Next Article in Journal
Thermal Stability Analysis of Surface Wave Assisted Bio-Photonic Sensor
Next Article in Special Issue
Reflective Terahertz Metasurfaces Based on Non-Volatile Phase Change Material for Switchable Manipulation
Previous Article in Journal
Local Correction of the Light Position Implemented on an FPGA Platform for a 6 Meter Telescope
Previous Article in Special Issue
All-Dielectric Metasurface Based on Complementary Split-Ring Resonators for Refractive Index Sensing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Research of Gate-Tunable Phase Modulation Metasurfaces Based on Epsilon-Near-Zero Property of Indium-Tin-Oxide

1
School of Mathematics & Information Technology, Hebei Normal University of Science & Technology, Qinhuangdao 066004, China
2
School of Electrical Engineering, Yanshan University, Qinhuangdao 066004, China
*
Author to whom correspondence should be addressed.
Photonics 2022, 9(5), 323; https://doi.org/10.3390/photonics9050323
Submission received: 8 April 2022 / Revised: 5 May 2022 / Accepted: 6 May 2022 / Published: 9 May 2022
(This article belongs to the Special Issue Active/Reconfigurable Metasurfaces)

Abstract

:
In this paper, we proposed a reflection phase electrically tunable metasurface composed of an Au/Al2O3/ITO/Au grating structure. This antenna array can achieve a broad phase shift continuously and smoothly from 0° to 320° with a 5.85 V applied voltage bias. Tunability arises from field-effect modulation of the carrier concentrations or accumulation layer at the Al2O3/ITO interface, which excites electric and magnetic resonances in the epsilon-near-zero region. To make the reflected phase tuning range as wide as possible, some of the intensity of the reflected light is lost due to the excited surface plasmon effect. Simulation results show that the effect of optimal phase modulation can be realized at a wavelength range of 1550 nm by modulating the carrier concentration in our work. Additionally, we utilized an identical 13-unit array metasurface to demonstrate its application to the beam steering function. This active optical metasurface can enable a new realm of applications in ultrathin integrated photonic circuits.

1. Introduction

With the advancement of ultra-compact optical devices, there is an urgent demand for new structures that can realize the control of the phase and amplitude of the light dynamically [1,2,3,4,5,6,7,8]. Photonics based on 2-dimensional nanostructures, such as metasurfaces, can potentially provide a feasible solution for this specific challenge [9,10,11,12]. Metasurfaces are composed of periodic geometry arrays of subwavelength elements such that each element can impose a phase shift and amplitude change on the transmitted, reflected, and scattered light [13,14,15,16]. Additionally, there are three essential functions of metasurfaces: beam steering [17], focusing of light [18], and polarization conversion [19]. For example, Liu et al. summarized micromachined tunable metamaterials with mechanical actuation based on mechanical reconfiguration, which has required mechanical movement in recent years [20]. However, this method depends on processing conditions, and the modulation process is not flexible. To respond to the requirement of dynamic tunability, our work is based on optically tunable reflective metasurfaces that have a more comprehensive application prospect than passive metasurfaces [21,22].
The material of choice for the active metasurfaces needs to be quickly and controllably integrated into nanophotonic structures, and if possible, the selected material should be CMOS-compatible, thus allowing integration with on-chip platforms. In a vast regime of wavebands, transparent conductive oxides (TCOs) have been demonstrated as an ideal material that meets the requirements for many reasons, such as the low resistivity, controllable electro-optical property, and the epsilon-near-zero (ENZ) property [23,24,25,26,27]. Indium Tin Oxide (ITO) belongs to this class of material, which is an n-type semiconductor material formed by doping In2O3 with Sn in a specific proportion. The advantages of using ITO as an active material in metasurfaces are manifolds: ITO films can modulate the optical refractive index in a plasmonic cavity configuration and show strong electro-optic solid effects in epsilon-near-zero (ENZ) behavior due to the carrier concentration can be changed in the range of 1020–1021 cm−3 [28,29,30]. To date, industrial advancements, such as magnetron sputtering systems, reactive evaporation, and electrochemical diffusion processes, have made a high yield and reliable process possible, contributing to the wide application of ITO [31,32].
Among the various physical mechanisms for regulating the complex refractive index of materials to achieve phase shifts, field-effect modulation is chosen to offer continuous tunability, shorter response time, and a relatively wide tuning range. It can also provide an amply significant carrier density change in doped semiconductors or conducting oxide, leading to a variable complex refractive index in the charge depletion or accumulation regions [33,34,35]. Based on the formation principle of charge depletion regions in doped semiconductors, field-effect modulation is a distinctly mature approach in the semiconductor process and is compatible with CMOS [36,37]. For example, Sherrott et al. demonstrated a 237-degree phase modulation range at an operating wavelength of 8.50 mm by using a gate-tunable graphene-gold resonator structure in 2017 [38]. Huang et al. demonstrated an active metasurface realizing the dynamic control of both amplitude and phase [39]. A phase shift of about 225 degrees can be observed at a wavelength around 1550 nm by applying a voltage of 4.6 V in their work. Inspired by these studies, the phase modulation range and the incident wavelength are optimized under a smaller voltage in our work.
In this paper, a dynamically tunable metasurface with a metal–medium–metal grating structure is proposed. We chose ITO as the active material to reach an expansive phase change due to the Fabry–Perot resonance and zero-crossing effect of the permittivity of the ITO accumulation layer. In addition, we simulate a dynamical beam steering device implemented with 30°, 20°, and 10° modulation functions. This optical array will find potential applications in the focusing lens, beam steering device, and polarimeters.

2. Theoretical Models and Design Considerations

A schematic representation of the proposed metasurface structure is shown in Figure 1. We designed tunable metasurfaces composed of an Au back-reflector, on top of which an ITO layer is deposited. This layer is followed by the deposition of a dielectric spacer Al2O3 layer and Au nanoantennas. In designing the metasurface element, the physical parameters for the ITO layers should be obtained [40,41]. In our calculations, we assume that the carrier concentration of the active material is N0 = 3.2 × 1020 cm−3.
When no external bias is applied, the real part permittivity of the ITO layer is positive at a wavelength of 1550 nm. Once under bias, a charge accumulation layer is formed in the ITO, reaching the epsilon-near-zero (ENZ) region with the voltage increasing, which results in a considerable electric field enhancement occurring in the accumulation layer for near-infrared wavelengths. The permittivity dispersion of generic TCO can be described by the Drude model:
ε = ε ω p 2 ω 2 + i ω γ
ω p 2 = N 0 e c 2 ε 0 m t
where ε = 3.95 is the high-frequency permittivity [42]; γ = 180 THz is the collision frequency; ω p is the plasma frequency, and ω is the angular frequency of incident light. ε 0 is the vacuum permittivity; m t = 0.35 × m 0 is the effective mass of electrons in ITO; m 0 is the electron rest mass [43]; e c is the electron charge, and the initial electron concentration N0 is set to be 3.2 × 1020 cm−3. Based on Formula (2), the plasma frequency is found to be 1.84 × 1014 rad·s−1.
Another parameter that determines performance is the choice of the plasmonic metal. The relative permittivity of Au is 1.59 [44,45], and our work function of Au is 5.1 eV, which is higher than that of ITO (4.4 eV) when the carrier concentration N0 = 3.2 × 1020 cm−3. In addition, selecting Au as a metal electrode in our metasurfaces ensures that the phase change is stable and continuous. The other is the choice of gate dielectric material. Alumina (Al2O3) exhibits high DC permittivity and almost perfect interfacial properties with Si-based substrates. It is often employed in field-effect transistor fields as high dielectric constant gate dielectric materials and has a high breakdown field of up to 10 MV/cm. The average thickness t a c c of the accumulation layer related to the carriers injected into the ITO layer in a standard metal–insulator–semiconductor structure [46] is given by:
t a c c = π 2 k B T ε 0 ε c N 0 e c 2
where k B = 1.38 × 10 23 J / K is the Boltzmann constant; T = 300 K is the Kelvin temperature; e c = 1.6 × 10−19C is the electron charge; ε 0 is the free space permittivity, and εc = 9.3 is the relative static permittivity of ITO [47]. Due to the aforementioned N0 = 3.2 × 1020 cm−3, the thickness of the accumulation layer is found to be t a c c = 0.7 nm. Then, we can find the carrier concentration Nacc in the accumulation layer. Assuming a uniform carrier distribution in the accumulation layer [48], the injected carrier concentration can be calculated as:
N a c c = V g ε 0 ε i n s e c × t a c c × t i n s
where tins = 5 nm is the thickness of the insulator spacer, ε i n s = 9 is the relative permittivity of Al2O3 [49], and Vg is the applied voltage.
Based on the above analysis, we can perform optical measurements to illustrate the phase change of incident light with the applied voltage. All the simulations were performed using the numerical simulation software COMSOL Multiphysics based on the finite element method (FEM).

3. Results and Discussion

The metal-insulator-semiconductor field-effect structure was utilized to observe the ENZ region and injected electron density in the ITO accumulation layer. We simulated reflectance and phase modulation of the periodical antenna structure under normal incidence illumination with transverse magnetic (TM) polarization (H-field along with the stripes) based on Figure 1. Electromagnetic waves and frequency domain modules were utilized in our simulation, and the boundary condition was set to a periodic boundary condition. For the ITO layer, we used extremely refined mesh division and free triangular mesh for other areas. Figure 2a,b show the distribution of electromagnetic fields at different applied voltages for a fixed wavelength (1550 nm). The metal grating excites surface plasmon resonance, causing the electric field to converge on the ITO layer, and then changing the dielectric constant of ITO at this time can make the phase changes. At 0 V, symmetrical electric fields between the Au antenna and Au backplane can be observed due to ITO optically behaving as a dielectric (ε > 0). In the case of an applied voltage of 2.6 V, the ENZ condition takes up the accumulation layer at the interface of the ITO and Al2O3, and the enhancement of the z-component of the electric field in the accumulation layer is magnified in Figure 2c. As one can see, the real part of the active layer’s permittivity (εacc) is 2.093, and the carrier concentration is 3.2 × 1020 cm−3 at no bias for the incident wavelength λ = 1550 nm in Figure 2d. Based on the carrier concentration model in ITO materials [50,51], it shows that the concentration of carriers within the active layer could rise to 1.1 × 1021 cm−3 with the increasing gate bias, which leads to the permittivity of the accumulation layer changing from a positive one to a negative one. For applied bias values larger than 3.0 V, the permittivity of the accumulation layer is negative, similar to metal. In the case of an applied bias around 2.6 V, εacc is close to zero (εacc = 0.041–0.58i), which stimulates the resonance of the local surface plasma, resulting in a steep change of phase shift. This is in line with our expectations of achieving the ENZ region. Figure 2e shows the reflectance and phase shift spectra as a function of the applied voltage. The dotted lines indicate the ENZ region in the accumulation layer at the Al2O3/ITO interface. With the gate bias increasing, the magnetic dipole plasmon resonance couples to the ENZ region in the ITO accumulation layer can induce a significant phase shift in reflection [52]. According to this feature, we can study the phase change that conforms to the present description.
As shown in Figure 3a, the designed element phase tuning range (the blue line) is more significant than 300° at λ = 1550 nm, which is obviously better than the phase modulation of the other wavelengths. At 2.65 V bias, the accumulation layer preliminarily reaches the ENZ region, leading to a phase shift of 247°. Based on Formula (5), the phase change is related to the applied voltage [53]. In the ENZ region, the electric field energy is highly confined inside the ITO thin film and fully interacts with the ITO layer. The locally concentrated electric field is very sensitive to changes in the surrounding electromagnetic environment parameters. Therefore, when doping ITO through the bias electrode due to the change in the ITO film’s conductivity, the light scattering phase will change sharply, resulting in electrical modulation of the phase. Figure 3b shows the variation between the reflectance of the metasurface and the applied voltage at different wavelengths. With the ENZ condition holding, the plasmonic resonance enhancement is generated in the active layer, leading to the minimum reflectance of 5.9% at 3.35 V shown in Figure 3b. Because critical coupling occurs at the Al2O3/ITO interface, the incident wave is absorbed, and the stronger the absorption, the more severe the phase mutation, i.e., the phase change is the most obvious when the amplitude is the lowest.
ϕ V = 2 π Δ n L λ
Where Δ n is the change in the refractive index of the material caused by the voltage, L is the waveguide length, and ϕ is the phase change. In this case, the reflectance of field-effect modulation is observed at different biases from −2 V to 6 V, which can be understood from the continuity of the normal component of the electric displacement at the field-effect modulation interface. In modern electronic communication, the integration of electronic scales usually requires a nanoscale, and the reduction of device size also requires further energy consumption reduction [52,53]. The designed antenna arrays realized smooth and extensive phase regulation over 300° at near-infrared wavelengths under a smaller voltage condition. Therefore, voltage reduction in our structure is of great significance for future device design.
As demonstrated in the above results, a significant phase change can be achieved in this gated active reflective metasurface. We further employed 13 independent unit cells to illustrate the function of beam steering. These tunable beam steering components have an identical antenna structure of Au/Al2O3/ITO/Au planar layers with a periodicity of 400 nm operating at a wavelength of 1550 nm. Thus, the 13-element metasurface operating over the 1550 nm band has a length of 10 μm. The array surrounded by air is at the bottom of the simulation region, and the boundary condition is set to the scattering boundary condition. The incident wave propagates vertically from the z-direction, and the electromagnetic beam shown in Figure 4 is reflected by the entire antenna structure. The maximum delay required to steer the antenna over an angle θ is given by [54]:
θ = arcsin φ n n k r d
where θ is the deflection angle of the reflected beam, k r = ω / c represents the relationship between the phase of the incident wave and the distance, n is the element number, d is the spacing of adjacent array elements, and φ n is the degrees of the adjacent unit’s phase shift. We define a positive scan angle when the antenna looks to the left of the boresight. For our present scenario (where n = 13 and θ = 10 deg), the maximum delays required for the angle +10 deg with respect to elements n = 0, 1, 2, … are 0, 16.1°, 31.8°, … The simulation results are consistent with our analysis above, which demonstrates the function of beam steering successfully, i.e., each element can satisfy the condition for a fixed phase shift when the incident wave frequency is fixed. The direction of the reflected beam toward 30°, 20°, and 10° has shown in Figure 4. This shows that the beam deflection of 30°, 20°, and 10° corresponds to a single element phase difference of 46.5°, 31.8°, and 16.1°, respectively. This efficient electronically tunable beam steering system will have potential value in the development of ultrathin on-chip applications and sensing devices.

4. Conclusions

In summary, gate-tunable metasurfaces based on Indium-Tin-Oxide were studied in our work. The simulated phase shift can reach 247 degrees by applying a 2.65 V gate bias at a wavelength of 1550 nm, whose maximum phase shift can reach 320 degrees at 5.85 V. This continuous and smooth phase modulation is more than 20 degrees greater than other mid-IR metasurface systems [38,39], which only need a smaller applied voltage. In addition, beam steering with different angles was demonstrated by controlling the phase shift of 13 unit cells, i.e., the phase shift of the adjacent unit moves by 16.1 degrees, and the reflected beam deflects by 10 degrees. This strong light–matter interaction can be explained by the critical coupling of incident waves to the metasurface. The incident light excites eigenmodes through radiative coupling, and the energy is converted into resistive losses under critical coupling conditions so that the reflections are sufficiently suppressed. Phase modulation is a transition state that occurs in under-damped and over-damped oscillations, which means that this transition inevitably passes through a critical coupling region. Therefore, a wide range of phase modulation will be strongly coupled with absorption strength, i.e., with the range of phase tunability increasing, the absorption intensity will also increase, and the phase tunability comes at the expense of energy loss. Due to the critical coupling condition of perfect absorption during the phase change process, in the part with the most severe phase change, the corresponding light absorption is the strongest, and the reflection amplitude is the smallest, at only about 1%. Such a two-dimensional array could enable guidance on devices, such as dynamic holograms, ultrathin lenses, nanoscale light modulators, and polarimeters.

Author Contributions

Q.C. and X.L. contributed to the conception of the study; Q.C. and S.G. contributed to the analysis and manuscript preparation; Q.C. and Z.L. performed the data analyses and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the Natural Science Foundation of Hebei Province grant (No: F2017203316) in China, the Science and Technology Program of Hebei (No: 206Z1703G) in China, the Science and Technology Project of Hebei Education Department (No: ZD2022087) in China, and the Fundamental Research Funds for Provincial Universities (No: 2021JK03) in China.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ahmadivand, A.; Gerislioglu, B.; Ramezani, Z. Gated Graphene Island-Enabled Tunable Charge Transfer Plasmon Terahertz Metamodulator. Nanoscale 2019, 11, 8091–8095. [Google Scholar] [CrossRef]
  2. Cheng, J.; Fan, F.; Chang, S. Recent Progress on Graphene-Functionalized Metasurfaces for Tunable Phase and Polarization Control. Nanomaterials 2019, 9, 398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Liu, Z.; Feng, W.; Long, Y.; Guo, S.; Liang, H.; Qiu, Z.; Fu, X.; Li, J. A Metasurface Beam Combiner Based on the Control of Angular Response. Photonics 2021, 8, 489. [Google Scholar] [CrossRef]
  4. Zhou, C.; Zhen, M.; Lu, P.; Teng, S. Compound Vector Light Generator Based on a Metasurface. Photonics 2021, 8, 243. [Google Scholar] [CrossRef]
  5. Kildishev, A.V.; Boltasseva, A.; Shalaev, V.M. Planar Photonics with Metasurfaces. Science 2013, 339, 1289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Hsiao, H.; Chu, C.H.; Tsai, D.P. Fundamentals and Applications of Metasurfaces. Small Methods 2017, 1, 1600064. [Google Scholar] [CrossRef] [Green Version]
  7. Luo, X.; Pu, M.; Guo, Y.; Li, X.; Ma, X. Electromagnetic Architectures: Structures, Properties, Functions and Their Intrinsic Relationships in Subwavelength Optics and Electromagnetics. Adv. Photonics Res. 2021, 2, 2100023. [Google Scholar] [CrossRef]
  8. Yang, X.; Hu, C.; Deng, H.; Rosenmann, D.; Czaplewski, D.A.; Gao, J. Experimental Demonstration of Near-Infrared Epsilon-near-Zero Multilayer Metamaterial Slabs. Opt. Express 2013, 21, 23631–23639. [Google Scholar] [CrossRef]
  9. Pang, K.; Alam, M.Z.; Zhou, Y.; Liu, C.; Reshef, O.; Manukyan, K.; Voegtle, M.; Pennathur, A.; Tseng, C.; Su, X.; et al. Adiabatic Frequency Conversion Using a Time-Varying Epsilon-Near-Zero Metasurface. Nano Lett. 2021, 21, 5907–5913. [Google Scholar] [CrossRef]
  10. Genevet, P.; Capasso, F. Holographic optical metasurfaces: A review of current progress. Rep. Prog. Phys. 2015, 78, 024401. [Google Scholar] [CrossRef]
  11. Genevet, P.; Capasso, F.; Aieta, F.; Khorasaninejad, M.; Devlin, R. Recent advances in planar optics: From plasmonic to dielectric metasurfaces. Optica 2017, 4, 139. [Google Scholar] [CrossRef]
  12. Yu, N.; Genevet, P.; Kats, M.A.; Aieta, F.; Tetienne, J.-P.; Capasso, F.; Gaburro, Z. Light Propagation with Phase Discontinuities: Generalized Laws of Reflection and Refraction. Science 2011, 334, 333–337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Shen, X.; Cui, T.J.; Martin-Cano, D.; Garcia-Vidal, F.J. Conformal Surface Plasmons Propagating on Ultrathin and Flexible Films. Proc. Natl. Acad. Sci. USA 2013, 110, 40–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Liang, Y.; Lin, H.; Lin, S.; Wu, J.; Li, W.; Meng, F.; Yang, Y.; Huang, X.; Jia, B.; Kivshar, Y. Hybrid Anisotropic Plasmonic Metasurfaces with Multiple Resonances of Focused Light Beams. Nano Lett. 2021, 21, 8917–8923. [Google Scholar] [CrossRef] [PubMed]
  15. Long, Z.; Liang, Y.; Feng, L.; Zhang, H.; Liu, M.; Xu, T. Low-Cost and High Sensitivity Glucose Sandwich Detection Using a Plasmonic Nanodisk Metasurface. Nanoscale 2020, 12, 10809–10815. [Google Scholar] [CrossRef] [PubMed]
  16. Sun, S.; Yang, K.-Y.; Wang, C.-M.; Juan, T.-K.; Chen, W.T.; Liao, C.Y.; He, Q.; Xiao, S.; Kung, W.-T.; Guo, G.-Y.; et al. High-Efficiency Broadband Anomalous Reflection by Gradient Meta-Surfaces. Nano Lett. 2012, 12, 6223–6229. [Google Scholar] [CrossRef]
  17. Khorasaninejad, M.; Chen, W.T.; Devlin, R.C.; Oh, J.; Zhu, A.Y.; Capasso, F. Metalenses at visible wavelengths: Diffraction-limited focusing and subwavelength resolution imaging. Science 2016, 352, 1190–1194. [Google Scholar] [CrossRef] [Green Version]
  18. Liu, A.Q.; Zhu, W.M.; Tsai, D.P.; Zheludev, N. Micromachined tunable metamaterials: A review. J. Opt. 2012, 14, 114009. [Google Scholar] [CrossRef] [Green Version]
  19. Sautter, J.; Staude, I.; Decker, M.; Rusak, E.; Neshev, D.N.; Brener, I.; Kivshar, Y.S. Active Tuning of All-Dielectric Metasurfaces. ACS Nano 2015, 9, 4308–4315. [Google Scholar] [CrossRef]
  20. Waters, R.F.; Hobson, P.A.; MacDonald, K.F.; Zheludev, N.I. Optically switchable photonic metasurfaces. Appl. Phys. Lett. 2015, 107, 917. [Google Scholar] [CrossRef] [Green Version]
  21. Naik, G.V.; Shalaev, V.M.; Boltasseva, A. Alternative Plasmonic Materials: Beyond Gold and Silver. Adv. Mater. 2013, 25, 3264–3294. [Google Scholar] [CrossRef] [PubMed]
  22. Babicheva, V.E.; Boltasseva, A.; Lavrinenko, A. Transparent conducting oxides for electro-optical plasmonic modulators. Nanophotonics 2015, 4, 436. [Google Scholar] [CrossRef]
  23. Melikyan, A.; Lindenmann, N.; Walheim, S.; Leufke, P.M.; Ulrich, S.; Ye, J.; Vincze, P.; Hahn, H.; Schimmel, T.; Koos, C.; et al. Surface plasmon polariton absorption modulator. Opt. Express 2011, 19, 8855–8869. [Google Scholar] [CrossRef] [PubMed]
  24. Oleksandr, B.; Nina, P.; Malgosia, K.; Nikolay, I.; Fedotov, V.A. Electrically Controlled Nanostructured Metasurface Loaded with Liquid Crystal: Toward Multifunctional Photonic Switch. Adv. Opt. Mater. 2015, 3, 674–679. [Google Scholar]
  25. Forouzmand, A.; Mosallaei, H. All-Dielectric C-Shaped Nanoantennas for Light Manipulation: Tailoring Both Magnetic and Electric Resonances to the Desire. Adv. Opt. Mater. 2017, 5, 1700147. [Google Scholar] [CrossRef]
  26. Liu, Z.-Q.; Liu, G.-Q.; Liu, X.-S.; Zhou, H.-Q.; Gu, G. Multispectral Broadband Light Transparency of a Seamless Metal Film Coated with Plasmonic Crystals. Plasmonics 2014, 9, 615–622. [Google Scholar] [CrossRef]
  27. Sorger, V.J.; Lanzillotti-Kimura, N.D.; Ma, R.-M.; Zhang, X. Ultra-Compact Silicon Nanophotonic Modulator with Broadband Response. Nanophotonics 2012, 1, 17–22. [Google Scholar] [CrossRef] [Green Version]
  28. Park, J.; Kang, J.-H.; Liu, X.; Brongersma, M.L. Electrically Tunable Epsilon-Near-Zero (ENZ) Metafilm Absorbers. Sci. Rep. 2015, 5, 15754. [Google Scholar] [CrossRef] [Green Version]
  29. Tahersima, M.H.; Ma, Z.; Gui, Y.; Miscuglio, M.; Sun, S.; Amin, R.; Dalir, H.; Sorger, V.J. Coupling-Controlled Dual ITO Layer Electro-Optic Modulator in Silicon Photonics. arXiv 2018, arXiv:arXiv1812.11458. [Google Scholar]
  30. Amin, R.; Maiti, R.; Carfano, C.; Ma, Z.; Tahersima, M.H.; Lilach, Y.; Ratnayake, D.; Dalir, H.; Sorger, V.J. 0.52 V mm ITO-based Mach-Zehnder modulator in silicon photonics. APL Photonics 2018, 3, 126104. [Google Scholar] [CrossRef]
  31. Feigenbaum, E.; Diest, K.; Atwater, H.A. Unity-Order Index Change in Transparent Conducting Oxides at Visible Frequencies. Nano Lett. 2010, 10, 2111–2116. [Google Scholar] [CrossRef] [PubMed]
  32. Tiwari, D.L.; Sivasankaran, K. Nitrogen-doped NDR behavior of double-gate graphene field-effect transistor. Superlattices Microstruct. 2019, 136, 106308.1–106308.8. [Google Scholar] [CrossRef]
  33. Schon, J.H. Field-Effect Modulation of the Conductance of Single Molecules. Science 2001, 294, 2138–2140. [Google Scholar] [CrossRef]
  34. Cao, Z.; Duan, B.; Yuan, S.; Yang, Y. New super junction LDMOS with surface and bulk electric field modulation by buffered step doping and multi floating buried layers—ScienceDirect. Superlattices Microstruct. 2017, 111, 221–229. [Google Scholar] [CrossRef]
  35. Dionne, J.A.; Diest, K.; Sweatlock, L.A.; Atwater, H.A. PlasMOStor: A Metal-Oxide-Si Field-Effect Plasmonic Modulator. Nano Lett. 2009, 9, 897–902. [Google Scholar] [CrossRef] [PubMed]
  36. Sherrott, M.C.; Hon, P.W.C.; Fountaine, K.T.; Garcia, J.C.; Ponti, S.M.; Brar, V.W.; Sweatlock, L.A.; Atwater, H.A. Experimental Demonstration of >230° Phase Modulation in Gate-Tunable Graphene-Gold Reconfigurable Mid-Infrared Metasurfaces. Nano Lett. 2017, 17, 3027–3034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Huang, Y.-W.; Lee, H.W.H.; Sokhoyan, R.; Pala, R.A.; Thyagarajan, K.; Han, S.; Tsai, D.P.; Atwater, H.A. Gate-Tunable Conducting Oxide Metasurfaces. Nano Lett. 2016, 16, 5319. [Google Scholar] [CrossRef] [Green Version]
  38. Amin, R.; George, J.K.; Sun, S.; de Lima, T.F.; Tait, A.N.; Khurgin, J.B.; Miscuglio, M.; Shastri, B.J.; Prucnal, P.R.; El-Ghazawi, T.; et al. ITO-based Electro-absorption Modulator for Photonic Neural Activation Function. APL Mater. 2019, 7, 081112. [Google Scholar] [CrossRef] [Green Version]
  39. Fatemi, R.; Abiri, B.; Khachaturian, A.; Hajimiri, A. High sensitivity active flat optics optical phased array receiver with a two-dimensional aperture. Opt. Express 2018, 26, 29983. [Google Scholar] [CrossRef]
  40. Zhewei, W.; Chaonan, C.; Ke, W.; Chong, H.; Ye, H. Transparent Conductive Oxides and Their Applications in Near Infrared Plasmonics. Phys. Status Solidi (A) 2019, 216, 1700794. [Google Scholar]
  41. Fei, Y.; Shim, E.; Zhu, A.Y.; Zhu, H.; Reed, J.C.; Cubukcu, E. Voltage tuning of plasmonic absorbers by indium tin oxide. Appl. Phys. Lett. 2013, 102, 1448. [Google Scholar]
  42. Campbell, S.D.; Ziolkowski, R.; Cao, J.; Laref, S.; Muralidharan, K.; Deymier, P. Anisotropic permittivity of ultra-thin crystalline Au films: Impacts on the plasmonic response of metasurfaces. Appl. Phys. Lett. 2013, 103, 075120. [Google Scholar] [CrossRef]
  43. Lynch, D.W. Comments on the Optical Constants of Metals and an Introduction to the Data for Several Metals. Handb. Opt. Constants Solids 1997, 1, 275–367. [Google Scholar]
  44. Vexler, M.I.; Tyaginov, S.E.; Shulekin, A.F.; Grekhov, I.V. Current-voltage characteristics of Al/SiO2/p-Si MOS tunnel diodes with a spatially nonuniform oxide thickness. Semiconductors 2006, 40, 1109–1115. [Google Scholar] [CrossRef]
  45. Li, S.-Q.; Sakoda, K.; Ketterson, J.B.; Chang, R.P.H. Broadband Resonances in ITO Nanorod Arrays. Eprint Arxiv 2015, 63, 694–698. [Google Scholar]
  46. Siahkal-Mahalle, B.H.; Abedi, K. The Effect of Carrier Distribution on Performance of ENZ-Based Electro-Absorption Modulator. Plasmonics 2020, 15, 1689–1697. [Google Scholar] [CrossRef]
  47. Etinger-Geller, Y.; Zoubenko, E.; Baskin, M.; Kornblum, L.; Pokroy, B. Thickness Dependence of the Physical Properties of Atomic-Layer Deposited Al2O3. J. Appl. Phys. 2019, 125, 185302. [Google Scholar] [CrossRef] [Green Version]
  48. Jin, L.; Chen, Q.; Liu, W.; Song, S. Electro-absorption Modulator with Dual Carrier Accumulation Layers Based on Epsilon-Near-Zero ITO. Plasmonics 2016, 11, 1087–1092. [Google Scholar] [CrossRef]
  49. Pshenichnyuk, I.A.; Kosolobov, S.S.; Drachev, V.P. Fine-Tuning of the Electro-Optical Switching Behavior in Indium Tin Oxide. Phys. Rev. B. 2021, 103, 115404. [Google Scholar] [CrossRef]
  50. Nemati, A.; Wang, Q.; Ang, N.S.S.; Hong, M.H.; Teng, J. Ultra-high extinction-ratio light modulation by electrically tunable metasurface using dual epsilon-near-zero resonances. Opto-Electron. Adv. 2021, 4, 200088-1–200088-11. [Google Scholar] [CrossRef]
  51. Horikawa, K.; Nakasuga, Y. Self-heterodyning optical waveguide beam forming and steering network integrated on Lithium Niobate substrate. Microw. Theory Tech. IEEE Trans. 1995, 43, 2395–2401. [Google Scholar] [CrossRef]
  52. Ding, K.; Shen, Y.; Ng, J.; Zhou, L. Equivalent-medium theory for metamaterials made by planar electronic materials. Epl 2013, 102, 28005. [Google Scholar] [CrossRef]
  53. Zhu, B.; Ren, G.; Zheng, S.; Lin, Z.; Jian, S. Nanoscale dielectric-graphene-dielectric tunable infrared waveguide with ultrahigh refractive indices. Opt. Express 2013, 21, 17089–17096. [Google Scholar] [CrossRef] [PubMed]
  54. Goutzoulis, A.P.; Davies, D.K.; Zomp, J.M. Hybrid electronic fiber optic wavelength-multiplexed system for true time delay steering of phased array antennas. Opt. Express 2005, 31, 2312–2322. [Google Scholar] [CrossRef]
Figure 1. (a) Sectional view of the unit cell of the gate-tunable metasurface composed of an Au back reflector and an ITO layer followed by a gate dielectric on top of which Au antennas are located. The thicknesses of the antenna array, the insulator spacer, the ITO layer, and the back reflector are tg = 30 nm, tins = 5 nm, tITO = 20 nm, and tb = 100 nm, respectively. The antenna dimensions Wb = 400 nm, and the electrode width is Wg = 230 nm. A voltage bias is applied between the ITO layer and the top antennas. The applied voltage biases result in forming an accumulation/depletion region in the ITO layer at the top, and the TM wave is incident perpendicular to the direction of the metasurface. (b) Schematic of the metasurface. The period of the metasurface is 400 nm.
Figure 1. (a) Sectional view of the unit cell of the gate-tunable metasurface composed of an Au back reflector and an ITO layer followed by a gate dielectric on top of which Au antennas are located. The thicknesses of the antenna array, the insulator spacer, the ITO layer, and the back reflector are tg = 30 nm, tins = 5 nm, tITO = 20 nm, and tb = 100 nm, respectively. The antenna dimensions Wb = 400 nm, and the electrode width is Wg = 230 nm. A voltage bias is applied between the ITO layer and the top antennas. The applied voltage biases result in forming an accumulation/depletion region in the ITO layer at the top, and the TM wave is incident perpendicular to the direction of the metasurface. (b) Schematic of the metasurface. The period of the metasurface is 400 nm.
Photonics 09 00323 g001
Figure 2. (a,b) is the spatial distribution of the z component of the electric field Ez for applied biases of 0 V and 2.6 V, and (c) is the magnified accumulation region near the Al2O3/ITO interface of 2.6 V applied bias at the wavelength of 1550 nm. The legend represents the magnitude of the electromagnetic field. Simulated (d) is the carrier concentration, and the permittivity of the ITO accumulation layer varies due to gating applied voltage at λ = 1550 nm. The black line shows carrier concentration variation, and the red line represents different permittivity from −2.8 to 3.7 at the Al2O3/ITO interface. Simulated (e) is the real part of the permittivity spectrum at different bias voltages and wavelengths. The legend represents the value of the dielectric constant at different wavelengths.
Figure 2. (a,b) is the spatial distribution of the z component of the electric field Ez for applied biases of 0 V and 2.6 V, and (c) is the magnified accumulation region near the Al2O3/ITO interface of 2.6 V applied bias at the wavelength of 1550 nm. The legend represents the magnitude of the electromagnetic field. Simulated (d) is the carrier concentration, and the permittivity of the ITO accumulation layer varies due to gating applied voltage at λ = 1550 nm. The black line shows carrier concentration variation, and the red line represents different permittivity from −2.8 to 3.7 at the Al2O3/ITO interface. Simulated (e) is the real part of the permittivity spectrum at different bias voltages and wavelengths. The legend represents the value of the dielectric constant at different wavelengths.
Photonics 09 00323 g002
Figure 3. (a) Phase modulation as a function of applied bias for different wavelengths, from 1400 nm to 1600 nm. (b) Simulated reflectance as a function of the applied bias voltage at different wavelengths.
Figure 3. (a) Phase modulation as a function of applied bias for different wavelengths, from 1400 nm to 1600 nm. (b) Simulated reflectance as a function of the applied bias voltage at different wavelengths.
Photonics 09 00323 g003
Figure 4. The simulated distributions of the electric field for tunable antennas capable of dynamical angular beam steering toward (a) 30°, (b) 20°, and (c) 10°, respectively. The scale bar is shown in the subfigure. 1 cm on the picture represents the actual distance of 2 μm. The legend represents the magnitude of the electromagnetic field. The polarization direction of the electric field is the z-direction.
Figure 4. The simulated distributions of the electric field for tunable antennas capable of dynamical angular beam steering toward (a) 30°, (b) 20°, and (c) 10°, respectively. The scale bar is shown in the subfigure. 1 cm on the picture represents the actual distance of 2 μm. The legend represents the magnitude of the electromagnetic field. The polarization direction of the electric field is the z-direction.
Photonics 09 00323 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, X.; Cheng, Q.; Guo, S.; Li, Z. Research of Gate-Tunable Phase Modulation Metasurfaces Based on Epsilon-Near-Zero Property of Indium-Tin-Oxide. Photonics 2022, 9, 323. https://doi.org/10.3390/photonics9050323

AMA Style

Li X, Cheng Q, Guo S, Li Z. Research of Gate-Tunable Phase Modulation Metasurfaces Based on Epsilon-Near-Zero Property of Indium-Tin-Oxide. Photonics. 2022; 9(5):323. https://doi.org/10.3390/photonics9050323

Chicago/Turabian Style

Li, Xin, Qiufan Cheng, Shiliang Guo, and Zhiquan Li. 2022. "Research of Gate-Tunable Phase Modulation Metasurfaces Based on Epsilon-Near-Zero Property of Indium-Tin-Oxide" Photonics 9, no. 5: 323. https://doi.org/10.3390/photonics9050323

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