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

Active metasurface modulator with electro-optic polymer using bimodal plasmonic resonance

Open Access Open Access

Abstract

Electrically tunable metasurfaces have gained special interest as they can realize ultrathin surface-normal modulators in planar geometries. In this paper, we demonstrate a novel metasurface modulator based on electro-optic (EO) polymer that utilizes bimodal resonance inside a metallic subwavelength grating to increase the modulation efficiency. When two metal-insulator-metal (MIM) resonant modes are excited simultaneously inside the grating, they couple strongly to generate a sharp dip in the reflected spectrum. As a result, efficient intensity modulation with 15-dB extinction ratio can be obtained at the resonant wavelength under a small refractive index change of 8.5 × 10−3, corresponding to modulation voltage of less than 10 V. Due to the low parasitic capacitance of EO polymer and high conductivity of metallic gratings which is also used as the electrodes, the RC bandwidth of the device should easily exceed 100 GHz, potentially applicable to high-speed surface-normal modulators.

© 2017 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Metasurfaces are two-dimensional metamaterials composed of dense arrays of metallic or dielectric nanostructures placed on planar substrates [1,2]. By configuring each element to operate as an efficient optical nano-antenna or resonator, the intensity and/or phase of incident light can be manipulated at a subwavelength scale [3–5]. With the recent development of nano-fabrication technologies, a variety of passive or devices, such as lenses, polarizers, absorbers, and filters, have been demonstrated [6–10]. Due to the inherent merits of ultrathin flat structure, flexibility in design, and ease of fabrication using planar lithographic tools, these metasurface-based devices are becoming promising to replace bulky optical components.

As the next logical step, there has been a growing interest to realize active metasurfaces that can be tuned electrically [11–21]. Several groups have demonstrated electro-optic tuning by embedding graphene [15,16], transparent conducting oxide (TCO) [17–20], and electro-optic (EO) polymer [21] as active materials inside metallic metasurfaces. In these demonstrations, metallic structures were preferred over dielectric counterparts, as they could also serve as highly conductive electrodes to potentially realize high-speed modulation. On the other hand, inherent plasmonic losses in metals have been the major obstacle in achieving high-quality (high-Q) resonance at near-infrared or shorter wavelengths. As a result, previous demonstrations were often limited to operation in mid-infrared wavelengths [15–18,20] or with insufficient efficiency [21].

In this work, we propose a novel EO-polymer-based metallic metasurface modulator operating at 1.55-μm wavelength, that utilizes bimodal resonance. Among various active materials, we select EO polymer for several reasons; it exhibits large EO coefficient over 200 pm/V [22–25], ultrahigh-speed modulation capability beyond 40 GHz [25,26], simple and low-cost spin-coating-based fabrication that enables flexible device design [25–28], and high material reliability proved in commercial products [26]. Unlike previous demonstrations based on similar schemes [17–21], we judiciously design the structure, so that two metal-insulator-metal (MIM) resonant modes are excited inside the EO polymer. By utilizing the strong coupling between these two modes, we demonstrate that a sharp dip emerges in the reflected spectrum, having significantly higher Q factor than that obtained when a single-mode Fabry-Perot (FP) resonance is employed. As a result, we numerically demonstrate a thin (<1 μm) surface-normal intensity modulator with 15-dB extinction ratio under a refractive index change of 8.5 × 10−3, corresponding to the modulation voltage of less than 10 V. Due to the highly conductive metallic electrodes and low parasitic capacitance, we estimate that the RC bandwidth should easily exceed 100 GHz.

2. Proposed structure and principle

The schematic of the proposed metasurface modulator is depicted in Fig. 1. The EO polymer with thickness of d is sandwiched between the bottom Au layer and the top thin Au grating layer. The grating period Λ is designed to be shorter than the wavelength of light, so that the diffraction is prohibited. The thickness (200 nm in this work) of the top Au layer is designed to be larger than the skin depth to eliminate direct coupling from the incident light into EO polymer. The bottom Au layer is also set to the same thickness, so that it operates as an almost perfect reflective mirror.

 figure: Fig. 1

Fig. 1 Schematic (a), cross-section (b), and (c) modulation principle of the proposed surface-normal metasurface modulator.

Download Full Size | PDF

As a plane-wave light with the electric field polarized perpendicularly to the grating (along x in Fig. 1) is normally incident to the device, it excites MIM waveguide modes propagating in ± x direction inside the Au/EO-polymer/Au layers [see Fig. 1(b)]. When these MIM waveguide modes satisfy the FP condition, the light is confined strongly inside the EO polymer and eventually absorbed by the metal. As a result, we obtain a resonant dip in the reflected spectrum as shown schematically in Fig. 1(c). It has been demonstrated that such resonance generally has relatively low Q factor due to the metallic losses of MIM modes [17–21]. We will, however, demonstrate in the next section that the Q factor can be increased significantly by judiciously designing the structure, so that two MIM FP modes are excited simultaneously and couple strongly.

By applying the poling voltage between the top and bottom Au layers, the EO polymer is poled vertically along z direction to induce electro-optic coefficient r13, r23 and r33 in x, y, and z directions, respectively [23]. Under a realistic poling voltage, r31 and r32 is approximately 1/3 of r33 [29–31]. When modulation voltage is applied using the same electrodes, we can, through Pockels effect, change the refractive index of the EO polymer and shift the resonant wavelength λ0, so that the intensity of reflected light is modulated at λ0, as shown in Fig. 1(c). The change in refractive index can be expressed as [29,30]

(ΔnxΔnyΔnz)=12(nx3r13ny3r23nz3r33)Emod.
Here, ni (i = x, y, z) are the i-diagonal components of the refractive index tensor of the EO polymer without modulation, and Emod ( = Vmod/d, where Vmod is applied voltage and d is EO polymer thickness) is the applied electric field along z.

The amount of wavelength shift ∆λ required to obtain desired extinction is inversely proportional to the Q factor of the resonant dip. Since ∆λ is proportional to the amount of refractive index change, we see from Eq. (1) that the required ∆λ is proportional to Vmod/d. Therefore, the required Vmod to achieve a given extinction ratio is proportional to d/Q. We thus define the figure-of-merit (FOM) of our modulator as Q/d; in order to reduce the modulation voltage, we need to optimize the structure to maximize Q/d.

3. Numerical results and discussion

We first simulate the reflectance of the proposed structure by the finite-difference time-domain (FDTD) method. For simplicity, we assume the width of the grating in y direction to be infinity, so that two-dimensional (2D) simulation is performed. A periodic boundary condition is employed in x direction while absorptive condition with perfect matched layers (PMLs) is applied at z boundaries. The dielectric permittivity of EO polymer is represented as a tensor, where the refractive indices (nx, ny and nz) are set to be 1.65 in the absence of modulation. The complex permittivity of Au is modeled by fitting Drude-Lorentzian poles to the experimental data [32].

3.1 Reflectance property

Figure 2 shows the calculated reflectance of the device with d = 500 nm, plotted as a function of grating period Λ and wavelength λ. For convenience, the grating gap g ( = Λ - a) is fixed to 400 nm. From Fig. 2, we see that a number of resonance valleys emerge along well-defined curves. From the electric field profiles shown in Fig. 2 insets, we presume that these curves correspond to the FP resonance conditions for the TM0 and TM1 MIM modes, respectively. To confirm this speculation, we plot in Fig. 3(a) the dispersion curves for the TM modes of Au/EO-polymer/Au MIM waveguide with d = 500 nm. Indeed, we can see that in the wavelength range from 0.9 to 1.8 µm, two propagating modes (TM0 and TM1) co-exist inside the MIM waveguide. By using the propagation constant of each mode, we can then derive the FP resonance conditions for respective modes as

βa+φ=mπ.
where β is the propagation constant of each mode as plotted in Fig. 3(a), and φ is the phase shift of the mode as it is reflected at the boundary between the EO polymer and the gap [see Fig. 1(b)]. While m is generally any positive integer, it needs to be an even number for the case considered here, due to the symmetric device structure and normal incident condition.

 figure: Fig. 2

Fig. 2 Reflectance as a function of grating period Λ and wavelength. Inset shows Re(Ey) distribution at two curves.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 (a) Dispersion relationship of Au/EO-polymer/Au MIM waveguide (b) Calculated Fabry-Perot resonance of MIM modes overlaid on the reflectance plot. d = 500 nm in two plots.

Download Full Size | PDF

The calculated FP resonance conditions are overlaid by white solid curves in Fig. 3(b). Here, TMn,m denotes the FP resonance condition for the TMn waveguide modes (n = 0, 1, 2,…) with index m ( = 2, 4, …) in x direction as defined in Eq. (2). We approximate φ to be a wavelength-independent constant, and fit to 1.53 rad (TM0,m) and 4.14 rad (TM1,m) to have best agreement with the FDTD results. While slight mismatch arises due to the fact that φ actually depends on wavelength, reasonable agreement is observed between the FDTD results and analytically derived FP resonance conditions.

More importantly, we notice from Fig. 3(b) that a sharp resonant dip emerges at Λ ∼1 μm and wavelength of around 1.5 μm, where TM0,2 and TM1,2 FP curves approach. Such feature is present at the anti-crossing point of a strongly coupled bimodal resonant system in general [33], and has been demonstrated in various other systems, such as high-contrast gratings (HCG) [8, 34, 35] and coupled ring resonators [36]. In our case, two MIM modes satisfy the resonant condition simultaneously and couple strongly at this anti-crossing point. Due to the destructive interference between these two modes, external coupling to the outgoing plane wave is suppressed, leading to increase in Q factor [33]. We can therefore realize high-Q resonance at the wavelength of our interest, by deliberately changing the EO polymer thickness d and the period Λ.

3.2 Intensity modulation property

As discussed in Section 2, it is essential to maximize the FOM, defined as Q/d, to reduce the modulation voltage. In order to find the optimal EO polymer thickness d, we plot in Fig. 4 the Q factor of TM0,2 FP resonance in the 1550-nm wavelength range as well as Q/d as a function of d. For different EO polymer thickness, we fine-tune the grating period Λ within the range from 1.00 to 1.04 μm, so that the TM0,2 FP mode exits inside the wavelength range between 1530 and 1570 nm. From Fig. 4, we see that Q factor increases periodically as TM0,2 resonance intersects with TMn,2 (n = 1, 2, …). For example, the Q factor increases to 87 to exhibits a strong peak at d = 500 nm, corresponds to the anti-crossing point of the TM0,2 and TM1,2, that we have discussed in Fig. 3(b). Additional peaks emerge as we increase d to 1030 and 1600 nm, which correspond to the anti-crossing of TM2,2 and TM3,2 modes, respectively. Due to the smaller loss of the MIM modes with increasing d, these peaks have higher Q factor. However, if we compare Q/d to evaluate the overall modulation efficiency, we see that a maximum Q/d of 0.174 nm−1 is obtained at d = 500 nm. We assess this Q/d value to be more than 5 times improvement from the simulated result presented in [21].

 figure: Fig. 4

Fig. 4 Calculated Q factor and figure of merit (Q/d) with different EO polymer thickness.

Download Full Size | PDF

From Fig. 4, we can conclude that the optimal EO polymer thickness to obtain largest Q/d is 500 nm. In this case, other parameters of the modulator are determined as Λ = 1 µm and a = 600 nm. Using this optimized structure, we calculate the change in reflection spectrum under modulation. Figure 5(a) shows the reflectance spectrum for various values of ∆nz. Note that ∆nz represents the refractive index change for the Ez component under modulation, and we assume ∆nx = ∆ny = ∆nz/3 as explained in Eq. (1). From Fig. 5(a), we see that the resonant wavelength shifts continuously from 1536 nm to 1546 nm as we change ∆nz from −8 × 10−3 to 8 × 10−3. As a result, at a signal wavelength of 1541 nm, the intensity of reflected light can be modulated as a function of ∆nz as shown in Fig. 5(b). For example, we can achieve 15-dB intensity modulation with 5-dB insertion loss by ∆nz of 8.5 × 10−3. Assuming the r33 coefficient of 200 pm/V [22–25], the required driving voltage in this case is calculated from Eq. (1) as Vmod = ± 4.7 V.

 figure: Fig. 5

Fig. 5 (a) Reflectance spectrum with ∆nz varied from −8 × 10−3 to 8 × 10−3, (b) Modulation characteristic of reflected light at wavelength of 1541 nm.

Download Full Size | PDF

Since the change of refractive index of EO polymer originates from electronic polarization, having an inherent response in the order of picoseconds [25,26], the bandwidth of the modulator should be limited by the RC delay of the electrodes. Unlike the previous active metasurfaces based on metal-oxide-semiconductor (MOS) capacitor [17–20], relatively thick EO polymer (d = 500 nm) with small radio-frequency (RF) permittivity (ε = 2.7) of our device results in ultra-small capacitance of 48 aF/µm2. Combined with the high conductivity (4.2 × 107 S⋅m−1) of the Au grating used as the electrode, the RC bandwidth is estimated to be easily over 100 GHz for a realistic case with the grating length l of 500 µm.

4. Conclusion

We have proposed and numerically demonstrated a novel ultrathin (< 1 μm) surface-normal EO-polymer-based metallic metasurface modulator operating at 1.55-µm wavelength. Significant increase in modulation efficiency was achieved by exploiting the strong coupling between two MIM modes inside the grating, which generated a sharp resonance at the anti-crossing point of their dispersion curves. An optimized structure with 500-nm-thick EO polymer and 1-μm-pitch Au grating demonstrated 15-dB intensity modulation under a refractive index change of 8.5 × 10−3, corresponding to the modulation voltage of ± 4.7 V. Owing to inherently high-speed response of EO polymer as well as small RC constant due to high conductivity of Au and small capacitance of EO polymer, this modulator can potentially operate at bandwidth exceeding 100 GHz. With relatively easy fabrication and scalability to two-dimensional arrays, the proposed device should be attractive for wide ranges of applications, such as high-density optical interconnects, free-space optical communication, and high-speed imaging.

Funding

Japan Science and Technology Agency (JST) PRESTO.

Acknowledgments

The author Jiaqi Zhang is supported by a scholarship under the State Scholarship Fund of China Scholarship Council.

References and links

1. N. Yu and F. Capasso, “Flat optics with designer metasurfaces,” Nat. Mater. 13(2), 139–150 (2014). [PubMed]  

2. N. I. Zheludev and Y. S. Kivshar, “From metamaterials to metadevices,” Nat. Mater. 11(11), 917–924 (2012). [PubMed]  

3. N. Yu, P. Genevet, M. A. Kats, F. Aieta, J. P. Tetienne, F. Capasso, and Z. Gaburro, “Light propagation with phase discontinuities: generalized laws of reflection and refraction,” Science 334(6054), 333–337 (2011). [PubMed]  

4. A. V. Kildishev, A. Boltasseva, and V. M. Shalaev, “Planar photonics with metasurfaces,” Science 339(6125), 1232009 (2013). [PubMed]  

5. Y. Zhao and A. Alu, “Manipulating light polarization with ultrathin plasmonic metasurfaces,” Phys.Rev. B 84 (20), 205428 (2011).

6. F. Aieta, P. Genevet, M. A. Kats, N. Yu, R. Blanchard, Z. Gaburro, and F. Capasso, “Aberration-free ultrathin flat lenses and axicons at telecom wavelengths based on plasmonic metasurfaces,” Nano Lett. 12(9), 4932–4936 (2012). [PubMed]  

7. E. Hasman, V. Kleiner, G. Biener, and A. Niv, “Polarization dependent focusing lens by use of quantized Pancharatnam–Berry phase diffractive optics,” Appl. Phys. Lett. 82(3), 328–330 (2003).

8. C. J. Chang-Hasnain, “High-contrast gratings as a new platform for integrated optoelectronics,” Semicond. Sci. Technol. 26 (2011), 014043 (2011).

9. L. Verslegers, P. B. Catrysse, Z. Yu, J. S. White, E. S. Barnard, M. L. Brongersma, and S. Fan, “Planar lenses based on nanoscale slit arrays in a metallic film,” Nano Lett. 9(1), 235–238 (2009). [PubMed]  

10. K. Aydin, V. E. Ferry, R. M. Briggs, and H. A. Atwater, “Broadband polarization-independent resonant light absorption using ultrathin plasmonic super absorbers,” Nat. Commun. 2(517), 517 (2011). [PubMed]  

11. N. I. Zheludev and Y. S. Kivshar, “From metamaterials to metadevices,” Nat. Mater. 11(11), 917–924 (2012). [PubMed]  

12. X. Su, C. Ouyang, N. Xu, W. Cao, X. Wei, G. Song, J. Gu, Z. Tian, J. F. O’Hara, J. Han, and W. Zhang, “Active metasurface terahertz deflector with phase discontinuities,” Opt. Express 23(21), 27152–27158 (2015). [PubMed]  

13. M. Yoo and S. Lim, “Active metasurface for controlling reflection and absorption properties,” Appl. Phys. Express 7(11), 112204 (2014).

14. P. C. Wu, N. Papasimakis, and D. P. Tsai, “Self-affine graphene metasurfaces for tunable broadband absorption,” Phys. Rev. Appl. 6(4), 044019 (2016).

15. Y. Yao, R. Shankar, M. A. Kats, Y. Song, J. Kong, M. Loncar, and F. Capasso, “Electrically tunable metasurface perfect absorbers for ultrathin mid-infrared optical modulators,” Nano Lett. 14(11), 6526–6532 (2014). [PubMed]  

16. M. C. Sherrott, P. W. C. Hon, K. T. Fountaine, J. C. Garcia, S. M. Ponti, V. W. Brar, L. A. Sweatlock, and H. A. Atwater, “Experimental demonstration of >230° phase modulation in gate-tunable graphene-gold reconfigurable mid-infrared metasurfaces,” Nano Lett. 17(5), 3027–3034 (2017). [PubMed]  

17. F. Yi, E. Shim, A. Y. Zhu, H. Zhu, J. C. Reed, and E. Cubukcu, “Voltage tuning of plasmonic absorbers by indium tin oxide,” Appl. Phys. Lett. 102(22), 221102 (2013).

18. J. Park, J. H. Kang, X. Liu, and M. L. Brongersma, “Electrically tunable epsilon-near-zero (ENZ) metafilm absorbers,” Sci. Rep. 5, 15754 (2015). [PubMed]  

19. Y.-W. Huang, H. W. H. Lee, R. Sokhoyan, R. A. Pala, K. Thyagarajan, S. Han, D. P. Tsai, and H. A. Atwater, “Gate-tunable conducting oxide metasurfaces,” Nano Lett. 16(9), 5319–5325 (2016). [PubMed]  

20. J. Park, J. H. Kang, S. J. Kim, X. Liu, and M. L. Brongersma, “Dynamic reflection phase and polarization control in metasurfaces,” Nano Lett. 17(1), 407–413 (2017). [PubMed]  

21. F. Ren, M. Li, Q. Gao, W. Cowell III, J. Luo, A. K.-Y. Jen, and A. X. Wang, “Surface-normal plasmonic modulator using sub-wavelength metal grating on electro-optic polymer thin film,” Opt. Commun. 352(1), 116–120 (2015).

22. Y. Enami, C. T. Derose, D. Mathine, C. Loychik, C. Greenlee, R. A. Norwood, T. D. Kim, J. Luo, Y. Tian, A. K.-Y. Jen, and N. Peyghambarian, “Hybrid polymer/sol–gel waveguide modulators with exceptionally large electro–optic coefficients,” Nat. Photonics 1(3), 180–185 (2007).

23. L. R. Dalton, P. A. Sullivan, and D. H. Bale, “Electric field poled organic electro-optic materials: state of the art and future prospects,” Chem. Rev. 110(1), 25–55 (2010). [PubMed]  

24. M. Li, S. Huang, X.-H. Zhou, Y. Zang, J. Wu, Z. Cui, J. Luoa, and A. K.-Y. Jen, “Poling efficiency enhancement of tethered binary nonlinear optical chromophores for achieving an ultrahigh n3r33 figure-of-merit of 2601 pm V−1,” J. Mater. Chem. C Mater. Opt. Electron. Devices 3, 6737–6744 (2015).

25. R. Palmer, S. Koeber, D. L. Elder, M. Woessner, W. Heni, D. Korn, M. Lauermann, W. Bogaerts, L. Dalton, W. Freude, J. Leuthold, and C. Koos, “High-speed, low drive-voltage silicon-organic hybrid modulator based on a binary-chromophore electro-optic material,” J. Lightwave Technol. 32(16), 2726–2734 (2014).

26. G. Raybon, J. Cho, A. Adamiecki, P. Winzer, L. H. H. Carvalho, J. C. R. F. Oliveira, A. Konczykowska, J.-Y. Dupuy, and F. Jorge, “180-Gb/s (90-GBd QPSK) single carrier transmitter using a polymer I/Q modulator,” in Proceedings of European Conference on Optical Communications (ECOC, 2016), pp. 31–33.

27. R. Himmelhuber, O. D. Herrera, R. Voorakaranam, L. Li, A. M. Jones, R. A. Norwood, J. Luo, A. K.-Y. Jen, and N. Peyghambarian, “A Silicon-polymer hybrid modulator - design, simulation and proof of principle,” J. Lightwave Technol. 31(24), 4067–4072 (2013).

28. F. Qiu, A. M. Spring, D. Maeda, M. A. Ozawa, K. Odoi, A. Otomo, I. Aoki, and S. Yokoyama, “A hybrid electro-optic polymer and TiO2 double-slot waveguide modulator,” Sci. Rep. 5, 8561 (2015). [PubMed]  

29. S. Inoue and A. Otomo, “Electro-optic polymer/silicon hybrid slow light modulator based on one-dimensional photonic crystal waveguides,” Appl. Phys. Lett. 103, 171101 (2013).

30. T. Yamada and A. Otomo, “Usefulness of transmission ellipsometric method for evaluation of electro-optic materials,” IEICE Trans. Electron. 98(2), 143–146 (2015).

31. D. M. Burland, R. D. Miller, and C. A. Walsh, “Second-order nonlinearity in poled-polymer systems,” Chem. Rev. 94(1), 31–75 (1994).

32. P. B. Johnson and R. W. Christy, “Optical constants of the noble metals,” Phys. Rev. B 6(12), 4370–4379 (1972).

33. J. Wiersig, “Formation of long-lived, scarlike modes near avoided resonance crossings in optical microcavities,” Phys. Rev. Lett. 97(25), 253901 (2006). [PubMed]  

34. C. J. Chang-Hasnain and W. Yang, “High-contrast gratings for integrated optoelectronics,” Adv. Opt. Photonics 4(3), 379–440 (2012).

35. Y. Kosugi, T. Yamada, A. Otomo, Y. Nakano, and T. Tanemura, “Surface-normal electro-optic-polymer modulator with silicon subwavelength grating,” IEICE Electron. Express 13(17), 1–9 (2016).

36. W. M. J. Green, M. J. Rooks, L. Sekaric, and Y. A. Vlasov, “Optical modulation using anti-crossing between paired amplitude and phase resonators,” Opt. Express 15(25), 17264–17272 (2007). [PubMed]  

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 Schematic (a), cross-section (b), and (c) modulation principle of the proposed surface-normal metasurface modulator.
Fig. 2
Fig. 2 Reflectance as a function of grating period Λ and wavelength. Inset shows Re(Ey) distribution at two curves.
Fig. 3
Fig. 3 (a) Dispersion relationship of Au/EO-polymer/Au MIM waveguide (b) Calculated Fabry-Perot resonance of MIM modes overlaid on the reflectance plot. d = 500 nm in two plots.
Fig. 4
Fig. 4 Calculated Q factor and figure of merit (Q/d) with different EO polymer thickness.
Fig. 5
Fig. 5 (a) Reflectance spectrum with ∆nz varied from −8 × 10−3 to 8 × 10−3, (b) Modulation characteristic of reflected light at wavelength of 1541 nm.

Equations (2)

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

( Δ n x Δ n y Δ n z ) = 1 2 ( n x 3 r 13 n y 3 r 23 n z 3 r 33 ) E mod .
β a + φ = m π .
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.