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

High-efficiency tunable circular asymmetric transmission using dielectric metasurface integrated with graphene sheet

Open Access Open Access

Abstract

High-efficiency tunable asymmetric transmission (AT) based on simple-constructed metasurface is highly desired for next generation optical polarization devices. Here, we numerically investigate high-efficiency and frequency-tunable circular AT effect in mid-infrared region by combining simple-shape silicon array with a graphene sheet. The asymmetric parameter of the dielectric nanostrip structure reaches 0.92 at 12.68 THz and the width of tunable spectra (AT > 0.7) is 1100 nm, which represent a major advance compared with previously reported AT. The AT behavior originates from extrinsic chirality induced by oblique illumination, and the high AT efficiency results from the constructive and deconstructive interferences of selectively excited electric and magnetic resonances. In addition, the working waveband of AT is shifted by dynamically modulating graphene’s Fermi energy, which offers a new degree of freedom to achieve multifunctions without refabricating structures. The proposed array system possessing the merits of high efficiency, simple inclusions and frequency-tunability has significant potentials for practical applications in polarization devices such as polarization sensor, polarizer, etc.

© 2017 Optical Society of America

1. Introduction

Asymmetric transmission phenomenon, like unidirectional conduction of diode, was discovered in planar chiral metamaterial first [1], and it is found that AT shows a circular conversion dichroism effect, i.e. different cross-polarization transmission coefficients. This AT behavior is quite distinct from 3D chiral response associated with different co-polarization transmission coefficients [2, 3]. Subsequently, AT has quickly become a topic of research interest as its promising to be used in polarization devices, such as rotators, polarization transformers, diode-like devices, polarizers and isolators [4, 5]. The effort to create artificial metamaterials with AT was focused on different types of planar metamaterials [6–8], but these earlier proposed single-layered chiral metamaterials seriously suffer from low efficiency. Later, helical metamaterials [9, 10], achiral twisted bi-layer [11–13], and multi-layered structures [14–18] have been proposed to enhance AT efficiency. In spite of gaining higher asymmetric parameters, most reported structures composed of complex metallic components face the challenges of large absorptive losses and costly fabrication. However, dielectric materials without intrinsic Ohmic losses could further enhance AT efficiency and simultaneously simplify the shapes of unit cells due to their Mie-type electric dipole resonance (EDR) and magnetic dipole resonance (MDR) [19]. Moreover, spatial symmetry breaking caused by oblique illumination will also extremely simplify the shape of nanoparticles compared with the one in intrinsic chiral metamaterials [20, 21]. Therefore, simple-shape dielectric metasurface based on extrinsic chirality becomes a good candidate for high-efficiency AT.

On the other hand, due to the demand of tunability in laser beam steering, axial scan focusing and tunable polarization control, some tunable devices based on flexible substrates [22], phase-changed materials [23], the control of incident angles [24], and graphene [25, 26] have been proposed. Among these tunable devices, only a few frequency-tunable AT effects have been reported with asymmetric parameters no more than 0.25 [25, 26]. In this paper, we achieve high-efficiency (0.92) and dynamically tunable AT in the mid-infrared (MIR) region by incorporating simple-shape dielectric array with a graphene sheet. The high AT efficiency, i.e. one near-unity cross-polarization transmission and the other near-zero cross-polarization transmission, originates from the difference in spectral superposition of EDRs and MDRs. Such a propagation direction-dependent polarization sensitive transmission effect results from extrinsic chirality induced by oblique illumination. In addition, the working waveband of AT is shifted by electrically controlling graphene’s Fermi energy without reoptimizing geometric parameters of structures. The high-efficiency and frequency-tunable AT effect based on simple-shape array is first reported, which provides a novel paradigm for polarization-sensitive devices, sensing, and other potential applications.

2. Design and simulations

To gain high-efficiency tunable AT with simple-shape metasurface, we first excite extrinsic chirality in symmetric array using oblique incident CPL. As a geometrical definition, chirality can be obtained from achiral structures simply by tilting their symmetry axis out of incident plane to yield the three dimensional asymmetry, which is so-called extrinsic chirality [3, 27]. To satisfy the extrinsically chiral excitation criteria, we break the mirror symmetry and C2 rotational symmetry of strip array by considering wavevector k direction of propagation neither parallel nor perpendicular to mirror planes or rotational axes. As shown in Fig. 1(a), our array is located in x-y plane and incident light propagating in the direction with θ representing the angle of k off -z axis and with the angle φ off x axis. In this way, the polarization components of obliquely incident CPL are not along x or y direction, instead, the polarization fields in x and y directions are contributed by the projections of these two polarization components. The incident angles θ=±45°,φ=±45° assures the maximum projections of incident electric field in x and y directions simultaneously, which means the largest symmetry breaking. If the normally incident CPL is decomposed into two polarization components along x and y directions with 90 degree phase delay, the transmission coefficients of x- and y-polarized incident light (tx and ty) are related to the electric and magnetic surface polarizabilities [28],

tx=1(k0/2)2αexαmy1+(k0/2)2αexαmyik0(αmyαex)/2,ty=1(k0/2)2αmxαey1+(k0/2)2αmxαey+ik0(αeyαmx)/2.
where k0 is the wavenumber in free space, the transmission of x- and y-polarized light can be controlled independently by constructing the surface electric and magnetic polarizabilities (αex, αey, αmx and αmy). In principle, both unit and zero circular polarization transmission can be achieved by constructing the asymmetric structure with different phase delay for CPL. For obliquely incident CPL, if the two orthogonal polarization components are regarded as normally incident light after coordinate transformation, the strip array becomes asymmetric, which means there are more freedom to control polarization transmissions. Thus we optimize microstructure by CST Microwave Studio as shown in Fig. 2(b), where a = 9 μm and b = 4.05 μm. The period of the silicon array is p = 14 μm in x and y directions, and the thicknesses of strips and substrate are 7.1 μm and 0.5 μm, respectively. Refractive index of silicon is 3.45 in MIR and substrate is assumed to be FR-4 (n = 2.04) [29, 30]. The simulation is based on the frequency-domain finite integration technology, using a unit cell boundary coupled with Floquet ports. Meanwhile, a Finite Difference Time Domain (FDTD) method software, FDTD Solutions from Lumerical, Inc. was also used to monitor field distributions.

 figure: Fig. 1

Fig. 1 (a) The scheme of optimize silicon nanostrip array placed on FR-4 substrate. (b) The top view of a unit cell, where a = 9 μm, b = 4.05 μm and p = 14 μm in both x and y directions.

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 (a), (b) Transmission coefficients of CPL with θ=±45°φ=±45°incidences, t,t+,t+,t++ represent left-to-left, left-to-right, right-to-left, right-to-right transmitted circularly polarized light, respectively. (c), (d) Transmission spectra with θ=±45°,φ=±45°incidences, (e), (f) Asymmetry parameters spectra withθ=±45°,φ=±45° incidences.

Download Full Size | PDF

Next, we combine graphene sheet with silicon array to gain the frequency-tunability. Graphene, an atom-thick sheet with two-dimensionally arranged carbon atoms, is regarded as an anisotropic material with different conductivities in- and out-of-plane [31, 32]. The in-plane conductivity is calculated using Random Phase Approximation [33, 34],

σ(ω)=2e2ωTπiω+iτ1log[2cosh(ωF2ωT)]+e24[H(ω2)+i2ωπ0H(ω'2)H(ω2)ω2ω'2dω']
where H (ω) = sinh (ω/ωT)/[cosh (ωFT) + cosh (ω/ωT)], ωF=EF/,ωT=κBT/, e andκBare the charge of an electron and Boltzmann constant, and =h/2π represents the reduced Planck constant. According to Eq. (2), graphene’s in-plane conductivity depends on the angular frequency of incident light ω, temperature T, Fermi energyEF and relaxation time τ. In our simulation, we assume T = 300 K, τ = 1 ps and Fermi energyEFranges from 0 eV to 0.6 eV, which requires about 100 V bias voltage in experiment. Graphene is modeled as an anisotropic material with out-of-plane (z direction) dielectric constant 2.25 and in-plane (x-y plane) permittivity derived from ε(ω) = 1 + iσ (ω)/ωε0t [35], where ε0 represents the permittivity in the vacuum, t = 1 nm is the thickness of graphene.

3. Results and discussion

Figure 2 shows the transmission coefficient spectra of silicon array under the excitations of obliquely incident CPL, where “→” and “←” denote incident angles θ=45°,φ=45°andθ=45°,φ=45°, respectively. Right-to-left polarization transmission coefficient (t+) is much higher than the left-to-right one (t+) near 12.68 THz, while the left-to-right transmission coefficient (t+) is higher than the right-to-left one (t+) near 14.32 THz in Fig. 2(a). Co-polarization transmission coefficients spectra (t,t++) marked by pink and black lines are also different especially near 14.32 THz, which denotes the existence of circular dichroism (CD). The appearance of CD due to symmetry breaking brings more applications (discussed below). The indexes (“+” and “”) of transmission efficient in Fig. 2(b) are exchanged each other compared with Fig. 2(a), this is because the change of incident angle induces the exchange of two enantiomeric arrangements.

For the transmitted intensity spectra of T=|t|2+|t+|2 and T+=|t++|2+|t+|2, two obvious peaks and valleys are observed at 12.68 THz and 14.32 THz in Fig. 2(c). The transmission of RCP light reaches 0.93 at 12.68 THz, which originates from artificially constructed Huygens surface characterized by simultaneous excitations of strong EDR and MDR. The electric field components of strong EDR and MDR with orthogonal electric and magnetic polarizabilities are in the same direction, thus the constructive superposition of electric field components from two strong resonances and incident wave leads to a near-unity RCP transmission. While the transmission of LCP light is near zero at the same spectral position because of the deconstructive superposition of electric field components from two weak resonances and the incident wave. Differently, LCP transmission is much higher than RCP transmission at 14.32 THz owing to the variation in excitations of resonance modes, which means opposite polarization-state selectivity in the other waveband. In addition, the spectra of LCP transmission backwards (θ=45°,φ=45°) in Fig. 2(d) is equal to that of RCP transmission forwards (θ=45°,φ=45°) in Fig. 2(d). This is because AT phenomenon originates from the fact that an incident CPL see different handedness for two opposite propagation directions.

To estimate AT efficiency, we calculate the spectra of asymmetry parameters defined as Δcir+=T+T=(|t++|2+|t+|2)(|t|2+|t+|2)=Δcir [7] in Fig. 2(e). The maximum asymmetry parameter of RCP light reaches 0.92 at 12.68 THz, which represents a significant improvement in efficiency and simple-shape configuration. For the case of opposite-direction incidence, the sign of asymmetry parameter for both LCP and RCP are changed upon reversal of propagation direction. As expected, the asymmetry parameter of RCP with the incident angle of θ=45°,φ=45° in Fig. 2(f) equals to that of LCP one with θ=45°,φ=45°incidence in Fig. 2(e). The sensitivity to incident direction at dual bands means two kinds of selectivities of polarization state at each waveband by controlling incident directions.

In order to verify AT phenomenon visually, we plot the distributions of electric field magnitude in the middle of nanoparticles at y-z and x-z planes under the excitations of RCP light at 12.68 THz. Figures 3(a) and 3(b) show the electric field distributions under RCP light incidence backwards (θ=45°,φ=45°). The field below and above periodic array (marked by black line frame) are homogeneous, which means RCP light passes through freely. However, there are obvious differences in field distributions below and above the black line marked metasurface when RCP light illuminates forwards (θ=45°,φ=45°), which confirms that RCP light cannot propagate along the opposite direction as shown in Figs. 3(c) and 3(d).

 figure: Fig. 3

Fig. 3 (a), (b), (c), (d) The distributions for electric field magnitude driven by RCP light backwards and forwards, respectively (by FDTD). (a'),(b'),(c'),(d') The field vectors distributions of a unit cell by CST, the black dashed line frame marks a strip in one unit cell.

Download Full Size | PDF

To analyze the contributions of different resonance modes to the giant AT effect, we give out the electric field vector distributions of a unit cell as magnified red dashed line frame marked. The electric field vectors flow from top-left to bottom-right as shown in Fig. 3(a') mimicking the radiation pattern of electric dipole in far field, which means strong EDR dominates in y-z plane. The EDR contributes to the electric polarizability components in y and -z directions. While the distribution of electric field vectors in Fig. 3(b') shows the cross section of a ring-like electric current in x-z plane, which denotes strong MDR with magnetic polarizability components in z and y directions. The components of electric polarizabilities in y and -z direction and the components of magnetic polarizabilities in z and y direction constitute a vectorial Huygens metasurface, which is characterized by high transmission [36, 37]. In other words, the simultaneous excitations of two-set components of EDR and MDR lead to near-unity RCP transmission in Fig. 2(c). Because the electric fields scattered by y-direction electric polarizability and z-direction magnetic polarizability are in the same direction (y direction), similarly, the electric fields scattered by y-direction electric polarizability and z-direction magnetic polarizability are in the same direction (-z direction) [37]. Therefore, the amplitude of the vectorial sum of two-direction electric fields and incident field is near-unity. However, the field vector distribution in Fig. 3(c') characterized by ring-like electric current in y-z plane denotes a weak MDR mode with small magnetic polarizability in x direction. The distribution of field vectors in top-right region outside the unit structure in Fig. 3(d') represent a weak electric resonance with a small electric polarizability in z direction. The electric field components from incident wave and two weak resonances are counteracted leading to near-zero RCP transmission in Fig. 2(d). Noted the selectively excited resonance modes of RCP forwards described in Figs. 3(c') and 3(d') also represents the resonance modes of LCP excitations backwards, which is also responsible for the near-zero LCP transmission in Fig. 2(c). Generally, the constructive and deconstructive superpositions of electric field components from incident wave and different resonance modes result in a significant difference in the transmission of RCP and LCP waves.

After discussing the giant AT effect, we continue to analyze the frequency-tunability by integrating graphene sheet to the former silicon array. As shown in Fig. 4(a), we place graphene sheet on substrate and fabricate nanostrip array on graphene. Asymmetry parameter spectra of this structure are plotted in Fig. 4(b), which exhibits a clear blue-shift when Fermi energy ranges from 0 eV to 0.6 eV. The asymmetry parameter above 0.7 in the tunable spectra ranges from 12.52 THz to 13.13 THz as dash area marked, which makes great progress in efficiency compared with the tunable AT in [3, 27]. The large blue shift (1100 nm) results from the changes in graphene’s conductivity induced by the rises of Fermi energies. Different conductivities bring different disturbed fields. These fields affect EDRs and MDRs at various degree when the dielectric microstructures are fabricated close to graphene sheet. Concretely, the changes in resonant frequency can be interpreted using perturbation theory as [38, 39]

Δωω0=VdV[(ΔμH0)H0*+(ΔεE0)E0*]VdV(μH0H0*+εE0E0*)=ΔWm+ΔWeWm+We
where ω0 is resonant angular frequency, E0and H0 are unperturbed electric and magnetic fields whose complex conjugates are represented by E0 and H0. Moreover, Δε and Δμ are the changes in dielectric permittivity and magnetic permeability regarded as the material perturbation, We andWm are the total electric and magnetic energies. The value of ∆ω relies on not only the material perturbation caused by grapheneΔε, but also the local enhancement of optical fields generated by dielectric arrayE0. The change in resonant frequency (∆ω) is large when a perturbation encounter with strong EDR and MDR simultaneously, because the enhanced fields will make great contribution in frequency-tunability as described in Eq. (3). Thus, we obtain frequency-tunability of giant AT by dynamically adjusting graphene’s Fermi energy.

 figure: Fig. 4

Fig. 4 (a) A diagram of silicon-graphene hybrid metasurface with frequency-tunable AT effects, (b) Solid and dashed lines represent the ATs of RCP and LCP light with different Fermi energies.

Download Full Size | PDF

4. Application in polarization devices: sensor and circular polarizer

The proposed AT microstructure possessing the merits of high efficiency, simple-shape inclusions and frequency-tunability provide a novel paradigm to achieve tunable polarization devices, such as polarization sensor and circular polarizer. Figure 5(a) shows polarization transmission spectra of silicon array from 12.33 THz to 14.50 THz, which is defined asTin+=(|t++|2+|t+|2)and Tin=(|t+|2+|t|2). Such a metasurface with giant AT can be used to distinguish the polarization of incident waves as a polarization sensor. In addition, CD (different t and t++) together with AT (different t+ and t+) makes Tin± different fromTout, which leads to another application, a circular polarizer. Figure 5(b) gives out the polarization transmission spectra of Tout+=(|t++|2+|t+-|2)and Tout=(|t+|2+|t|2), whose definition of transmission is different from the polarization transmission in Fig. 5(a). So the transmission peak at 14.23 THz (0.93) becomes higher than that at 12.75 THz, which means the polarizer can gain LCP or RCP light efficiently at both wavebands by controlling incident angles. Furthermore, the frequency-tunability based on graphene offers our polarization sensor and circular polarizer the capacity to serve as dynamically tunable polarization devices.

 figure: Fig. 5

Fig. 5 (a) Transmission spectra for sensing polarization state of incident light as a polarization sensor, (b) Transmission spectra for generating desired CPL as a circular polarizer.

Download Full Size | PDF

5. Summary and Conclusion

In conclusion, a high-efficiency and frequency-tunable AT based on simple-shape inclusions operating in MIR is demonstrated numerically in this paper. The AT effect originates from extrinsic chirality induced by oblique illumination in a symmetric array system, which not only provides flexible way to overcome complex chiral structures, but also shows even stronger AT than the intrinsic one. Specifically, our achiral strip array passes 93% RCP/LCP light and nearly blocks all LCP/RCP light with the incident angle of θ=±45°,φ=±45°at 12.68 THz. In addition, the working wavebands of AT can be shifted by adjusting graphene’s Fermi energy, which offers a certain metasurface more functions without refabricating structures. In general, the realization of tunable circular AT effect in MIR makes great progress in efficiency and simple-shape construction, which facilitates many significant applications in polarization devices.

References and links

1. V. A. Fedotov, P. L. Mladyonov, S. L. Prosvirnin, A. V. Rogacheva, Y. Chen, and N. I. Zheludev, “Asymmetric propagation of electromagnetic waves through a planar chiral structure,” Phys. Rev. Lett. 97(16), 167401 (2006). [CrossRef]   [PubMed]  

2. L. Wu, Z. Yang, Y. Cheng, Z. Lu, P. Zhang, M. Zhao, R. Gong, X. Yuan, Y. Zheng, and J. Duan, “Electromagnetic manifestation of chirality in layer-by-layer chiral metamaterials,” Opt. Express 21(5), 5239–5246 (2013). [CrossRef]   [PubMed]  

3. E. Plum, V. Fedotov, and N. Zheludev, “Extrinsic electromagnetic chirality in metamaterials,” J. Opt. A, Pure Appl. Opt. 11(7), 074009 (2009). [CrossRef]  

4. D. Leung, B. Rahman, and K. Grattan, “Numerical analysis of asymmetric silicon nanowire waveguide as compact polarization rotator,” IEEE Photonics J. 3(3), 381–389 (2011). [CrossRef]  

5. G. Kenanakis, A. Xomalis, A. Selimis, M. Vamvakaki, M. Farsari, M. Kafesaki, C. M. Soukoulis, and E. N. Economou, “Three-dimensional infrared metamaterial with asymmetric transmission,” ACS Photonics 2(2), 287–294 (2015). [CrossRef]  

6. E. Plum, V. Fedotov, and N. Zheludev, “Planar metamaterial with transmission and reflection that depend on the direction of incidence,” Appl. Phys. Lett. 94(13), 131901 (2009). [CrossRef]  

7. R. Singh, E. Plum, C. Menzel, C. Rockstuhl, A. Azad, R. Cheville, F. Lederer, W. Zhang, and N. Zheludev, “Terahertz metamaterial with asymmetric transmission,” Phys. Rev. B 80(15), 153104 (2009). [CrossRef]  

8. Z. Li, M. Gokkavas, and E. Ozbay, “Manipulation of asymmetric transmission in planar chiral nanostructures by anisotropic loss,” Adv. Opt. Mater. 1(7), 482–488 (2013). [CrossRef]  

9. W. Fan, Y. Wang, R. Zheng, D. Liu, and J. Shi, “Broadband high efficiency asymmetric transmission of achiral metamaterials,” Opt. Express 23(15), 19535–19541 (2015). [CrossRef]   [PubMed]  

10. J. Kaschke, L. Blume, L. Wu, M. Thiel, K. Bade, Z. Yang, and M. Wegener, “A helical metamaterial for broadband circular polarization conversion,” Adv. Opt. Mater. 3(10), 1411–1417 (2015). [CrossRef]  

11. J. Shi, X. Liu, S. Yu, T. Lv, Z. Zhu, H. Feng Ma, and T. Jun Cui, “Dual-band asymmetric transmission of linear polarization in bilayered chiral metamaterial,” Appl. Phys. Lett. 102(19), 191905 (2013). [CrossRef]  

12. D. Liu, Z. Xiao, X. Ma, Q. Ma, X. Xu, and Z. Wang, “Asymmetric transmission of chiral metamaterial slab with double L resonators,” Opt. Commun. 338, 359–365 (2015). [CrossRef]  

13. Y. H. Wang, J. Shao, J. Li, Z. Liu, J. Li, Z.-G. Dong, and Y. Zhai, “Broadband high-efficiency transmission asymmetry by a chiral bilayer bar metastructure,” J. Appl. Phys. 117(17), 173102 (2015). [CrossRef]  

14. C. Pfeiffer, C. Zhang, V. Ray, L. J. Guo, and A. Grbic, “High performance bianisotropic metasurfaces: asymmetric transmission of light,” Phys. Rev. Lett. 113(2), 023902 (2014). [CrossRef]   [PubMed]  

15. Z. Wei, Y. Cao, Y. Fan, X. Yu, and H. Li, “Broadband polarization transformation via enhanced asymmetric transmission through arrays of twisted complementary split-ring resonators,” Appl. Phys. Lett. 99(22), 221907 (2011). [CrossRef]  

16. J. Liu, Z. Li, W. Liu, H. Cheng, S. Chen, and J. Tian, “High-efficiency mutual dual-band asymmetric transmission of circularly polarized waves with few-layer anisotropic metasurfaces,” Adv. Opt. Mater. 4(12), 2028–2034 (2016). [CrossRef]  

17. H. Ma and R.-X. Wu, “Asymmetric transmission of linearly polarized wave in all-dielectric chiral metamaterial,” in Progress in Electromagnetic Research Symposium (PIERS) (IEEE, 2016), pp. 658–661.

18. B. Tang, Z. Li, Z. Liu, F. Callewaert, and K. Aydin, “Broadband asymmetric light transmission through tapered metallic gratings at visible frequencies,” Sci. Rep. 6(1), 39166 (2016). [CrossRef]   [PubMed]  

19. T. Cao, C.-W. Wei, L.-B. Mao, and S. Wang, “Tuning of giant 2D-chiroptical response using achiral metasurface integrated with graphene,” Opt. Express 23(14), 18620–18629 (2015). [CrossRef]   [PubMed]  

20. L. Zhu, J. Kapraun, J. Ferrara, and C. J. Chang-Hasnain, “Flexible photonic metastructures for tunable coloration,” Optica 2(3), 255–258 (2015). [CrossRef]  

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

22. M. Chen, W. Sun, J. Cai, L. Chang, and X. Xiao, “Frequency-tunable mid-infrared cross polarization converters based on graphene metasurface,” Plasmonics 12(3), 1–7 (2016).

23. J. Ding, B. Arigong, H. Ren, J. Shao, M. Zhou, Y. Lin, and H. Zhang, “Mid-infrared tunable dual-frequency cross polarization converters using graphene-based L-shaped nanoslot array,” Plasmonics 10(2), 351–356 (2015). [CrossRef]  

24. I. De Leon, M. J. Horton, S. A. Schulz, J. Upham, P. Banzer, and R. W. Boyd, “Strong, spectrally-tunable chirality in diffractive metasurfaces,” Sci. Rep. 5(1), 13034 (2015). [CrossRef]   [PubMed]  

25. J. Zhao, J. Zhang, Z. Zhu, X. Yuan, and S. Qin, “Tunable asymmetric transmission of THz wave through a graphene chiral metasurface,” J. Opt. 18(9), 095001 (2016). [CrossRef]  

26. Z. Li, W. Liu, H. Cheng, S. Chen, and J. Tian, “Tunable dual-band asymmetric transmission for circularly polarized waves with graphene planar chiral metasurfaces,” Opt. Lett. 41(13), 3142–3145 (2016). [CrossRef]   [PubMed]  

27. S. S. Kruk, A. N. Poddubny, D. A. Powell, C. Helgert, M. Decker, T. Pertsch, D. N. Neshev, and Y. S. Kivshar, “Polarization properties of optical metasurfaces of different symmetries,” Phys. Rev. B 91(19), 195401 (2015). [CrossRef]  

28. C. L. Holloway, M. A. Mohamed, E. F. Kuester, and A. Dienstfrey, “Reflection and transmission properties of a metafilm: With an application to a controllable surface composed of resonant particles,” IEEE Trans. Electromagnetic Compatibility 47(4), 853–865 (2005). [CrossRef]  

29. S. Thongrattanasiri, F. H. Koppens, and F. J. G. De Abajo, “Complete optical absorption in periodically patterned graphene,” Phys. Rev. Lett. 108(4), 047401 (2012). [CrossRef]   [PubMed]  

30. J. Ding, B. Arigong, H. Ren, J. Shao, M. Zhou, Y. Lin, and H. Zhang, “Mid-infrared tunable dual-frequency cross polarization converters using graphene-based L-shaped nanoslot array,” Plasmonics 10(2), 351–356 (2015). [CrossRef]  

31. V. Kravets, A. Grigorenko, R. Nair, P. Blake, K. Novoselov, and A. Geim, “Spectroscopic ellipsometry of graphene and an exciton-shifted van Hove peak in absorption,” Phys. Rev. B 81(15), 155413 (2010). [CrossRef]  

32. F. Lin, Z. Zhu, X. Zhou, W. Qiu, C. Niu, J. Hu, K. Dahal, Y. Wang, Z. Zhao, Z. Ren, D. Litvinov, Z. Liu, Z. M. Wang, and J. Bao, “Orientation control of graphene flakes by magnetic field: broad device applications of macroscopically aligned graphene,” Adv. Mater. 29(1), 1604453 (2017). [PubMed]  

33. N. K. Emani, A. V. Kildishev, V. M. Shalaev, and A. Boltasseva, “Graphene: a dynamic platform for electrical control of plasmonic resonance,” Nanophotonics 4(1), 214–223 (2015). [CrossRef]  

34. L. Falkovsky and S. Pershoguba, “Optical far-infrared properties of a graphene monolayer and multilayer,” Phys. Rev. B 76(15), 153410 (2007). [CrossRef]  

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

36. M. Decker, I. Staude, M. Falkner, J. Dominguez, D. N. Neshev, I. Brener, T. Pertsch, and Y. S. Kivshar, “High-efficiency dielectric Huygens’ surfaces,” Adv. Opt. Mater. 3(6), 813–820 (2015). [CrossRef]  

37. H. Jiang, W. Zhao, and Y. Jiang, “All-dielectric circular polarizer with nearly unit transmission efficiency based on cascaded tensor Huygens surface,” Opt. Express 24(16), 17738–17745 (2016). [CrossRef]   [PubMed]  

38. J. A. Kong, Electromagnetic Wave Theory (John Wiley and Sons, 1986)

39. A. Raman and S. Fan, “Perturbation theory for plasmonic modulation and sensing,” Phys. Rev. B 83(20), 205131 (2011). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 (a) The scheme of optimize silicon nanostrip array placed on FR-4 substrate. (b) The top view of a unit cell, where a = 9 μm, b = 4.05 μm and p = 14 μm in both x and y directions.
Fig. 2
Fig. 2 (a), (b) Transmission coefficients of CPL with θ=±45°φ=±45°incidences, t , t + , t + , t ++ represent left-to-left, left-to-right, right-to-left, right-to-right transmitted circularly polarized light, respectively. (c), (d) Transmission spectra with θ=±45°, φ=±45°incidences, (e), (f) Asymmetry parameters spectra with θ=±45°, φ=±45° incidences.
Fig. 3
Fig. 3 (a), (b), (c), (d) The distributions for electric field magnitude driven by RCP light backwards and forwards, respectively (by FDTD). (a'), (b'), (c'), (d') The field vectors distributions of a unit cell by CST, the black dashed line frame marks a strip in one unit cell.
Fig. 4
Fig. 4 (a) A diagram of silicon-graphene hybrid metasurface with frequency-tunable AT effects, (b) Solid and dashed lines represent the ATs of RCP and LCP light with different Fermi energies.
Fig. 5
Fig. 5 (a) Transmission spectra for sensing polarization state of incident light as a polarization sensor, (b) Transmission spectra for generating desired CPL as a circular polarizer.

Equations (3)

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

t x = 1 ( k 0 /2) 2 α ex α my 1+ ( k 0 /2) 2 α ex α my i k 0 ( α my α ex )/2 , t y = 1 ( k 0 /2) 2 α mx α ey 1+ ( k 0 /2) 2 α mx α ey +i k 0 ( α ey α mx )/2 .
σ(ω)= 2 e 2 ω T π i ω+i τ 1 log[ 2cosh( ω F 2 ω T ) ]+ e 2 4 [ H( ω 2 )+i 2ω π 0 H( ω' 2 )H( ω 2 ) ω 2 ω ' 2 dω' ]
Δω ω 0 = V dV[(Δ μ H 0 ) H 0 * +(Δ ε E 0 ) E 0 * ] V dV (μ H 0 H 0 * +ε E 0 E 0 * ) = Δ W m +Δ W e W m + W e
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.