Next Article in Journal
Hand Exoskeleton—Development of Own Concept
Previous Article in Journal
Supporting Learning Analytics Adoption: Evaluating the Learning Analytics Capability Model in a Real-World Setting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimized Catenary Metasurface for Detecting Spin and Orbital Angular Momentum via Momentum Transformation

1
State Key Laboratory of Optical Technologies on Nano-Fabrication and Micro-Engineering, Institute of Optics and Electronics, Chinese Academy of Sciences, Chengdu 610209, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
3
Key Laboratory for Information Science of Electromagnetic Waves (MoE), Fudan University, Shanghai 200433, China
4
Research Center on Vector Optical Fields, Institute of Optics and Electronics, Chinese Academy of Sciences, Chengdu 610209, China
5
Tianfu Xinglong Lake Laboratory, Chengdu 610209, China
*
Author to whom correspondence should be addressed.
Submission received: 21 January 2023 / Revised: 23 February 2023 / Accepted: 28 February 2023 / Published: 3 March 2023
(This article belongs to the Special Issue Advances in Orbital Angular Momentum (OAM))

Abstract

:
Theoretically, the topological charge l in the vortex can be any integer or fraction, thus the vortex carrying different topological charges can form an infinitely orthogonal orbital angular momentum state space, which can provide new dimensional resources for optical communication. However, high-capacity optical communication requires low delay, thus real-time detection of the OAM is significant for communication. Metasurfaces have the characteristics of low loss, ultra-thin, easy integration, and flexible phase controls, which provide a meaningful way to realize integrated OAM generation and detection. Here, an optimized streamlined metasurface (OSM) is presented, which can detect high-order vortex beams in a single, simple, and rapid manner by photon momentum transformation (PMT). Since different vortices are converted into focusing modes with distinct azimuthal coordinates on a transverse plane through PMT, a single measurement can determine OAMs in an ample mode space. In addition, the OSM can detect more and higher order OAMs compared with a discrete metasurface (DM) at the same size, due to its better wavefront sampling capabilities. With the merits of an ultra-compact device size, simple optical structure, and outstanding vortex recognition ability, our approach may underpin the development of integrated optics and quantum systems.

1. Introduction

There are two main types of angular momentum: spin angular momentum (SAM) [1,2] based on polarization vector rotation and orbital angular momentum (OAM) [3] based on wavefront phase structure rotation. In 1992, Allen et al. [4] precisely deduced that the orbital angular momentum in optical vortices comes from the phase factor of  exp ( i l ϕ ) , where l is the topological charge,  ϕ  is the azimuth angle, and each photon carries orbital angular momentum of  l . Since then, OAM has drawn increasing attention in many realms, including super-resolution imaging [5], quantum information [6], optical communications [7,8,9], and signal processing [10,11,12,13], which is one of the important topics for SAM and OAM recognition. Moreover, different OAM modes can transmit the information as separate channels in a division multiplexing communication system, thus it is crucial to encode and decode different OAM modes separately. Therefore, detecting topological charges and identifying different OAM modes play a vital role in the application of vortex beams. Therefore, detecting topological charges and identifying different OAM modes play a crucial role in the application of the OAM beams.
Generally, vortex beams can be manipulated in a variety of ways, such as spiral phase plates [14,15], glued hollow axicons [16], birefringent elements [17,18], orbital–angular-momentum holography [19], microscopic ring resonators [20], spatiotemporal dynamics of plasmonic vortices [21], spatial light modulators (SLMs), ultrathin and planar metasurfaces [22,23], peritrophic multiplexing metasurfaces [24], and azimuthal-quadratic diffraction optical elements [25]. OAM identification is challenging due to the missing phase information for traditional intensity detectors. The early OAM detection was proposed based on interference with the mirror or reference wavefront [26,27], diffraction of specific apertures [28], and mutual projection of forked diffraction grating [6]. Nowadays, a variety of vortices have been successfully detected, such as single-photon vortices [27], supercontinuum femtosecond vortices, noninteger vortices [28], and entanglement vortices [4]. However, multiple repeated projection measurements and calculating the number of interference fringes are inefficient for detecting possible OAM modes, especially for higher-order OAM modes [29].
In addition, the traditional vortex optical recognition system is often complicated and bulky, and the recognition method of the OAM is usually complex, which brings certain limitations to the practical application. A more straightforward system will lead to more excellent stability from an engineering point of view. In recent years, a variety of functional devices with excellent properties have emerged, such as metastructures [30], nanoparticles [31], graphene [32], plasmas [33], and multi-functional structures [34], with the rapid development of new technologies. In addition, with the same secondary phase form of a radial converging lens, a more convenient design has been proposed, namely, angular lens [35], which focuses different OAM modes to corresponding azimuth positions in the focal plane at a constant intensity. It can significantly reduce the optical system’s complexity, improve the system’s stability, and increase the identification range of the vortex topological charge.
Based on the principle of symmetry transformations via quadratic phase [36], Guo et al. used a spin-decoupling metasurface to realize the sorting detection of topological charges of multiple incidents of left and right vortices simultaneously [37]. However, composed of discrete unit cells, the spin-decoupled metasurface suffered from the problem of insufficient phase sampling, which reduced the detection range of the OAM. Here, an optimized streamlined metasurface (OSM) [38] was adopted to improve detection capabilities and achieve the same size to increase the detection range of the OAM, especially for higher-order OAMs. In addition, compared with the l0 = 40 conducted by the discrete metasurface (DM) at present, the detection range of the OAMs by the OSM in this paper was improved by about 6.5 times.

2. Theory and Methods

The OSM is an optimized dielectric nanostructure of catenary [39,40]. Compared with the DM, the OSM can achieve continuous wavefront sampling in 0–2π, with improved diffraction efficiency and continuous phase distribution in a wider spectral band [38]. Here, based on the photonic spin hall effect [41] and catenary optics [22], an OSM structure was designed. It can transform different vortices in a large mode space into focusing modes of different azimuth coordinates in the transverse plane by photon momentum transformation (PMT), thus, OAMs can be determined through a single-shot measurement. PMT is an optical transformation that transforms the received OAMs’ mode to the focusing modes on the focal plane with different angular positions [35,37], which can be expressed as:
M ( r , ϕ ) = g ( r ) exp ( l 0 ϕ 2 / 2 )
where:
{ g ( r ) = 1 g ( r ) = 0 r i r r o o t h e r w i s e
Φ ( ϕ ) = l 0 2 ϕ 2
where M is the phase mask, l0 is the quadratic phase coefficient,  ϕ  is the azimuthal coordinate defined as tan−1 (y/x), and ri and ro represent the inner and outer radius of the ring phase mask. By flattening the polar coordinates into Cartesian coordinates, the azimuth-quadratic phase contour can be converted into a cylindrical quadratic phase contour.  ϕ  can also be expressed as:
ϕ = s s 0
where s0 = πr0 and r0 = (ri + ro)/2, respectively, and s is the focusing position of the detection position from the origin position [37]. Then, after the vortex light with topological charge l acts on the OSM, the phase distribution of the outgoing light is as follows:
Φ ( ϕ ) + l ϕ = l 0 2 ( s s 0 ) 2 + l s s 0 = l 0 2 s 0 2 ( s + l s 0 l 0 ) 2 - l 2 2 l 0 = Φ ( ϕ ) - l 2 2 l 0
where  ϕ  denotes a new azimuthal coordinate. Neglecting the azimuth-independent term on the right-hand side of the equation, we find that there is only a mode-dependent rotation through the PMT:
s = l s 0 l 0
According to Equation (4), the theoretical focus  ϕ l  can be expressed as:
ϕ l = l l 0 ,   ϕ l [ - π ,   π ]
Here,  ϕ l  is the radian corresponding to the focus and the origin (the focus position when the topological charge is zero) of the vortex light with the topological charge l passing through the OSM. Then, the focusing spot is symmetric around the azimuth coordinate  ϕ = 0 . As the topological charge of the vortex increases from zero to the positive and negative ends, the focusing mode will take the zero point as the origin and rotate to both ends with an azimuth angle spaced l/l0 apart. By recording the azimuth coordinates of the focusing mode, the received OAM mode can be determined as follows.
l = l 0 ϕ l ,   ϕ l [ - π ,   π ]
According to equation  f = k 0 r 0 2 l 0  [37], where k0 is the wavenumber in free space, r0 = (ri + ro)/2, and the focal length is independent of the topological charge of the incident OAM mode. Therefore, different OAM modes can be detected in the same transverse plane of the focal length.
According to Equation (4), the detection of vortex optical topological charge is correlated with l0, and l0 is positively correlated with the phase gradient that the designed metasurface can accommodate. As shown in Figure 1a, the 2π periodic phase distribution within the metasurface changes rapidly from right to left, reducing the structural space that accommodates 2π phase adjustment. To ensure the phase control ability of DM, it is necessary for the metasurface retention space to accommodate sufficient nanopillars for 2π phase adjustment, which limits the maximum l0 and the maximum detectable OAMs. However, the single streamline line of the OSM has the 2π phase adjustment ability in both large and small Λ periods. When the structure period gradually decreases, the OSM can still accommodate more columns, effectively improving the distinguishing power of DM and significantly increasing the detection range and maximum detection mode of the OAMs. Then, the OSM is an equal width streamlined metasurface optimized according to the catenary formula. The catenary equation [22] is as follows:
y = Λ π ln ( | sec ( π x / Λ ) | )
In full-model design, there is no difference between the OSM and DM in essence, and both fill the corresponding structure according to the phase distribution. However, DM requires the interaction of multiple nanocolumns to adjust the 2π phase, while the OSM requires only one structure for the 2π phase condition due to its unique design, which has significant advantages when the space range is limited. In addition, with the feasibility of manufacturing guaranteed, better phase adjustment ability can be obtained, and the streamline line is truncated at 0.75Λ (Λ is the catenary span). A vertical structure with width w2 was added to the truncated region to increase the streamlined metasurface optimized for phase continuity and amplitude uniformity. Simultaneously, the w2 structure was canceled when the distance between two adjacent catenary lines in the circumferential direction was less than 0.1 rad. This design maximizes the coefficient l0 while ensuring adequate wavefront sampling and fabrication feasibility, making it have practical application potential.

3. Results

3.1. Analysis and Simulation

The wavefront sampling of DM phase modulation is insufficient, and the detection capability of the OAMs is low, especially for higher-order OAMs. Here, we give explanations from theoretical derivation. If the phase coverage range of a single rectangular nanocolumn structure is  2 π n  and the period is  p , the difference can be expressed as:
| Φ   ( s 2 ) - Φ   ( s 1 ) | = l 0 2 s 0 2 | s 2 2 - s 1 2 | = 2 π n , s 2 s 1 > p
From Equation (10), it can be deduced that:
| s 1 | < 2 π s 0 2 n l 0 p p 2
According to Formulas (4) and (11), it can be expressed as:
| ϕ max | < 2 π s 0 n l 0 p p 2 s 0
When  | ϕ | > | ϕ max | , the topology detection cannot be realized by the structure.
As shown in Figure 1a, the vortex passes through the metasurface with ideal phase distribution and form focusing spots at corresponding positions in the focal plane. The focus spots are distributed in a circular direction, and the course rotates clockwise. Then, it can be inferred that the whole optical system is relatively simple, and the differentiation of the different OAM modes is also relatively simple and convenient. The topological charge of incident light can be inferred only by comparing the location of the focusing light spot.
Here, the differences between the OSM and DM in sorting OAMs are further studied. The separation of LCP vortex light was calculated through the OSM and DM phase plates with l0 = 200. Then, for a better comparison, a numerical simulation of the OSM and DM at l = −620 to l = 620, Δl = 10 was conducted here, and the simulation results were statistically sorted out to form Figure 1b,c. As can be seen in Figure 1b,c, the number of the OAM modes detected by the OSM is much larger than that detected by DM structure. Figure 1c indicates the simulation results are consistent with Equation (12), i.e., according to the square result, when  ϕ l  exceeds ±1.36 radians, OAMs detection fails using the DM structure. However, due to the advantages of its structure, the OAMs’ mode can be effectively detected by the OSM structure when DM detection fails, which significantly improves the detection range of the OAMs. Furthermore, in the same detection range, the OSM can also generate stronger focusing spots than DM, as shown in Figure 1d (normalization is performed here for the convenience of comparison). According to the above information, the OSM has higher efficiency than DM for PTM, which makes it have a larger mode resolution range for OAMs, and improves the practical application value in related fields.

3.2. Model Establishment and Simulation

Then, an l0 = 200 OSM was designed to ensure the ease of production, and the left circular polarization (LCP) OAM was simulated and sequenced at wavelength 633 nm. Figure 2a illustrates the three-dimensional structure of the OSM. The Λ of each column of the catenary in the OSM decreases gradually from right to left. Considering the detection ability of the OSM and the feasibility of the device fabrication, the l0 was set to 200, and the corresponding minimum Λ value of the metasurface was about 300 nm. The OSM was designed as a streamlined array with H = 600 nm and interval d1 = 700 nm, where w1 = 100 nm was the optimized streamlined catenary. To ensure the feasibility of manufacturing, the streamline was truncated by d2 = 0.75Λ, and vertical bars of width w2 = 100 nm were added to the truncated region to increase phase continuity and amplitude uniformity. Furthermore, a ring OSM with ri = 50 μm and ro = 80 μm was selected as an example of a proof-of-concept demonstration. The opening of the streamline catenary in the upper semicircle of the metasurface was outward, while the opening of the lower semicircle was inward. Then, the streamline catenary Λ gradually decreases with the circumference from right to left, and the Λ position reaches the minimum at the left radius of the inner circle. In addition, TiO2 was selected as the constituent material of the OSM due to its high refractive index and low loss at visible frequency, and it realized this larger aspect ratio processing in metalenses at visible wavelengths [42], meta-optics [43], and spin-decoupled metasurfaces [37]. Then, Figure 2b shows the ordering of the OAMs by PTM. The designed OSM transforms vortex beams with different topological charges into circular focusing modes in the transverse focal plane by Pancharatnam–Berry phase regulation. Thus, OAMs can be determined through a single measurement, giving them simple and fast sorting capabilities that may find potential applications in quantum communication.
Then, FDTD simulation was used to simulate the designed OSM, and the simulation results are shown in Figure 3a (the LCP is filtered, and the RCP is retained to obtain a sharper focused image). Here, the FDTD simulation results of the different OAMs with topological charges between −600 and 550 through the PMT were shown, where the adjacent interval Δl was 50. According to the simulation results, it can be seen that all the focusing modes under the same focal length are located on the same circumference, and the spot position rotates clockwise as the topological charge increases from negative to positive. Moreover, when the topological charge is in the range of l = ±2.5l0, the focus spot is double-arc cross-focusing, and the intensity of the focus spot is stronger. Then, when the topological charge changes in the range of ±2.5l0 ~±3.14l0, the focus gradually becomes a single arc, and the intensity gradually decreases. However, the position of the focal spot corresponding to the theoretical value does not change, thus it is still effective for OAM detection in this range.
To verify the accuracy of the OSM simulation, the corresponding relationship between simulation results and theoretical calculated values is given here, as shown in Figure 3b. Then, we can obtain that the azimuth is linearly related to the topological charge, and the difference between the azimuth coordinates of FDTD simulation and the theoretical calculated values is not more than 0.001 rad. For the coefficient l0 = 200 of the OSM, the azimuth interval between the intensity peaks of the two superimposed adjacent modes is 1/l0 = 0.005 rad. Therefore, the topological charge can be unambiguously predicted as follows:
l predict = l 0 ϕ l
Theoretically, the designed OSM can detect the intensity modes of 1256 OAMs’ modes with topological charges ranging from −628 to 628, indicating an excellent OAM mode detection ability. In summary, the OSM has a wide mode resolution range for single isolated OAMs. In addition, the position of the focusing spot formed by the OAM through the OSM has a very high coincidence degree with the theory, which is conducive to reverse inferring the topological charge of incident vortex light from the position of the focal spot in a practical application and dramatically improves the detection accuracy.

3.3. Resolution of Superimposed OAMs

The resolution of multiple superimposed OAM modes was further investigated here. As shown in Figure 4a, the overlap area of two adjacent focusing spots decreases with the increase in Δl, their mutual interference gradually decreases, and their resolution gradually increases. When the topological charge interval is seven, the intensity curve of two-phase OAMs reduces to 38.2%, and the two focal spots can be distinguished clearly. Based on the interval Δl = 7, 180 different OAM modes ((200 × 2π)/7 + 1 = 180) can be detected simultaneously on the circle, and the large mode detection capability is demonstrated. In addition, decreasing the size of the focusing spot or increasing the radius of the focusing ring is expected to improve the angular resolution, such as adding a super-oscillatory phase on the original metasurface or designing a helical focusing mode.
To verify the validity of the OSM detection on the whole circumference, in this paper, FDTD was used to simulate the single topological charge, and then all the single simulation results were superimposed together to obtain all the focusing modes, as shown in Figure 4b. To receive a more uniform focusing light intensity, the theoretical focal length ( f = k 0 r 0 2 l 0 ) is increased by 50 μm as a new detection focal length. In addition, the intensity of the single OAM is normalized to compare two adjacent focusing spots more conveniently. As is shown in Figure 4c, for OAM with different topological charges, the azimuth Angle difference between adjacent normalized intensity peaks is about 0.05 rad, and adjacent peaks are separated obviously. Then, it can be concluded that 1256 topological charges can be detected by the OSM when measuring a single topological charge. Moreover, the OAMs of 180 different topological charges can be detected simultaneously in multiple topological detections. This improvement in the detection efficiency indicates that the OSM can detect large-mode OAMs efficiently and quickly, suggesting the application of the OSM on integrated optical systems that require high-capacity communications.

4. Discussion

In the present article, we first investigate the theoretical differences between the OSM and DM, showing that the OSM is easier to function in space-limited situations due to the unique properties of its individual microstructure when the full-model size is equal. Then, the streamlined structure has greater sampling efficiency than the discrete structure, thus the vortex light forms a more concentrated focus spot after passing through the metasurface, which is beneficial to improve the mode resolution ability. When DM arranges the modes of the OAMs, there is a large space in the left part of the focusing plane that cannot generate the focusing spot, reducing the maximum detection range and maximum detection value. The area in the left part of the focal plane where focal spots cannot be generated is related to the single unit structure in DM and the l0 designed in the full model, as shown in Equation (12). When the l0 of the full model is constant, the period p of the unit structure is smaller, and the  ϕ max  is larger.
However, this problem does not exist in the OSM, which can effectively use the space in the focal plane and form the focused spot throughout the circumference, as shown in Figure 4b. The above advantages make the OSM significantly increase the detection range and maximum detection value of the OAMs’ modes and do not reduce the mode resolution when multiple superimposed OAMs. It cannot resolve two superimposed vortices with a mode interval smaller than Δl = 5, and it is consistent with the described spin-decoupled metasurface [37]. Still, its designed l0 is only 40, and the OAMs’ modes resolution range is within ±100 topology charge, which is far below the maximum mode detection range in this paper. Due to the characteristics of the catenary metasurface, it can only regulate LCP (or RCP) at the same time. The OSM is designed for incident vortex for LCP and is not suitable for the vortex of RCP in this paper. In addition, there are some limitations on the resolution of adjacent OAM, and strict alignment is also required during detection.
Then, we demonstrated a resolving power of Δl = 1 when only a single isolated OAM mode is received at a time, there was no cross-talk between adjacent modes. However, when multiple overlapping vortices are received simultaneously, the intensity overlap between them leads to reduced resolution power. As shown in Figure 4a, FDTD simulation results show that they cannot resolve two superimposed vortices with a mode interval smaller than Δl = 5, which means that at most 180 superimposed vortices can be simultaneously detected. Similar mode overlaps and resolution limitations have been observed in the angular lens-based OAM sorting methods [35]. Since the perimeter of the ring in which the focus spot is located in the focal plane is fixed, as is the size of the focus spot, the resolution can be further improved by decreasing the focus spot (adding a super-oscillatory phase) or enlarging the perimeter of the ring (designing it into a spiral arrangement pattern). In addition, a modified azimuthal-quadratic phase mask can also effectively improve the mode resolution power, which is expressed as  ϕ l = n 2 l l 0  [37] and indicates that the azimuth difference of adjacent OAM modes becomes n2 times the original one. The detectable mode range is proportional to  l 0 / n 2 , while the resolving power is inversely proportional to  l 0 / n 2 . With  l 0 / n 2  decreasing, the mode resolving power for superimposed vortices increases, but it reduces the range of detectable modes. Therefore, the coefficients l0 and n can be selected in practical applications according to the requirements of detection range and resolution.
For broadband PMTs and vortex sorting, we carried out simple simulation verifications for four wavelengths of 480, 532, 633, and 780 nm, although the focal length is wavelength dependent, and the focal depth is long enough that we can choose a transverse plane across all the focal planes. The results are not significantly different from the spin-decoupled metasurface [37], and it is possible to simultaneously detect wideband vortices covering the entire visible band, which greatly increases its potential application value. For example, there are two main ways to use vortex light in optical communication systems: one is to use vortex beams of the different OAM states as different carriers to transmit information (OAM division multiplexing); another method is to use different OAM states in the vortex as symbols of information transmission for data transmission (OAM coding). Therefore, the system that can easily and quickly identify the large capacity of the vortex with different OAM states has important potential application value for increasing the capacity of optical communication systems.

5. Conclusions

For the OSM structure, a single streamlined line can cover 2π phase control while multiple units are required in the DM structure. Therefore, in the limited space of the metasurface, the OSM can arrange more columns than the DM to achieve larger phase gradient regulation, thus significantly increasing the detection range and maximum detection value of the OAM. Moreover, the OSM also improves the problem of insufficient wavefront sampling of the DM, thus the detection resolution of the superimposed OAMs is improved. Then, when the detection range of a single OAM is greatly increased, a certain resolution of multiple superimposed OAMs can also be maintained. In addition, similar to the DM, the OSM can simultaneously detect wideband vortices covering the whole visible band. Then, in addition to being ultra thin and ultra light, the OSM also has an overall ultra-small size, making it highly integrated. At the same time, due to the large mode resolution ability of the OAMs, they can easily and quickly resolve high-capacity information, which makes it a potential application value in the medium application of optical quantum computers.
Then, we investigated the mode resolution of a single OAM and multiple superimposed OAMs and demonstrated its huge OAM mode resolution power. Furthermore, the mode detection method of the OAMs is shown. The whole optical system is straightforward, and the mode resolution method is also straightforward and fast. To predict the topological charge of the vortex, it is only necessary to determine the location of the focusing spot formed on the focal plane after the vortex light passes through the metasurface. This simple and effective system will have high stability in the practical application, thus it can adapt to various working environments and increase its potential application value.
It is worth mentioning that more structure arrays can be arranged on the circumference of the metasurface by adding the inner and outer radii of the full model, i.e., a more significant phase gradient can be controlled, and the scope of the OAMs’ mode sorting can be further expanded. On the contrary, if the system needs to be integrated with limited space, the inner and outer circle radius of the metasurface can be reduced to meet the requirements of the system. Of course, part of the range of the OAMs’ mode sorting will be lost. Moreover, if the demand for multiple superimposed OAMs’ mode resolution is high, increasing the mode resolution power can be achieved by appropriately decreasing the coefficient l0. The scheme is simple, practical, convenient, and flexible and can be applied to many application scenarios, such as large-capacity OAM communication and highly integrated optical quantum systems.
In summary, we showed a comparison of the OAM discrimination efficiency between the OSM and DM, proving that the OSM has higher mode discrimination efficiency than DM. Then, we have demonstrated a single geometric TiO2 OSM-based OAM mode discrimination method via PMT of the azimuthal-quadratic phase metasurface, where vortex beams are converted into focusing modes rotating clockwise on the screen as topological charges increase. Finally, we showed that the proposed approach could be extended to the ordering of superimposed OAMs with a proper mode interval. The results show that the OSM can significantly increase the mode detection range of the OAMs but does not reduce the ability to rank multiple superimposed OAMs’ modes. These results may have many important applications in momentum measurement of spin and angular momentum and detection of phase and polarization singularities.

Author Contributions

G.F. completed the optimization of the graphic structure and designed the final full model for simulation, data analysis, and processing. S.C. provided the initial simulation code. L.X.; Y.W.; F.Z. and Y.L. participated in the revision of the first draft. writing—review and editing, Q.H., X.L. and Y.G. All authors participated in the discussion and contributed to improving the manuscript. M.P. and X.L. co-supervised and directed the study. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant Nos. 61975210 and U20A20217) and the Natural Science Foundation of Sichuan (2021ZYCD001, 2020YFJ0001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No link to the archived dataset is publicly available.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Poynting, J.H. The wave motion of a revolving shaft, and a suggestion as to the angular momentum in a beam of circularly polarised light. Proc. R. Soc. London Ser. A Contain. Pap. A Math. Phys. Character 1909, 82, 560–567. [Google Scholar]
  2. Beth, R.A. Mechanical Detection and Measurement of the Angular Momentum of Light. Phys. Rev. 1936, 50, 115–125. [Google Scholar] [CrossRef]
  3. Yao, A.M.; Padgett, M.J. Orbital angular momentum: Origins, behavior and applications. Adv. Opt. Photon 2011, 3, 161–204. [Google Scholar] [CrossRef] [Green Version]
  4. Allen, L.; Beijersbergen, M.W.; Spreeuw, R.J.C.; Woerdman, J.P. Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes. Phys. Rev. A 1992, 45, 8185–8189. [Google Scholar] [CrossRef] [PubMed]
  5. Park, S.C.; Park, M.K.; Kang, M.G. Super-resolution image reconstruction: A technical overview. IEEE Signal Process. Mag. 2003, 20, 21–36. [Google Scholar] [CrossRef] [Green Version]
  6. Mair, A.; Vaziri, A.; Weihs, G.; Zeilinger, A. Entanglement of the orbital angular momentum states of photons. Nature 2001, 412, 313–316. [Google Scholar] [CrossRef] [Green Version]
  7. Shen, Y.; Wang, X.; Xie, Z.; Min, C.; Fu, X.; Liu, Q.; Gong, M.; Yuan, X. Optical vortices 30 years on: OAM manipulation from topological charge to multiple singularities. Light. Sci. Appl. 2019, 8, 1–29. [Google Scholar] [CrossRef] [Green Version]
  8. Wang, J.; Yang, J.-Y.; Fazal, I.M.; Ahmed, N.; Yan, Y.; Huang, H.; Ren, Y.; Yue, Y.; Dolinar, S.; Tur, M.; et al. Terabit free-space data transmission employing orbital angular momentum multiplexing. Nat. Photonics 2012, 6, 488–496. [Google Scholar] [CrossRef]
  9. Willner, A.E.; Huang, H.; Yan, Y.; Ren, Y.; Ahmed, N.; Xie, G.; Bao, C.; Li, L.; Cao, Y.; Zhao, Z.; et al. Optical communications using orbital angular momentum beams. Adv. Opt. Phot. 2015, 7, 66–106. [Google Scholar] [CrossRef] [Green Version]
  10. Fang, X.; Ren, H.; Gu, M. Orbital angular momentum holography for high-security encryption. Nat. Photonics 2020, 14, 102–108. [Google Scholar] [CrossRef]
  11. Zhou, H.; Sain, B.; Wang, Y.; Schlickriede, C.; Zhao, R.; Zhang, X.; Wei, Q.; Li, X.; Huang, L.; Zentgraf, T. Polarization-encrypted orbital angular momentum multiplexed metasurface holography. ACS Nano 2020, 14, 5553–5559. [Google Scholar] [CrossRef] [PubMed]
  12. Ruffato, G.; Massari, M.; Romanato, F. Multiplication and division of the orbital angular momentum of light with diffractive transformation optics. Light. Sci. Appl. 2019, 8, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wen, Y.; Chremmos, I.; Chen, Y.; Zhang, Y.; Yu, S. Arbitrary Multiplication and Division of the Orbital Angular Momentum of Light. Phys. Rev. Lett. 2020, 124, 213901. [Google Scholar] [CrossRef] [PubMed]
  14. Khonina, S.N.; Kotlyar, V.V.; Shinkaryev, M.V.; Soi-Fer, V.A.; Uspleniev, G.V. The Phase Rotor Filter. J. Mod. Opt. 1992, 39, 1147–1154. [Google Scholar] [CrossRef]
  15. Beijersbergen, M.W.; Coerwinkel, R.P.C.; Kristensen, M.; Woerdman, J.P. Helical-wavefront laser beams produced with a spiral phase plate. Opt. Commun. 1994, 112, 321–327. [Google Scholar] [CrossRef]
  16. Bouchard, F.; Mand, H.; Mirhosseini, M.; Karimi, E.; Boyd, R.W. Achromatic orbital angular momentum generator. New J. Phys. 2014, 16, 123006. [Google Scholar] [CrossRef] [Green Version]
  17. Biener, G.; Niv, A.; Kleiner, V.; Hasman, E. Formation of helical beams by use of Pancharatnam–Berry phase optical elements. Opt. Lett. 2002, 27, 1875–1877. [Google Scholar] [CrossRef]
  18. Marrucci, L.; Manzo, C.; Paparo, D. Optical Spin-to-Orbital Angular Momentum Conversion in Inhomogeneous Anisotropic Media. Phys. Rev. Lett. 2006, 96, 163905. [Google Scholar] [CrossRef] [Green Version]
  19. Meng, W.; Hua, Y.; Cheng, K.; Li, B.; Liu, T.; Chen, Q.; Luan, H.; Gu, M.; Fang, X. 100 Hertz frame-rate switching three-dimensional orbital angular momentum multiplexing holography via cross convolution. Opto-Electronic Sci. 2022, 1, 220004. [Google Scholar] [CrossRef]
  20. Zhang, X.; Liang, G.; Feng, D.; Zhou, L.; Guo, Y. Ultra-broadband metasurface holography via quasi-continuous nano-slits. J. Phys. D Appl. Phys. 2019, 53, 104002. [Google Scholar] [CrossRef]
  21. Yuan, X.; Xu, Q.; Lang, Y.; Jiang, X.; Xu, Y.; Chen, X.; Han, J.; Zhang, X.; Han, J.; Zhang, W. Tailoring spatiotemporal dynamics of plasmonic vortices. Opto-Electronic Adv. 2023, 220133. [Google Scholar] [CrossRef]
  22. Pu, M.; Li, X.; Ma, X.; Wang, Y.; Zhao, Z.; Wang, C.; Hu, C.; Gao, P.; Huang, C.; Ren, H.; et al. Catenary optics for achromatic generation of perfect optical angular momentum. Sci. Adv. 2015, 1, e1500396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Guo, Y.; Pu, M.; Zhao, Z.; Wang, Y.; Jin, J.; Gao, P.; Li, X.; Ma, X.; Luo, X. Merging Geometric Phase and Plasmon Retardation Phase in Continuously Shaped Metasurfaces for Arbitrary Orbital Angular Momentum Generation. ACS Photon. 2016, 3, 2022–2029. [Google Scholar] [CrossRef]
  24. Chen, J.; Wang, D.; Si, G.; Teo, S.L.; Wang, Q.; Lin, J. Planar peristrophic multiplexing metasurfaces. Opto-Electronic Adv. 2023, 220141. [Google Scholar] [CrossRef]
  25. Lv, Y.; Shang, Z.; Fu, S.; Huang, L.; Gao, L.; Gao, C. Sorting orbital angular momentum of photons through a multi-ring azimuthal-quadratic phase. Opt. Lett. 2022, 47, 5032. [Google Scholar] [CrossRef]
  26. Harris, M.; Hill, C.; Vaughan, J. Optical helices and spiral interference fringes. Opt. Commun. 1994, 106, 161–166. [Google Scholar] [CrossRef]
  27. Leach, J.; Padgett, M.; Barnett, S.; Franke-Arnold, S.; Courtial, J. Measuring the Orbital Angular Momentum of a Single Photon. Phys. Rev. Lett. 2002, 88, 257901. [Google Scholar] [CrossRef]
  28. de Araujo, L.E.E.; Anderson, M.E. Measuring vortex charge with a triangular aperture. Opt. Lett. 2011, 36, 787–789. [Google Scholar] [CrossRef] [Green Version]
  29. Liu, Z.; Gao, S.; Xiao, W.; Yang, J.; Huang, X.; Feng, Y.; Li, J.; Liu, W.; Li, Z. Measuring high-order optical orbital angular momentum with a hyperbolic gradually changing period pure-phase grating. Opt. Lett. 2018, 43, 3076–3079. [Google Scholar] [CrossRef]
  30. Estakhri, N.M.; Edwards, B.; Engheta, N. Inverse-designed metastructures that solve equations. Science 2019, 363, 1333–1338. [Google Scholar] [CrossRef]
  31. Lalegani, Z.; Ebrahimi, S.S.; Hamawandi, B.; La Spada, L.; Batili, H.; Toprak, M. Targeted dielectric coating of silver nanoparticles with silica to manipulate optical properties for metasurface applications. Mater. Chem. Phys. 2022, 287, 126250. [Google Scholar] [CrossRef]
  32. Akbari, M.; Shahbazzadeh, M.J.; La Spada, L.; Khajehzadeh, A. The Graphene Field Effect Transistor Modeling Based on an Optimized Ambipolar Virtual Source Model for DNA Detection. Appl. Sci. 2021, 11, 8114. [Google Scholar] [CrossRef]
  33. Greybush, N.J.; Pacheco-Peña, V.; Engheta, N.; Murray, C.B.; Kagan, C.R. Plasmonic Optical and Chiroptical Response of Self-Assembled Au Nanorod Equilateral Trimers. ACS Nano 2019, 13, 1617–1624. [Google Scholar] [CrossRef] [PubMed]
  34. Lincoln, R.L.; Scarpa, F.; Ting, V.P.; Trask, R.S. Multifunctional composites: A metamaterial perspective. Multifunct. Mater. 2019, 2, 043001. [Google Scholar] [CrossRef]
  35. Sahu, R.; Chaudhary, S.; Khare, K.; Bhattacharya, M.; Wanare, H.; Jha, A.K. Angular lens. Opt. Express 2018, 26, 8709–8718. [Google Scholar] [CrossRef]
  36. Pu, M.; Li, X.; Guo, Y.; Ma, X.; Luo, X. Nanoapertures with ordered rotations: Symmetry transformation and wide-angle flat lensing. Opt. Express 2017, 25, 31471–31477. [Google Scholar] [CrossRef]
  37. Guo, Y.; Zhang, S.; Pu, M.; He, Q.; Jin, J.; Xu, M.; Zhang, Y.; Gao, P.; Luo, X. Spin-decoupled metasurface for simultaneous detection of spin and orbital angular momenta via momentum transformation. Light. Sci. Appl. 2021, 10, 1–12. [Google Scholar] [CrossRef]
  38. Zhang, F.; Pu, M.; Li, X.; Ma, X.; Guo, Y.; Gao, P.; Yu, H.; Gu, M.; Luo, X. Extreme-Angle Silicon Infrared Optics Enabled by Streamlined Surfaces. Adv. Mater. 2021, 33, e2008157. [Google Scholar] [CrossRef]
  39. Luo, X.-G.; Pu, M.-B.; Li, X.; Ma, X.-L. Broadband spin Hall effect of light in single nanoapertures. Light. Sci. Appl. 2017, 6, e16276. [Google Scholar] [CrossRef] [Green Version]
  40. Wang, D.; Hwang, Y.; Dai, Y.; Si, G.; Wei, S.; Choi, D.; Gómez, D.E.; Mitchell, A.; Lin, J.; Yuan, X. Broadband High-Efficiency Chiral Splitters and Holograms from Dielectric Nanoarc Metasurfaces. Small 2019, 15, e1900483. [Google Scholar] [CrossRef]
  41. Yin, X.; Ye, Z.; Rho, J.; Wang, Y.; Zhang, X. Photonic spin Hall effect at metasurfaces. Science 2013, 339, 1405–1407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. 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] [PubMed] [Green Version]
  43. Li, Z.; Lin, P.; Huang, Y.-W.; Park, J.-S.; Chen, W.T.; Shi, Z.; Qiu, C.-W.; Cheng, J.-X.; Capasso, F. Meta-optics achieves RGB-achromatic focusing for virtual reality. Sci. Adv. 2021, 7, 4458. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Comparison of the OSM and DM for detection and separation of the different OAMs’ modes at 633 nm. (a). PMT diagram of LCP vortex light passing through a phase plate when l0 = 200, ri = 50 μm, and ro = 80 μm. (b). OSM (l0 = 200) simulation results of the OAMs at l = −620 to l = 620, Δl = 10. (c). Simulation results of DM (l0 = 200) with period p = 600 nm and under the same conditions as (b). According to the square result, the detection fails when  ϕ l  exceeds ±1.36 rad. (d). Intensity curves along the white dotted circle line in (b,c).
Figure 1. Comparison of the OSM and DM for detection and separation of the different OAMs’ modes at 633 nm. (a). PMT diagram of LCP vortex light passing through a phase plate when l0 = 200, ri = 50 μm, and ro = 80 μm. (b). OSM (l0 = 200) simulation results of the OAMs at l = −620 to l = 620, Δl = 10. (c). Simulation results of DM (l0 = 200) with period p = 600 nm and under the same conditions as (b). According to the square result, the detection fails when  ϕ l  exceeds ±1.36 rad. (d). Intensity curves along the white dotted circle line in (b,c).
Applsci 13 03237 g001
Figure 2. The design of the OSM and PTM detection diagram. (a). Schematic diagram of the OSM with SiO2 substrate. The OSM graphic structure consists of TiO2 streamline arrays with H = 600 nm, d1 = 700 nm, w1 = 100 nm, ri = 50 μm, and ro = 80 μm. The streamlines are truncated (d2 = 0.75Λ) to guarantee the feasibility of fabrication. The vertical structure with width w2 = 100 nm was added in the truncated region to improve the phase continuity and amplitude uniformity. (b). Schematic diagram of the OSM detection, OAMs sequencing by spin-controlled PMT. Vortex beams with different spins are converted into circular focusing modes on the screen in the transverse focal plane with topological charge-dependent azimuth rotation.
Figure 2. The design of the OSM and PTM detection diagram. (a). Schematic diagram of the OSM with SiO2 substrate. The OSM graphic structure consists of TiO2 streamline arrays with H = 600 nm, d1 = 700 nm, w1 = 100 nm, ri = 50 μm, and ro = 80 μm. The streamlines are truncated (d2 = 0.75Λ) to guarantee the feasibility of fabrication. The vertical structure with width w2 = 100 nm was added in the truncated region to improve the phase continuity and amplitude uniformity. (b). Schematic diagram of the OSM detection, OAMs sequencing by spin-controlled PMT. Vortex beams with different spins are converted into circular focusing modes on the screen in the transverse focal plane with topological charge-dependent azimuth rotation.
Applsci 13 03237 g002
Figure 3. Test results of different topological OAM in simulation at 633 nm. (a). Detection results of the OAM topological charges from −600 to 550 in simulation (Δl = 50). (b). Comparison of azimuth coordinates of the OAM focusing peak of different topological charges between theoretical and FDTD simulation. (c). Topological charge of the OAM predicted by azimuth coordinate theory.
Figure 3. Test results of different topological OAM in simulation at 633 nm. (a). Detection results of the OAM topological charges from −600 to 550 in simulation (Δl = 50). (b). Comparison of azimuth coordinates of the OAM focusing peak of different topological charges between theoretical and FDTD simulation. (c). Topological charge of the OAM predicted by azimuth coordinate theory.
Applsci 13 03237 g003
Figure 4. Resolution efficiency of the OSM of wavelength 633 nm. (a). Focusing mode of two superimposed OAMs’ modes with different mode intervals simulated by FDTD. The OAMs interval increases from Δl = 4 to Δl = 7, and the left figure shows the intensity mode of the two topological charge focusing spots, and the right figure shows the intensity curve of the corresponding dashed white line. (b). FDTD simulation results of PMT with 124 superimposed OAM modes and mode interval Δl = 10. (c). Normalized intensity curve along the white dotted round line in b.
Figure 4. Resolution efficiency of the OSM of wavelength 633 nm. (a). Focusing mode of two superimposed OAMs’ modes with different mode intervals simulated by FDTD. The OAMs interval increases from Δl = 4 to Δl = 7, and the left figure shows the intensity mode of the two topological charge focusing spots, and the right figure shows the intensity curve of the corresponding dashed white line. (b). FDTD simulation results of PMT with 124 superimposed OAM modes and mode interval Δl = 10. (c). Normalized intensity curve along the white dotted round line in b.
Applsci 13 03237 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fu, G.; Chen, S.; He, Q.; Xiong, L.; Wen, Y.; Zhang, F.; Lu, Y.; Guo, Y.; Pu, M.; Luo, X. Optimized Catenary Metasurface for Detecting Spin and Orbital Angular Momentum via Momentum Transformation. Appl. Sci. 2023, 13, 3237. https://0-doi-org.brum.beds.ac.uk/10.3390/app13053237

AMA Style

Fu G, Chen S, He Q, Xiong L, Wen Y, Zhang F, Lu Y, Guo Y, Pu M, Luo X. Optimized Catenary Metasurface for Detecting Spin and Orbital Angular Momentum via Momentum Transformation. Applied Sciences. 2023; 13(5):3237. https://0-doi-org.brum.beds.ac.uk/10.3390/app13053237

Chicago/Turabian Style

Fu, Guoquan, Siran Chen, Qiong He, Lingxing Xiong, Yifeng Wen, Fei Zhang, Yuran Lu, Yinghui Guo, Mingbo Pu, and Xiangang Luo. 2023. "Optimized Catenary Metasurface for Detecting Spin and Orbital Angular Momentum via Momentum Transformation" Applied Sciences 13, no. 5: 3237. https://0-doi-org.brum.beds.ac.uk/10.3390/app13053237

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