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

Generating ultraviolet perfect vortex beams using a high-efficiency broadband dielectric metasurface

Open Access Open Access

Abstract

Due to the topological charge-independent doughnut spatial structure as well as the association of orbital angular momentums, perfect vortex beams promise significant advances in fiber communication, optical manipulation and quantum optics. Inspired by the development of planar photonics, several plasmonic and dielectric metasurfaces have been constructed to generate perfect vortex beams, instead of conventional bulky configuration. However, owing to the intrinsic Ohmic losses and interband electron transitions in materials, these metasurface-based vortex beam generators only work at optical frequencies up to the visible range. Herein, using silicon nitride nanopillars as high-efficiency half-wave plates, broadband and high-performance metasurfaces are designed and demonstrated numerically to directly produce perfect vortex beams in the ultraviolet region, by combining the phase profiles of spiral phase plate, axicon and Fourier transformation lens based on geometric phase. The conversion efficiency of the metasurface is up to 86.6% at the design wavelength. Moreover, the influence of several control parameters on perfect vortex beam structures is discussed. We believe that this ultraviolet dielectric generator of perfect vortex beams will find many significant applications, such as high-resolution spectroscopy, optical tweezer and on-chip communication.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Due to the helical phase profile with a unique capacity of carrying orbital angular momentum (OAM), optical vortex beams have received a wide applications in optics manipulation and information processing and particularly in optical communication [16]. OAM-based mode division multiplexing scheme effectively enhances information carrying capacity of photons, satisfying the demand of high data transmission for optical communication system. However, the annular intensity pattern highly relies on the topological charge (TC, number of twist in a wave front per unit wavelength) of vortex beams, resulting in serious difficulty in coupling the vortex beams with different TCs into single air-core fiber [7,8]. Thus it is worthwhile to control the ring diameter of vortex intensity profile at will not only for optics communication but also for nano-particle manipulation.

The concept of “perfect” vortex (PV) beams was introduced by Ostrovsky and co-workers to solve this problem, in which annular intensity diameter remains unchanged no matter what TC is [9]. Another scheme was demonstrated in theory and experiment to generate PV beams, deriving from Fourier transformation of Bessel-Gaussian (BG) beams [10]. However, these methods involve plenty of free-space bulky reflective or refractive devices including spatial light modulator, axicon, mirror and lens, which inevitably lead to optical system complexity and inevitably impede applications of PV beams in OAM-based miniaturized photonic circuits. Recently, metasurface, a two-dimensional version of metamaterial, has been demonstrated that it processes a unique advantage in full control of electromagnetic waves at will and exhibits remarkable superiority in flexibility and integration over conventional optical elements with wavelength-dependent phase shift based on optical path difference [1115]. Therefore, large amounts of bulky optical components are reproduced in the form of plane configurations using metasurfaces, such as lenses, wave plates, beamsplitters, holograms and arbitrary vector beam generators [1625]. Furthermore, this design flexibility endows the metasurface platform with an inbuilt advantage of integrating multiple diversified functionalities into one single device [2631].

Based on the theory that PV beams can be generated from Fourier transmission of BG beams, Wu et al. first constructed three geometric phase metasurfaces to replace spiral phase plate, axicon and Fourier transform lens for the generation of PV beams [32]. Subsequently, Zhang et al. combined the phase profiles of these three optical elements into one plasmonic metasurface to directly produce PV beams [33]. Recently, polarization-controllable PV beams are achieved by simply switching incident polarization between two orthogonal directions [34]. However the above metasurface-based PV generators only well operate in the optical region and shut off at shorter wavelength due to the high Ohmic losses of materials in the ultraviolet range. Here we demonstrate numerically broadband, high-efficiency PB phase metasurfaces in the ultraviolet range, which are composed of dielectric rectangular nanopillars and capable of converting directly circularly polarized incident light into PV beams. The results indicate that the conversion efficiency reaches up to 86.6% and meanwhile remains over 65% in the mid-ultraviolet region. Furthermore, we explore the influence of several control parameters, including axicon period and lens focal length, on the annular ring radius of PV beams.

2. Theory and design

The PV beam is first proposed as an ideal model of optical vortices, where the radius of vortex ring is independent on its TC m. Experimentally, this special structured beam can be described approximately as

$$E(\rho ,\varphi ) = \exp \left[ { - \frac{{{{(\rho - {\rho_0})}^2}}}{{\Delta {\rho^2}}}} \right]\exp (im\varphi )$$
where (ρ, φ) are the polar coordinates; ρ0 and Δρ are the radium and width of annular profile of the PV beam. Among different techniques for the generation of PV beams, the simplest scheme is to directly operate the Bessel-Gauss (BG) beam with Fourier transformation using a Fourier lens, and the amplitude distribution of the generated perfect vortex at the focal plane has the following form [12,13].
$$E(\rho ,\varphi ) = {i^{l - 1}}\frac{{{w_g}}}{{{w_0}}}\exp \left[ { - \frac{{{{({\rho^2} - {\rho_0})}^2}}}{{w_0^2}}} \right]{J_l}\left( {\frac{{2{\rho_0}\rho }}{{w_0^2}}} \right)\exp (im\varphi )$$
where wg is the Gaussian beam waist to confine the BG beam, Jl is the lth order modified Bessel function of first kind, and 2w0 is the annular width of intensity pattern. Obviously, the modified Bessel function and the Gaussian function cooperate with each other to shape the amplitude of vortex beams. This physical mechanism suggests that a Gaussian beam will be converted into a PV beam as it passes through a spiral phase plate, an axicon and a Fourier lens one by one. Here the spiral phase plate transforms the Gaussian beam into the high-order Laguerre-Gaussian (LG) beam with a certain TC; then the axicon with conical surface converts the LG beam into the corresponding BG mode; at last the Fourier lens with spherical phase pattern implements the Fourier transformation.

For miniaturization and integration of photonic systems, we will use one metasurface to replace the three bulky refractive optical elements. The desired phase profile Φ(x, y) encoded on the metasurface should be a superposition of their phase distributions (as illustrated in Fig. 1), which is expressed as

$${\Phi _{\textrm{PV}}}(x,y) = {\Phi _{\textrm{spiral}}}(x,y) + {\Phi _{\textrm{axicon}}}(x,y) + {\Phi _{\textrm{lens}}}(x,y)$$
$${\Phi _{\textrm{spiral}}}(x,y)\textrm{ = }m \cdot \arctan \left( {\frac{x}{y}} \right)$$
$${\Phi _{\textrm{axicon}}}(x,y) ={-} \textrm{2}\pi \frac{{\sqrt {{x^2} + {y^2}} }}{d}$$
$${\Phi _{\textrm{lens}}}(x,y) ={-} \pi \frac{{({{x^2} + {y^2}} )}}{{{\lambda _\textrm{d}}f}}$$
where (x, y) is the Cartesian coordinates in the beam cross section, d represents the axicon period, f denotes the local length of Fourier lens, and λd is the design wavelength.

 figure: Fig. 1.

Fig. 1. Phase profile of metasurface-based PV beam generator as a superposition of spiral phase plate, axicon and Fourier transformation lens.

Download Full Size | PDF

In order to to fully control electromagnetic waves for the generation of PV beams, the building block of optical metasurfaces is capable of providing phase shift spanning the whole 0 to 2π range while maintaining high-efficiency uniform scattering amplitude. The geometric phase (also known as Pancharatnam–Berry (PB) phase) schemes are widely utilized to produce the phase responses [13,16,18,3537]. For an arbitrary anisotropic scatterer (such as nanobar or nanoslit) under the excitation of left-hand circularly polarized (LCP) incidence, its scattered complex field can be described as follow [14]:

$${E_\textrm{s}}(\varphi ) = \frac{{{t_\textrm{o}} + {t_\textrm{e}}}}{2}{E_\textrm{R}} + \frac{{{t_\textrm{o}} - {t_\textrm{e}}}}{2}\exp (i2\varphi ){E_\textrm{L}}$$
where to and te are the coefficients of forward scattering for incident light linearly polarized along the two principal axes of the anisotropic scatterer, respectively. The first term in Eq. (7) represents co-polarized scattered component with the same handedness as the incident light, and the second term represents cross-polarized scattered waves with opposite handedness and an additional geometric phase of 2φ, where φ represents the rotation angle. In other words, the latter provides fully continuous phase control from 0 to 2π by simply rotating the scatterer from 0 to 180°. Thus we will employ the anisotropic high-contrast rectangular nanopillar as the building block (as shown in Fig. 2(a)), which has been extensively used in the design of metasurface with different functionality [3739].

 figure: Fig. 2.

Fig. 2. Optical properties of building block array. (a) Schematic illustrations of metasurface and its building block, a high-aspect-ratio Si3N4 nanopillar with height H, width W and length L. The nanopillars are periodically arranged on a SiO2 substrate to form a square lattice with unit cell dimensions P × P. The nanopillars are rotated locally by an azimuth angle θ along the x-axis to account for the phase profile of metasurface under excitation of circularly polarized light, according to the PB phase. (b) Co-polarized, cross-polarized (also named as polarization conversion efficiency) and total transmission coefficients of nanopillar array under the LCP incidence. (c) Mie scattering coefficients of optical multipoles of the corresponding individual nanopillar excited by the LCP incident light, including the electric dipole, magnetic dipole, electric quadrupole and magnetic quadrupole. The inset describes the angle distribution of scattering fields of the nanopillar at wavelength of 295 nm, where the arrow represents the propagation direction of incident waves. (d) Magnetic components Hy and Hx existing in the nanopillar array excited by x-polarized (the left) and y-polarized (the right) incident light of the 295 nm wavelength. Simplified sketches of antiferromagnetic modes (denoted by red arrows) with even and odd diploes are shown in the insets.

Download Full Size | PDF

Equation (7) indicates that the high-performance PB phase metasurfaces require that the building blocks possess the capacity of high polarization conversion efficiency, in other words, the scatterers work as high-efficiency half-wave plates. Due to negligible absorption loss, all-dielectric metasurfaces, such as silicon (Si) in the near infrared range and titanium dioxide (TiO2) in the visible region, have been illustrated remarkable performance in comparison with plasmonic metasurfaces [1719,40]. In this work, silicon nitride (Si3N4) with high refractive index in our considered spectrum is selected as dielectric material to construct ultraviolet PV generator, owing to its large bandgap of about 5.1 eV [41,42]. The optical constant of Si3N4 is adopted from the Ref. [43]. To numerically investigate optical performance of the PV generators under the incidence of LCP wave, the simulation of metasurfaces is implemented using finite-difference time-domain (FDTD) method [15]. Periodic boundary conditions are applied at lateral boundaries of elementary cell (in the x and y directions) for simulating the periodic array, and perfectly matched layers (PMLs) are adopted to the propagation direction (the z direction) to eliminate nonphysical reflections.

In order to construct high-performance PV beam generators, the geometric parameters of Si3N4 nanopillar including its height, width, length and period have been carefully optimized for maximizing the circular polarization conversion efficiency and the corresponding values are set to be H = 600 nm, L = 130 nm, W = 50 nm and P = 200 nm. The total, co-polarized and cross-polarized transmittances of the nanopillar array are calculated and displayed in Fig. 2(b). It is clearly seen that most LCP incident light passes through the waveguide-like high-contrast nanopillars in the broadband ultraviolet range of 250-400 nm. More importantly, the co-polarized components are suppressed and high polarization conversion efficiency is achieved spanning the whole mid-ultraviolet range.

To reveal the physical mechanism under this broadband transmission of the dielectric nanopillar, we revisit the resonant Mie scattering of high-index dielectric nanoparticles from the perspective of multipolar interference. The light scattered from arbitrary obstacles can be decomposed into electromagnetic multipoles that have even or odd parity under spatial inversion [44], thus the scattering efficiency in the backward direction may be written as a sum of even components and odd components [45,46]

$${\sigma _\textrm{b}}(\lambda ) = {\left|{\sum\nolimits_{n = 1}^\infty {(2n + 1){{( - 1)}^n}({b_\textrm{n}}(\lambda ) - {a_\textrm{n}}(\lambda ))} } \right|^2}$$
where n corresponds the parity, an and bn represent the scattering coefficients of electric and and magnetic multipoles, respectively. As each pair of coefficients are equal, that is an = bn, the backscattering can be totally eliminated. For Example, the electric and magnetic dipoles in a magnetodielectric nanosphere interfere destructively in the backward direction, resulting in zero scattering in narrow spectral range. This method has been demonstrated for Huygens’ metasurfaces [47]. Likewise, as more higher-order multipoles overlap and satisfy the above condition, not only will the zero backward scattering still appear, but also the operating bandwidth will be broadened. Therefore the multipolar scattering efficiency and the far-field scattering pattern (λ = 300 nm) of the Si3N4 nanopillar are calculated and plotted in Fig. 2(c). It is evident that several high-order multipolar modes are also excited in the ultraviolet spectral range, and their destructive interference produces minima in the backward scattering direction (see the inset of Fig. 2(c)). In addition, although this zero-backward condition for the complete anti-reflection effect is not fully satisfied in the whole targeted spectrum, the magnetic dipole as well as the electric and magnetic quadrupoles establish a relative balance, resulting in a broadband transparent window in the ultraviolet range.

Equation (7) indicates that a high-performance PB-phase building block must operate as an efficient half-wave plate, which means that the transmission coefficients (to, te) should have a phase delay of π to reach maximum under the orthogonal linearly polarized excitations. To unveil the mechanism behind the high polarization conversion efficiency of this resonator, we plot the magnetic-field distribution profiles at the same wavelength (λ = 300 nm) for both polarizations to display the propagation of optical wave in the waveguide-like nanopillar. As shown in Fig. 2(d), the induced magnetic pattern Hy indicates that five circle displacement currents with alternative clockwise and anti-clockwise directions are excited by x-polarized incidence, as five antiparallel magnetic dipoles are vertically located along the propagation direction. Likewise, six antiparallel magnetic dipoles appear in the z axis under y-polarized incidence. Therefore, this waveguide-like mode with even and odd antiparallel magnetic dipoles induces a phase delay of π, making the nanopillar a high-efficiency half-wave plate [48].

3. Results and discussion

Based on the above discussion, the optimized Si3N4 nanopillars as basic elements are utilized to build transmission-type metasurfaces for the generation of PV beam with the TC parameters m = 1, 3 and 5, where the axicon period is chosen to be d = 3 µm, and the focal length of Fourier lens is selected to be f = 40 µm at the design wavelength λd = 300 nm. These high-efficiency identical phase shifters are arrayed periodically on the glass substrate with spatially varying rotation angle φ(x, y) = ΦPV(x, y)/2 (as shown in Fig. 2(a)). All numerical simulations of the designed PV beam generators are carried out using the FDTD method, where PML boundary conditions are applied in the x, y, and z axes. Figure 3(a) and (b) show the far-field phase distribution patterns and the corresponding intensity profiles of cross-polarized beams in the x-y plane (z = 30 µm) for different TCs. As illustrated in Fig. 3(a), the number of spiral branches in the phase patterns indicates that the vortex beams with different TCs are produced by the dielectric metasurfaces, and are equivalent to the designed TCs. More importantly, the radius of annular intensity rings keeps almost unchanged for different TCs, verifying that this metasurface-based scheme can generate ultraviolet PV beams. To clearly exhibit the evolution process of PV beams, we plot the intensity patterns in the x-z plane, as shown in Fig. 3(c). It can be seen that the incident LCP plane wave is first converted into a hollow non-diffractive Bessel-like beam with a certain TC as it passes through the metasurface, acting as an axicon functionality. Then the beam focuses at the plane of z = 22 µm, forming a donut-shaped intensity pattern with the zero field at the center of focal plane. Note that the ring radius is gradually broadened as TC increases at this position. Finally the optical wave diverges at the free space and evolves a perfect annular ring profile. As the PV beam further propagates, the radius of ring becomes larger, but the ring shape remains the same.

 figure: Fig. 3.

Fig. 3. Far-field profiles of generated PV beams at the design wavelength (300 nm). (a) Phase distributions and (b) intensity patterns in the x-y plane as well as intensity patterns in the x-z plane of PV beams with different topological charges.

Download Full Size | PDF

To explore the dependence of annular ring radius on the axicon period and focal length of Fourier lens, we design six sets of metasurfaces with different axicon periods (d = 2, 3 and 4 µm) and lens focal lengths (f = 30, 40 and 50 µm), respectively. The simulated intensity profiles of PV beams in the x-z plane are plotted in Fig. 4. From Fig. 4(a), we can observe that the parameter of axicon period dramatically influences the annular ring radius, which gets smaller as d increases from 2 to 4 µm. This result has good consistence with the conclusion of conventional free-space approach [49]. Because the axicon period is inversely proportional to its base angle, the focal length of PV beam generator will increase gradually with the increase of axicon period. Likewise, we also investigate the influence of focal length of Fourier lens on the annular ring radius, as shown in Fig. 4(b). The results indicate that, although the change of the parameter f causes a slight decrease in the ring radius, the focal length of the metasurface markedly elongates when d is increased from 30 to 50 µm. Despite we do not provide the quantitative relation between the ring radius of PV beam and these characteristic parameters d and f, this discussion is useful in designing metasurface-based PV beam generators.

 figure: Fig. 4.

Fig. 4. Intensity profiles of PV beams at the design wavelength (300 nm), generated by the metasurfaces with (a) different axicon periods (d =2, 3 and 4 µm) or (b) lens focal lengths (f = 30, 40 and 50 µm). For the previous metasurfaces, the focal length of Fourier lens is set to be 40 µm; for the latter ones, the period of axicon is set to be 3 µm.

Download Full Size | PDF

In order to further demonstrate the broadband operation of PB metasurface-based PV beam generators, we also simulate the above designed metasurfaces at three wavelengths of 285, 300 and 315 nm, and plot their intensity distribution patterns at the propagation distance z = 40 µm, as shown in Fig. 5. It is intuitively observed that at a certain wavelength, the annular ring radius of PV beam is almost unchanged for different TCs under a fixed axicon period. This broadband speciality is attributed to the dispersionless geometric effect of the PB phase of anisotropic nanoscatterers. For the axicon period, it can be rewritten as d = 2πλd/(λ∇Φaxicon(x, y)), where the phase gradient ∇Φaxicon(x, y) is also wavelength-independent according to the geometric phase, which causes the decrease of actual period of axicon and results that the ring radius is slightly enlarged as the incident wavelength increases [50]. The observed annular intensity rings of the PV beams have good quality at all wavelengths, indicating the broadband operation capability of the PV beam generators. As a diffractive optical device, not all cross-polarized transmitted beams are converted into PV beams by metasurfaces, thus the conversion efficiency, defined as the ratio of the intensity of bright ring to the intensity of incident beam [34], is adopted to characterize the performance of the broadband PV beam generators. The efficiency of the metasurface with TC m = 1 is estimated as a function of wavelength spanning the mid-ultraviolet region, as shown in Fig. 6. Here, the conversion efficiency of the PV generator is defined as the ratio of the cross-polarized transmitted power passing through the focal spot in a circle whose radius is three times the full width at half-maximum (FWHM) [51]. This metasurface has a conversion efficiency up to 86.6% at the design wavelength λd = 300 nm, which remains above 65% in the whole spectral range. Furthermore, it can be seen that the evolvement trend of the conversion efficiency of the beam convertor with wavelength is well consistent with that of polarization conversion efficiency of the building block array.

 figure: Fig. 5.

Fig. 5. Intensity profiles of PV beams at propagation distance of z = 40 µm at three wavelengths of 285, 295 and 315 nm. From left column to right column, TC m is 1, 3 and 5.

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. Conversion efficiency of the PV beam generator in the mid-ultraviolet spectral rang. The selected metasurface has topological charge m = 1, axicon period d = 3 µm, lens focal length f = 40 µm and design wavelength λd = 300 nm, respectively.

Download Full Size | PDF

4. Conclusions

In summary, with the help of destructive interference of electric and magnetic multipoles as well as the different parity of antiparallel magnetic dipoles in both orthogonal polarizations, the anisotropic rectangular nanopillar can work as a broadband and high-efficiency half-wave plate in the ultraviolet rang. The orientation-varying building blocks are employed to construct the metasurface-based PV beam generators based on geometric phase. We demonstrate that the metasurfaces are capable of producing ultraviolet PV beams with TC-independent annular intensity profiles. Moreover, we find that the axicon period has dramatically influences on the annular ring radius of PV beams in contrast to the lens focal length. In addition, the simulated results indicate that the metasurfaces can operate well in the broad spectral range and the conversion efficiency reach up to 86.6% at the design wavelength. We believe the ultraviolet PV beam generators will find significant applications, ranging from optical tweezer and high-resolution spectroscopy to on-chip communication.

Funding

Natural Science Foundations of China (11704208, 12174351); Key Research Project for Science and Technology of the Education Department of Henan Province (22A140027).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but maybe obtained from the authors upon reasonable request.

References

1. D. G. Grier, “A revolution in optical manipulation,” Nature 424(6950), 810 (2003). [CrossRef]  

2. M. Padgett and R. Bowman, “Tweezers with a twist,” Nat. Photonics 5(6), 343–348 (2011). [CrossRef]  

3. L. Allen, M. W. Beijersbergen, R. Spreeuw, and J. P. Woerdman, “Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes,” Phys. Rev. A 45(11), 8185–8189 (1992). [CrossRef]  

4. R. Fickler, R. Lapkiewicz, W. N. Plick, M. Krenn, C. Schaeff, S. Ramelow, and A. Zeilinger, “Quantum Entanglement of High Angular Momenta,” Science 338(6107), 640–643 (2012). [CrossRef]  

5. A. M. Yao and M. J. Padgett, “Orbital angular momentum: origins, behavior and applications,” Adv. Opt. Photonics 3(2), 161–204 (2011). [CrossRef]  

6. A. E. Willner, H. Huang, Y. Yan, Y. Ren, N. Ahmed, G. Xie, C. Bao, L. Li, Y. Cao, Z. Zhao, J. Wang, M. P. J. Lavery, M. Tur, S. Ramachandran, A. F. Molisch, N. Ashrafi, and S. Ashrafi, “Optical communications using orbital angular momentum beams,” Adv. Opt. Photonics 7(1), 66–106 (2015). [CrossRef]  

7. H. Yan, E. Zhang, B. Zhao, and K. Duan, “Free-space propagation of guided optical vortices excited in an annular core fiber,” Opt. Express 20(16), 17904–17915 (2012). [CrossRef]  

8. S. Li and W. Jian, “Multi-Orbital-Angular-Momentum Multi-Ring Fiber for High-Density Space-Division Multiplexing,” IEEE Photonics J. 5(5), 7101007 (2013). [CrossRef]  

9. A. S. Ostrovsky, C. Rickenstorff-Parrao, and V. Arrizón, “Generation of the “perfect” optical vortex using a liquid-crystal spatial light modulator,” Opt. Lett. 38(4), 534–536 (2013). [CrossRef]  

10. P. Vaity and L. Rusch, “Perfect vortex beam: Fourier transformation of a Bessel beam,” Opt. Lett. 40(4), 597–600 (2015). [CrossRef]  

11. 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). [CrossRef]  

12. X. Ni, N. K. Emani, A. V. Kildishev, A. Boltasseva, and V. M. Shalaev, “Broadband Light Bending with Plasmonic Nanoantennas,” Science 335(6067), 427 (2012). [CrossRef]  

13. L. Huang, X. Chen, H. Mühlenbernd, G. Li, B. Bai, Q. Tan, G. Jin, T. Zentgraf, and S. Zhang, “Dispersionless Phase Discontinuities for Controlling Light Propagation,” Nano Lett. 12(11), 5750–5755 (2012). [CrossRef]  

14. H. T. Chen, A. J. Taylor, and N. Yu, “A review of metasurfaces: physics and applications,” Rep. Prog. Phys. 79(7), 076401 (2016). [CrossRef]  

15. M. Wan, P. Ji, R. Wang, X. Zhang, M. Tian, S. Yuan, L. Zhang, and J. He, “Ultraviolet wavefront manipulation using topological insulator metasurfaces based on geometric phase,” Opt. Commun. 487, 126812 (2021). [CrossRef]  

16. X. Chen, L. Huang, H. Mühlenbernd, G. Li, B. Bai, Q. Tan, Q. Jin, C.-W. Qiu, S. Zhang, and T. Zentgraf, “Dual-polarity plasmonic metalens for visible light,” Nat. Commun. 3(1), 1198 (2012). [CrossRef]  

17. A. Arbabi, Y. Horie, A. J. Ball, M. Bagheri, and A. Faraon, “Subwavelength-thick lenses with high numerical apertures and large efficiency based on high-contrast transmitarrays,” Nat. Commun. 6(1), 7069 (2015). [CrossRef]  

18. M. Khorasaninejad, W. T. Chen, R. C. Devlin, J. Oh, A. Y. Zhu, and F. Capasso, “Metalenses at visible wavelengths: Diffraction-limited focusing and subwavelength resolution imaging,” Science 352(6290), 1190–1194 (2016). [CrossRef]  

19. W. Liu W, Z. Li, H. Cheng, C. Tang, J. Li, S. Zhang, S. Chen, and J. Tian, “Metasurface Enabled Wide-Angle Fourier Lens,” Adv. Mater. 30(23), 1706368 (2018). [CrossRef]  

20. 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]  

21. A. Y. Zhu, W. T. Chen, M. Khorasaninejad, J. Oh, A. Zaidi, I. Mishra, R. C. Devlin, and F. Capasso, “Ultra-compact visible chiral spectrometer with meta-lenses,” APL Photonics 2(3), 036103 (2017). [CrossRef]  

22. M. Khorasaninejad, W. T. Chen, J. Oh, and F. Capasso, “Super-Dispersive Off-Axis Meta-Lenses for Compact High Resolution Spectroscopy,” Nano Lett. 16(6), 3732–3737 (2016). [CrossRef]  

23. X. Li, L. Chen, Y. Li, X. Zhang, M. Pu, Z. Zhao, X. Ma, Y. Wang, M. Hong, and X. Luo, “Multicolor 3D meta-holography by broadband plasmonic modulation,” Sci. Adv. 2(11), e1601102 (2016). [CrossRef]  

24. N. K. Grady, J. E. Heyes, R. C. Dibakar, Z. Yong, M. T. Reiten, A. K. Azad, A. J. Taylor, D. A. R. Dalvit, and H.-T. Chen, “Terahertz Metamaterials for Linear Polarization Conversion and Anomalous Refraction,” Science 340(6138), 1304–1307 (2013). [CrossRef]  

25. P. Yu, J. Li, C. Tang, H. Cheng, Z. Liu, Z. Li, Z. Liu, C. Gu, J. Li, S. Chen, and J. Tian, “Controllable optical activity with non-chiral plasmonic metasurfaces,” Light: Sci. Appl. 5(7), e16096 (2016). [CrossRef]  

26. S. M. Kamali, E. Arbabi, A. Arbabi, Y. Horie, M. Faraji-Dana, and A. Faraon, “Angle-Multiplexed Metasurfaces: Encoding Independent Wavefronts in a Single Metasurface under Different Illumination Angles,” Phys. Rev. X 7, 041056 (2017). [CrossRef]  

27. J. P. B. Mueller, N. A. Rubin, R. C. Devlin, B. Groever, and F. Capasso, “Metasurface Polarization Optics: Independent Phase Control of Arbitrary Orthogonal States of Polarization,” Phys. Rev. Lett. 118(11), 113901 (2017). [CrossRef]  

28. E. Maguid, I. Yulevich, D. Veksler, V. Kleiner, M. L. Brongersma, and E. Hasman, “Photonic spin-controlled multifunctional shared-aperture antenna array,” Science 352(6290), 1202–1206 (2016). [CrossRef]  

29. W. Liu, Z. Li, Z. Li, H. Cheng, C. Tang, J. Li, S. Chen, and J. Tian, “Energy-Tailorable Spin-Selective Multifunctional Metasurfaces with Full Fourier Components,” Adv. Mater. 31, 1901729 (2019). [CrossRef]  

30. S. Chen, W. Liu, Z. Li, H. Cheng, and J. Tian J, “Metasurface-Empowered Optical Multiplexing and Multifunction,” Adv. Mater. 32(3), 1805912 (2020). [CrossRef]  

31. Z. Li, W. Liu, H. Cheng, D. Choi, S. Chen, and J. Tian, “Spin-Selective Full-Dimensional Manipulation of Optical Waves with Chiral Mirror,” Adv. Mater. 32(26), 1907983 (2020). [CrossRef]  

32. Y. Liu, Y. Ke, J. Zhou, Y. Liu, H. Luo, S. Wen, and D. Fan, “Generation of perfect vortex and vector beams based on Pancharatnam-Berry phase elements,” Sci. Rep. 7(1), 44096 (2017). [CrossRef]  

33. Y. Zhang, W. Liu, G. Jie, and X. Yang, “Generating Focused 3D Perfect Vortex Beams By Plasmonic Metasurfaces,” Adv. Opt. Mater. 6(4), 1701228 (2018). [CrossRef]  

34. J. Xie, H. Guo, S. Zhuang, and J. Hu, “Polarization-controllable perfect vortex beam by a dielectric metasurface,” Opt. Express 29(3), 3081–3089 (2021). [CrossRef]  

35. S. Wang, Z.-L. Deng, Y. Wang, Q. Zhou, X. Wang, Y. Cao, B.-O. Guan, X. Xiao, and X. Li, “Arbitrary polarization conversion dichroism metasurfaces for all-in-one full Poincaré sphere polarizers,” Light: Sci. Appl. 10(1), 24 (2021). [CrossRef]  

36. W. Dai, Y. Wang, R. Li, Y. Fan, G. Qu, Y. Wu, Q. Song, J. Han, and S. Xiao, “Achieving Circularly Polarized Surface Emitting Perovskite Microlasers with AllDielectric Metasurfaces,” ACS Nano 14(12), 17063–17070 (2020). [CrossRef]  

37. W. Luo, S. Sun, H.X. Xu, Q. He, and L. Zhou, “Transmissive Ultrathin Pancharatnam-Berry Metasurfaces with nearly 100% Efficiency,” Phys. Rev. Applied 7(4), 044033 (2017). [CrossRef]  

38. Z. Li, M. H. Kim, W. Cheng, Z. Han, S. Shrestha, A.C. Overvig, M. Lu, A. Stein, A. M. Agarwal, M. Lončar, and N. Yu, “Controlling propagation and coupling of waveguide modes using phase-gradient metasurfaces,” Phys. Rev. Appl. 12(7), 675–683 (2017). [CrossRef]  

39. S. Chen, Z. Li, W. Liu, H. Cheng, and J. Tian, “From Single-Dimensional to Multidimensional Manipulation of Optical Waves with Metasurfaces,” Adv. Mater. 31(16), 1802458 (2019). [CrossRef]  

40. R. C. Devlin, M. Khorasaninejad, W. T. Chen, J. Ohand, and F. Capasso, “Broadband high-efficiency dielectric metasurfaces for the visible spectrum,” Proc. Natl. Acad. Sci. 113(38), 10473–10478 (2016). [CrossRef]  

41. S. Kanwal, J. Wen, B. Yu, D. Kumar, X. Chen, Y. Kang, C. Bai, and D Zhang, “High-Efficiency, Broadband, Near Diffraction-Limited, Dielectric Metalens in Ultraviolet Spectrum,” Nanomaterials 10(3), 490 (2020). [CrossRef]  

42. H. Ahmed, A. A. Rahim, H. Maab, M. M. Ali, and S Naureen, “Polarization insensitive all-dielectric metasurfaces for the ultraviolet domain,” Opt. Mater. Express 10(4), 1083–1091 (2020). [CrossRef]  

43. E. D. Palik and G. Ghosh, Electronic Handbook of Optical Constants of Solids: User Guide (Academic, Cambridge1985).

44. S. Kruk, B. Hopkins, I. I. Kravchenko, A. Miroshnichenko, D. N. Neshev, and Y. S. Kivshar, “Invited Article: Broadband highly efficient dielectric metadevices for polarization control,” APL Photonics 1(3), 030801 (2016). [CrossRef]  

45. W. Liu and Y. S. Kivshar, “Generalized Kerker effects in nanophotonics and meta-optics [Invited],” Opt. Express 26(10), 13085–13105 (2018). [CrossRef]  

46. C. F. Bohren and D. R. Huffman, Absorption and Scattering of Light by Small Particles (Wiley: New York, 1998).

47. C. Pfeiffer, N. K. Emani, A. M. Shaltout, A. Boltasseva, V. M. Shalaev, and A. Grbic, “Efficient Light Bending with Isotropic Metamaterial Huygens’ Surfaces,” Nano Lett. 14(5), 2491–2497 (2014). [CrossRef]  

48. K. Huang, J. Deng, H. S. Leong, S. Yap, R. B. Yang, J. Teng, and H. Liu, “Ultraviolet Metasurfaces of 80% Efficiency with Antiferromagnetic Resonances for Optical Vectorial Anti-Counterfeiting,” Laser Photonics Rev. 13(5), 1800289 (2019). [CrossRef]  

49. S. Fu, T. Wang, and C. Gao, “Perfect optical vortex array with controllable diffraction order and topological charge,” J. Opt. Soc. Am. A 33(9), 1836–1842 (2016). [CrossRef]  

50. W. T. Chen, M. Khorasaninejad, A. Y. Zhu, J. Oh, R. C. Devlin, A. Zaidi, and F. Capasso, “Generation of wavelength-independent subwavelength Bessel beams using metasurfaces,” Light: Sci. Appl. 6(5), e16259 (2017). [CrossRef]  

51. Z. Zhang, T. Li, X. Jiao, G. Song, and Y. Xu, “High-Efficiency All-Dielectric Metasurfaces for the Generation and Detection of Focused Optical Vortex for the Ultraviolet Domain,” Appl. Sci. 10(16), 5716 (2020). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but maybe obtained from the authors upon reasonable request.

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 (6)

Fig. 1.
Fig. 1. Phase profile of metasurface-based PV beam generator as a superposition of spiral phase plate, axicon and Fourier transformation lens.
Fig. 2.
Fig. 2. Optical properties of building block array. (a) Schematic illustrations of metasurface and its building block, a high-aspect-ratio Si3N4 nanopillar with height H, width W and length L. The nanopillars are periodically arranged on a SiO2 substrate to form a square lattice with unit cell dimensions P × P. The nanopillars are rotated locally by an azimuth angle θ along the x-axis to account for the phase profile of metasurface under excitation of circularly polarized light, according to the PB phase. (b) Co-polarized, cross-polarized (also named as polarization conversion efficiency) and total transmission coefficients of nanopillar array under the LCP incidence. (c) Mie scattering coefficients of optical multipoles of the corresponding individual nanopillar excited by the LCP incident light, including the electric dipole, magnetic dipole, electric quadrupole and magnetic quadrupole. The inset describes the angle distribution of scattering fields of the nanopillar at wavelength of 295 nm, where the arrow represents the propagation direction of incident waves. (d) Magnetic components Hy and Hx existing in the nanopillar array excited by x-polarized (the left) and y-polarized (the right) incident light of the 295 nm wavelength. Simplified sketches of antiferromagnetic modes (denoted by red arrows) with even and odd diploes are shown in the insets.
Fig. 3.
Fig. 3. Far-field profiles of generated PV beams at the design wavelength (300 nm). (a) Phase distributions and (b) intensity patterns in the x-y plane as well as intensity patterns in the x-z plane of PV beams with different topological charges.
Fig. 4.
Fig. 4. Intensity profiles of PV beams at the design wavelength (300 nm), generated by the metasurfaces with (a) different axicon periods (d =2, 3 and 4 µm) or (b) lens focal lengths (f = 30, 40 and 50 µm). For the previous metasurfaces, the focal length of Fourier lens is set to be 40 µm; for the latter ones, the period of axicon is set to be 3 µm.
Fig. 5.
Fig. 5. Intensity profiles of PV beams at propagation distance of z = 40 µm at three wavelengths of 285, 295 and 315 nm. From left column to right column, TC m is 1, 3 and 5.
Fig. 6.
Fig. 6. Conversion efficiency of the PV beam generator in the mid-ultraviolet spectral rang. The selected metasurface has topological charge m = 1, axicon period d = 3 µm, lens focal length f = 40 µm and design wavelength λd = 300 nm, respectively.

Equations (8)

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

E ( ρ , φ ) = exp [ ( ρ ρ 0 ) 2 Δ ρ 2 ] exp ( i m φ )
E ( ρ , φ ) = i l 1 w g w 0 exp [ ( ρ 2 ρ 0 ) 2 w 0 2 ] J l ( 2 ρ 0 ρ w 0 2 ) exp ( i m φ )
Φ PV ( x , y ) = Φ spiral ( x , y ) + Φ axicon ( x , y ) + Φ lens ( x , y )
Φ spiral ( x , y )  =  m arctan ( x y )
Φ axicon ( x , y ) = 2 π x 2 + y 2 d
Φ lens ( x , y ) = π ( x 2 + y 2 ) λ d f
E s ( φ ) = t o + t e 2 E R + t o t e 2 exp ( i 2 φ ) E L
σ b ( λ ) = | n = 1 ( 2 n + 1 ) ( 1 ) n ( b n ( λ ) a n ( λ ) ) | 2
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.