Next Article in Journal
ADG-SleepNet: A Symmetry-Aware Multi-Scale Dilation-Gated Temporal Convolutional Network with Adaptive Attention for EEG-Based Sleep Staging
Previous Article in Journal
Multiple Bifurcation Analysis in a Discrete-Time Predator–Prey Model with Holling IV Response Function
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quasi-Periodic Hyperbolic Metamaterials Composed of Graphene and Dielectric

1
School of Electronic and Information Engineering, Hubei University of Science and Technology, Xianning 437100, China
2
College of Mechanical and Vehicle Engineering, Chongqing University, Chongqing 400044, China
3
College of Information Engineering, Henan University of Science and Technology, Luoyang 471023, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Symmetry 2025, 17(9), 1460; https://doi.org/10.3390/sym17091460
Submission received: 21 July 2025 / Revised: 11 August 2025 / Accepted: 18 August 2025 / Published: 5 September 2025
(This article belongs to the Section Physics)

Abstract

The hyperbolic properties are demonstrated in Cantor photonic multilayers, which exhibit quasi-periodicity and consist of graphene and dielectric layers, forming a centrosymmetric structure. Different from the traditionally periodic hyperbolic metamaterials, here we introduce the concept of layer weight in quasi-periodic photonic multilayers; the equivalent permittivity tensor presents a hyperbolic dispersion with a specific weight of dielectric thickness. Subsequently, the dispersion relation transforming from hyperbolicity to ellipticity has been explored. The tunability of the hyperbolic properties can be enhanced by varying the Fermi energy of the graphene, the thickness of the dielectric, the relaxation time of graphene, and the number of graphene monolayers. The reflectance and transmittance of light waves in the quasi-periodic hyperbolic metamaterials depend on the hyperbolically dispersive speciality strongly as well. This research provides theoretical guidance for the design and optimization of photonic devices based on hyperbolic metamaterials.

1. Introduction

Unlike naturally occurring materials, metamaterials [1] are a type of macroscopic functional material formed by arranging artificial sub-wavelength structural units in a regular pattern. These units typically range in size from 10 nm to 100 nm, with critical dimensions much smaller than the operating wavelength. By altering the shape, size, arrangement, and material composition of these units, the dielectric constant and magnetic permeability of metamaterials can be adjusted over a broad range, thereby enabling special optical properties hard to achieve with conventional materials and unconventional control of the light field. Hyperbolic Metamaterials (HMMs) are a class of anisotropic metamaterials that exhibit unique electromagnetic properties. In the permittivity or permeability tensor of HMMs, one principal component has an opposite sign compared to the other two principal components, leading to hyperbolic isofrequency contours [2,3]. HMMs are typically fabricated using multilayer film structures or nanowire array structures [4,5]. These materials find applications in various fields, including fundamental research [6,7,8], biological imaging [9,10], optical communication [11,12], and quantum optics [2]. Specifically, the metal–dielectric multilayer HMM can function as optical superlenses by precisely tailoring the optical field [13]. It converts the scattered evanescent waves into propagating waves within the anisotropic medium, thereby magnifying sub-wavelength objects and projecting high-resolution images in the far field. Researchers have also conducted a systematic study on the spectral characteristics of periodically arranged nanostructured HMM absorbers [14]. Simultaneously, they proposed a nanoscale HMM-based prism-coupled waveguide sensor (PCWS) operating in the near-infrared band [15]. Additionally, optical biosensors based on HMMs have been constructed to explore their potential for early cancer cell detection [16]. Another study further proposed a polarization converter assisted by a centrally split S-shaped anisotropic reflective superstructure [17].
In recent years, graphene, as a special two-dimensional material, has received extensive attention due to its excellent electrical and optical properties. The surface conductivity of graphene can be flexibly adjusted by means of external gate voltages, electromagnetic fields, or chemical doping [18,19]. This tunability allows graphene to exhibit excellent optical response characteristics in the near-infrared to terahertz frequency bands [20,21,22,23,24,25,26]. By mixing or stacking graphene with common dielectrics, hyperbolic metamaterials with tunable permittivity can be constructed, and the structure is also called graphene-based hyperbolic metamaterials (GHMMs). This kind of material can provide new ideas for the design of novel functional devices [27,28,29,30]. In the field of metamaterial research, the exploration of graphene-based structures has continuously deepened. Ivan V. Iorsh et al. proposed metamaterials based on the periodic multilayer graphene structure and demonstrated that the tuning from elliptical dispersion to hyperbolic dispersion can be achieved through an external gate voltage in the Terahertz (THz) frequency [29]. Tavana S. et al. proposed a novel structure that sandwiches GHMM between two one-dimensional periodic photonic crystals. Research indicates that this structure can be utilized to fabricate terahertz devices, with its performance tunable by varying the incident angle and the chemical potential of graphene [30]. Previous studies focused on bandgap formation in Cantor multilayers [31,32,33], whereas this work emphasizes tunable hyperbolic dispersion in graphene-based metamaterials. Different from the traditionally periodic hyperbolic metamaterials, here we achieved the hyperbolic characteristics in the Cantor quasi-periodic structure composed of graphene and dielectric. The effective permittivity tensor was calculated through the specific weight distribution of the dielectric thickness. Furthermore, the tunability of the hyperbolic properties can be enhanced by varying the Fermi energy of the graphene, the thickness of the dielectric, and the number of graphene monolayers. Additionally, the impact of hyperbolic dispersion on the reflectance and transmittance of the light beam is characterized. This effect can provide theoretical guidance for the design of the optical switch.

2. Theoretical Model and Method

The structure composed of graphene and dielectric is illustrated in Figure 1. The photonic multilayers are arranged according to the Cantor sequence [34,35], which is a typical quasi-periodic sequence. The quasi-periodic sequence is a special type of non-periodic sequence, but it has stricter long-range order than completely random or disordered non-periodic sequences. Mathematically, the iterative rule of the Cantor sequence is defined as follows: S0 = A, S1 = ABA, S2 = S1(BBB)S1, ……, SN = SN−1(BBB)N−1SN−1, ……, where (BBB)N−1 represents 3N−1 instances of B; and N = 2, 3, 4, … denotes the sequence number. Here, A and B represent the graphene monolayer and dielectric, respectively. BaF2 is selected as the dielectric. BaF2 is a commonly used material in laser optics and is regarded as a lossless dielectric [35]. In the figure, the Cantor sequence number is N = 2, leading to the formation of the structure ABA(BBB)ABA, which exhibits centrosymmetry. The transverse magnetic (TM) wave is incident on the graphene layer along the Z-axis. Specifically, εg and εd denote the permittivity of the graphene monolayer and dielectric, respectively, while tg and td represent their respective thicknesses.
The surface conductivity of graphene (σ = σintra + σinter) can be obtained using the Kubo formula [36]. Among them, the intraband conductivity σintra and the interband conductivity σinter are expressed as follows:
σ intra = i e 2 K B T π 2 ( ω + i / τ ) E F K B T + 2 ln ( 1 + e E F K B T ) ,
σ int er = i e 2 4 π 2 ln 2 E F ( ω + i / τ ) 2 E F + ( ω + i / τ ) ,
where e, , and KB are the electronic charge, the reduced Planck constant, and the Boltzmann constant, respectively. EF, τ, and T are the Fermi energy of graphene, the relaxation time, and the ambient temperature, respectively. ω represents the angular frequency of the incident light.
The effective dielectric constant of the graphene [37] can be expressed as
ε g = 1 + i σ t g ω ε 0 ,
where ε0 represents the vacuum permittivity. Since the optical wavelength λ in the near-infrared band satisfies λ >> tg + td, the effective medium theory can be adopted to describe the wave propagation based on graphene [38]. The permittivity tensor of the structure takes a diagonalized form and is expressed as [ε] = [εxx, εyy, εzz], where εxx = εyy = ε and εzz = ε⟂. Here, ε and ε represent the parallel and vertical components of the relative permittivity concerning the graphene layers, respectively. These components are formulated as follows:
ε x x = ε y y = t g 1 ε g + t d 1 ε d t g 1 + t d 1 ,
ε z z = ε d ε g ( t g 1 + t d 1 ) ε d t g 1 + t d 1 ε g ,
where tg1 = Tgtg and td1 = Tdtd. Tg and Td correspond to the proportional ratios of graphene and dielectric layers to the total layer count, respectively. For N = 2, the structure of Figure 1 consists of nine layers, including four graphene layers and five dielectric layers. The proportional coefficients are Tg = 4/9 and Td = 5/9, respectively. Based on this, the equivalent thickness of the entire multilayer structure can be derived.
For higher orders (N > 2), the counts of graphene and dielectric layers are determined by the Cantor sequence SN = SN−1(BBB)N−1SN−1. Correspondingly, the coefficients Tg and Td are derived. This framework ultimately enables the systematic analysis of hyperbolic dispersion behavior. Additionally, the dielectric constant formula obtained can also be normalized to the periodic structure. This method enhances the robustness of the structure design—even if the local arrangement deviates from the ideal quasi-periodic sequence due to fabrication errors, the overall performance can still be stabilized. When the size of the basic unit is less than one-tenth of the wavelength, this model can be used to perform equivalent modeling of the structure. This model can also be applied to other metal–dielectric structures.
For the propagation of TM-polarized light, the dispersion surface can be expressed as
k x 2 ε z z + k z 2 ε x x = k 0 2 .
kx and kz denote the wave vectors along the X-axis and Z-axis, respectively, while k0 represents the wave vector in free space. When εxxεzz < 0, the multilayer structure exhibits hyperbolic dispersion. Conversely, when εxxεzz > 0, the multilayer structure demonstrates elliptical dispersion.
For a multilayer structure, the transfer matrix method [39] can be used to analyze the reflection coefficient r and transmission coefficient t, and the matrix elements are functions of the permittivity of the structure. For a specific layer l, when an electromagnetic wave is incident on the structure, the transfer matrix can be expressed as
M l = cos δ l i η l sin δ l i η l sin δ l cos δ l ,
where δ l = 2 π n l t l cos θ / λ , (l = d,g). ηd and ηg are the effective optical admittances in the dielectric and graphene, respectively. The incident light is denoted as λ, while nd and ng are the refractive indices of the dielectric and graphene, respectively. These refractive indices can be derived from n d = ε d and n g = ε g , respectively. The structure consists of nine layers in total in Figure 1. The transfer matrix of the multilayer system is obtained by sequentially multiplying the transfer matrices of each individual layer, described as follows:
M = l = 1 9 M l = A o B o C o D o .
Then, the reflection coefficient of the structure can be calculated as
r = A o η 0 + B o η 0 η 2 N + 1 C o D o η 2 N + 1 A o η 0 + B o η 0 η 2 N + 1 + C o + D o η 2 N + 1 .
This reflection coefficient can be further decomposed into r = |r|exp(r), where |r| represents the amplitude, and Φr denotes the complex angle. The reflectance and transmittance of the structure are then determined by R = |r|2 and T = |t|2.

3. Numerical Results and Discussions

By adjusting the Fermi energy of graphene through electric fields or chemical doping, the carrier concentration in graphene can be significantly increased at terahertz frequencies. Under such conditions, electron energy transitions predominantly occur via “in-band absorption”, and the electron transport behavior becomes more analogous to that of “free electrons”, aligning with the characteristics described by the Drude model [40]. This variation in electron behavior leads to changes in the electrical conductivity of graphene, which subsequently affects the dielectric constant of the structure and further influences its dispersion characteristics.
Figure 2a,b, respectively, represent the changes in the real and imaginary parts of the graphene conductivity in the parameter space composed of the Fermi energy and frequency. In the calculations, the temperature is set as T = 300 K, and the thickness of the dielectric BaF2 is td = 20 nm, and its permittivity is εd = 2.2. The number of graphene monolayers in a single graphene layer is Ng = 1. The thickness of the graphene monolayer is tg = 0.5 nm, and the relaxation time is set as τ = 1 ps in order to take into account the scattering loss from the acoustic phonons [40,41]. The Fermi energy EF varies from 0.4 eV to 1.0 eV, and the frequency f of the light arises from 10 THz to 100 THz, corresponding to a wavelength range of 3–30 μm. The dimensions of the structure in this manuscript are much smaller than the working wavelength, fully meeting the requirements of the effective medium theory (EMT) for sub-wavelength scales. It can be seen that at low frequencies, the real and imaginary parts of the conductivity increase with the increase in Fermi energy, with intraband transitions dominating, and the conductivity follows a Drude response. Graphene is lossy. At high frequencies, the values of the real and imaginary parts of the conductivity are very small and change slowly with the increase in frequency, indicating that graphene has less loss.
Figure 3 shows the permittivity components in the parameter space composed of the Fermi energy and the frequency. It is assumed that the TM-polarized light is normally incident from the air. Figure 3a represents the real part of the vertical components εzz of the relative permittivity. The real part of εzz hardly varies with the frequency, and its value is always positive. Figure 3b shows the real part of the vertical components εxx of the relative permittivity. The imaginary part of εzz is very small at low frequencies, which indicates the low loss at this region.
Figure 3c shows the real part of the parallel components εxx of the relative permittivity, and the solid line represents the contour line with a value of 0. In the low-frequency range, the real part of εxx is negative, and εxx and εzz have opposite signs, resulting in εxxεzz < 0. Under this condition, the dispersion curve of the structure is hyperbolic. Here, a variable denoted as fc is introduced to represent the characteristic frequency at which a transition occurs from hyperbolic to elliptical dispersion. When the frequency is further increased to a certain characteristic frequency fc, with εxx > 0, the dispersion changes from hyperbolic to elliptical. Therefore, the sign change in the dielectric constant components is achieved through the epsilon-near-zero (ENZ) point [42]. When EF increases from 0.4 eV to 1.0 eV, the characteristic frequency fc increases from 47.72 THz to 76.90 THz. This means that the characteristic frequency of the hyperbolic dispersion shifts to the high-frequency region as EF increases. The imaginary part of εxx in Figure 3d decreases as the frequency increases. When f > 47.72 THz, Im[εxx] < 0.001, indicating that graphene has very low loss. Therefore, the hyperbolic properties of the structure can be flexibly adjusted by the Fermi energy of graphene.
Next, calculate the influence of different factors on the dispersion relationship. The dispersion relation formula in Equation (6) is transformed as follows:
( k x / k 0 ) 2 ε z z + ( k z / k 0 ) 2 ε x x = 1 .
Set the range of kx/k0, solve for kz/k0 in the equation, and the dispersion curve showing the variation of kz/k0 with kx/k0 can be obtained.
Figure 4a displays the dispersion curves at various Fermi energies (EF = 0.4 eV, 0.6 eV, 0.8 eV, and 1.0 eV for f = 45 THz). All dispersion curves are hyperbolic since εxxεzz < 0. As EF rises from 0.4 eV to 1.0 eV, the position and slope of the curves in the ky/k0 direction change. The slope of the curves represents the group velocity, and the group velocities of electrons are different for different Fermi energies.
In Figure 4b, the dispersion contours exhibit an elliptical shape at 58 THz when EF = 0.4 eV, under which condition εxxεzz > 0. For other values of EF, the dispersion contours remain hyperbolic. In Figure 4c, elliptical dispersion is observed at 68 THz for EF = 0.4 eV and 0.6 eV, whereas hyperbolic dispersion occurs at EF = 0.8 eV and 1.0 eV. Specifically, in Figure 4d, the structure retains a hyperbolic dispersion characteristic only at EF = 1.0 eV for f = 76.89 THz. Consequently, the Fermi energy can significantly alter the dispersion relation at high frequencies.
Figure 5a–d show the variation of the transmittance of the structure with the imaginary part of the conductivity (Im(σ)) of graphene under different Fermi energies (EF = 0.4 eV, 0.6 eV, 0.8 eV, and 1.0 eV). Im(σ) is a key parameter characterizing the optical loss of graphene, and the larger the imaginary part, the stronger the loss. As shown in the figures, increasing the Fermi energy leads to a rise in Im(σ) and a concurrent decrease in transmittance. Specifically, at EF = 0.4 eV, the maximum Im(σ) is 0.74 mS, corresponding to a transmittance of 0.75; at EF = 1.0 eV, the maximum Im(σ) increases to 1.86 mS, with transmittance dropping to 0.33, representing an approximate 2.5-fold increase in loss.
The hyperbolic dispersion characteristics are analyzed at a fixed wavelength. Subsequently, the reflectivity and transmissivity are analyzed through full-wave simulation. Figure 6a displays the reflectance of the structure under different Fermi energies (EF = 0.4 eV, 0.6 eV, 0.8 eV, and 1.0 eV). The reflectance gradually decreases with increasing frequency f and approaches zero at a characteristic frequency fc, which shifts toward higher frequencies as EF increases. Correspondingly, the transmittance in Figure 6b rises with the frequency and peaks near 1 at the same fc, which also shifts higher as EF increases. This behavior arises from the near-zero condition of the permittivity (εxx ≈ 0) at fc, where the effective wave impedance of the graphene–dielectric hybrid structure matches that of the surrounding medium. This impedance matching minimizes interfacial reflection losses, leading to maximal transmittance. The transmittance undergoes a change between the “high–low” states by altering the Fermi energy, thereby enabling the on–off function of the optical switch.
Assuming 10 THz is the selected operating frequency, tuning the gate voltage allows control over the Fermi energy of graphene, enabling the transmittance at this frequency to stably switch between two distinct states: the transmittance exceeding 0.75 is defined as the “on state” (for efficient optical signal transmission), while a transmittance below 0.33 is defined as the “off state” (for significant optical signal blocking). Using these values, the modulation depth is calculated as |0.75 − 0.33|/0.75 = 56%.
In addition to being influenced by the Fermi energy, the effective permittivity is also significantly dependent on the thickness of the dielectric. Figure 7 presents the permittivity components in the parameter space defined by the thickness of the dielectric and the frequency. The thickness td varies from 20 nm to 50 nm, and the frequency f varies from 10 THz to 100 THz. Other parameters are as follows: T = 300 K, τ = 1 ps, tg = 0.5 nm, Ng = 1, and EF = 0.4 eV.
The real part of εzz in Figure 7a is always positive and slowly decreases with the increase in td at a certain frequency. The imaginary part of εzz in Figure 7b is positive and gradually increases with the increase in frequency, but its value is very small. The real part of εxx, as shown in Figure 7c, is negative in the low-frequency range. In this case, since εxxεzz < 0, the dispersion curve exhibits a hyperbolic characteristic. The characteristic frequency fc decreases from 47.72 THz to 30.84 THz with td increasing from 20 nm to 50 nm, and the hyperbolic region shifts towards lower frequencies. Figure 7d shows the imaginary part of εxx, and the value is sensitive to values of low frequencies. Consequently, altering the thickness of the dielectric enables the flexible adjustment of the hyperbolic characteristics of the quasi-periodic structure.
In Figure 8a, all dispersion curves are hyperbolic (open contours) since εxx < 0 < εzz at 30 THz, allowing propagation of large-kx waves. In Figure 8b at 47.7 THz, most dispersion contours become elliptical for smaller td as εxx becomes positive. Only at td = 50 nm does the medium remain hyperbolic. The alteration in frequency not only modifies the shape of the curve but also results in distinct regulation patterns of dielectric thickness on dispersion characteristics.
The relaxation time of graphene is a characteristic time scale that describes the time required for electrons to recover from a non-equilibrium state to a thermal equilibrium state after being disturbed by an electric field, light field, etc. Specifically, it is the average free time between two consecutive electron-scattering events. Changing the relaxation time of graphene also exhibits a hyperbolic dispersion characteristic.
Figure 9 shows the variation of the relative dielectric constant components of the structure with frequency under different graphene relaxation times. The graphene relaxation time values are τ = 0.008 ps, 0.01 ps, 0.02 ps, and 1 ps. The frequency range is 40 THz–60 THz. Other parameters are the following: T = 300 K, εd = 2.2, EF = 0.4 eV, td = 20 nm, Ng = 1. Figure 9a represents the real part of the vertical component of the relative dielectric constant εzz. The real part of εzz increases with the increase in the frequency, but the real part values do not change much under different relaxation times. Figure 9b represents the imaginary part of εzz. Its value gradually decreases with the increase in relaxation time. The real part of the parallel component of the relative dielectric constant εxx is shown in Figure 9c. The characteristic frequencies fc corresponding to the relaxation times of 0.008 ps, 0.01 ps, 0.02 ps, and 1 ps are 43.39 THz, 44.99 THz, 47.04 THz, and 47.71 THz, respectively. The hyperbolic region moves to the high-frequency region with the increase in relaxation time. When the frequency is greater than fc, the dispersion curve of the structure is elliptical. Figure 9d represents the imaginary part of εxx. It decreases with the increase in frequency and gradually decreases in value. Therefore, the hyperbolic characteristics of the structure can be flexibly regulated by changing the frequency and graphene relaxation time.
When the frequency f = 43.38 THz, the dispersion curves of the structure all present a hyperbolic shape for τ = 0.008 ps, 0.01 ps, 0.02 ps, and 1 ps. As shown in Figure 10a, as the graphene relaxation time increases, the opening angle of the isochronous hyperbola becomes larger, which can support propagation modes with high wave vectors. When the frequency is raised to 47.70 THz, as shown in Figure 10b, only when τ = 1 ps, the dispersion curve is a hyperbola, while in other cases, it is an elliptical dispersion. As the graphene relaxation time increases, the minor axis of the elliptical isochronous contour becomes longer, demonstrating strong anisotropy.
The number of graphene monolayers is also an important factor affecting the hyperbolic properties of the structure. Figure 11 shows the permittivity components for different numbers of graphene monolayers. The number of graphene monolayers Ng varies from 1 to 4 in a single graphene layer, and the frequency f varies from 40 THz to 50 THz. Other parameters are as follows: T = 300 K, τ = 1 ps, td = 20 nm, and EF = 0.4 eV. The real and the imaginary parts of εzz are shown in Figure 11a,b, respectively. The values are positive and almost independent of low frequencies. The effect of the number of graphene monolayers on the real part of εxx is investigated in Figure 11c. The characteristic frequency fc decreases from 47.71 THz to 47.11 THz with Ng increasing from 1 to 4, and the hyperbolic region shifts towards lower frequencies. The imaginary part of εxx in Figure 11d decreases as the frequency increases and has a very small value. Therefore, changing the number of graphene monolayers can also regulate the hyperbolic characteristics.
All dispersion curves exhibit a hyperbolic shape in Figure 12a at 47.10 THz. The opening angle of the isofrequency hyperbola gradually decreases as Ng increases from 1 to 4. At 47.70 THz, as shown in Figure 12b, the dispersion contour becomes elliptical for Ng values of 2, 3, and 4 due to the condition εxxεzz > 0. The minor axis of the elliptical isofrequency contour becomes progressively shorter with increasing Ng, demonstrating strong anisotropy. The number of graphene monolayers can effectively regulate the dispersion characteristics of the metamaterial.
Figure 13a presents the electric field distribution within the structure, simulated by the FDTD method implemented in COMSOL Multiphysics 5.5 software. Figure 13b shows the electric field distribution, obtained through the transmission matrix method using MATLAB R2022b software. The horizontal axis corresponds to the direction of the graphene and dielectric layers arrangement, while the vertical axis represents the normalized intensity of the horizontal electric field in the Z direction. The entire structure is much shorter than the wavelength, and the electric field intensity decreases along the Z direction. As shown in the figures, the electric field distributions obtained from both simulation methods are consistent.
For practical device integration, graphene and dielectric layers can be alternately stacked via transfer techniques for constructing the multilayer structure. Furthermore, graphene can also be directly deposited onto the surface of dielectric layers via CVD [43,44]. CVD is a mature technique widely utilized in micro- and nanofabrication, enabling precise control of the thickness and uniformity of dielectric layers. Recent advances in techniques such as PECVD growth of graphene films, dielectric-layer-assisted transfer techniques, gradient surface energy (GSE) modulation methods, and CVD growth of graphene [45,46,47,48] validate the feasibility of fabricating such nanostructures. Regarding dynamic functionality, the optical response of graphene can be changed by electrically regulating the Fermi energy of graphene (via iron doping or external gate voltage [18]).

4. Conclusions

In conclusion, tunable hyperbolic dispersion is achieved in the Cantor-sequence multilayers of graphene and dielectric. The effective medium theory was used to derive the anisotropic permittivity of the structure. With a specific weighting of the dielectric thickness, the equivalent permittivity tensor exhibits hyperbolic dispersion. The frequency range of hyperbolic dispersion can be effectively tuned by adjusting the Fermi energy of graphene, the thickness of the dielectric, and the number of graphene monolayers. The characteristic frequency of hyperbolic dispersion shifts toward higher frequencies with the Fermi energy and the relaxation time of graphene increasing. Conversely, the characteristic frequency shifts toward lower frequencies with the thickness of the dielectric and the number of graphene monolayers increasing. Reflectance and transmittance spectra were analyzed to identify frequency bands supporting low-loss hyperbolic modes. It is hoped that this research can offer valuable theory for designing and optimizing photonic devices based on hyperbolic metamaterials.

Author Contributions

Conceptualization, D.Z. and Z.Q.; Data curation, M.Z., H.N., and X.C.; Formal analysis, J.Y., H.N., X.C., and Z.Y.; Funding acquisition, M.Z.; Investigation, M.Z. and J.Y.; Methodology, D.Z.; Project administration, Z.Y. and Y.A.; Resources, J.Y. and D.Z.; Software, M.Z.; Supervision, Y.A.; Validation, Y.A.; Writing—original draft, M.Z.; Writing—review and editing, J.Y. and Y.A. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by the Young Talent Project of the Scientific and Technological Research Plan of the Educational Commission of Hubei Province (Q20232808), the Scientific Research Project of Hubei University of Science and Technology (BK202323), the Hubei Provincial Outstanding Young and Middle-Aged Scientific and Technological Innovation Team Program for Higher Education Institutions (T2024025).

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Du, Q.-J.; Liu, J.-S.; Wang, K.-J.; Yi, X.-N.; Yang, H.-W. Dual-band terahertz left-handed metamaterial with fishnet structure. Chin. Phys. Lett. 2011, 28, 014201. [Google Scholar] [CrossRef]
  2. Poddubny, A.; Iorsh, I.; Belov, P.; Kivshar, Y. Hyperbolic metamaterials. Nat. Photonics 2013, 7, 948–957. [Google Scholar] [CrossRef]
  3. Noginov, M.; Lapine, M.; Podolskiy, V.; Kivshar, Y. Focus issue: Hyperbolic metamaterials. Opt. Express 2013, 21, 14895–14897. [Google Scholar] [CrossRef] [PubMed]
  4. Huo, P.; Zhang, S.; Liang, Y.; Lu, Y.; Xu, T. Hyperbolic metamaterials and metasurfaces: Fundamentals and applications. Adv. Opt. Mater. 2019, 7, 1801616. [Google Scholar] [CrossRef]
  5. Euvé, L.-P.; Pham, K.; Petitjeans, P.; Pagneux, V.; Maurel, A. Experimental demonstration of negative refraction of water waves using metamaterials with hyperbolic dispersion. Phys. Rev. Fluids 2024, 9, L112801. [Google Scholar] [CrossRef]
  6. Gao, H.; Zhang, X.; Wang, M.; Ding, C.; Qu, N.; Yang, B.; Zhao, M. Broadband type-I hyperbolicity independent of carrier density in RuOCl2 crystals. Phys. Rev. B 2024, 109, 115432. [Google Scholar] [CrossRef]
  7. Song, J.; Lu, L.; Cheng, Q.; Luo, Z. Three-body cat transfer between anisotropic magneto-dielectric hyperbolic metamaterials. J. Heat Transf. 2018, 140, 082005. [Google Scholar] [CrossRef]
  8. Hong, J.; Son, H.; Kim, C.; Mun, S.-E.; Sung, J.; Lee, B. Absorptive metasurface color filters based on hyperbolic metamaterials for a CMOS image sensor. Opt. Express 2021, 29, 3643–3658. [Google Scholar] [CrossRef] [PubMed]
  9. Liu, X.; Kong, W.; Wang, C.; Pu, M.; Li, Z.; Yuan, D.; Li, X.; Ma, X.; Luo, X. Hyperbolic metamaterial-assisted structured illumination microscopy using periodic sub-diffraction speckles. Opt. Mater. Express 2022, 12, 3108–3117. [Google Scholar] [CrossRef]
  10. Janaszek, B.; Tyszka-Zawadzka, A.; Szczepanski, P. Graphene-based hyperbolic metamaterial acting as tunable THz power limiter. IEEE J. Sel. Top. Quant. 2023, 29, 470029. [Google Scholar] [CrossRef]
  11. Novikov, V.B.; Leontiev, A.P.; Napolskii, K.S.; Murzina, T.V. Nonlocality-driven switchable fast-slow light effect in hyperbolic metamaterials in epsilon-near-zero regime. Phys. Rev. B 2022, 106, 165415. [Google Scholar] [CrossRef]
  12. Majumder, S.; Bhadra, S.K. Enhancement of spontaneous emission induced by all-dielectric hyperbolic metamaterial at quasi-Dirac points. Opt. Commun. 2017, 393, 113–117. [Google Scholar] [CrossRef]
  13. Baqir, M.A.; Choudhury, P.K. Hyperbolic Metamaterial-Based UV Absorber. IEEE Photonics Technol. Lett. 2017, 29, 1548–1551. [Google Scholar] [CrossRef]
  14. Baqir, M.A.; Farmani, A.; Fatima, T.; Raza, M.R.; Shaukat, S.F.; Mir, A. Nanoscale, tunable, and highly sensitive biosensor utilizing hyperbolic metamaterials in the near-infrared range. Appl. Opt. 2018, 57, 9447–9454. [Google Scholar] [CrossRef] [PubMed]
  15. Baqir, M.A.; Choudhury, P.K. Hyperbolic metamaterial-based optical biosensor for detecting cancer cells. IEEE Photonics Technol. Lett. 2023, 35, 183–186. [Google Scholar] [CrossRef]
  16. Baqir, M.; Altintas, O.; Zakir, S.; Saqlain, M.; Karaaslan, M.; Ali, M.M.; Choudhury, P. Centrally split S-shaped anisotropic reflective metasurface-assisted polarization converter. J. Opt. Soc. Am. B 2025, 42, 1054–1059. [Google Scholar] [CrossRef]
  17. Zare, S.; Tajani, B.Z.; Edalatpour, S. Effect of nonlocal electrical conductivity on near-field radiative heat transfer between graphene sheets. Phys. Rev. B 2022, 105, 125416. [Google Scholar] [CrossRef]
  18. Poumirol, J.-M.; Liu, P.Q.; Slipchenko, T.M.; Nikitin, A.Y.; Martin-Moreno, L.; Faist, J.; Kuzmenko, A.B. Electrically controlled terahertz magneto-optical phenomena in continuous and patterned graphene. Nat. Commun. 2017, 8, 14626. [Google Scholar] [CrossRef]
  19. Wang, L.; Paul, N.K.; Hihath, J.; Gomez-Diaz, J.S. Giant and broadband THz and IR emission in drift-biased graphene-based hyperbolic nanostructures. Appl. Phys. Lett. 2023, 122, 191703. [Google Scholar] [CrossRef]
  20. Li, Z.; Liang, W.Y.; Chen, W.H. Switchable hyperbolic metamaterials based on the graphene-dielectric stacking structure and optical switches design. Europhys. Lett. 2017, 120, 37001. [Google Scholar] [CrossRef]
  21. Li, N.; Li, Y.; Yu, D.; Song, H.; Zhang, Q.; Zhou, S.; Fu, S.; Wang, X. Spatial shifts on a hyperbolic metasurface of graphene grating/topological insulators. Sci. Rep. 2024, 14, 29130. [Google Scholar] [CrossRef]
  22. Hao, S.; Wang, J.; Fanayev, I.; Khakhomov, S.; Li, J. Hyperbolic metamaterial structures based on graphene for THz super-resolution imaging applications. Opt. Mater. Express 2023, 13, 247–262. [Google Scholar] [CrossRef]
  23. Lu, L.; Zhang, B.; Li, B.; Song, J.; Luo, Z.; Cheng, Q. Weak magnetic field-controlled near-field radiative heat transfer between nanoparticle-based metamaterials. Opt. Lett. 2022, 47, 4087–4090. [Google Scholar] [CrossRef]
  24. Yang, J.; Xie, Z.; Ni, H.; Qin, Z.; Chen, X.; Zhao, M.; Zhao, D. Tunable GH shift based on hyperbolic metamaterials composed of graphene and dielectric. Sci. Rep. 2025, 15, 7276. [Google Scholar] [CrossRef]
  25. Yang, J.; Xie, Z.; Ni, H.; Chen, X.; Qin, Z.; Zhao, D.; Zhao, M. Giant spatial Goos-Hänchen shift achieved in superconducting hyperbolic metamaterials with graphene. Opt. Express 2025, 33, 19359–19370. [Google Scholar] [CrossRef]
  26. Othman, M.A.K.; Guclu, C.; Capolino, F. Graphene-based tunable hyperbolic metamaterials and enhanced near-field absorption. Opt. Express 2013, 21, 7614–7632. [Google Scholar] [CrossRef] [PubMed]
  27. Gric, T.; Hess, O. Tunable surface waves at the interface separating different graphene-dielectric composite hyperbolic metamaterials. Opt. Express 2017, 25, 11466. [Google Scholar] [CrossRef] [PubMed]
  28. Iorsh, I.V.; Mukhin, I.S.; Shadrivov, I.V.; Belov, P.A.; Kivshar, Y.S. Hyperbolic metamaterials based on multilayer graphene structures. Phys. Rev. B 2013, 87, 075416. [Google Scholar] [CrossRef]
  29. Tavana, S.; Bahadori-Haghighi, S.; Ye, W.N. Tunable and ultra-narrowband multifunctional terahertz devices using anisotropic graphene based hyperbolic metamaterials. Sci. Rep. 2024, 14, 31303. [Google Scholar] [CrossRef]
  30. Eyni, Z.; Milanchian, K. Optical properties of 1D quasiperiodic structures containing graphene-based hyperbolic metamaterials. Opt. Quantum Electron. 2023, 55, 892. [Google Scholar] [CrossRef]
  31. Tavana, S.; Bahadori-Haghighi, S.; Ye, W.N. Tunable and giant spatial Goos-Hanchen shifts in a parity-time symmetric Cantor photonic crystals incorporated with a centered graphene layer. Phys. Scr. 2023, 98, 055511. [Google Scholar]
  32. Liu, J.; Shen, J.; Zhao, D.; Zhang, P.; Zubair, M. Photonic passbands induced by optical fractal effect in Cantor dielectric multilayers. PLoS ONE 2022, 17, e0268908. [Google Scholar] [CrossRef]
  33. Chen, X.; Zhang, P.; Zhong, D.; Dong, J. Tunable Dual- and Multi-Channel Filter Based on Cantor Photonic Crystals Embedded with Graphene. IEEE Access 2023, 11, 8433–8440. [Google Scholar] [CrossRef]
  34. King, T.C. Optical properties of symmetry breaking of the self-similar order in triadic Cantor photonic crystals. Appl. Opt. 2023, 62, 3632–3636. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Shi, Y.; Liang, C.H. Broadband tunable graphene-based metamaterial absorber. Opt. Mater. Express 2016, 6, 3036–3044. [Google Scholar] [CrossRef]
  36. Xu, J.; Fu, X.; Peng, Y.; Wang, S.; Zheng, Z.; Zou, X.; Qian, S.; Jiang, L.; Wang, P. Enhancement and manipulation of group delay based on topological edge state in one-dimensional photonic crystal with graphene. Opt. Express 2021, 29, 30348–30356. [Google Scholar] [CrossRef] [PubMed]
  37. Jiang, L.; Wang, Q.; Xiang, Y.; Dai, X.; Wen, S. Electrically tunable Goos-Hänchen shift of light beam reflected from a grapheneondielectric surface. IEEE Photonics J. 2013, 5, 6500108. [Google Scholar] [CrossRef]
  38. Pianelli, A.; Kowerdziej, R.; Dudek, M.; Sielezin, K.; Olifierczuk, M.; Parka, J. Graphene-based hyperbolic metamaterial as a switchable reflection modulator. Opt. Express 2020, 28, 6708–6718. [Google Scholar] [CrossRef]
  39. Zhao, M.; Chen, X.; Liu, Q.; Liu, J.; Liu, J.; Wang, Y.; Chau, Y.-F.C. Optical fractal in cryogenic environments based on distributed feedback Bragg photonic crystals. PLoS ONE 2023, 18, e0291863. [Google Scholar] [CrossRef] [PubMed]
  40. Vakil, A.; Engheta, N. Transformation Optics Using Graphene. Science 2011, 332, 1291–1294. [Google Scholar] [CrossRef] [PubMed]
  41. Sreekanth, K.V.; Luca, A.D.; Strangi, G. Negative refraction in graphene-based hyperbolic metamaterials. Appl. Phys. Lett. 2013, 103, 023107. [Google Scholar] [CrossRef]
  42. Smith, E.M.; Chen, J.; Hendrickson, J.R.; Cleary, J.W.; Dass, C.; Reed, A.N.; Vangala, S.; Guo, J. Epsilon-near-zero thin-film metamaterials for wideband near-perfect light absorption. Opt. Mater. Express 2020, 10, 2439–2446. [Google Scholar] [CrossRef]
  43. He, X.; Srinivasan, S.T.; Gupta, S.; Patel, R.M.; Wang, Q.; Davis, L.J.; Forouhar, S. Ultrahigh-power semiconductor lasers and their applications. In Semiconductor Lasers III; SPIE: Bellingham, WA, USA, 1998; Volume 3547, pp. 86–101. [Google Scholar]
  44. Ismach, A.; Druzgalski, C.; Penwell, S.; Schwartzberg, A.; Zheng, M.; Javey, A.; Bokor, J.; Zhang, Y. Direct chemical vapor deposition of graphene on dielectric surfaces. Nano Lett. 2010, 10, 1542–1548. [Google Scholar] [CrossRef]
  45. Zhou, F.; Shan, J.; Cui, L.; Qi, Y.; Hu, J.; Zhang, Y.; Liu, Z. Direct plasma-enhanced-Chemical-Vapor-Deposition syntheses of vertically oriented graphene films on functional insulating substrates for wide-range applications. Adv. Funct. Mater. 2022, 32, 2202026. [Google Scholar] [CrossRef]
  46. Liao, J.; Zhao, Y.; Chen, X.; Hu, Z.; Bu, S.; Zhu, Y.; Lu, Q.; Shang, M.; Wu, H.; Li, F.; et al. Dielectric-assisted transfer using single-crystal antimony oxide for two-dimensional material devices. Nat. Electron. 2025, 8, 309–321. [Google Scholar] [CrossRef]
  47. Song, Y.; Zou, W.; Lu, Q.; Lin, L.; Liu, Z. Graphene transfer: Paving the road for applications of chemical vapor deposition graphene. Small 2021, 17, 2007600. [Google Scholar] [CrossRef]
  48. Zhao, Y.; Song, Y.; Hu, Z.; Wang, W.; Chang, Z.; Zhang, Y.; Lu, Q.; Wu, H.; Liao, J.; Zou, W.; et al. Large-area transfer of two-dimensional materials free of cracks, contamination and wrinkles via controllable conformal contact. Nat. Commun. 2022, 13, 4409. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the aperiodic structure composed of graphene and dielectric.
Figure 1. Schematic diagram of the aperiodic structure composed of graphene and dielectric.
Symmetry 17 01460 g001
Figure 2. The conductivity of graphene in the parameter space composed of the Fermi energy and the frequency. (a) The real part of σ; (b) The imaginary part of σ.
Figure 2. The conductivity of graphene in the parameter space composed of the Fermi energy and the frequency. (a) The real part of σ; (b) The imaginary part of σ.
Symmetry 17 01460 g002
Figure 3. The relative permittivity in the parameter space composed of the Fermi energy and the frequency. (a) The real part of εzz; (b) The imaginary part of εzz; (c) The real part of εxx; (d) The imaginary part of εxx.
Figure 3. The relative permittivity in the parameter space composed of the Fermi energy and the frequency. (a) The real part of εzz; (b) The imaginary part of εzz; (c) The real part of εxx; (d) The imaginary part of εxx.
Symmetry 17 01460 g003
Figure 4. Dispersion curves at different Fermi energies for the frequency f = 45 THz (a), f = 58 THz (b), f = 68 THz (c), and f = 76.89 THz (d). The other parameters are the same as those in Figure 3.
Figure 4. Dispersion curves at different Fermi energies for the frequency f = 45 THz (a), f = 58 THz (b), f = 68 THz (c), and f = 76.89 THz (d). The other parameters are the same as those in Figure 3.
Symmetry 17 01460 g004
Figure 5. Variation of transmittance with the imaginary part of conductivity of graphene at different Fermi energies. (a) EF = 0.4 eV; (b) EF = 0.6 eV; (c) EF = 0.8 eV; (d) EF = 1.0 eV.
Figure 5. Variation of transmittance with the imaginary part of conductivity of graphene at different Fermi energies. (a) EF = 0.4 eV; (b) EF = 0.6 eV; (c) EF = 0.8 eV; (d) EF = 1.0 eV.
Symmetry 17 01460 g005
Figure 6. (a) The reflectance at different Fermi energies; (b) The transmittance at different Fermi energies. The other parameters are set as follows: T = 300 K, τ = 1 ps, td = 20 nm, and Ng = 1.
Figure 6. (a) The reflectance at different Fermi energies; (b) The transmittance at different Fermi energies. The other parameters are set as follows: T = 300 K, τ = 1 ps, td = 20 nm, and Ng = 1.
Symmetry 17 01460 g006
Figure 7. The relative permittivity in the parameter space composed of the thickness of the dielectric and the frequency. (a) The real part of εzz; (b) The imaginary part of εzz; (c) The real part of εxx; (d) The imaginary part of εxx.
Figure 7. The relative permittivity in the parameter space composed of the thickness of the dielectric and the frequency. (a) The real part of εzz; (b) The imaginary part of εzz; (c) The real part of εxx; (d) The imaginary part of εxx.
Symmetry 17 01460 g007
Figure 8. Dispersion curves at different thicknesses of the dielectric for the frequency f = 30 THz (a) and f = 47.71 THz (b). The other parameters are the same as those in Figure 7.
Figure 8. Dispersion curves at different thicknesses of the dielectric for the frequency f = 30 THz (a) and f = 47.71 THz (b). The other parameters are the same as those in Figure 7.
Symmetry 17 01460 g008
Figure 9. The variation of the relative permittivity with frequency under different relaxation times of graphene. (a) The real part of εzz; (b) The imaginary part of εzz; (c) The real part of εxx; (d) The imaginary part of εxx.
Figure 9. The variation of the relative permittivity with frequency under different relaxation times of graphene. (a) The real part of εzz; (b) The imaginary part of εzz; (c) The real part of εxx; (d) The imaginary part of εxx.
Symmetry 17 01460 g009
Figure 10. Dispersion curves at different relaxation times of graphene for the frequency f = 43.38 THz (a) and f = 47.70 THz (b). The other parameters are the same as those in Figure 9.
Figure 10. Dispersion curves at different relaxation times of graphene for the frequency f = 43.38 THz (a) and f = 47.70 THz (b). The other parameters are the same as those in Figure 9.
Symmetry 17 01460 g010
Figure 11. The relative permittivity components varying with the frequency for different numbers of graphene monolayers. (a) The real part of εzz; (b) The imaginary part of εzz; (c) The real part of εxx; (d) The imaginary part of εxx.
Figure 11. The relative permittivity components varying with the frequency for different numbers of graphene monolayers. (a) The real part of εzz; (b) The imaginary part of εzz; (c) The real part of εxx; (d) The imaginary part of εxx.
Symmetry 17 01460 g011
Figure 12. Dispersion curves at different numbers of graphene monolayers for the frequency f = 47.10 THz (a) and f = 47.70 THz (b). The other parameters are the same as those in Figure 11.
Figure 12. Dispersion curves at different numbers of graphene monolayers for the frequency f = 47.10 THz (a) and f = 47.70 THz (b). The other parameters are the same as those in Figure 11.
Symmetry 17 01460 g012
Figure 13. (a) The electric field distribution simulated via the Finite-Difference Time-Domain (FDTD) method in the structure; (b) The electric field distribution simulated via the Transfer matrix method in the structure.
Figure 13. (a) The electric field distribution simulated via the Finite-Difference Time-Domain (FDTD) method in the structure; (b) The electric field distribution simulated via the Transfer matrix method in the structure.
Symmetry 17 01460 g013
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

Zhao, M.; Yang, J.; Zhao, D.; Ni, H.; Chen, X.; Qin, Z.; Yu, Z.; Ao, Y. Quasi-Periodic Hyperbolic Metamaterials Composed of Graphene and Dielectric. Symmetry 2025, 17, 1460. https://doi.org/10.3390/sym17091460

AMA Style

Zhao M, Yang J, Zhao D, Ni H, Chen X, Qin Z, Yu Z, Ao Y. Quasi-Periodic Hyperbolic Metamaterials Composed of Graphene and Dielectric. Symmetry. 2025; 17(9):1460. https://doi.org/10.3390/sym17091460

Chicago/Turabian Style

Zhao, Miaomiao, Junfu Yang, Dong Zhao, Hao Ni, Xiaoling Chen, Zhongli Qin, Zhiyong Yu, and Yingquan Ao. 2025. "Quasi-Periodic Hyperbolic Metamaterials Composed of Graphene and Dielectric" Symmetry 17, no. 9: 1460. https://doi.org/10.3390/sym17091460

APA Style

Zhao, M., Yang, J., Zhao, D., Ni, H., Chen, X., Qin, Z., Yu, Z., & Ao, Y. (2025). Quasi-Periodic Hyperbolic Metamaterials Composed of Graphene and Dielectric. Symmetry, 17(9), 1460. https://doi.org/10.3390/sym17091460

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