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

Robust metamaterial-based antireflection coating for surface plasmon polariton resonance

Open Access Open Access

Abstract

A metasurface antireflection coating (meta-ARC) consisting of a gold disk array (GDA) and a dielectric layer greatly provides the flexibility to reduce the undesired reflection of light at the interface of the incident medium (air) and the surface plasmon polariton (SPP) structures such as a metallic hole array (MHA). Due to the possibility of changing the coupling strength between two different kinds of resonances (in GDA and MHA) caused by fabrication imperfections, we investigate the impacts of resonance coupling to the performance of antireflection by varying the shape of gold disks in meta-ARC and the alignment-shift between GDA and MHA. Simulation results show that (1) the amplitude of transmitted light through meta-ARC (with fabrication-induced shape-variation in GD) and MHA is only changed by small amount as compared to the meta-ARC with ideal-shaped GDs, (2) the transmission through MHA can be improved in the range of 44% (maximally alignment-shifted GDs) up to 83% (perfectly aligned GDs) by suppressing the reflected light owing to the meta-ARC, which indicates that MHA transmission light can be enhanced by more than 44% regardless of any imperfections in the fabrication of GDs while the SPP resonance wavelengths remain invariant in both cases. Our work can pave the way for a robust ARC method, leading to efficiently improving the performance of plasmonic optoelectronic devices when the meta-ARC is integrated.

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

1. Introduction

Few-layer metasurfaces have recently attracted tremendous interests to achieve emergent functionality to control and manipulate light including perfect absorption [1–3], antireflection [4–6], polarization conversion and amplitude-phase control of light [7]. These few-layer metasurfaces are usually composed of two or more layers of metallic structures with different geometric designs. To understand the electromagnetic property of the few-layer metasurface, the entire structure can be considered as a single-layer homogenous thin film with certain effective permittivity (ε) and permeability (μ) that are responsible for the functionality (e.g. perfect absorption [2]). Recent studies show that each layer in the few-layer metasurface acts as a homogenous thin film (meta-film) that allows light to propagate through or to be reflected at the interfaces between meta-films [3,4,7]. Light reflected at each layer results in constructive or destructive interferences at certain interface of the structure and thereby bring forth the desired functionalities. It has been shown that the transmission and reflection of light from the entire few-layer metasurface can be very well derived by assembling all layers of meta-films using their effective permittivity and permeability [3,4,8]. In the multilayer metafilm model, the inter-layer resonance coupling is usually neglected because the operating frequency is usually far-away from the resonance (if any) of each layer. However, the resonance in the metafilm can still affect the behavior of neighboring-layers through capacitive or inductive coupling, especially when the separation between layers are very small.

We choose a metallic disk array (MDA) metasurface that acts as an anti-reflection coating layer atop a metallic hole array (MHA). It has been shown that such MDA layer can significantly increase the extraordinary optical transmission (EOT) of the MHA [4] by eliminating the undesired reflection. EOT is well kwon as a phenomenon that the transmission of light though subwavelength apertures (e.g. MHA) is greatly enhanced by the surface plasmon polariton (SPP) resonance at the interface between metallic structure and dielectric [9–13]. The enhanced interaction between light and active layer of optoelectronic devices due to the EOT has been widely used to achieve improved performance or extended functionality, for instance, color and polarization sensitivity, dynamic-range enhancement, and so on [14–21]. Our recent work has shown that the EOT can be greatly increased by suppression of the reflection using conventional thin-film antireflection coating (ARC) method [22] or metamaterial-based antireflection coating (meta-ARC) [4]. In this work, we focus on the impact of inter-layer coupling effect on the performance of meta-ARC that arises due to alignment-shift between the gold disk array (GDA) and MHA, and the shape variation of gold disks (GDs). Such alignment-shift and shape variation are inevitable due to the imperfections of samples during the nano-fabrication process.

2. Geometry and dimensions of meta-ARC

Our meta-ARC consists of a planar gold disk array (GDA) atop the benzocyclobutene (BCB) layer, which is placed on top of a gold film with an array of circular apertures (MHA) as illustrated in Fig. 1. All structures were fabricated with the standard semiconductor process (spin coater & photolithography). The geometrical parameters of GDA and MHA are as follows: pitch (p) = 1.8 μm, aperture diameter in MHA (dMHA) = 0.9 μm, disk diameter in GDA (dGDA) = 1.26 μm, thickness of MHA (tMHA) and GDA (tGDA) = 0.05 μm, BCB thickness (tBCB) = 0.35 μm. The fabrication process has been described in detail elsewhere [4].

 figure: Fig. 1

Fig. 1 Scanning electron microscope (SEM) images and illustration of meta-ARC (GDA on BCB layer) on MHA structure with fabrication induced alignment-shift and shape-variation. (a) Developed photoresist pattern, followed by e-beam deposition of gold. (b) Nonzero sidewall angled GDs (using a lift-off process with acetone to remove the photoresist layer shown in (a). (c) Cross-sectional view of alignment-shifted GDA to MHA. (d) Schematic view and geometrical parameters of meta-ARC on MHA.

Download Full Size | PDF

3. Fabrication-induced imperfections

As shown in Figs. 1(a)-(c), the shape-variation of GDs (i.e., additional gold thickness tring and nonzero sidewall angle θ) is attributed to the gold deposition on the circular holes of the developed photoresist sample (not vertical sidewall) that defines GD structure or e-beam deposition geometry (distance from the gold source and collimation), followed by a lift-off process. Such fabrication process inevitably results in a sidewall-angled GD shown in Fig. 1(b). Another imperfection, the spatial deviation of GDs is associated with the alignment-shift between GDA and MHA as shown in Fig. 1(c), which results from the alignment errors to be about 0.5 μm (Karl SUSS MA/BA6 Mask Aligner was used to expose the photoresist). We performed numerical simulations to investigate the effect of shape-variation and alignment-shift using CST Microwave Studio, which utilizes a finite integration technique. Figure 1(d) shows the configuration of meta-ARC:MHA used in simulations, where the shape-variation and alignment-shift are described by parameters (tring,θ) and (a,b), respectively.

We vary the shape of GDs with additional gold thickness tring and sidewall-angle θ, and numerically simulated the transmission of the meta-ARC:MHA sample (it can be seen by SEM images that tring ~0.02 μm and θ45°). Figure 2 shows the obtained transmission spectra for (tring, θ) = (0 μm, 0°), (0.02 μm, 0°) and (0.02 μm, 45°) in addition to the difference in transmission between ideal-shaped GDs and imperfect-shaped GDs given by Δi=[(TiT(0,0))/T(0,0)]×100, where i= (0.02 μm, 0°) and (0.02 μm, 45°). At two peaks corresponding to the first (second) order SPP resonance, 6.32 μm (4.38 μm), the transmission difference is found to be −0.1% (−0.26%) and 1.0% (−3.2%) for Δ(0.02,0) and Δ(0.02,45), which is owing to very small ratio of fabrication-induced shape-variation to resonance wavelength (for instance, 0.02 μm / 6.3 μm 0.003).

 figure: Fig. 2

Fig. 2 Simulated transmission spectra due to shape-variation of GDs. Δi (symbols) represents the difference in transmission between meta-ARC:MHA with ideal-shaped and imperfectly-shaped GDs. Shape-variation is described with (tring, θ), where tring is additional gold thickness (0 and 0.02 μm) and θ is the sidewall-angle (0° and 45°) of GDs as displayed in Fig. 1(d). The color convention is as follows: black for (0 μm, 0°), red for (0.02 μm, 0°) and blue for (0.02 μm, 45°).

Download Full Size | PDF

In order to estimate the influence of alignment-shift of GDs on the performance of meta-ARC, we observe the difference of transmission enhancement factors (TEF) between experiment and simulation with various (a, b) as shown in Fig. 3. TEF is defined as TmetaARC:MHA/TMHA with TmetaARC:MHA (TMHA) being the transmission of meta-ARC:MHA (MHA). (a, b) is varied from perfectly aligned (a=0p and b=0p) to completely alignment-shifted GDs (a=0.5p and b=0.5p), where p is the periodicity of GDA. Detailed experimental results have been reported in Ref [4]. In Fig. 3, black (blue) symbols represent the alignment-shifted GDs with a0.5 μm and 0b0.9 μm with y-axis on the left side (a0.7 μm and 0b0.9 μm with y-axis on the right side), respectively. In addition, TEF difference at transmission peak wavelength of 6.32 μm (the first-order SPP) and 4.38 μm (the second-order SPP) related to the first-order (second-order) SPP resonance is presented in the Fig. 3(a) and 3(b), respectively. As (a, b) increases from (0, 0) to (0.9 μm, 0.9 μm), three smallest difference of TEF at the first (second) SPP resonance can be found to be 0.1 (0.05), 0.08 (−0.15) and 0.07 (−0.08) at (a, b) = (0.3 μm, 0.7 μm), (0.3 μm, 0.9 μm) and (0.5 μm, 0.5 μm), respectively. Normally incident x-polarized light was used for simulation as indicated in Fig. 1(d). However, in the experiment, an unpolarized incident light was used to measure the transmission of meta-ARC:MHA (i.e., average of transmission intensity for x- and y-polarizations). Therefore, it is of great importance to consider the averaged difference of TEF with (a, b) and (b,a), i.e., averaged difference of [(0.3, 0.7), (0.7, 0.3)], [(0.3, 0.9), (0.9, 0.3)] and [(0.5, 0.5), (0.5, 0.5)]. Simulated TEF with alignment-shifted GDs by 0.5 μm in x- and y-direction is found to be closest to the measured TEF. This analysis indicates a good quantitative agreement with alignment-shifted GD measured by SEM image in Fig. 1(c).

 figure: Fig. 3

Fig. 3 Difference of transmission enhancement factors (TEFs) between experiment and simulation at (a) first-order and (b) second-order SPP resonance as a function of (a, b) resulting from the alignment-shifted GDA and MHA.

Download Full Size | PDF

4. E-field distribution and discussion

To better understand the influence of the spatial deviation of GDA, we investigate how the transmission, x- and z-components of electric field intensity change for selected four types of samples, namely the MHA on GaAs substrate (indicated as Sample A in Fig. 4, black), a BCB layer coated MHA (Sample B, green), alignment-shifted meta-ARC on MHA (Sample C, red; Sample D, blue), and a perfectly aligned meta-ARC on MHA (Sample E, light blue), where a 0.35 μm thick BCB layer is applied to the Samples B-E. Here, two extreme cases for the alignment-shifted GDs are (i) (a, b) = (0.5p, 0.5p) (i.e., GDs of Sample C are relatively off from the center of apertures in MHA layer by a half periodicity in both x- and y-directions) and (ii) (a, b) = (0.5p,0p) (i.e., GDs of Sample D are at a distance of 0.5p apart in only x-direction from the center of holes in MHA layer). The simulated transmission spectra of the above-mentioned Samples A-E are plotted in Fig. 4(a). The transmission enhancement ratio (defined as Tj/TA, where j = B, C, D, E) at the first-order SPP resonance is found to be ~1.12, ~1.44, ~1.61 and ~1.83, respectively, in addition that the transmission peak of sample A is only slightly shifted by less than 0.09 μm (Δλ/λ=0.09 μm / 6.27 μm ≈1.4%) even with BCB layer and meta-ARC layer. The advantage drawn from these results is that our meta-ARC can be readily utilized with MHA integrated optoelectronic device to improve the device-performance because the peak wavelength of MHA resulting from SPP resonance is not shifted and the fabrication tolerance (spatial variation of GDs) can be allowed in some degree (transmission of Sample C, the maximally alignment-shifted GDs atop the BCB layer on MHA, is obtained with ~44% increase, as compared with Sample A).

 figure: Fig. 4

Fig. 4 (a) Simulated normal incidence transmission spectra of MHA on GaAs substrate (Sample A, black), a BCB coated MHA (Sample B, green), the maximally alignment-shifted meta-ARC on MHA (Sample C, red: (a, b) = (0.9 μm, 0.9 μm)), the alignment-shifted meta-ARC on MHA (Sample D, blue: (a, b) = (0.9 μm, 0 μm)), and a perfectly aligned meta-ARC on MHA (Sample E, light blue: (a, b) = (0 μm, 0 μm)). Here, a 0.35 μm BCB layer was used for sample B-E. Inset shows that the schematic views of sample A-E. The dash and solid black lines denote the MHA and GDA position, respectively. (b) Diagram of sample structure to calculate the electric field intensity (|Eij(z)|2Iij(z), where i=x,  z and j= Sample A-E). (c) Calculated |Ez| at the half thickness of BCB layer (z=20.2 μm), y/p=0, 0.25, 0.35, 0.5 along x-direction and the simulated Ez distributions of sample C-E at each of the first-order SPP resonance wavelength in the plane y=0. (d) Enhancement ratio of x-component of electric field intensity at the first-order SPP resonance, at the position of x = 0 μm and y = 0 μm in z -direction, as compared with Ix(z) of Sample A (Ixj(z)/IxA(z), where j = B, C, D and E samples) (e) z-component of electric field intensity at x = 0.54 μm and y = 0 μm in z-direction for all samples (Izj(z), where j = Samples A-E), mainly displayed with a range of penetration depth (~1/e2). Inset shows that the enhancement ratio of transmission (vs. Sample A, MHA) is well matched with one calculated using Ix (z=0 μm).

Download Full Size | PDF

Next, we investigate x- and z-components of electric field at the first-order SPP resonance along z-direction of GaAs substrate (Ex(z)) and along x-direction in the middle of BCB layer (Ez(x)) when the direction of polarized incoming light is parallel to x-axis. First, we observe Ez(x) for spatial deviation of GDA (Samples C-E) at z=20.2 μm (at the half thickness of BCB layer, as indicated in Fig. 4(b)) and y/p=0, 0.25, 0.35, 0.5 (at the center, edge of hole in MHA, at the edge of the disk in GDA, the half pitch, respectively). The simulated Ez(x) of Samples D, E (C) tends to decrease (increase) as y/p increases. This tendency can be qualitatively interpreted by the accumulated charges and electric field strength at various y-position: the maximum accumulated charges and electric field intensity are around the hole / disk edges, and they are decreased as the distance from x-axis increases (location-dependent due to spatial variation of GDA). Moreover, Ez can be considered as the coupling (η) between GDA layer and MHA layer, which is clearly seen in Ez distribution (in the plane y=0) of Sample C-E as displayed in the right panel of Fig. 4(c). For y/p=0, as the spatial deviation gradually decreases (i.e., x, y=p/2 for Sample C; x=p/2 for Sample D; x, y=0 for Sample E), the interaction between MHA and GDA is stronger (ηE>ηD>ηC), which results in enhancing / suppressing the transmission / reflection. The channel-waveguide properties of the hole are known to influence the transmission properties, i.e., changing the coupling strength between localized SP (GDA) and propagating SP (MHA) layers is attributed to enhanced transmission. In addition, the amplitude and phase for Sample E can be calculated using the multiple-layer model (based on a transfer matrix method) developed in our previous publications, satisfying the perfect antireflection condition. Second, Figs. 4(d) and 4(e) show the calculated x- and z-component electric field intensity (|Eij(z)|2Iij(z), where i=x, z and j = Samples A-E), respectively. The IxA(z) at the center of MHA’s aperture (x = 0 μm and y = 0 μm) from z = 20 μm to z = 0 μm as displayed in Fig. 4(d) (the interface between MHA and GaAs substrate is set to 20 μm in z-axis) and ~13.69 V2/μm2 was obtained at z=0 μm. Also, the enhancement ratio is defined as Ixj/IxA, where j = Samples B-E and it is found to be ~1.11 (B), ~1.43 (C), ~1.60 (D), ~1.81 (E) at z=0 μm. The Izj(z) at x= 0.54 μm and y=0 μm (j = Samples A-E) and ~1/e2 decay are apparent from Fig. 4(e). The inset of Fig. 4(e) shows that the ratio of the transmitted light intensity of Samples B-E to that of Sample A (=Tj/TA, black circle) is in excellent agreement with the enhancement ratio calculated using Ex (=Ixj/IxA, red cross).

5. Conclusion

We have numerically investigated the impacts of resonance coupling between GDA and MHA layers on the performance of meta-ARC at the SPP resonance wavelengths through alignment-shifts and shape-variations. As compared to the ideal-shaped GDs, the change in transmission intensity due to fabrication-induced shape-variation of GDs (with a slightly angled sidewall and additionally deposited thin gold) was found to be ~0.1% and ~0.26% at the first and second order SPP resonance wavelengths, respectively. Also, the first order SPP-transmitted light can be increased by more than 44% even with the maximum change of resonance coupling between GDA and MHA. Our results drawn from this work can be used to as a basis to realize the practical, easy-to-fabricate, structurally tunable antireflection coating layer for next generation of optoelectronic devices associated with a variety of plasmonic architectures.

Funding

The KRISS portion of this work was supported by Nano-Material Fundamental Technology Development Program (2018069993) through the National Research Foundation of Korea (NRF) funded by Ministry of Science and ICT, the KRISS grant GP2019-0015-03 and the AOARD grant FA2386-14-1-4094 funded by the U.S. government (AFOSR/AOARD). The USF portion of this work was supported by the Alfred P. Sloan Research Fellow grant FG-BR2013-123, the KRISS grant GP2018-0023 and the AOARD grant FA2386-18-1-4104 funded by the U.S. government (AFOSR/AOARD).

References

1. C. M. Watts, X. Liu, and W. J. Padilla, “Metamaterial Electromagnetic Wave Absorbers,” Adv. Mater. 24(23), OP98–OP120 (2012). [PubMed]  

2. N. I. Landy, S. Sajuyigbe, J. J. Mock, D. R. Smith, and W. J. Padilla, “Perfect Metamaterial Absorber,” Phys. Rev. Lett. 100(20), 207402 (2008). [CrossRef]   [PubMed]  

3. K. Bhattarai, S. Silva, K. Song, A. Urbas, S. J. Lee, Z. Ku, and J. Zhou, “Metamaterial perfect absorber analyzed by a meta-cavity model consisting of multilayer metasurfaces,” Sci. Rep. 7(1), 10569 (2017). [CrossRef]   [PubMed]  

4. J. Jeon, K. Bhattarai, D.-K. Kim, J. O. Kim, A. Urbas, S. J. Lee, Z. Ku, and J. Zhou, “A Low-loss Metasurface Antireflection Coating on Dispersive Surface Plasmon Structure,” Sci. Rep. 6(1), 36190 (2016). [CrossRef]   [PubMed]  

5. H.-T. Chen, J. Zhou, J. F. O’Hara, F. Chen, A. K. Azad, and A. J. Taylor, “Antireflection Coating Using Metamaterials and Identification of Its Mechanism,” Phys. Rev. Lett. 105(7), 073901 (2010). [CrossRef]   [PubMed]  

6. A. Kabiri, E. Girgis, and F. Capasso, “Buried Nanoantenna Arrays: Versatile Antireflection Coating,” Nano Lett. 13(12), 6040–6047 (2013). [CrossRef]   [PubMed]  

7. H. Cheng, Z. Liu, S. Chen, and J. Tian, “Emergent Functionality and Controllability in Few-Layer Metasurfaces,” Adv. Mater. 27(36), 5410–5421 (2015). [CrossRef]   [PubMed]  

8. H.-T. Chen, “Interference theory of metamaterial perfect absorbers,” Opt. Express 20(7), 7165–7172 (2012). [CrossRef]   [PubMed]  

9. C. Genet and T. W. Ebbesen, “Light in tiny holes,” Nature 445(7123), 39–46 (2007). [CrossRef]   [PubMed]  

10. W. L. Barnes, A. Dereux, and T. W. Ebbesen, “Surface plasmon subwavelength optics,” Nature 424(6950), 824–830 (2003). [CrossRef]   [PubMed]  

11. W. Fan, S. Zhang, B. Minhas, K. J. Malloy, and S. R. J. Brueck, “Enhanced Infrared Transmission through Subwavelength Coaxial Metallic Arrays,” Phys. Rev. Lett. 94(3), 033902 (2005). [CrossRef]   [PubMed]  

12. R. Stanley, “Plasmonics in the mid-infrared,” Nat. Photonics 6(7), 409–411 (2012). [CrossRef]  

13. K. Bhattarai, S. Silva, A. Urbas, S. J. Lee, Z. Ku, and J. Zhou, “Angle-Dependent Spoof Surface Plasmons in Metallic Hole Arrays at Terahertz Frequencies,” IEEE J. Sel. Top. Quantum Electron. 23(4), 1–6 (2017). [CrossRef]  

14. S. J. Lee, Z. Ku, A. Barve, J. Montoya, W.-Y. Jang, S. R. J. Brueck, M. Sundaram, A. Reisinger, S. Krishna, and S. K. Noh, “A monolithically integrated plasmonic infrared quantum dot camera,” Nat. Commun. 2(1), 286 (2011). [CrossRef]   [PubMed]  

15. Z. Ku, W.-Y. Jang, J. Zhou, J. O. Kim, A. V. Barve, S. Silva, S. Krishna, S. R. J. Brueck, R. Nelson, A. Urbas, S. Kang, and S. J. Lee, “Analysis of subwavelength metal hole array structure for the enhancement of back-illuminated quantum dot infrared photodetectors,” Opt. Express 21(4), 4709–4716 (2013). [CrossRef]   [PubMed]  

16. H. A. Atwater and A. Polman, “Plasmonics for improved photovoltaic devices,” Nat. Mater. 9(3), 205–213 (2010). [CrossRef]   [PubMed]  

17. J. Rosenberg, R. V. Shenoi, T. E. Vandervelde, S. Krishna, and O. Painter, “A multispectral and polarization-selective surface-plasmon resonant midinfrared detector,” Appl. Phys. Lett. 95(16), 161101 (2009). [CrossRef]  

18. L. Rao, D. Yang, L. Zhang, T. Li, and S. Xia, “Design and experimental verification of terahertz wideband filter based on double-layered metal hole arrays,” Appl. Opt. 51(7), 912–916 (2012). [CrossRef]   [PubMed]  

19. C.-C. Chang, Y. D. Sharma, Y.-S. Kim, J. A. Bur, R. V. Shenoi, S. Krishna, D. Huang, and S.-Y. Lin, “A Surface Plasmon Enhanced Infrared Photodetector Based on InAs Quantum Dots,” Nano Lett. 10(5), 1704–1709 (2010). [CrossRef]   [PubMed]  

20. F. Zhao, C. Zhang, H. Chang, and X. Hu, “Design of Plasmonic Perfect Absorbers for Quantum-well Infrared Photodetection,” Plasmonics 9(6), 1397–1400 (2014). [CrossRef]  

21. C. Zhang, H. Chang, F. Zhao, and X. Hu, “Design principle of Au grating couplers for quantum-well infrared photodetectors,” Opt. Lett. 38(20), 4037–4039 (2013). [CrossRef]   [PubMed]  

22. M.-S. Park, K. Bhattarai, D.-K. Kim, S.-W. Kang, J. O. Kim, J. Zhou, W.-Y. Jang, M. Noyola, A. Urbas, Z. Ku, and S. J. Lee, “Enhanced transmission due to antireflection coating layer at surface plasmon resonance wavelengths,” Opt. Express 22(24), 30161–30169 (2014). [CrossRef]   [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 (4)

Fig. 1
Fig. 1 Scanning electron microscope (SEM) images and illustration of meta-ARC (GDA on BCB layer) on MHA structure with fabrication induced alignment-shift and shape-variation. (a) Developed photoresist pattern, followed by e-beam deposition of gold. (b) Nonzero sidewall angled GDs (using a lift-off process with acetone to remove the photoresist layer shown in (a). (c) Cross-sectional view of alignment-shifted GDA to MHA. (d) Schematic view and geometrical parameters of meta-ARC on MHA.
Fig. 2
Fig. 2 Simulated transmission spectra due to shape-variation of GDs. Δ i (symbols) represents the difference in transmission between meta-ARC:MHA with ideal-shaped and imperfectly-shaped GDs. Shape-variation is described with ( t ring , θ), where t ring is additional gold thickness (0 and 0.02 μm) and θ is the sidewall-angle (0° and 45°) of GDs as displayed in Fig. 1(d). The color convention is as follows: black for (0 μm, 0°), red for (0.02 μm, 0°) and blue for (0.02 μm, 45°).
Fig. 3
Fig. 3 Difference of transmission enhancement factors (TEFs) between experiment and simulation at (a) first-order and (b) second-order SPP resonance as a function of ( a, b) resulting from the alignment-shifted GDA and MHA.
Fig. 4
Fig. 4 (a) Simulated normal incidence transmission spectra of MHA on GaAs substrate (Sample A, black), a BCB coated MHA (Sample B, green), the maximally alignment-shifted meta-ARC on MHA (Sample C, red: ( a, b) = (0.9 μm, 0.9 μm)), the alignment-shifted meta-ARC on MHA (Sample D, blue: ( a, b) = (0.9 μm, 0 μm)), and a perfectly aligned meta-ARC on MHA (Sample E, light blue: ( a, b) = (0 μm, 0 μm)). Here, a 0.35 μm BCB layer was used for sample B-E. Inset shows that the schematic views of sample A-E. The dash and solid black lines denote the MHA and GDA position, respectively. (b) Diagram of sample structure to calculate the electric field intensity ( | E i j (z) | 2 I i j (z), where i=x,  z and j= Sample A-E). (c) Calculated | E z | at the half thickness of BCB layer ( z=20.2 μm), y/p=0, 0.25, 0.35, 0.5 along x-direction and the simulated E z distributions of sample C-E at each of the first-order SPP resonance wavelength in the plane y=0. (d) Enhancement ratio of x-component of electric field intensity at the first-order SPP resonance, at the position of x = 0 μm and y = 0 μm in z -direction, as compared with I x (z) of Sample A ( I x j (z)/ I x A (z), where j = B, C, D and E samples) (e) z-component of electric field intensity at x = 0.54 μm and y = 0 μm in z-direction for all samples ( I z j (z), where j = Samples A-E), mainly displayed with a range of penetration depth ( ~1/ e 2 ). Inset shows that the enhancement ratio of transmission (vs. Sample A, MHA) is well matched with one calculated using I x  (z=0 μm).
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.