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

Super-sensitive tunable planar lens based on graphene hyperbolic metamaterials

Open Access Open Access

Abstract

We theoretically study the topological transition of dispersion types and propose a tunable planar lens based on graphene hyperbolic metamaterials (HMMs). By tuning the chemical potential (μc) of graphene, the dispersion relation of the HMM is topologically switchable between ellipse (μc<0.6 eV) and hyperbola (μc>0.6 eV) where positive and negative refractions occur respectively. Especially, for μc>0.6 eV, a Gaussian light beam is negatively refracted twice and focuses at a far-field point finally, acting well as a planar lens. Furthermore, its focal length l can be sensitively tuned by controlling μc, and Δl reaches 260 μm (from 528 to 268 μm) while μc varies with only 0.05 eV (from 0.65 to 0.7 eV). The physical reason is attributed to the different anisotropy degrees of EFCs for different μc. Such a compact, high-speed, and sensitively tunable planar lens holds great promise in photonic integration, photonic imaging, and directional coupling applications.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Metamaterials have attracted great attention due to their abundant properties and potential applications in both science and technology. Many interesting phenomena, such as slow light [1], negative refraction [2], super-collimation [3,4], and superprism effect [5–7], have been explored in theoretical and experimental works. Hyperbolic metamaterials (HMMs) are highly anisotropic metamaterials whose effective electric and/or magnetic tensors have opposite-sign principal components. This special feature results in its unusual hyperbolic dispersion and lots of unique properties. HMMs are usually constructed with metal nanowire-array [8] or layered metal-dielectric structures [9]. Recently, graphene has attracted great attention due to its excellent optoelectronic properties. Its conductivity is very sensitive to external field, chemical potential, Fermi energy, gate voltage, magnetic field, chemical doping, or optical pump [10–19], so that its optoelectronic properties can be effectively and precisely tuned [20]. Graphene has been suggested as an alternative of metal to construct HMMs to confine light [21], guide surface plasmon polaritons [22], and manipulate wavefronts [23,24]. Obviously, graphene-based metamaterials show advantages over metal-based metamaterials which have unavoidable material loss and cannot be tuned after fabrication [25]. Recently, graphene metamaterials have been experimentally realized in the mid-infrared range [15]. A diversity of unusual applications, such as subwavelength imaging, super-lenses, hyperbolic waveguides, and slow light devices [21,26–28], have been proposed.

In recent years, tunable photonic devices are highly desired because they play more and more important roles in integrated photonic/optical circuits. Among them, tunable metamaterial lens is an elementary device and has attracted great attention. Unlike conventional lens design dealing with the shape of a uniform material to satisfy a special performance requirement, a metamaterial lens focuses on structure design and arranging the refractive index profiles. So far, two typical methods have been proposed to realize such lens. The first method is to change the geometrical structure mechanically such as the epsilon-near-zero lens [29] and the MEMS-based metasurface lens [30,31]. This method has the weaknesses of complex structure, slow speed, as well as low stability [32]. The second method is to modify the refractive index by using electro-optic [33] or thermo-optic effect [34]. It seems that this method is more promising because its response is much quicker. However, in order to obtain index change large enough for generating available difference of dispersion properties, it requires some extreme conditions (e.g., very strong external electric field or very high temperature), which significantly reduce the application promise of such metamaterial lens. What is more, even under such extreme conditions, it is still difficult to obtain a large focal-length tuning range. Therefore, it is highly desirable to develop new mechanisms to achieve a large focal-length tuning range of lens without requirements of extreme conditions.

In this paper, we propose to construct a tunable graphene-based HMM and study its switchable topological dispersion properties as well as its lens application. It is found that when the chemical potential (μc) of graphene sheets varies across the critical value of 0.6 eV, the dispersion relation of the HMM switches between ellipse and hyperbola, leading to transition between positive and negative refractions accordingly. Especially, by using the negative refraction effect when μc>0.6 eV, we design a compact planar HMM lens which can make a Gaussian beam redirect twice and focus at a far-field point in air finally. Moreover, its focal length can be sensitively tuned by controlling μc, and its focal-length tuning range as large as 260 μm can be easily achieved which requires only 0.05 eV change of μc from 0.65 to 0.7 eV. Additionally, the resolution of this lens is also discussed.

2. Model of Graphene HMM

Figure 1(a) shows the designed HMM structure consisting of alternative graphene sheets and dielectric layers with εd = 2.2 along the z-axis. The thickness of the dielectric layer is d = 100 nm. Graphene is a very thin material layer with a thickness as small as one atom, and it possesses extremely high carrier mobility of 15000 cm2/(V·s) which theoretically makes graphene-based devices with high response speed in the order of picosecond possible. Owing to negligible thickness of graphene sheet as compared with the dielectric layer, the period of HMM can be regarded the same as d. In this study, the working frequency is f = 30 THz (i.e., λ = 10 μm). Because of λ>>d = 100 nm, the designed structure can be modeled as an anisotropic effective medium by effective medium theory [35]. A Gaussian light beam is normally incident on the structure and propagates along the x-axis. The x-z view of the structure is presented in Fig. 1(b).

 figure: Fig. 1

Fig. 1 (a) Schematic diagram of the HMM consisting of periodic graphene/dielectric layers. (b) The x-z view of the designed structure.

Download Full Size | PDF

In our model, graphene is characterized by the following surface conductivity σ(f,μc,Γ,Τ) model according to the well-known Kubo formula [10],

σ(f,μc,Γ,Τ)=kBΤe2j2π2(πf+Γj)(μckBΤ+2ln(eμckBΤ+1))+2e2j(πf+Γj)π20fd(E)fd(E)4(πf+Γj)24(E)2dE
Where fd(E) = {exp[(E-μc)/(kBT)] + 1}−1 is the Fermi-Dirac distribution with the energy E. f is the working frequency and e is the charge of the electron. Τ = 300 K is the temperature. kB and ћ are Boltzmann’s constant and reduced Planck’s constant, respectively. Γ is the phenomenological scattering rate (which is assumed to be 0.1 meV) and independent of the energy E. μc is the chemical potential of graphene sheets.

To realize the tunable planar lens to be discussed later, the chemical potentials μc of all graphene sheets needed to be changed simultaneously. One effective way to do this is using external electrostatic field biasing Ebias by electrical gating. The relation between Ebias and μcis deduced as [10],

Ebias=2eπ2vF2ε0εd[(kBT)2μc/kBTμc/kBTxex+1dx+kBTμcln(eμc/kBT+1)+kBTμcln(eμc/kBT+1)]
Where vF9.5×105m/s is the Fermi velocity, ε0 is the vacuum permittivity. The other parameters are the same as those in Eq. (1).

According to effective medium theory [35,36], the effective relative permittivity tensor εeff of the HMM is described as,

εeff=[ε000ε000ε]
Where the subscript || and ⊥ indicate parallel and vertical components to the graphene sheets, respectively [37]. The specific representations of ε|| and ε are shown as follows,

{ε=εdjσ(f,μc,Γ,Τ)2πfε0dε=εd

Here ε is a positive real number because of εd = 2.2>0. As for ε||, Fig. 2(a) clearly shows that Re(ε||) varies from positive to negative values as μc increases from 0 to 1 eV within a wide frequency range; while Fig. 2(b) indicates that Im(ε||) is near zero under conditions of μc≥0.2 eV and f>27 THz. Specifically, for the working frequency f = 30 THz, the curves of Re(ε||)-μc and Im(ε||)-μc are shown in Fig. 2(c). We can see that for 0<μc<0.2 eV, both of Re(ε||) and Im(ε||) decrease nonlinearly as μc increases. Obviously, the existence of non-zero Im(ε||) originated from losses of graphene layers is unfavorable for the design of tunable planar lens to be discussed later. But for μc≥0.2 eV, the Re(ε||)-μc curve is quite linear and it changes from positive to negative values at the critical value of μc = 0.6 eV where Re(ε||) = 0, while Im(ε||) keeps to be zero. This property is quite useful for the design of super-sensitive planar HMM lens later. Therefore, we will choose μc≥0.2eV in our later study unless there are additional specifications, and use ε|| instead of Re(ε||) because of the near-zero Im(ε||) for simplicity.

 figure: Fig. 2

Fig. 2 (a, b) Re(ε||) and Im(ε||) as functions of f and μc. (c) Re(ε||)-μc and Im(ε||)-μc curves at f = 30 THz.

Download Full Size | PDF

3. Switchable topological dispersion properties of HMM

In this work, the dispersion of the HMM is computed after the homogenization of the HMM has been performed by using effective medium theory. Considering the symmetry of the 1D periodic structure shown in Fig. 1(a), the dispersion relation after homogenization for extraordinary waves (i.e. TM waves) can be written as [37],

kx2ε+kz2ε=k02
Where kx and kz are x and z components of the wave vector respectively, and k0 is the wave number in vacuum determined by k0=2πfμ0ε0 with the vacuum permeability μ0 and the vacuum permittivity ε0. For f = 30 THz, Fig. 2(c) shows that ε|| is positive (or negative) when μc is smaller (or larger) than the critical value of 0.6 eV, resulting in an elliptic (or hyperbolic) topological type of dispersion relation. This property provides an effective way to realize the topological transition of dispersion types and makes a tunable graphene-HMM-based planar lens possible.

For better understanding the transmission behaviors of electromagnetic (EM) waves, equi-frequency contour (EFC) analysis is employed to predict the refraction behaviors at an interface. A EFC is a curve consisting of all the points with the same frequency in k-space. As is well known, the general Snell’s law n1sinθ1 = n2sinθ2 (n1 and n2 are refractive indices, while θ1 and θ2 are the incident and refractive angles) is only applicable to the case of isotropic medium. As for a periodic structure with anisotropic dispersion relation, it is necessary to develop it to be the extended Snell’s law by the following three steps: (1) The tangential components of the incident and refractive wave vectors (ki and kr) parallel to the interface should be conserved (i.e., ki|| = kr||), because of |k0|n1sinθ1 = |k0|n2sinθ2 (k0 is the wave vector in vacuum), i.e., kisinθ1 = krsinθ2. (2) The propagation direction of EM wave at certain point of a EFC is along its gradient direction, because the energy velocity in a lossless periodic structure is the same as the group velocity vector defined as vg = ▽kω(k). (3) According to the energy conservation law, the direction of the group velocity should point to the direction away from the source, so that the angle between the direction of the incident wave and that of the group velocity is acute. As a result, we are able to qualitatively determine the propagation direction of the refraction light at an interface.

The transmission behaviors for two topological types of EFCs of the graphene HMM are further studied. Equation (5) shows that as ε|| varies from positive to negative values, the EFC diagram changes from an ellipse to a hyperbola, which are clearly shown by the yellow EFC curves in Figs. 3(a1) and 3(b1). For the case of μc = 0.2 eV in Fig. 3(a1), ε|| is 1.58, which together with ε = 2.2 makes the EFC to be an ellipse according to Eq. (5). Such an elliptic EFC shows slightly anisotropy. Based on the extended Snell’s law, the direction of energy velocity in HMM deviates a little from that of phase velocity, i.e., the energy direction S2 is non-parallel to the wave vector k2 inside the HMM. The FDTD simulations in Fig. 3(a2) clearly shows that a Gaussian light beam incident from region 1 (i.e. air) undergoes two times of positive refractions at interfaces 1/2 and 2/3, and finally diverges to propagate in air region 3, which agrees well with the EFC analysis in Fig. 3(a1). While for the case of μc = 0.65 eV (ε|| = −0.15), the results are very different. In Fig. 3(b1), the EFC of HMM becomes a hyperbola with strong anisotropy, so that S2 inside HMM deviates dramatically from k2. According to the extended Snell’s law, negative refraction will occur at the interface between HMM and air. Since a Gaussian light beam always contains a range of wave vectors but not a single one, negative refraction occurs for each wave vector at the air-HMM interface, leading to a redirected light beam to focus at a focal point. As shown in Fig. 3(b2), due to the strong anisotropy of the HMM, the Gaussian beam first focuses at a focal point quite near the left air-HMM interface inside the HMM, then propagates with quite large divergent angle, and finally passes through the right HMM-air interface to focus again at a far-field point in air region 3. These results agree quite well with the predictions in Fig. 3(b1). In other words, a planar lens can be realized provided that the HMM possesses hyperbolic EFCs under the condition of μc>0.6 eV. It should be noticed that from the viewpoint of applications, only the second focal point in air region 3 is the key performance parameter to characterize the planar HMM lens, so that we will further investigate the properties of such focal point in the following text.

 figure: Fig. 3

Fig. 3 EFC diagrams and Finite-Difference Time-Domain (FDTD) simulation results at f = 30 THz. (a1), (a2), (a3) and (b1), (b2), (b3) are for μc = 0.2 and 0.65 eV, respectively. ki and Si represent the directions of wave vector and energy velocity in region i (i = 1, 2, 3). The HMM is located between x = 50 and 150 μm [See the enlargements of region 2 in Figs. 3(a3) and 3(b3)]. A Gaussian beam with a waist width of 10 μm is located at x = 10 μm.

Download Full Size | PDF

4. Discussions of tunable sensitive HMM lens

From the aforementioned analyses, it is found that the shape of the hyperbolic EFC described by Eq. (5) sensitively depends on μc. Here we take four representative μc to quantitatively study the influence of the chemical potential on the HMM lens. The EFC diagrams and FDTD simulation results are shown in Fig. 4. It is noticed that all of the considered μc here are larger than 0.6 eV where ε|| = 0, leading to negative ε||, i.e., ε|| = −0.15, −0.53, −0.9, and −1.28 for μc = 0.65, 0.75, 0.85, and 0.95 eV, respectively. Hence all of their corresponding EFCs are hyperbolic according to Eq. (5). Due to the negative refraction behaviors occurring at 1/2 and 2/3 interfaces, a Gaussian light beam finally focuses in region 3 to form a focal point, as shown in Fig. 4. Though the incident angle θ1 keeps the same, the energy direction of EM wave inside the HMM (i.e. S2) varies sensitively as μc increases, resulting in a sensitive change of focal lengths of the planar HMM lens in region 3. These theoretical analyses are well verified by the FDTD simulations shown in Figs. 4(a2)-4(d2) whose focal lengths are 528, 170, 91, and 52 μm, respectively.

 figure: Fig. 4

Fig. 4 EFC diagrams (a1-d1) and the corresponding |E| distributions (a2-d2) for μc = 0.65, 0.75, 0.85, and 0.95 eV, respectively. The focal points are denoted by the red dashed lines. The HMM region 2 (denoted by two white dashed lines) locates between x = 50 and 150 μm.

Download Full Size | PDF

Next, we will discuss the sensitivity of the focal length of the HMM lens. Figure 4 shows that S1 and S3 always keep unchanged whatever μc is, whereas S2 is sensitive to μc. Therefore, the variation of focal length mainly depends on the direction of S2. In a word, one can tune the focal length of the HMM lens by changing μc conveniently. Figure 5(a) presents more details about the relationship between the focal length l and the chemical potential μc. With increasing μc, the focal length l decreases dramatically at first (when μc<0.75 eV) and then slowly (when μc>0.75 eV). For examples, for the case that μc is near the critical point μc = 0.6 eV, the decrement of the focal length is as large as Δl = 260 μm (from 528 to 268 μm, about 26 times of the incident wavelength of 10 μm) when μc increases with only 0.05 eV (from 0.65 to 0.7 eV); while for the case that μc is much large than 0.6 eV, Δl is only 8 μm (from 52 to 44 μm) when μc increases with 0.05 eV (from 0.95 to 1 eV). The physical reason is attributed to the different anisotropy degrees of EFCs for different μc. The variation of anisotropy degree for 0.65<μc<0.7 eV is much larger than that for 0.95<μc<1 eV. This sensitive dependence characteristics of focal length on chemical potential provides an effective means to realize tunable compact planar HMM lens.

 figure: Fig. 5

Fig. 5 (a) The relationship between the focal length l of the HMM lens and the chemical potential μc. (b) Dependence of transmittance on μc. The inset shows the enlargement of the red rectangle. (c) The |E| profile along the y direction at focal points in region 3 for μc = 0.65, 0.75, 0.85, and 0.95 eV respectively.

Download Full Size | PDF

We further study the influence of μc on the transmittance of the HMM lens. Transmittance is calculated by the ratio of the power flowing through the 2/3 interface to the power of incident Gaussian light beam. As shown in Fig. 5(b), for 0.2≤μc≤0.54eV or 0.64≤ μc≤1 eV, the HMM lens can act as a concave or convex lens, and the transmittance maintains high (>92%) and stable, meaning that the tunable HMM lens can work well within a wide chemical potential range. However, for 0.54<μc<0.64 eV, there exists a minimum transmittance close to zero at the critical point μc = 0.6 eV where ε|| = 0. In this special case of μc = 0.6 eV, the hyperbolic EFC degenerates to a horizontal line parallel to the kx-axis, causing the corresponding S2 to be along the y-axis. In other words, all of the EM waves inside the HMM propagates upwards and downwards, causing no energy to propagate rightwards, thus the transmittance is almost zero. As μc deviates from 0.6 eV, the transmittance increases rapidly. No matter the EFC is elliptical or hyperbolic, more and more energy transmits rightwards, leading to higher and higher transmittance. It should be noticed that as μc increases from 0 to 0.2 eV, the transmittance increases quickly from 70% to 92% and then keeps stable, as shown by the inset in Fig. 5(b). By further referring to the Im(ε||)-μc curve in Fig. 2(c) where Im(ε||) characterizing the losses of graphene decreases to near-zero with increasing μc from 0 to 0.2 eV, we can conclude that the transmittance increases quickly to be stable when the losses of graphene reduces to near-zero. That is to say, the losses of graphene will reduce the transmittance of the structure, which is unfavorable for the design of tunable HMM lens. Thus we choose μc≥0.2 eV where the losses of graphene can be neglected for our planar lens design.

We also investigate the resolution of the HMM lens. Figure 5(c) shows the |E| profiles along the y direction at focal points in region 3 for four different μc. The resolution is characterized by the full width half maximum (FWHM) of the curve. The FWHMs are 25.6, 20, 19, and 18.8 μm for μc = 0.65, 0.75, 0.85, and 0.95 eV, respectively. Obviously, the resolution decreases quickly at first and then slowly as μc increases from 0.65 to 0.95 eV. This means that higher μc is beneficial for enhancing the resolution of the lens, but the resolution will gradually reach a saturation value after μc = 0.85 eV.

5. Conclusions

To summarize, we have constructed a tunable graphene-based HMM and studied its switchable topological dispersion properties and lens application. There exists a critical value of chemical potential (μc = 0.6 eV) leading to ε|| = 0, and the EFC can be topologically switched between ellipse (μc<0.6 eV) and hyperbola (μc>0.6 eV) conveniently. Accordingly, positive or negative refraction can be realized for elliptic or hyperbolic EFCs respectively. Based on these properties, we design a compact tunable planar HMM lens. For μc>0.6 eV, FDTD simulations results show that a Gaussian light beam is negatively refracted two times and focuses at a far-field point in air region finally, which agrees well with the theoretical predictions of EFC analysis. Furthermore, the influence of μc on the focal length l is studied and the results show that the focal length varies sensitively for μc<0.75 eV and smoothly for μc>0.75 eV. The physical reason is attributed to the different anisotropy degrees of EFCs for different μc. These results have potential in photonic integration, photonic imaging, and directional coupling applications.

Funding

National Natural Science Foundation of China (NSFC) (11504114, 11434017); Science and Technology Program of Guangzhou (201904010105); National Key R&D Program of China (2018YFA 0306200); Guangdong Innovative and Entrepreneurial Research Team Program (2016ZT06C594); Natural Science Foundation of Guangdong Province, 2019; Fundamental Research Funds for the Central Universities, (x2wl/D2191420); Teaching and Research Reform Project of SCUT (x2wl/Y1190281); Student Research Program of SCUT (x2wl/C9192092).

References

1. A. W. Zeng, M. X. Gao, and B. Guo, “Slow light in a hyperbolic metamaterial waveguide cladded with arbitrary nonlinear dielectric materials,” Appl. Phys. B 124(7), 146 (2018). [CrossRef]  

2. S. S. Islam, M. R. I. Faruque, M. T. Islam, and M. T. Ali, “A new wideband negative refractive index metamaterial for dual-band operation,” Appl. Phys., A Mater. Sci. Process. 123(4), 252 (2017). [CrossRef]  

3. V. Purlys, L. Maigyte, D. Gailevicius, M. Peckus, R. Gadonas, and K. Staliunas, “Super-collimation by axisymmetric photonic crystals,” Appl. Phys. Lett. 104(22), 221108 (2014). [CrossRef]  

4. X. Lin, X. Zhang, L. Chen, M. Soljačić, and X. Jiang, “Super-collimation with high frequency sensitivity in 2D photonic crystals induced by saddle-type van Hove singularities,” Opt. Express 21(25), 30140–30147 (2013). [CrossRef]   [PubMed]  

5. S. Pahlavan and V. Ahmadi, “Novel optical demultiplexer design using oblique boundary in hetero photonic crystals,” IEEE Photonic. Tech. L. 29(6), 511–514 (2017). [CrossRef]  

6. W. Li, X. Zhang, X. Lin, and X. Jiang, “Enhanced wavelength sensitivity of the self-collimation superprism effect in photonic crystals via slow light,” Opt. Lett. 39(15), 4486–4489 (2014). [CrossRef]   [PubMed]  

7. B. Gao, Z. Shi, and R. W. Boyd, “Design of flat-band superprism structures for on-chip spectroscopy,” Opt. Express 23(5), 6491–6496 (2015). [CrossRef]   [PubMed]  

8. T. A. Morgado, S. I. Maslovski, and M. G. Silveirinha, “Ultrahigh casimir interaction torque in nanowire systems,” Opt. Express 21(12), 14943–14955 (2013). [CrossRef]   [PubMed]  

9. M. Kim, S. Kim, and S. Kim, “Optical bistability based on hyperbolic metamaterials,” Opt. Express 26(9), 11620–11632 (2018). [CrossRef]   [PubMed]  

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

11. A. Vakil and N. Engheta, “Transformation optics using graphene,” Science 332(6035), 1291–1294 (2011). [CrossRef]   [PubMed]  

12. I. V. Iorsh, I. S. Mukhin, I. V. Shadrivov, P. A. Belov, and Y. S. Kivshar, “Hyperbolic metamaterials based on multilayer graphene structures,” Phys. Rev. B Condens. Matter Mater. Phys. 87(7), 075416 (2012). [CrossRef]  

13. H. Deng, F. Ye, B. A. Malomed, X. Chen, and N. C. Panoiu, “Optically and electrically tunable Dirac points and Zitterbewegung in graphene-based photonic superlattices,” Phys. Rev. B Condens. Matter Mater. Phys. 91(20), 201402 (2015). [CrossRef]  

14. H. Deng, X. Chen, B. A. Malomed, N. C. Panoiu, and F. Ye, “Tunability and robustness of Dirac points of photonic nanostructures,” IEEE J. Sel. Top. Quantum Electron. 22(5), 5000509 (2016). [CrossRef]  

15. Y. C. Chang, C. H. Liu, C. H. Liu, S. Zhang, S. R. Marder, E. E. Narimanov, Z. Zhong, and T. B. Norris, “Realization of mid-infrared graphene hyperbolic metamaterials,” Nat. Commun. 7(1), 10568 (2016). [CrossRef]   [PubMed]  

16. T. Gric and O. Hess, “Tunable surface waves at the interface separating different graphene-dielectric composite hyperbolic metamaterials,” Opt. Express 25(10), 11466–11476 (2017). [CrossRef]   [PubMed]  

17. T. Guo, L. Zhu, P. Y. Chen, and C. Argyropoulos, “Tunable terahertz amplification based on photoexcited active graphene hyperbolic metamaterials,” Opt. Mater. Express 8(12), 3941–3952 (2018). [CrossRef]  

18. P. Y. Chen and J. Jung, “PT Symmetry and Singularity-Enhanced Sensing Based on Photoexcited Graphene Metasurfaces,” Phys. Rev. Appl. 5(6), 064018 (2016). [CrossRef]  

19. M. Sakhdari, M. Farhat, and P. Y. Chen, “PT-symmetric metasurfaces: wave manipulation and sensing using singular points,” New J. Phys. 19(6), 065002 (2018). [CrossRef]  

20. B. Sensale-Rodriguez, R. Yan, M. M. Kelly, T. Fang, K. Tahy, W. S. Hwang, D. Jena, L. Liu, and H. G. Xing, “Broadband graphene terahertz modulators enabled by intraband transitions,” Nat. Commun. 3(1), 780 (2012). [CrossRef]   [PubMed]  

21. A. Tyszka-Zawadzka, B. Janaszek, and P. Szczepański, “Tunable slow light in graphene-based hyperbolic metamaterial waveguide operating in SCLU telecom bands,” Opt. Express 25(7), 7263–7272 (2017). [CrossRef]   [PubMed]  

22. H. Lu, X. Gan, D. Mao, and J. Zhao, “Graphene-supported manipulation of surface plasmon polaritons in metallic nanowaveguides,” Photon. Res. 5(3), 162–167 (2017). [CrossRef]  

23. C. Wang, W. Liu, Z. Li, H. Cheng, Z. Li, S. Chen, and J. Tian, “Dynamically tunable deep subwavelength high-order anomalous reflection using graphene metasurfaces,” Adv. Opt. Mater. 6(3), 1701047 (2018). [CrossRef]  

24. H. Zhao, Z. Chen, F. Su, G. Ren, F. Liu, and J. Yao, “Terahertz wavefront manipulating by double-layer graphene ribbons metasurface,” Opt. Commun. 402, 523–526 (2017). [CrossRef]  

25. L. Menon, W. T. Lu, A. L. Friedman, S. P. Bennett, D. Heiman, and S. Sridhar, “Negative index metamaterials based on metal-dielectric nanocomposites for imaging applications,” Appl. Phys. Lett. 93(12), 123117 (2008). [CrossRef]  

26. T. Zhang, L. Chen, and X. Li, “Graphene-based tunable broadband hyperlens for far-field subdiffraction imaging at mid-infrared frequencies,” Opt. Express 21(18), 20888–20899 (2013). [CrossRef]   [PubMed]  

27. P. Wang, C. Tang, Z. Yan, Q. Wang, F. Liu, J. Chen, Z. Xu, and C. Sui, “Graphene-based superlens for subwavelength optical imaging by graphene plasmon resonances,” Plasmonics 11(2), 515–522 (2016). [CrossRef]  

28. H. Xu, L. Wu, X. Dai, Y. Gao, and Y. Xiang, “Tunable infrared plasmonic waveguides using graphene based hyperbolic metamaterials,” Optik (Stuttg.) 127(20), 9640–9646 (2016). [CrossRef]  

29. M. N. Cia, M. Beruete, M. Sorolla, and N. Engheta, “Lensing system and Fourier transformation using epsilon-near-zero metamaterials,” Phys. Rev. B Condens. Matter Mater. Phys. 86(16), 165130 (2012). [CrossRef]  

30. T. Roy, S. Zhang, I. L. W. Jung, M. Troccoli, F. Capasso, and D. Lopez, “Dynamic metasurface lens based on MEMS Technology,” Apl Photonics 3(2), 021302 (2018). [CrossRef]  

31. E. Arbabi, A. Arbabi, S. M. Kamali, Y. Horie, M. Faraji-Dana, and A. Faraon, “MEMS-tunable dielectric metasurface lens,” Nat. Commun. 9(1), 812 (2018). [CrossRef]   [PubMed]  

32. X. Shang, A. M. Trinidad, P. Joshi, J. D. Smet, D. Cuypers, and H. D. Smet, “Tunable optical beam deflection via liquid crystal gradient refractive index generated by highly resistive polymer film,” IEEE Photonics J. 8(3), 6500411 (2016). [CrossRef]  

33. D. Su, X. Y. Zhang, Y. L. Ma, F. Shan, J. Y. Wu, X. C. Fu, L. J. Zhang, K. Q. Yuan, and T. Zhang, “Real-time electro-optical tunable hyperlens under subwavelength scale,” IEEE Photonics J. 10(1), 4600109 (2018). [CrossRef]  

34. M. Tinker and J.-B. Lee, “Thermal and optical simulation of a photonic crystal light modulator based on the thermo-optic shift of the cut-off frequency,” Opt. Express 13(18), 7174–7188 (2005). [CrossRef]   [PubMed]  

35. Y. Xiang, X. Dai, J. Guo, H. Zhang, S. Wen, and D. Tang, “Critical coupling with graphene-based hyperbolic metamaterials,” Sci. Rep. 4(1), 5483 (2014). [CrossRef]   [PubMed]  

36. T. Gric and O. Hess, “Controlling hybrid-polarization surface plasmon polaritons in dielectric-transparent conducting oxides metamaterials via their effective properties,” J. Appl. Phys. 122(19), 193105 (2017). [CrossRef]  

37. Z. Li, W. Liang, and W. Chen, “Switchable hyperbolic metamaterials based on the graphene-dielectric stacking structure and optical switches design,” A Letters Journal Exploring the Frontiers of Physics 120(3), 37001 (2017).

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 (a) Schematic diagram of the HMM consisting of periodic graphene/dielectric layers. (b) The x-z view of the designed structure.
Fig. 2
Fig. 2 (a, b) Re(ε||) and Im(ε||) as functions of f and μc. (c) Re(ε||)-μc and Im(ε||)-μc curves at f = 30 THz.
Fig. 3
Fig. 3 EFC diagrams and Finite-Difference Time-Domain (FDTD) simulation results at f = 30 THz. (a1), (a2), (a3) and (b1), (b2), (b3) are for μc = 0.2 and 0.65 eV, respectively. k i and S i represent the directions of wave vector and energy velocity in region i (i = 1, 2, 3). The HMM is located between x = 50 and 150 μm [See the enlargements of region 2 in Figs. 3(a3) and 3(b3)]. A Gaussian beam with a waist width of 10 μm is located at x = 10 μm.
Fig. 4
Fig. 4 EFC diagrams (a1-d1) and the corresponding |E| distributions (a2-d2) for μc = 0.65, 0.75, 0.85, and 0.95 eV, respectively. The focal points are denoted by the red dashed lines. The HMM region 2 (denoted by two white dashed lines) locates between x = 50 and 150 μm.
Fig. 5
Fig. 5 (a) The relationship between the focal length l of the HMM lens and the chemical potential μc. (b) Dependence of transmittance on μc. The inset shows the enlargement of the red rectangle. (c) The |E| profile along the y direction at focal points in region 3 for μc = 0.65, 0.75, 0.85, and 0.95 eV respectively.

Equations (5)

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

σ ( f , μ c , Γ , Τ ) = k B Τ e 2 j 2 π 2 ( π f + Γ j ) ( μ c k B Τ + 2 ln ( e μ c k B Τ + 1 ) ) + 2 e 2 j ( π f + Γ j ) π 2 0 f d ( E ) f d ( E ) 4 ( π f + Γ j ) 2 4 ( E ) 2 d E
E b i a s = 2 e π 2 v F 2 ε 0 ε d [ ( k B T ) 2 μ c / k B T μ c / k B T x e x + 1 d x + k B T μ c ln ( e μ c / k B T + 1 ) + k B T μ c ln ( e μ c / k B T + 1 ) ]
ε e f f = [ ε 0 0 0 ε 0 0 0 ε ]
{ ε = ε d j σ ( f , μ c , Γ , Τ ) 2 π f ε 0 d ε = ε d
k x 2 ε + k z 2 ε = k 0 2
Select as filters


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