Plasmonic metasurfaces manipulating the two spin components from spin–orbit interactions of light with lattice field generations
-
Ruirui Zhang
, Manna Gu
, Yuqin Zhang
Abstract
Artificial nanostructures in metasurfaces induce strong spin–orbit interactions (SOIs), by which incident circularly polarized light can be transformed into two opposite spin components. The component with an opposite helicity to the incident light acquires a geometric phase and is used to realize the versatile functions of the metasurfaces; however, the other component, with an identical helicity, is often neglected as a diffused background. Here, by simultaneously manipulating the two spin components originating from the SOI in plasmonic metasurfaces, independent wavefields in the primary and converted spin channels are achieved; the wavefield in the primary channel is controlled by tailoring the dynamic phase, and that in the converted channel is regulated by designing the Pancharatnam–Berry phase in concurrence with the dynamic phase. The scheme is realized by generating optical lattice fields with different topologies in two spin channels, with the metasurfaces composed of metal nanoslits within six round-apertures mimicking the multi-beam interference. This study demonstrates independent optical fields in a dual-spin channel based on the SOI effect in the metasurface, which provides a higher polarization degree of freedom to modify optical properties at the subwavelength scale.
1 Introduction
Photonic spin–orbit interactions (SOIs) are optical phenomena in which the spin affects the spatial distribution and propagation of light [1]; it is related to the geometric phase and interaction between the spin angular momentum (SAM) and orbital angular momentum (OAM). Recently, SOI has gained significant interest in the exploration and understanding of a variety of unusual physical effects. The optical spin Hall effect [2], [3], [4] is a fundamental phenomenon, originating from SOI, which exhibits a spin-dependent transverse beam shift at the medium interface or through inhomogeneous media. Another important effect of SOI is spin-to-orbit conversions (SOC) [5], which is realized in the focusing of nonparaxial beams and scattering from small particles [6]. Metamaterials and metasurfaces, as inhomogeneous and anisotropic media, also demonstrate strong SOI effects, and provide powerful spin-dependent control of wavefronts [7, 8] and easier implementation of SOC. More interestingly, in the surface mode of evanescent waves, the optical quantum spin Hall effect [9] appears and exhibits strong spin-momentum locking, which originates from the transverse spin locked to a unidirectional propagation, and is similar to the surface state of the topological insulator. Overall, the effects in SOIs have offered numerous important applications in broad areas, such as nanodevice design [10], precision metrology [11], topological insulators [12], and quantum information processing [13, 14].
As a paradigm of the enhanced SOI effect, metasurfaces are planar interfaces composed of artificial nanostructures with various dimensions and orientations [15, 16]; they regulate the amplitude, phase, and polarization of the output light via the interaction between the SAM and nanostructures [1]. Owing to their extraordinary capabilities in reshaping optical wavefronts into desired profiles, metasurfaces have been applied to metalenses [17, 18], holography [19], [20], [21], vortex generators [22, 23], and polarization controllers [24], [25], [26]. An incident circularly polarized beam, of helicity σ (σ = ±1) for left and right circular polarization (LCP and RCP), transmitted through the metasurface is decomposed into two components [27]. One component has the same helicity as the incident light, referred to as the primary spin component, and the other has the opposite helicity with the PB phase Φg = 2σθ, referred to as the converted spin component, which is the field to be manipulated in most common metasurface design. The converted spin components under the incident helicity ±σ are conjugate, owing to the opposite PB phase. The combination of the dynamic phase and PB phase is a novel and potent SOI scheme for helicity-switchable metasurfaces, which enables the imposition of independent phases on the converted component under reversal of incident helicity σ, realizing an independent control of the output field. This scheme has notably extended the applications of metasurfaces to spin selective holography [28, 29] and vector holography [30], vector beam generation with SOC [22, 31], full-Stokes polarization camera design [32], and achromatic imaging and lensing [33, 34].
Produced by the interference of multiple laser beams, the photonic lattice fields have periodic intensity and polarization patterns, and they were originally used as the periodic potentials to form artificial cold atom lattices [35]. The forms of the lattice fields depend on the intersection geometry and the phase configuration of the beams, and the familiar types of optical lattice fields of different orders include the honeycomb, Kagome, and hexagon vortex lattices [36]. In recent years, the creation of photonic lattice fields has become an interesting novel topic, and the fields can be translated into materialized lattices, which have enabled the newly emerged field of topological photonics [37, 38] and other areas of research [39, 40]. Exemplarily, the lattice fields enable the construction of pseudospin states [41] with different degrees of freedom, such as OAM [42], polarization [43], and crystal lattice symmetry [41, 44], to realize edge states without breaking time-reversal symmetry [45]. The topology of the nanoscale lattice fields can be controlled by SOI in metasurfaces and nanoslit polygons based on either the PB phase or its combination with the dynamic phase. With the PB phase metasurfaces, topology is achieved for the hexagonal vortex lattice through the quasi-Talbot effect [46], and the Kagome lattice through an additional prism deflection phase [47]. With the combination of the PB phase and dynamic phase, control over lattice topology is realized with more flexibility. By arraying metasurface nanostructures on truncated spirals, all four lattices, including the two other forms of lattices (namely, hexagonal and honeycomb), are acquired [48]. By repositioning a slit side or diagonal pair of slit sides of a hexagon to modulate the dynamic phase, Tsesses et al. realized the photonic skyrmion lattice and its transition from bubble-type to Néel-type [49], and have regulated different lattice topologies [50].
In the previous studies for generating lattice fields, only the converted spin component is considered, and two different fields of this component are obtained under the corresponding illuminations of the two spin helicities [46], [47], [48, 50], whereas the primary spin component has been typically neglected and scattered uncontrollably as futile background noise. With the dynamic phase to directly control field of the primary component, and the PB phase cooperatively to control the converted component, this work realizes the independent output of two spin fields under certain incident spin helicity, making a novel contribution to the literature by utilizing the primary spin component as a new channel for generating desired field. For implementation demonstration, the metasurfaces are designed with nanoslits distributed within six round apertures in the analogy of multi-beam interference to generate hexagon-based lattices, enabling lattice creation in the two spin channels. The dynamic phase is introduced by a spiral, on which the centers of the round-apertures are located, and the lattice topology in the primary spin channel is regulated into the desired form by adjusting the pitch of the spiral. Considering its combination with the dynamic phase, the PB phase from the varied orientation of the nanoslit in each round-aperture is designed accordingly to provide a phase shift between adjacent beams, and the lattice topology of the converted spin channel is controlled. Furthermore, when incident helicity σ is reversed to −σ, a different lattice of the converted component is acquired, demonstrating a helicity-dependent characteristic of the lattice topology, while the lattice topology of the primary component remains unchanged. The lattice fields of the two components exist in two respective spin channels, which overlap spatially, and they are experimentally acquired by a polarization filter (PF) composed of a quarter-wave plate and polarizer. The control over the dual-spin-channel lattices demonstrates that the SOI effect in metasurfaces provides the polarization degree of freedom for the manipulation of light field at the nanoscale, which is of significance to a variety of applications, in areas such as integrated photonic systems [51], topological photonics [52, 53], and ultra-thin quantum metadevices [54, 55].
2 Principle analysis and structure design
The metasurface is designed, based on the SOI effect, to manipulate the light field into two channels – of primary and converted spin – with the generation of optical lattice fields used as a practical demonstration; Figure 1(a) shows a schematic of the metasurface. N round apertures, composed of nanoslits with identical dimensions, lie on a spiral depicted by R = R0 + mλsppθ/2π in polar coordinates (R, θ), where R0 is the primary radius of the spiral, λspp is the surface plasmon polariton (SPP) wavelength, and δ r = mλspp describes the spiral pitch with integer m denoting the geometric order of the spiral. The azimuthal and radial coordinates of the center of the jth aperture are θ j = (2j + 1)π/N and R j = R0 + δ r θ j /2π, respectively, with j being an integer from 0 to N − 1, and the position vector of the central-slit is R j (R j , θ j ). The metasurface is intuitively designed to contain N = 6 apertures at the vertices of a hexagon. While the arrangement of the apertures provides the geometry for multi-beam-interference, the SOI of the nanoslits with the incident light will create two spin channels and provide different phase fronts for the expected wavefields.

Demonstration of principle and method for metasuface designs (a) Schematic of the metasurface. The radius of each aperture is a = 1.75 μm, the spiral at which the apertures are located has a primary radius R0 = 12 μm, and the thickness of the gold film is h = 200 nm. The nanoslits are arranged in a square lattice with lattice constant P R = P θ = λspp. The length of the slit L = 250 nm and width W = 110 nm. (b) Illustration of the slit orientation and local polar coordinates. (c) Illustration of the polarization filtering processing. χ and γ denote the fast axis and transmission axis directions of the QWP and polarizer, respectively. (d) Poincaré sphere for illustration of the conversion of the polarization state and geometric phase change of the converted spin component.
We now analyze the wave fields launched by a single nanoslit of the metasurface under incident circular polarization. As illustrated in Figure 1(b), within the jth aperture, a nanoslit is located at P(r, α) with position vector
r
in local polar coordinates (r, α) and the origin at the aperture center (R
j
, θ
j
). All nanoslits in the aperture have identical orientation angles φ
j
= α + φ(α), where φ
j
is a constant defined as the angle between the slit and x-axis in the local coordinate system, and φ(α) is the angle between the slit and radial unit vector
Using the expression of the unit vector
where r and α, labeled in E j (r, α; σ) on the left side of the equation, indicates that the SPP field from the nanoslit is described in the local polar coordinate system (r, α). Notably, the incident circular polarization E in, given by Eq. (1), transforming into the linear polarization of the outgoing wavefield E j (r, α; σ) with a phase shift, may enable control over the propagation and distribution of light, exhibiting a strong SOI effect.
To investigate the light field at observation point Q(ρ, β) near the center of the metasurface, we assume a global polar coordinate system (ρ, β) with unit vector (
The single slit at point P(r, α) can be regarded as a secondary source contributing to the wavefield E (ρ, β; σ) at point Q(ρ, β), with σ again representing the helicity of the illuminating light. Then, E (ρ, β; σ) represents the interference of the wavelets from all the slits in the metasurface, and it can be calculated as
where k sp is the wave vector of the wave propagating from the slit to point Q(ρ, β). Analogously, ρ and β, labeled in E (ρ, β; σ) on the left side of Eq. (4), represent the case of the wavefield being expressed in the global polar coordinate system (ρ, β). Equation (4) considers that the slits in each aperture are uniformly distributed, and the superposition of the wavelets has been changed from discrete slits to the integral over the area in the aperture. In the calculation of the position vector l = ρ − R j − r of point P relative to point Q, both the radius a of the aperture and radial coordinate ρ are significantly smaller than the radius R j , i.e., a ≪ R j and ρ ≪ R j . Then, the phase retardation in Eq. (4) is approximated as k sp⸱ l ≈ 2π[R j + r cos(α − θ j ) − ρ cos(β − θ j )]/λspp. Note that φ j = α + φ(α) reduces to φ j = θ j + π/2 from the geometry shown in Figure 1(b). Considering the variation of R j with θ j , and substituting Eqs. (2) and (3) into Eq. (4), we obtain
where the generating function exp[ikr cos(α − θ j )] of the integer-order Bessel function is expanded, and J k (kspr) represents the first type of Bessel function with order k. The integrals inside the second summation in Eq. (5) vanish if k ≠ 0, and E (ρ, β; σ) is derived as
with
where
The spatial distributions of the component fields E
L
(ρ, β; σ) and E
R
(ρ, β; σ) are dependent on two factors: one is the geometry of the arrangement of nanoslits in the metasurface, appearing as the phase
The wavefields of the primary and converted spin components can be manipulated in the two independent spin channels; however, they are spatially overlapped and indistinguishable, resulting in a mixed spatial output. As shown in Figure 1(c), we propose a separation of the two component fields in the circular polarization filtering system, which consists of a quarter-wave plate (QWP) and LP. In the global polar coordinate system (ρ, β), the operation of the QWP on a light field is expressed by its Jones matrix
where the upper bar designates the matrix, and χ is the fast axis angle of the QWP with respect to the x-axis. The Jones matrix of the LP is given by
where γ is the angle of the transmitting axis of the LP with respect to the x-axis. In the following discussion, we set the fast axis of the QWP in the direction of the y-axis for substantiated analysis, i.e., χ = π/2, and the operation of the QWP on the light field
E
(ρ, β; σ) at point Q given by Eqs. (6) and (7) creates the wavefield
E
qw(ρ, β; σ) =
where the two matrices
and
are the Jones vectors of the light wave linearly polarized at π/4 and in −π/4, respectively, in the (ρ, β) coordinate system. In deriving Eq. (10), we have used
and
where l1 = σ + m − 1 and l2 = σ + m + 1 are the orders of the lattices transmitted as linear polarizations
The different lattice topologies of the LCP and RCP components depend on the phases Φ1j = l1θ j and Φ2j = l2θ j corresponding to each aperture, if the geometry of the aperture arrangement is definite. Specifically, in the six-aperture regime of N = 6, the lattices include the hexagon-type topologies of hexagonal, hexagonal vortex, Kagome, and honeycomb [36], when the lattice order l = 0, ±1, ±2, and ±3, respectively, with l designating l1 and l2. For a lattice of order l, the wavefields of two adjacent apertures require a phase increment ΔΦ = lΔθ = 2lπ/N, with N = 6 and Δθ = θ j − θj−1 = 2π/N, or the cumulative phase difference is ΔΦt = 2lπ with aperture index j successively iterating through all the N apertures. Notably, the phase increments ΔΦ or ΔΦt are related to the three parameters σ, m, and the rotation of the slits. Considering the LCP incidence, the lattice order l1 of the primary spin component is directly determined by the dynamic phase ΔΦd = 2mπ related to the spiral pitch mλspp, achieving l1 = m. In contrast, the lattice topology of the converted spin component depends on both the dynamic phase ΔΦd and geometric phase ΔΦg = 4σπ from the variation of the slit orientations, and the lattice order is given by l2 = 2σ + m. This indicates that the lattice topology of the primary component is independent of the spin helicity σ and geometrical phase, and that of the converted component is spin-dependent and spin-asymmetric.
Then, the lattice topologies are specifically related to the nanoslit arrangement in a metasurface sample. We first consider a metasurface, denoted as sample A, with the six apertures located in a circle, resulting in parameter m = 0. From Eq. (13), the primary spin component is the LCP wavefield filtered as
We further examine the other case of the six apertures located on a spiral of geometrical order m = 1, with the metasurface denoted as sample B. Under the LCP incidence with σ = 1, the cumulative dynamic phase ΔΦd = 2mπ = 2π, owing to the spiral pitch mλspp, leads to the lattice order l1 = 1 for the primary spin component, which is the LCP wavefield filtered as
Parameters of samples A and B, denoting a metasurface with m = 0 and m = 1, respectively.
Samples | A | A | B | B |
|
||||
Geometric order of spiral m | 0 | 0 | 1 | 1 |
Incident spin state σ | 1 | −1 | 1 | −1 |
Lattice order l1 | 0 | −2 | 1 | −1 |
Lattice order l2 | 2 | 0 | 3 | 1 |
Phase increment l12π/N | 0 | −2π/3 | π/3 | −π/3 |
Phase increment l22π/N | 2π/3 | 0 | π | π/3 |
In addition, from the above analysis, we can see that the method may also be extended to design metasurfaces to generate more complicated structured and arrayed light fields, which are of greater significance. When the metasurfaces are used to generate more general light field, different complicated patterns can be generated in the two channels of the primary and the converted spin components, respectively, under certain circularly polarized light illumination; but when illumination is changed to other circularly polarized light, the opposite spin dependence of the geometric phases can result in limitations in the independent and arbitrary generation of the pattern of the converted spin component due to the asymmetric SOI [57, 60]. However, more sophisticated design of propagation phase and geometric phase profiles might overcome such limitations.
3 Simulation results and discussion
To demonstrate the generation of the dual-spin-channel lattices using symmetric and asymmetric SOI, we numerically simulated the lattice wavefields of samples A and B with the aperture centers distributed in a circle and spiral, respectively. The parameters of samples A and B are given in Table 1, and the simulations were implemented using a commercial finite-difference time-domain (FDTD) software (Lumerical Solutions). The simulation area was 30 μm × 30 μm × 4 μm, and the size of the Yee cell was 4 nm. Figure 2(a) and (b) schematically show light of LCP and RCP, respectively, with a wavelength of 532 nm irradiated on the metasurfaces from the substrate side; the wavelength of the SPP waves is λspp = 470 nm he overlapping dual-spin-channel lattice intensities of the primary and converted spin components can be observed in the interference areas at z = 1.5 μm on the air side.

Simulation results of the dual-spin channel lattices induced by metasurfaces of samples A and B: (a) and (b) schematic of the generation of the overlapping lattice fields under the incidence of LCP and RCP, respectively. (c1) and (e1) Sketch map of sample A. (d1) and (f1) Sketch map of sample B. (c2–c6) and (d2–d6) Simulated results of samples A and B, respectively, illuminated by LCP light. (e2–e6) and (f2–f6) Simulated results of samples A and B, respectively, illuminated by RCP light.
Figure 2(c1) and (e1) depict the schematic maps of metasurface sample A, and Figure 2(d1) and (f1) show the map of sample B, with the incidence of LCP and RCP, respectively. We first consider the results for sample A with the symmetric SOI. Under LCP illumination, the parameters are m = 0 and σ = 1, and the results are shown in Figure 2(c2–c6). Figure 2(c2) shows the unfiltered intensity pattern of a triangle lattice. Then, the polarization filtering operation is numerically simulated to extract the wavefields of the primary and converted spin components from the FDTD data, achieving the wavefields of
Under RCP illumination, the incident helicity was reversed to σ = −1. The unfiltered intensity pattern of the overlapped spin components is shown in Figure 2(e2); it still appears to be a triangle lattice. The intensity pattern of the primary spin component, which is the RCP field of
We now consider the results of sample B with an asymmetric SOI. For the case of the sample under illumination of LCP light, with parameters m = 1 and σ = 1, the results are shown in Figure 2(d2−d6). Therein, Figure 2(d2) shows the unfiltered intensity pattern with the primary and converted spin components overlapped. Figure 2(d3) is the intensity pattern of the primary spin component, which is the LCP field of
RCP light is used to illuminate sample B, with m = 1 and σ = −1, and the overlapped intensity pattern of the primary and converted components is shown in Figure 2(f2); interestingly, it depicts the pattern of a hexagonal vortex lattice. The lattice of the primary spin component is the RCP field of
4 Experimental results
Figure 3 shows the experimental results of the lattice patterns for samples A and B under LCP and RCP illuminations, where Figure 3(a–d) show the intensity patterns of the overlapping lattices without polarization filtering; Figure 3(e–h) and Figure 3(i–l) depict the intensity patterns of the primary and converted spin components, respectively. Each inset shows the corresponding magnified view of the central lattice unit. These results show that all lattice fields are well generated. Figure 3(m–p) and Figure 3(q–t) present the phase distributions of the primary and converted spin components, respectively. From Figure 3(e) and (f) and Figure 3(i) and (j), the intensities for sample A with an incidence of LCP and RCP are shown as the changeless hexagonal and Kagome lattices for the primary and converted spin components, respectively, demonstrating the symmetric SOI of the sample. The topologies of 0 and ±2 can be validated by the phase maps in Figure 3(m), (n) and (q), (r), respectively; i.e., identical phase distributions for the two hexagonal lattices of the primary spin components and opposite phase distributions for the two Kagome lattices of the converted spin components, except for the phase difference of π/2 between the two components with opposite spin states. Figure 3(g) and (h) display the intensity patterns of the hexagonal vortex lattice in the primary spin channel for sample B with an incidence of LCP and RCP, respectively, indicating spin-independent hexagonal vortex lattice topology of an order of 1. Figure 3(k) and (l) depict the honeycomb lattice and hexagonal vortex lattice of the converted spin component for the sample, indicating spin-dependent topologies of +3 and −1. Though on the whole, the experimental results validate the conclusions drawn from the above theoretical analysis and simulations, we also notice that there are some discrepancies between the experimental results in Figure 3 and the simulated results in Figure 2. These differences are mainly due to the unavoidable factors which may deteriorate the quality of the lattice field in the experimental performances, such as the errors of nanoslit sizes in fabricating metasurfaces, and the imperfectness of optical elements, and misalignment in the optical setup for the lattice field measurement.

Experimental results for samples A and B under illumination of LCP and RCP light: (a)–(d) Measured intensity patterns for the overlapping lattices of the mixed spin components (IM). (e)–(h) and (i)–(l) Measured intensity distributions for the primary spin components (IP) and converted spin components (IC), respectively, and the enlarged views of the center lattice unit, labeled in white circles, are shown in the insets. (m)–(p) and (q)–(h) Experimental phase maps (ΦP and ΦC) corresponding to the two spin components. All the black open circle arrows represent the polarization of the incident light.
5 Methods
Dual-spin-channel lattices were observed in experiments with the two metasurface samples perforated in a gold film using focused ion beam milling. The gold film was deposited on a glass substrate of 500 μm thickness. Figure 4(a) and (b) shows the scanning electron microscopy (SEM) images of samples A and B, the parameters of which are shown in Table 1. Figure 4(c) shows a schematic of the experimental setup of a Mach–Zehnder interferometer consisting of two beam splitters (BS1 and BS2) and two planar mirrors (M1 and M2). A laser, at a wavelength of 532 nm, and QWP (QWP1) were utilized to obtain circular polarized light to illuminate the sample from its substrate side. The intensity patterns of the overlapping lattices at the position Z = 1.5 μm behind the metasurface were magnified by a microscope objective lens (MO, numerical aperture = 0.9/100×) and then captured by a scientific complementary metal–oxide–semiconductor (s-CMOS) camera (Zyla-5.5, Andor, 16 bit). The PF, composed of a QWP (QWP2) and LP were placed before the s-CMOS to filter the primary or converted spin component from the overlapping lattices. A uniform spherical wave, serving as a reference beam, was obtained using a pinhole spatial filter (PSF), with which the beam of the lattice interfered. The intensity and interference patterns were recorded by the s-CMOS camera, and the phase distributions of the imaged lattices were reconstructed from the interference patterns.

SEM images for samples and experimental setup: (a) and (b) show the SEM images of samples A and B, respectively. (c) Schematic diagram of the experimental setup.
6 Discussions and conclusions
The present metasurface designs have concentrated on the normal incidence of the illuminating light. When the incident angle of the illuminating light is changed to θinc, the additional phase factor exp[ikspx sin θinc] is imposed on the incident light, and it will cause asymmetric spatial changes in the propagation phase differences for light waves of the six holes, and resultantly the phase differences of light fields are mismatched for the generation of designed lattices. This means that the designed metasurfaces are angle-dependent. However, when the wavelength of illuminating light is changed under normal incidence, the nanoslits with the symmetrical spatial arrangement in the apertures will correspond to the optical path changes in proportion for changed wavelength due to the phase factor
For further discussions, here we mention several classical designs of metasurfaces related to the combination of dynamic and geometric phases. In the earlier work, Xiao et al. [63] performed the independent manipulations of the plasmonic fields matching the two spin components converted from the illuminating lights by controlling the constructive interference to form the target fields. In the work by Mueller et al. [28], the adjustable dimensions of the dielectric nanopillars in the metasurfaces were designed to convert the chirality of two orthogonal polarization states with dynamic phase imposed simultaneously, and the dynamic and geometric phases was combined to control the light fields of the two converted polarizations, with two images achieved at the polarization channels. As a parallelism in terahertz regimes to the above work, the delicate metasurfaces with meta-atoms capable of converting chirality of the circular polarization and imposing dynamic phase were designed recently by Wang et al. [64], and the manipulations of the spoof surface plasmons corresponding to either converted circular polarizations were demonstrated by beam deflection and focusing. Compared to studies in the literature, the present work makes use of dynamic phase that is related to the optical path, rather than dynamic phase related to the dimension of the nanostructures of the metasurfaces. The functionality of a nanoslit as linear polarizer is directly used to generate the different lattice fields in the two spin channels and under illuminations of the two different circular polarizations. In contrast to the two images of the converted spin fields achieved in most of the literature, the present work has made use of the asymmetric SOI, and achieved another lattice field of either primary spin in addition to the two lattice fields of converted spins.
In summary, based on the SOI effect, the primary and converted spin components transformed from circularly polarized light are used as two spin channels for independent manipulations of desired light fields. The converted component is used extensively for light field manipulations, and its unusual and novel properties have been studied intensely; however, metasurfaces are becoming a novel platform of unprecedented convenience and flexibility for miniatured and planar element designs. The primary spin field, which is well controlled by the dynamic phase as the light field propagating in ordinary diffractions, has been overlooked to some degree. Moreover, it is often considered as a futile background signal, and some effort has been made to eliminate it in recent studies [65]. The present study has demonstrated the manipulation of the two spin component fields in a single-layer plasmonic metasurface. By mimicking multi-beam interference using rotational nanoslit arrays arranged in six round apertures, different hexagon-based lattices are generated in the two component channels. The lattice topologies are controlled by cooperatively modulating the PB phase 2σθ j and dynamic phase mθ j associated with the rotation of the nanoslits and the location of the apertures on the spiral, respectively. Particularly, the lattice topologies of the primary spin component can be easily regulated by adjusting the spiral pitch, whereas the lattice topologies of the converted spin component can be determined by the rotation of the slits and spiral pitch. Moreover, they may be symmetrically or asymmetrically helicity-dependent. In practice, spatially overlapped lattice fields of the two spin components are separated by a PF. Notably, the primary spin component has been unintentionally used in some previous studies, such as vector beam generation with metasurfaces, where the fields of the primary and converted components are vectorially superposed without being independently manipulated [48]. We intend for the present study to provide more flexible and efficient manipulations on light fields and allow for micro-manipulations, topological photonics, and on-chip quantum optical applications. Particularly, we hope that our design to be generalized to the Moiré metasurfaces [66, 67] for the dynamic spin manipulations through the rotations of metasurface layers, which has become one of the most important fields in recent metasurface applications, and in addition, such generalization for the Moiré patterns of honeycomb lattices might be related to the magic angle effect in graphene which has been paid attention to in topological photonics recently.
Funding source: National Natural Science Foundation of China 10.13039/501100001809
Award Identifier / Grant number: 11904212
Award Identifier / Grant number: 12004215
Award Identifier / Grant number: 62175134
-
Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: This work was supported by the National Natural Science Foundation of China (Grant No. 62175134, 11904212, 12004215).
-
Conflict of interest statement: The authors declare no conflict of interest.
References
[1] K. Y. Bliokh, F. J. Rodríguez-Fortuño, F. Nori, et al.., “Spin–orbit interactions of light,” Nat. Photonics, vol. 9, pp. 796–808, 2015. https://doi.org/10.1038/nphoton.2015.201.Search in Google Scholar
[2] K. Y. Bliokh and Y. P. Bliokh, “Conservation of angular momentum, transverse shift, and spin Hall effect in reflection and refraction of an electromagnetic wave packet,” Phys. Rev. Lett., vol. 96, 2006, Art no. 073903. https://doi.org/10.1103/PhysRevLett.96.073903.Search in Google Scholar PubMed
[3] X. Yin, Z. Ye, J. Rho, et al.., “Photonic spin Hall effect at metasurfaces,” Science, vol. 339, pp. 1405–1407, 2013. https://doi.org/10.1126/science.1231758.Search in Google Scholar PubMed
[4] S. Xiao, J. Wang, F. Liu, et al.., “Spin-dependent optics with metasurfaces,” Nanophotonics, vol. 6, pp. 215–234, 2017. https://doi.org/10.1515/nanoph-2016-0121.Search in Google Scholar
[5] L. Marrucci, C. Manzo, and D. Paparo, “Optical spin-to-orbital angular momentum conversion in inhomogeneous anisotropic media,” Phys. Rev. Lett., vol. 96, 2006, Art no. 163905. https://doi.org/10.1103/physrevlett.96.163905.Search in Google Scholar PubMed
[6] L. Han, S. Liu, P. Li, et al.., “Catalystlike effect of orbital angular momentum on the conversion of transverse to three-dimensional spin states within tightly focused radially polarized beams,” Phys. Rev. A, vol. 97, 2018, Art no. 053802. https://doi.org/10.1103/physreva.97.053802.Search in Google Scholar
[7] A. S. Rana, I. Kim, M. A. Ansari, et al.., “Planar achiral metasurfaces-induced anomalous chiroptical effect of optical spin isolation,” ACS Appl. Mater. Interfaces, vol. 12, pp. 48899–48909, 2020. https://doi.org/10.1021/acsami.0c10006.Search in Google Scholar PubMed
[8] H. S. Khaliq, I. Kim, J. Kim, et al.., “Manifesting simultaneous optical spin conservation and spin isolation in diatomic metasurfaces,” Adv. Opt. Mater., vol. 9, 2021, Art no. 2002002. https://doi.org/10.1002/adom.202002002.Search in Google Scholar
[9] K. Y. Bliokh, D. Smirnova, and F. Nori, “Quantum spin Hall effect of light,” Science, vol. 348, p. 1448, 2015. https://doi.org/10.1126/science.aaa9519.Search in Google Scholar PubMed
[10] S. Sun, Q. He, S. Xiao, et al.., “Gradient-index meta-surfaces as a bridge linking propagating waves and surface waves,” Nat. Mater., vol. 11, pp. 426–431, 2012. https://doi.org/10.1038/nmat3292.Search in Google Scholar PubMed
[11] L. Zhang, A. Datta, and I. A. Walmsley, “Precision metrology using weak measurements,” Phys. Rev. Lett., vol. 114, 2015, Art no. 210801. https://doi.org/10.1103/physrevlett.114.210801.Search in Google Scholar
[12] X.-L. Qi and S.-C. Zhang, “Topological insulators and superconductors,” Rev. Mod. Phys., vol. 83, p. 1057, 2011. https://doi.org/10.1103/revmodphys.83.1057.Search in Google Scholar
[13] L. Marrucci, E. Karimi, S. Slussarenko, et al.., “Spin-to-orbital conversion of the angular momentum of light and its classical and quantum applications,” J. Opt., vol. 13, 2011, Art no. 064001. https://doi.org/10.1088/2040-8978/13/6/064001.Search in Google Scholar
[14] F. Flamini, N. Spagnolo, and F. Sciarrino, “Photonic quantum information processing: a review,” Rep. Prog. Phys., vol. 82, p. 016001, 2018. https://doi.org/10.1088/1361-6633/aad5b2.Search in Google Scholar PubMed
[15] S. Sun, Q. He, J. Hao, et al.., “Electromagnetic metasurfaces: physics and applications,” Adv. Opt Photon, vol. 11, pp. 380–479, 2019. https://doi.org/10.1364/aop.11.000380.Search in Google Scholar
[16] J. Kim, Y. Yang, T. Badloe, et al.., “Geometric and physical configurations of meta-atoms for advanced metasurface holography,” Info. Mat., vol. 3, pp. 739–754, 2021. https://doi.org/10.1002/inf2.12191.Search in Google Scholar
[17] M. L. Tseng, H.-H. Hsiao, C. H. Chu, et al.., “Metalenses: advances and applications,” Adv. Opt. Mater., vol. 6, 2018, Art no. 1800554. https://doi.org/10.1002/adom.201800554.Search in Google Scholar
[18] J.-S. Park, S. Zhang, A. She, et al.., “All-glass, large metalens at visible wavelength using deep-ultraviolet projection lithography,” Nano Lett., vol. 19, pp. 8673–8682, 2019. https://doi.org/10.1021/acs.nanolett.9b03333.Search in Google Scholar PubMed
[19] H. Ren, G. Briere, X. Fang, et al.., “Metasurface orbital angular momentum holography,” Nat. Commun., vol. 10, pp. 1–8, 2019. https://doi.org/10.1038/s41467-019-11030-1.Search in Google Scholar PubMed PubMed Central
[20] X. Ding, Z. Wang, G. Hu, et al.., “Metasurface holographic image projection based on mathematical properties of Fourier transform,” PhotoniX, vol. 1, pp. 1–12, 2020. https://doi.org/10.1186/s43074-020-00016-8.Search in Google Scholar
[21] M. A. Naveed, M. A. Ansari, I. Kim, et al.., “Optical spin-symmetry breaking for high-efficiency directional helicity-multiplexed metaholograms,” Microsyst. Nanoeng., vol. 7, pp. 1–9, 2021. https://doi.org/10.1038/s41378-020-00226-x.Search in Google Scholar PubMed PubMed Central
[22] E. Wang, L. Shi, J. Niu, et al.., “Multichannel spatially nonhomogeneous focused vector vortex beams for quantum experiments,” Adv. Opt. Mater., vol. 7, 2019, Art no. 1801415. https://doi.org/10.1002/adom.201801415.Search in Google Scholar
[23] W. Liu, Z. Li, Z. Li, et al.., “Energy-tailorable spin-selective multifunctional metasurfaces with full fourier components,” Adv. Mater., vol. 31, 2019, Art no. 1901729.10.1002/adma.201901729Search in Google Scholar PubMed
[24] S. Boroviks, R. A. Deshpande, N. A. Mortensen, et al.., “Multifunctional metamirror: polarization splitting and focusing,” ACS Photonics, vol. 5, pp. 1648–1653, 2017. https://doi.org/10.1021/acsphotonics.7b01091.Search in Google Scholar
[25] Z. Shi, A. Y. Zhu, Z. Li, et al.., “Continuous angle-tunable birefringence with freeform metasurfaces for arbitrary polarization conversion,” Sci. Adv., vol. 6, 2020, Art no. eaba3367. https://doi.org/10.1126/sciadv.aba3367.Search in Google Scholar PubMed PubMed Central
[26] D. Wang, F. Liu, and T. Liu, “Efficient generation of complex vectorial optical fields with metasurfaces,” Light Sci. Appl., vol. 10, pp. 1–14, 2021. https://doi.org/10.1038/s41377-021-00504-x.Search in Google Scholar PubMed PubMed Central
[27] M. Kang, T. Feng, H.-T. Wang, et al.., “Wave front engineering from an array of thin aperture antennas,” Opt. Express, vol. 20, pp. 15882–15890, 2012. https://doi.org/10.1364/oe.20.015882.Search in Google Scholar
[28] J. P. B. Mueller, N. A. Rubin, R. C. Devlin, et al.., “Metasurface polarization optics: independent phase control of arbitrary orthogonal states of polarization,” Phys. Rev. Lett., vol. 118, 2017, Art no. 113901. https://doi.org/10.1103/physrevlett.118.113901.Search in Google Scholar
[29] D. Wang, Y. Hwang, Y. Dai, et al.., “Broadband high-efficiency chiral splitters and holograms from dielectric nanoarc metasurfaces,” Small, vol. 15, 2019, Art no. 1900483. https://doi.org/10.1002/smll.201900483.Search in Google Scholar PubMed
[30] Z.-L. Deng, J. Deng, X. Zhuang, et al.., “Diatomic metasurface for vectorial holography,” Nano Lett., vol. 18, pp. 2885–2892, 2018. https://doi.org/10.1021/acs.nanolett.8b00047.Search in Google Scholar PubMed
[31] R. C. Devlin, A. Ambrosio, N. A. Rubin, et al.., “Arbitrary spin-to-orbital angular momentum conversion of light,” Science, vol. 358, pp. 896–901, 2017. https://doi.org/10.1126/science.aao5392.Search in Google Scholar PubMed
[32] N. A. Rubin, G. D’Aversa, P. Chevalier, et al.., “Matrix Fourier optics enables a compact full-Stokes polarization camera,” Science, vol. 365, 2019, Art no. eaax1839. https://doi.org/10.1126/science.aax1839.Search in Google Scholar PubMed
[33] S. Wang, P. C. Wu, V.-C. Su, et al.., “A broadband achromatic metalens in the visible,” Nat. Nanotechnol., vol. 13, pp. 227–232, 2018. https://doi.org/10.1038/s41565-017-0052-4.Search in Google Scholar PubMed
[34] W. T. Chen, A. Y. Zhu, J. Sisler, et al.., “A broadband achromatic polarization-insensitive metalens consisting of anisotropic nanostructures,” Nat. Commun., vol. 10, p. 355, 2019. https://doi.org/10.1038/s41467-019-08305-y.Search in Google Scholar PubMed PubMed Central
[35] I. Bloch, “Ultracold quantum gases in optical lattices,” Nat. Phys., vol. 1, pp. 23–30, 2005. https://doi.org/10.1038/nphys138.Search in Google Scholar
[36] M. Kumar and J. Joseph, “Generating a hexagonal lattice wave field with a gradient basis structure,” Opt. Lett., vol. 39, pp. 2459–2462, 2014. https://doi.org/10.1364/ol.39.002459.Search in Google Scholar PubMed
[37] L. Lu, J. D. Joannopoulos, and M. Soljačić, “Topological photonics,” Nat. Photonics, vol. 8, pp. 821–829, 2014. https://doi.org/10.1038/nphoton.2014.248.Search in Google Scholar
[38] M. Kim, Z. Jacob, and J. Rho, “Recent advances in 2D, 3D and higher-order topological photonics,” Light Sci. Appl., vol. 9, pp. 1–30, 2020. https://doi.org/10.1038/s41377-020-0331-y.Search in Google Scholar PubMed PubMed Central
[39] C. Monroe, “Quantum information processing with atoms and photons,” Nature, vol. 416, pp. 238–246, 2002. https://doi.org/10.1038/416238a.Search in Google Scholar PubMed
[40] M. P. MacDonald, G. C. Spalding, and K. Dholakia, “Microfluidic sorting in an optical lattice,” Nature, vol. 426, pp. 421–424, 2003. https://doi.org/10.1038/nature02144.Search in Google Scholar PubMed
[41] L.-H. Wu and X. Hu, “Scheme for achieving a topological photonic crystal by using dielectric material,” Phys. Rev. Lett., vol. 114, p. 223901, 2015. https://doi.org/10.1103/physrevlett.114.223901.Search in Google Scholar PubMed
[42] T. Jiang, M. Xiao, W.-J. Chen, et al.., “Experimental demonstration of angular momentum-dependent topological transport using a transmission line network,” Nat. Commun., vol. 10, p. 434, 2019. https://doi.org/10.1038/s41467-018-08281-9.Search in Google Scholar PubMed PubMed Central
[43] J.-W. Dong, X.-D. Chen, H. Zhu, et al.., “Valley photonic crystals for control of spin and topology,” Nat. Mater., vol. 16, pp. 298–302, 2017. https://doi.org/10.1038/nmat4807.Search in Google Scholar PubMed
[44] X. Zhu, H.-X. Wang, C. Xu, et al.., “Topological transitions in continuously deformed photonic crystals,” Phys. Rev. B, vol. 97, p. 085148, 2018. https://doi.org/10.1103/physrevb.97.085148.Search in Google Scholar
[45] T. Ozawa, H. M. Price, A. Amo, et al.., “Topological photonics,” Rev. Mod. Phys., vol. 91, p. 015006, 2019. https://doi.org/10.1103/revmodphys.91.015006.Search in Google Scholar
[46] H. Gao, Y. Li, L. Chen, et al.., “Quasi-Talbot effect of orbital angular momentum beams for generation of optical vortex arrays by multiplexing metasurface design,” Nanoscale, vol. 10, pp. 666–671, 2018. https://doi.org/10.1039/c7nr07873k.Search in Google Scholar PubMed
[47] Z. Li, H. Liu, X. Zhang, et al.., “Metasurface of deflection prism phases for generating non-diffracting optical vortex lattices,” Opt. Express, vol. 26, pp. 28228–28237, 2018. https://doi.org/10.1364/oe.26.028228.Search in Google Scholar PubMed
[48] R. Zhang, Y. Zhang, L. Ma, et al.., “Nanoscale optical lattices of arbitrary orders manipulated by plasmonic metasurfaces combining geometrical and dynamic phases,” Nanoscale, vol. 11, pp. 14024–14031, 2019. https://doi.org/10.1039/c9nr03381e.Search in Google Scholar PubMed
[49] S. Tsesses, E. Ostrovsky, K. Cohen, et al.., “Optical skyrmion lattice in evanescent electromagnetic fields,” Science, vol. 361, pp. 993–996, 2018. https://doi.org/10.1126/science.aau0227.Search in Google Scholar PubMed
[50] S. Tsesses, K. Cohen, E. Ostrovsky, et al.., “Spin-orbit interaction of light in plasmonic lattices,” Nano Lett., vol. 19, pp. 4010–4016, 2019. https://doi.org/10.1021/acs.nanolett.9b01343.Search in Google Scholar PubMed
[51] S. Chen, Y. Zhang, Z. Li, et al.., “Empowered layer effects and prominent properties in few-layer metasurfaces,” Adv, Opt. Mater., vol. 7, p. 1801477, 2019. https://doi.org/10.1002/adom.201801477.Search in Google Scholar
[52] S. Barik, A. Karasahin, C. Flower, et al.., “A topological quantum optics interface,” Science, vol. 359, pp. 666–668, 2018. https://doi.org/10.1126/science.aaq0327.Search in Google Scholar PubMed
[53] M. A. Gorlach, X. Ni, D. A. Smirnova, et al.., “Far-field probing of leaky topological states in all-dielectric metasurfaces,” Nat. Commun., vol. 9, no. 1, pp. 1–8, 2018. https://doi.org/10.1038/s41467-018-03330-9.Search in Google Scholar PubMed PubMed Central
[54] P. K. Jha, X. Ni, C. Wu, et al.., “Metasurface-enabled remote quantum interference,” Phys. Rev. Lett., vol. 115, p. 025501, 2015. https://doi.org/10.1103/PhysRevLett.115.025501.Search in Google Scholar PubMed
[55] K. Wang, J. G. Titchener, S. S. Kruk, et al.., “Quantum metasurface for multiphoton interference and state reconstruction,” Science, vol. 361, no. 6407, pp. 1104–1108, 2018. https://doi.org/10.1126/science.aat8196.Search in Google Scholar PubMed
[56] Y. Q. Zhang, X. Y. Zeng, L. Ma, et al.., “Manipulation for superposition of orbital angular momentum states in surface plasmon polaritons,” Adv. Opt. Mater., vol. 7, p. 1900372, 2019. https://doi.org/10.1002/adom.201900372.Search in Google Scholar
[57] F. Zhang, M. Pu, X. Li, et al.., “All-dielectric metasurfaces for simultaneous giant circular asymmetric transmission and wavefront shaping based on asymmetric photonic spin-orbit interactions,” Adv. Funct. Mater., vol. 27, p. 1704295, 2017. https://doi.org/10.1002/adfm.201704295.Search in Google Scholar
[58] G. Milione, S. Evans, D. A. Nolan, et al.., “Higher order Pancharatnam-Berry phase and the angular momentum of light,” Phys. Rev. Lett., vol. 108, no. 19, p. 190401, 2012. https://doi.org/10.1103/physrevlett.108.190401.Search in Google Scholar PubMed
[59] L. Sun, B. Bai, J. Wang, et al.., “Probing the photonic spin–orbit interactions in the near field of nanostructures,” Adv. Funct. Mater., vol. 29, no. 32, 2019, Art no. 1902286. https://doi.org/10.1002/adfm.201902286.Search in Google Scholar
[60] X. G. Luo, “Subwavelength Artificial structures: opening a new era for engineering optics,” Adv. Mater., vol. 31, 2019, Art no. 1804680. https://doi.org/10.1002/adma.201804680.Search in Google Scholar PubMed
[61] G. Gbur and T. D. Visser, “Coherence vortices in partially coherent beams,” Opt. Commun., vol. 222, pp. 117–125, 2003. https://doi.org/10.1016/s0030-4018(03)01606-7.Search in Google Scholar
[62] G. J. Gbur, Singular Optics, Boca Raton, FL, CRC Press, 2016.10.1201/9781315374260Search in Google Scholar
[63] S. Xiao, F. Zhong, H. Liu, et al.., “Flexible coherent control of plasmonic spin-Hall effect,” Nat. Commun., vol. 6, p. 8360, 2015. https://doi.org/10.1038/ncomms9360.Search in Google Scholar PubMed PubMed Central
[64] Z. Wang, S. Q. Li, X. Q. Zhang, et al.., “Excite spoof surface plasmons with tailored wavefronts using high-efficiency terahertz metasurfaces,” Adv. Sci., vol. 7, 2020, Art no. 2000982. https://doi.org/10.1002/advs.202000982.Search in Google Scholar PubMed PubMed Central
[65] R. Fu, L. Deng, Z. Guan, et al.., “Zero-order-free meta-holograms in a broadband visible range,” Photon. Res., vol. 8, p. 723, 2020. https://doi.org/10.1364/prj.387397.Search in Google Scholar
[66] Y. Luo, C. H. Chu, S. Vyas, et al.., “Varifocal metalens for optical sectioning fluorescence microscopy,” Nano Lett., vol. 21, pp. 5133–5142, 2021. https://doi.org/10.1021/acs.nanolett.1c01114.Search in Google Scholar PubMed
[67] X. D. Cai, R. Tang, H. Y. Zhou, et al.., “Dynamically controlling terahertz wavefronts with cascaded metasurfaces,” Adv. Photonics, vol. 3, p. 036003, 2021. https://doi.org/10.1117/1.ap.3.3.036003.Search in Google Scholar
© 2021 Ruirui Zhang et al., published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- Reviews
- Super-resolution imaging: when biophysics meets nanophotonics
- Structuring and functionalization of non-metallic materials using direct laser interference patterning: a review
- Review on fractional vortex beam
- Research Articles
- Tunable Faraday rotation of ferromagnet thin film in whole visible region coupled with aluminum plasmonic arrays
- On-chip nanophotonic broadband wavelength detector with 2D-Electron gas
- Quasi-BIC laser enabled by high-contrast grating resonator for gas detection
- The effects of bending on plasmonic modes in nanowires and planar structures
- A label-free optical system with a nanohole array biosensor for discriminating live single cancer cells from normal cells
- Scalable and compact photonic neural chip with low learning-capability-loss
- A learning based approach for designing extended unit cell metagratings
- Key role of surface plasmon polaritons in generation of periodic surface structures following single-pulse laser irradiation of a gold step edge
- Invisibility concentrator based on van der Waals semiconductor α-MoO3
- High-fidelity nano-FTIR spectroscopy by on-pixel normalization of signal harmonics
- Plasmonic metasurfaces manipulating the two spin components from spin–orbit interactions of light with lattice field generations
- Flat telescope based on an all-dielectric metasurface doublet enabling polarization-controllable enhanced beam steering
Articles in the same Issue
- Frontmatter
- Reviews
- Super-resolution imaging: when biophysics meets nanophotonics
- Structuring and functionalization of non-metallic materials using direct laser interference patterning: a review
- Review on fractional vortex beam
- Research Articles
- Tunable Faraday rotation of ferromagnet thin film in whole visible region coupled with aluminum plasmonic arrays
- On-chip nanophotonic broadband wavelength detector with 2D-Electron gas
- Quasi-BIC laser enabled by high-contrast grating resonator for gas detection
- The effects of bending on plasmonic modes in nanowires and planar structures
- A label-free optical system with a nanohole array biosensor for discriminating live single cancer cells from normal cells
- Scalable and compact photonic neural chip with low learning-capability-loss
- A learning based approach for designing extended unit cell metagratings
- Key role of surface plasmon polaritons in generation of periodic surface structures following single-pulse laser irradiation of a gold step edge
- Invisibility concentrator based on van der Waals semiconductor α-MoO3
- High-fidelity nano-FTIR spectroscopy by on-pixel normalization of signal harmonics
- Plasmonic metasurfaces manipulating the two spin components from spin–orbit interactions of light with lattice field generations
- Flat telescope based on an all-dielectric metasurface doublet enabling polarization-controllable enhanced beam steering