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

Planar dielectric cylindrical lens at 800 nm and the role of fabrication imperfections

Open Access Open Access

Abstract

Conventional optical components have been proposed to realize high-quality line focusing with uniform intensity distribution such as cylindrical lenses, segmented wedge-arrays, or a combination of prisms and spherical mirrors. Numerous factors such as the manufacturing tolerances or the need for precise alignment of conventional lenses cause wave front aberrations that impact the performance of optical systems. These aforementioned limitations affect the uniformity of the intensity distribution and the intercept factor of lenses. Here, we experimentally demonstrate an integrable planar dielectric cylindrical lens made of titanium dioxide for uniform line focusing and discuss the sensitivity of its performance to fabrication imperfections originating from non-ideal geometrical parameters. The lens has a numerical aperture of 0.247, an intercept factor of 0.85, and an efficiency of 79% at 800 nm.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

In recent years, efforts to miniaturize and improve the performance of microscale sized optical lenses have allowed the development of different imaging systems for the characterization of complex media. These systems, now accessible on almost all commercial spectrometers, have opened new fields of investigation, making it easier to characterize inhomogeneous samples. One can probe into two or even three-dimensional objects of macroscopic size (at the scale of mm or cm), while keeping a spatial resolution of the order of a micrometer, limited mainly by diffraction. Cylindrical lenses [1], random phase plates [2], segmented wedge arrays [3], and scanning mirrors [4] are often used as line focus generators for applications such as Raman spectroscopy [5] or X-rays imaging [6]. Metasurfaces have been investigated as potential alternatives for integrated optical free space components [7-11]. Metasurfaces are subwavelength nanostructured devices that enable the control of optical wave fronts, polarization, and phase to build a large variety of flat optical components [12-29]. To date, a large collection of metasurfaces have been proposed including metasurfaces that focus an incident circularly polarized light to line [29–34]. Efforts to maximize the efficiency of metasurfaces need not only to perfectly implement the theoretical geometries but also to account for the sensitivity of the results to fabrication imperfections. Indeed most experimentally reported metasurfaces have exhibited performances smaller than the theoretically predicted ones. We propose and experimentally demonstrated a polarization independent planar cylindrical lens that focuses a normal incident wave to a line and systematically analyze the sensitivity of the devices to fabrication. Using numerical simulations, we investigate the influence of relevant geometrical parameters on the performance of the device and classify the parameters with respect to their influence on the intercept factor of the lens. The proposed lens generates line focusing with intercept factor larger than 0.85. The line-focusing planar lens paves the way to applications such hyperspectral Raman spectroscopy, line-focusing planar solar collectors, or wave probes.

2. Numerical simulation

To achieve the desired line-focusing, the phase profile of the wave front as a function of position 𝑥 along the metasurface lens must satisfy the parabolic equation [23]:Φ=k0(fx2+f2) where k0 is the free space wave-vector, x is the distance from the considered element to the center of the metasurface lens and f is the focal length. The metasurface is made of TiO2 cylinders and operates at 800 nm wavelength.

The inset of Fig. 1 (a) shows the dielectric metasurfaces unit cell used to implement line focusing. It consists of a TiO2 cylindrical resonator on a SiO2 (hs = 440 nm) – silver (hg = 100 nm) multilayer. Silicon dioxide (SiO2) is the spacer and silver (Ag) [35] serves as ground plane for the lens to work in reflection. The thickness (h) of the cylindrical resonators is fixed to 250 nm and the radius (R) changes to achieve the required phase. The metasurface is periodic along the y direction with a sub-wavelength unit cell (px = py = 510 nm). Due to the low loss of glass and TiO2 at the desired frequency (λ = 800 nm), the metasurface is considered lossless. Using ellipsometry, the refractive indices of TiO2 and the SiO2 were measured and found to be ~2.325 and ~1.49 respectively (average of 5 measurements). To obtain the required 2π phase-shift (see Fig. 1), the scattering parameters of the resonators are calculated using the commercial software CST with periodic boundary conditions. Figure 1(a) shows that the 2π phase-shift is achieved by changing the radius from 90 nm to 200 nm. We designed a cylindrical metasurface lens with a focal length of 400 µm and a size of 200 µm by 200 µm. By interpolating Fig. 1 and using the phase equation, the radius of each element fitting the required phase is obtained (see Fig. 1 (b)). To have a line focus along the y direction, elements in the y-direction are identical with a periodicity py, and, the radius of resonators is varied only along the x direction.

 figure: Fig. 1

Fig. 1 (a) Phase shift for different radii and schematic of the unit cylindrical elements. P (period), h (thickness of the resonator), R (radius), hs (thickness of the spacer), hg (thickness of the ground plane). Materials: TiO2 (brown), SiO2 (blue), Ag (grey). (b) Phase distribution of a planar cylindrical lens under normal incidence with a 40 µm focal length and a 200 µm by 200 µm aperture size (blue). Required radius (green).

Download Full Size | PDF

3. Fabrication

Figure 2 illustrates the fabrication flowchart of the structures using top-down deposition and etching processes. The dielectric planar cylindrical lens is fabricated on a BK7 glass substrate [36]. It should be noted that the BK7 substrate has a crystallographic mesh parameter incompatible with that of gold or silver, resulting in a quasi-zero adhesion to the substrate. Therefore, we deposit, using electron beam metal deposition, a thin layer (5 nm) of Germanium (Ge) on the substrate since Ge improves adhesion and surface smoothness. A silver metal layer of 100 nm (larger than the skin depth of light in the visible) is then deposited without opening the chamber by changing the crucible [Fig. 2(a)]. Then, a 440 nm thick SiO2 layer was deposited on the Ag layer using PECVD (Plasma-enhanced chemical vapor deposition) as shown in Fig. 2(b). An electron beam resist (PMMA), is spun on the sample followed by baking at 170 °C in an oven. The metasurfaces pattern is written in the resist using electron beam lithography (EBL) [Fig. 2(c)] on an area of 200 µm by 200 µm and subsequently developed in solution to remove the exposed e-beam resist (EBR). This pattern is the inverse of our final metasurfaces. The exposed sample is transferred to an atomic layer deposition (ALD) chamber producing amorphous TiO2 while avoiding the contamination of the ALD chamber by the EBR. The ALD process (TiCl4 precursor) deposits 400 nm of TiO2 so that all features are filled with TiO2 [Fig. 2(d)]. The residual TiO2 film that coats the top surface of the resist is removed by RIE (reactive-ion-etching) process [Fig. 2(e)]. After removing PMMA, periodic TiO2 metasurfaces were obtained [Fig. 2(f)].

 figure: Fig. 2

Fig. 2 Fabrication process: (a) metal deposition, (b) SiO2 deposition using PECVD, (c) E-beam lithography (d) Thin layer TiO2 deposition using ALD, (e) TiO2 RIE process to expose the underlying PMMA pattern, (f) removal of the PMMA. (g) Top view SEM images of metasurfaces and zoom-in containing 6 x 5 cylindrical structures imaged in the xy plane and clearly evidencing the gradient in structures size. (h) Atomic Force Microscope (AFM) measurements of the shape at the top a cylinder.

Download Full Size | PDF

Scanning electron microscope (SEM) images, presented in Fig. 2(g), shows successful fabrication of the structures with good quality. The fabrication process of Fig. 2 results in a height of TiO2 slightly higher on top of PMMA, resulting in curved TiO2 particles when PMMA is removed. The flatness of the top of the fabricated cylindrical elements is characterized using AFM measurements (Fig. 2(h)). These measurements confirm a curvature of the top of the cylinders resulting from the ALD deposition process with δ = 25 nm.

Non-ideal particle shapes or fabrication imperfections are unavoidable, and, they may limit the performance of optical devices. In order to quantify the impact of fabrications imperfections on line-focusing for the fabricated metasurface, numerical simulations were performed. The investigated geometrical parameters are the curvature (δ) on the top of the cylinders and the radius of the cylinders (R). Figure 3 (a) shows the simulated phase distribution for periodic structures with different δ and R. We set R = 200 nm and δ = 0 as phase reference. Figure 3 (b) shows phase from R = 90 nm to R = 200 nm with δ = 0 nm and δ = 25 nm. One can observe that the curvature affects less resonators of small radius. The phase variation as function curvature (δ) for different R is presented in Fig. 3 (c). One can see that the maximum phase variation is about 30° at about R = 158 nm. Figure 3 (d) presents the profile of Fig. 3 (c) at R = 150 nm and clearly shows that the phase variation increases with δ. To further investigate the impact of such fabrication imperfections on the metasurface efficiency, the intercept factor is analyzed. The intercept factor is defined as the ratio of the integrated power within the focal spot with fabrication imperfections to the one with no fabrication imperfections [23]. Fabrication imperfections are modelled as a random fabrication noise (ε |ΔR|) that adds to the ideal dimensions. Hence, the total phase of each element is given by, R_real = R + ε |ΔR|, where R_real is the real dimension with fabrication imperfections, R is the ideal dimension. ε is a random number between −0.5 and 0.5, chosen from a uniform distribution. |ΔR| is the magnitude of fabrication imperfection. Dimension mismatches will cause phase mismatches that reduce the intercept factor. Modeling each resonator as a point source with the given parabolic phase with phase noise, we calculated 100 metasurfaces with random fabrication imperfections and obtained an average intercept factor. Figure 3 (e) presents the average intercept factor for different curvature (δ) and different magnitude of fabrication imperfection (|ΔR|) and Fig. 3 (f) is the profile with δ = 0 nm and δ = 25 nm. The intercept factor is more sensitive to |ΔR| and it drops to about 0.85 when magnitude of fabrication imperfection is about 10 nm.

 figure: Fig. 3

Fig. 3 (a) Phase distribution for a periodic structure for as a function of the radius (R) and the curvature (δ). (b) The phase profile of (a) at δ = 0 (nm) and δ = 25(nm). (c) The phase variation due to the curvature (δ) with different radius (R). (d) The phase variation with different δ at R = 150 nm. (e) Average intercept factor as function of fabrication imperfections (|ΔR|) and curvature (δ). (f) Average intercept factor as function of fabrication imperfection (|ΔR|) at δ = 0 nm and δ = 25 nm.

Download Full Size | PDF

The calculated intercept factor in Fig. 3 (f) shows that the metasurfaces are robust towards fabrication imperfections (radius and curvature of holes). This robustness can be understood by analyzing the confinement of optical modes in the resonators. Figure 4 presents the field distribution in resonators of various size at λ = 800 nm. One can observe that while the field is barely confined in the curved region for small radius (radius = 90 nm), the field is strongly confined in the center of the curved region when the radius is 160 nm. This explains the small sensitivity of smaller particles to the curvature of the resonators.

 figure: Fig. 4

Fig. 4 Total electric field (x-polarization) for different radii without fabrication imperfections (δ = 0 nm): (a) Radius R = 90 nm, (b) R = 160 nm, (c) R = 200 nm. Total electric field distribution for different radii with fabrication imperfections (δ = 25 nm): (d) Radius R = 90 nm, (e) R = 160 nm, (f) R = 200 nm.

Download Full Size | PDF

4. Experiment

To optically characterize the fabricated lenses, a custom setup was used. The experimental setup is composed of two main systems dedicated to illumination and imaging. The illumination system comprises a supercontinuum laser (NKT photonics) and an acousto-optic tunable filter (Super K) to select the operating wavelength (800 nm). A beam at λ = 800 nm was transmitted through a broadband polarizing beam (50:50) splitter to illuminate the lens. The focused reflected beam is collected by the imaging system via the beam splitter. The imaging system consists of an extra-long working distance objective (x50), a corresponding lens, and a camera. Due to the limited space between the objective and the lens, an extra-long working distance (WD = 10 mm) was used. The lens was also mounted on a stage to adjust the distance between the lens and objective (50 × objective).

A clear and uniform line focus is observed for both the x and y polarization demonstrating polarization independence. The measured intensity profile illustrates the line focus at the focal distance (0.4 mm). The small discrepancies between numerical simulations and experimental results are mainly attributed to the fabrication imperfections inducing slight non-uniformity in experimental results. To better quantify the uniformity of the focus, we calculated full width at half maximum (FWHM) of light intensity along x for different values of y. The average and standard deviation of the FWHM are then calculated and the uniformity is defined as 1 – standard deviation / average. Using plots in Fig. 5, the average FWHM is 1.85 and the standard deviation is 0.12 resulting in field uniformity of 93%. In our design at the wavelength of 800 nm, the numerical aperture (NA) is 0.247, and the diffraction limit is thus d = 1.65 µm.

 figure: Fig. 5

Fig. 5 (a), (b), and (c) present numerical simulations of the normalized electric field (x-y plane) for x, y polarizations and normalized electric field in x-z plane respectively. (d) and (e) present the experimental normalized electric field intensity at the focus in the xy plane for x and y polarizations. (f) compares the normalized intensity profile along x in numerical simulation and in experiment.

Download Full Size | PDF

5. Summary

We have experimentally demonstrated a polarization independent planar dielectric cylindrical lens made of TiO2 for line focusing. The fabrication method results in curved cylindrical resonators with a depth of about 25 nm that was experimentally characterized using an atomic force microscope (AFM). We have quantified the impact of fabrication imperfections (including both the curved shape and non-ideal radius) on the performance of the device. The intercept factor of the lens, i.e., the ratio of the amount of energy at the focus with and without fabrication imperfections revealed that the radius of resonators plays a more critical role than the curvature at the top of the resonators. The reported lens not only provides line focusing but also uniform intensity distribution along a line with a width mainly limited by diffraction. This work will pave the way to detailed analysis of nanoscale devices accounting for various geometrical imperfections originating from nano-structuring and to a selection of architectures based on sensitivities for various applications.

Funding

U.S. Department of Energy (Award Number DE-EE0007341).

Acknowledgements

The work was performed in part at the San Diego Nanotechnology Infrastructure, a member of the National Nanotechnology Coordinated Infrastructure, which is supported by NSF grant ECCS-1542148.

References

1. W. Chen, S. Wang, C. Mao, B. Chen, and A. Xu, “Cylinder lens array line focus system for x-ray laser experiments,” in Conference on Lasers and Electro-Optics (1990).

2. S. N. Dixit, I. M. Thomas, B. W. Woods, A. J. Morgan, M. A. Henesian, P. J. Wegner, and H. T. Powell, “Random phase plates for beam smoothing on the Nova laser,” Appl. Opt. 32(14), 2543–2554 (1993). [CrossRef]   [PubMed]  

3. B. Lü, B. Zhang, and B. Cai, “Focusing of a Gaussian Schell-model beam through a circular lens,” J. Mod. Opt. 42(2), 289–298 (1995). [CrossRef]  

4. H. Schenk, P. Dürr, T. Haase, D. Kunze, U. Sobe, H. Lakner, H. Kück, and J. Select, “Large deflection micromechanical scanning mirrors for linear scans and pattern generation,” Top. Quantum Electron. 6(5), 715–722 (2000). [CrossRef]  

5. K. Watanabe, A. F. Palonpon, N. I. Smith, L. D. Chiu, A. Kasai, H. Hashimoto, S. Kawata, and K. Fujita, “Structured line illumination Raman microscopy,” Nat. Commun. 6(1), 10095 (2015). [CrossRef]   [PubMed]  

6. Y. Higuchi and J. Hanari, “Fingerprint image input apparatus including a cylindrical lens,” US Patents. 5, (1992).

7. F. Falcone, T. Lopetegi, M. A. G. Laso, J. D. Baena, J. Bonache, M. Beruete, R. Marqués, F. Martín, and M. Sorolla, “Babinet Principle Applied to the Design of Metasurfaces and Metamaterials,” Phys. Rev. Lett. 93(19), 197401 (2004). [CrossRef]   [PubMed]  

8. B. Kanté, J. M. Lourtioz, and A. de Lustrac, “Experimental demonstration of a nonmagnetic metamaterial cloak at microwave frequencies,” Phys. Rev. B 80(20), 205120 (2009). [CrossRef]  

9. N. Yu and F. Capasso, “Flat optics with designer metasurfaces,” Nat. Mater. 13, 139 (2014).

10. A. Arbabi, Y. Horie, M. Bagheri, and A. Faraon, “Dielectric metasurfaces for complete control of phase and polarization with subwavelength spatial resolution and high transmission,” Nat. Nanotechnol. 10(11), 937–943 (2015). [CrossRef]   [PubMed]  

11. F. Aieta, M. A. Kats, P. Genevet, and F. Capasso, “Multiwavelength achromatic metasurfaces by dispersive phase compensation,” Science 347(6228), 1342–1345 (2015). [CrossRef]   [PubMed]  

12. A. V. Kildishev, A. Boltasseva, and V. M. Shalaev, “Planar Photonics with Metasurfaces,” Science 339(6125), 1232009 (2013). [CrossRef]   [PubMed]  

13. R. Paniagua-Domínguez, Y. F. Yu, E. Khaidarov, S. Choi, V. Leong, R. M. Bakker, X. Liang, Y. H. Fu, V. Valuckas, L. A. Krivitsky, and A. I. Kuznetsov, “A Metalens with a Near-Unity Numerical Aperture,” Nano Lett. 18(3), 2124–2132 (2018). [CrossRef]   [PubMed]  

14. P. Lalanne and P. Chavel, “Metalenses at visible wavelengths: past, present, perspectives,” Laser Photonics Rev. 3(3), 1600295 (2017). [CrossRef]  

15. N. Yu, F. Aieta, P. Genevet, M. A. Kats, Z. Gaburro, and F. Capasso, “A Broadband, Background-Free Quarter-Wave Plate Based on Plasmonic Metasurfaces,” Nano Lett. 12(12), 6328–6333 (2012). [CrossRef]   [PubMed]  

16. A. Pors, M. G. Nielsen, and S. I. Bozhevolnyi, “Broadband plasmonic half-wave plates in reflection,” Opt. Lett. 38(4), 513–515 (2013). [CrossRef]   [PubMed]  

17. Y. Yang, W. Wang, P. Moitra, I. I. Kravchenko, D. P. Briggs, and J. Valentine, “Dielectric Meta-Reflect array for Broadband Linear Polarization Conversion and Optical Vortex Generation,” Nano Lett. 14, 1394 (2014). [CrossRef]   [PubMed]  

18. W. T. Chen, A. Y. Zhu, V. Sanjeev, M. Khorasaninejad, Z. Shi, E. Lee, and F. Capasso, “A broadband achromatic metalens for focusing and imaging in the visible,” Nat. Nanotechnol. 13(3), 220–226 (2018). [CrossRef]   [PubMed]  

19. S. Wang, P. C. Wu, V. C. Su, Y. C. Lai, M. K. Chen, H. Y. Kuo, B. H. Chen, Y. H. Chen, T. T. Huang, J. H. Wang, R. M. Lin, C. H. Kuan, T. Li, Z. Wang, S. Zhu, and D. P. Tsai, “A broadband achromatic metalens in the visible,” Nat. Nanotechnol. 13(3), 227–232 (2018). [CrossRef]   [PubMed]  

20. X. Ni, Z. J. Wong, M. Mrejen, Y. Wang, and X. Zhang, “An ultrathin invisibility skin cloak for visible light,” Science 349, 1310 (2015).

21. L. Y. Hsu, T. Lepetit, and B. Kanté, “Extremely thin dielectric metasurface for carpet cloaking,” Prog. Electromagnetics Res. 152, 33–40 (2015). [CrossRef]  

22. L. Hsu, M. Dupré, A. Ndao, J. Yellowhair, and B. Kanté, “Local phase method for designing and optimizing metasurface devices,” Opt. Express 25(21), 24974–24982 (2017). [CrossRef]   [PubMed]  

23. L. Hsu, M. Dupré, A. Ndao, and B. Kanté, “From parabolic-trough to metasurface-concentrator: assessing focusing in the wave-optics limit,” Opt. Lett. 42(8), 1520–1523 (2017). [CrossRef]   [PubMed]  

24. C. Simovski, D. Morits, P. Voroshilov, M. Guzhva, P. Belov, and Y. Kivshar, “Enhanced efficiency of light-trapping nanoantenna arrays for thin-film solar cells,” Opt. Express 21(S4Suppl 4), A714–A725 (2013). [CrossRef]   [PubMed]  

25. A. Ndao, R. Salut, M. Suarez, and F. I. Baida, “Plasmonless polarization-selective metasurfaces in the visible range,” J. Opt. 20(4), 045003 (2018). [CrossRef]  

26. L. Cong, S. Tan, R. Yahiaoui, F. Yan, W. Zhang, and R. Singh, “Experimental demonstration of ultrasensitive sensing with terahertz metamaterial absorbers: A comparison with the metasurfaces,” Appl. Phys. Lett. 106(3), 031107 (2015). [CrossRef]  

27. R. Schmidt, A. Slobozhanyuk, P. Belov, and A. Webb, “Flexible and compact hybrid metasurfaces for enhanced ultra high field in vivo magnetic resonance imaging,” Sci. Rep. 7(1), 1678 (2017). [CrossRef]   [PubMed]  

28. J.-H. Park, A. Kodigala, A. Ndao, and B. Kanté, “Hybridized metamaterial platform for nano-scale sensing,” Opt. Express 25(13), 15590–15598 (2017). [CrossRef]   [PubMed]  

29. Ashok Kodigala, Thomas Lepetit, and Boubacar Kanté, “Exceptional points in three-dimensional plasmonic nanostructures,” Phys. Rev. B 94, 201103 (2016).

30. X. Yin, T. Steinle, L. Huang, T. Taubner, M. Wuttig, T. Zentgraf, and H. Giessen, “Beam switching and bifocal zoom lensing using active plasmonic metasurfaces,” Sci. Appl. 6(7), el17016 (2017). [CrossRef]  

31. D. Wen, F. Yue, M. Ardron, and X. Chen, “Multifunctional metasurface lens for imaging and Fourier transform,” Sci. Rep. 6(1), 27628 (2016). [CrossRef]   [PubMed]  

32. S. M. Kamali, A. Arbabi, E. Arbabi, Y. Horie, and A. Faraon, “Decoupling optical function and geometrical form using conformal flexible dielectric metasurfaces,” Nat. Commun. 7, 11618 (2016). [CrossRef]   [PubMed]  

33. P. Wang, N. Mohammad, and R. Menon, “Chromatic-aberration-corrected diffractive lenses for ultra-broadband focusing,” Sci. Rep. 6(1), 21545 (2016). [CrossRef]   [PubMed]  

34. J. Cheng, S. Jafar-Zanjani, and H. Mosallaei, “All-dielectric ultrathin conformal metasurfaces: lensing and cloaking applications at 532 nm wavelength,” Sci. Rep. 6(1), 38440 (2016). [CrossRef]   [PubMed]  

35. E. D. Palik, Handbook of Optical Constants of Solids (Academic, 1985).

36. R. C. Devlin, M. Khorasaninejad, W. T. Chen, J. Oh, and F. Capasso, “Broadband high-efficiency dielectric metasurfaces for the visible spectrum,” Proc. Natl. Acad. Sci. U.S.A. 113(38), 10473–10478 (2016). [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 (5)

Fig. 1
Fig. 1 (a) Phase shift for different radii and schematic of the unit cylindrical elements. P (period), h (thickness of the resonator), R (radius), hs (thickness of the spacer), hg (thickness of the ground plane). Materials: TiO2 (brown), SiO2 (blue), Ag (grey). (b) Phase distribution of a planar cylindrical lens under normal incidence with a 40 µm focal length and a 200 µm by 200 µm aperture size (blue). Required radius (green).
Fig. 2
Fig. 2 Fabrication process: (a) metal deposition, (b) SiO2 deposition using PECVD, (c) E-beam lithography (d) Thin layer TiO2 deposition using ALD, (e) TiO2 RIE process to expose the underlying PMMA pattern, (f) removal of the PMMA. (g) Top view SEM images of metasurfaces and zoom-in containing 6 x 5 cylindrical structures imaged in the xy plane and clearly evidencing the gradient in structures size. (h) Atomic Force Microscope (AFM) measurements of the shape at the top a cylinder.
Fig. 3
Fig. 3 (a) Phase distribution for a periodic structure for as a function of the radius (R) and the curvature (δ). (b) The phase profile of (a) at δ = 0 (nm) and δ = 25(nm). (c) The phase variation due to the curvature (δ) with different radius (R). (d) The phase variation with different δ at R = 150 nm. (e) Average intercept factor as function of fabrication imperfections (|ΔR|) and curvature (δ). (f) Average intercept factor as function of fabrication imperfection (|ΔR|) at δ = 0 nm and δ = 25 nm.
Fig. 4
Fig. 4 Total electric field (x-polarization) for different radii without fabrication imperfections (δ = 0 nm): (a) Radius R = 90 nm, (b) R = 160 nm, (c) R = 200 nm. Total electric field distribution for different radii with fabrication imperfections (δ = 25 nm): (d) Radius R = 90 nm, (e) R = 160 nm, (f) R = 200 nm.
Fig. 5
Fig. 5 (a), (b), and (c) present numerical simulations of the normalized electric field (x-y plane) for x, y polarizations and normalized electric field in x-z plane respectively. (d) and (e) present the experimental normalized electric field intensity at the focus in the xy plane for x and y polarizations. (f) compares the normalized intensity profile along x in numerical simulation and in experiment.
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.