google Scholar] [CrossRef] [Green Version]
  50. Tang, P.; Li, J.; Du, L.; Liu, Q.; Peng, Q.; Zhao, J.; Zhu, B.; Li, Z.; Zhu, L. Ultrasensitive specific terahertz sensor based on tunable plasmon induced transparency of a graphene micro-ribbon array structure. Opt. Express 2018, 26, 30655. [Google Scholar] [CrossRef]
  51. Liu, C.; Liu, P.; Yang, C.; Lin, Y.; Zha, S. Dynamic electromagnetically induced transparency based on a metal-graphene hybrid metamaterial. Opt. Mater. Express 2018, 8, 1132. [Google Scholar] [CrossRef]
  52. Mei, J.; Shu, C.; Yang, P. Tunable electromagnetically induced transparency in graphene metamaterial in two perpendicular polarization directions. Appl. Phys. B 2019, 125, 130. [Google Scholar] [CrossRef]
  53. Chen, T.; Liang, D.; Jiang, W. A tunable terahertz graphene metamaterial sensor based on dual polarized plasmon-induced transparency. IEEE Sens. J. 2022, 22, 14084–14090. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic view of the proposed graphene metamaterial in stereo. (b) Top view of the cell of the graphene metamaterial structure with geometric parameters.
Figure 1. (a) Schematic view of the proposed graphene metamaterial in stereo. (b) Top view of the cell of the graphene metamaterial structure with geometric parameters.
Photonics 10 00830 g001
Figure 2. (a) When   μ c = 1 eV, the real part of the effective refractive index varies with frequency. (b) When μ c = 1 eV, the imaginary part of the effective refractive index varies with frequency.
Figure 2. (a) When   μ c = 1 eV, the real part of the effective refractive index varies with frequency. (b) When μ c = 1 eV, the imaginary part of the effective refractive index varies with frequency.
Photonics 10 00830 g002
Figure 3. (a) Z-directional distribution of the normalized electric field in the structure proposed in the transmission spectra (bd) of a single PIT excitation process at frequencies of 3.63, 4.68, and 5.61 THz, respectively, when μ c = 1 eV.
Figure 3. (a) Z-directional distribution of the normalized electric field in the structure proposed in the transmission spectra (bd) of a single PIT excitation process at frequencies of 3.63, 4.68, and 5.61 THz, respectively, when μ c = 1 eV.
Photonics 10 00830 g003
Figure 4. Fitting curve of Lorentz coupling model and FDTD simulation results.
Figure 4. Fitting curve of Lorentz coupling model and FDTD simulation results.
Photonics 10 00830 g004
Figure 5. (a) Transmission spectrum of the PIT window with the variation of graphene chemical potential. (b) The variation of transmission peak and transmission dip with the chemical potential of graphene.
Figure 5. (a) Transmission spectrum of the PIT window with the variation of graphene chemical potential. (b) The variation of transmission peak and transmission dip with the chemical potential of graphene.
Photonics 10 00830 g005
Figure 6. (a) Transmission spectrum of PIT window with carrier mobility. (b) The peak and trough values of transmission versus carrier mobility.
Figure 6. (a) Transmission spectrum of PIT window with carrier mobility. (b) The peak and trough values of transmission versus carrier mobility.
Photonics 10 00830 g006
Figure 7. Transmission spectra of PIT windows for different cases (a) length of long transverse graphene strip (strip1) L1, (c) length of short transverse graphene strip (strip2) L2; (e) width of long transverse graphene strip (strip1) W1; (g) width of short transverse graphene strip (strip2) of width W2. Figures (b,d,f,h), show the corresponding transmission spectral isopleth maps for different cases.
Figure 7. Transmission spectra of PIT windows for different cases (a) length of long transverse graphene strip (strip1) L1, (c) length of short transverse graphene strip (strip2) L2; (e) width of long transverse graphene strip (strip1) W1; (g) width of short transverse graphene strip (strip2) of width W2. Figures (b,d,f,h), show the corresponding transmission spectral isopleth maps for different cases.
Photonics 10 00830 g007aPhotonics 10 00830 g007b
Figure 8. (a) Transmission spectra corresponding to different geometrical parameters S. (b) Equivalence plots of transmission spectra corresponding to different geometrical parameters S.
Figure 8. (a) Transmission spectra corresponding to different geometrical parameters S. (b) Equivalence plots of transmission spectra corresponding to different geometrical parameters S.
Photonics 10 00830 g008
Figure 9. (a) Transmission spectra with different refractive indices. (b) Fitted curves between transmission peaks and refractive indices.
Figure 9. (a) Transmission spectra with different refractive indices. (b) Fitted curves between transmission peaks and refractive indices.
Photonics 10 00830 g009
Figure 10. (a) Transmission spectra of different polarization angles. (b) Equivalence plots of transmission spectra corresponding to different polarization angles. (c) Transmission spectra of different incidence angles. (d) Equivalence plots of transmission spectra corresponding to different incidence angles.
Figure 10. (a) Transmission spectra of different polarization angles. (b) Equivalence plots of transmission spectra corresponding to different polarization angles. (c) Transmission spectra of different incidence angles. (d) Equivalence plots of transmission spectra corresponding to different incidence angles.
Photonics 10 00830 g010
Table 1. The transmission curves for different geometrical parameters S in Figure 8 are fitted with the parameter values obtained from the Lorentzian oscillation coupling model.
Table 1. The transmission curves for different geometrical parameters S in Figure 8 are fitted with the parameter values obtained from the Lorentzian oscillation coupling model.
S (μm) ω 1   ( THz ) ω 2   ( THz ) γ 1   ( THz ) γ 2   ( THz ) k ˜   ( THz )
06.6934.0790.1840.57360.1687
0.14.7953.4110.53350.24570.05305
0.25.1063.5680.45990.32450.07762
0.35.3113.6240.45490.34910.01136
0.45.3913.6710.39590.3890.0941
0.55.4153.7380.37770.40530.2985
0.65.53.710.36730.41510.1359
0.75.4993.7440.35930.42230.2584
0.85.53.7590.35540.42580.2941
0.95.4853.7740.35470.42650.3357
1.05.53.7430.35940.42220.2549
Table 2. Comparison between the proposed PIT graphene metamaterial and previous work.
Table 2. Comparison between the proposed PIT graphene metamaterial and previous work.
StructureMetamaterialWorking BandSensitivity (THz/RIU)Active TunabilityRef.
StripAl0.60–2.000.31None[17]
StripGraphene1.00–6.000.36Electric-tuning[50]
U-shaped & StripAu-Graphene0.30–1.800.44Electric-tuning[51]
Ring & stripGraphene2.00–5.001.00Electric-tuning[39]
Strip & H-shapedGraphene0.50–1.00/Electric-tuning[52]
Ring & Split-ringGraphene1.00–4.001.10Electric-tuning[53]
StripGraphene0.50–8.001.25Electric-tuningThis work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhu, A.; Bu, P.; Cheng, L.; Hu, C.; Mahapatra, R. High-Sensitivity Sensor Based on Diametrical Graphene Strip Plasma-Induced Transparency. Photonics 2023, 10, 830. https://doi.org/10.3390/photonics10070830

AMA Style

Zhu A, Bu P, Cheng L, Hu C, Mahapatra R. High-Sensitivity Sensor Based on Diametrical Graphene Strip Plasma-Induced Transparency. Photonics. 2023; 10(7):830. https://doi.org/10.3390/photonics10070830

Chicago/Turabian Style

Zhu, Aijun, Pengcheng Bu, Lei Cheng, Cong Hu, and Rabi Mahapatra. 2023. "High-Sensitivity Sensor Based on Diametrical Graphene Strip Plasma-Induced Transparency" Photonics 10, no. 7: 830. https://doi.org/10.3390/photonics10070830

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop