Open Access
18 August 2022 Classical and generalized geometric phase in electromagnetic metasurfaces
Yinghui Guo, Mingbo Pu, Fei Zhang, Mingfeng Xu, Xiong Li, Xiaoliang Ma, Xiangang Luo
Author Affiliations +
Abstract

The geometric phase concept has profound implications in many branches of physics, from condensed matter physics to quantum systems. Although geometric phase has a long research history, novel theories, devices, and applications are constantly emerging with developments going down to the subwavelength scale. Specifically, as one of the main approaches to implement gradient phase modulation along a thin interface, geometric phase metasurfaces composed of spatially rotated subwavelength artificial structures have been utilized to construct various thin and planar meta-devices. In this paper, we first give a simple overview of the development of geometric phase in optics. Then, we focus on recent advances in continuously shaped geometric phase metasurfaces, geometric–dynamic composite phase metasurfaces, and nonlinear and high-order linear Pancharatnam–Berry phase metasurfaces. Finally, conclusions and outlooks for future developments are presented.

1.

Introduction

In optics and electromagnetics, the geometric phase originates from the spin–orbit interaction (SOI) of light and describes the relationship between phase change and polarization conversion when light is transmitted or reflected through an anisotropic medium[1]. The general form of the geometric phase was developed by Berry in 1984[2]. He found that when a quantum system in an eigenstate is slowly transported around a circuit C by varying parameters R in its Hamiltonian H(R), it will acquire a geometrical phase factor exp(iγC). Since the pioneering work of Berry, the geometric phase has been applied in various fields of physics and expanded the understanding of state evolutions in different parameter spaces. The most common formulations of the geometric phase are known as the Aharonov–Bohm (AB) phase for electrons and the Pancharatnam–Berry (PB) phase for photons. Figure 1(a) shows a representative case of the AB effect. When a charged particle passes around a long solenoid, the wave function experiences a phase shift as a result of the enclosed magnetic field, despite the magnetic field being negligible in the region through which the particle passes and the particle’s wave function being negligible inside the solenoid[3]. Subsequently, Chiao et al. considered the manifestations of this phase factor for a photon in a state of adiabatically invariant helicity, and an effective optical activity for a helical optical fiber was predicted[4]. In 1987, Aharonov and Anandan noted that the appearance of the geometric phase did not necessarily go through an adiabatic process, and the geometric phase factor can be defined for any cyclic evolution of a quantum system[5]. Another important manifestation of the geometric phase in solids is known as the Zak phase[6], which underlies the existence of protected edge states.

Fig. 1

(a) Schematic illustration of AB effect, where an electron encircles a magnetic flux Φ confined to a thin, long solenoid. Although the magnetic field is zero in the vicinity of the superposed wave packets, the vector potential is non-zero outside the solenoid. Thus, the electronic wave packets acquire relative phases of exp(ieΦ/ħ), causing their interference patterns to change. (b) Polarization state representation on Poincaré sphere. (c) PB phase on Poincaré sphere, which equals half the solid angle (Ω) subtended at the origin by the area enclosed by the closed curves (C1 and C2) on the Poincaré sphere.

PI_1_1_R03_f001.png

In 1956, Pancharatnam noticed a phase shift arising when the polarization of a photonic beam was varied in a cyclic manner, when he studied the interference between optical waves of different polarizations[7]. As illustrated in Fig. 1(b), the polarization state of the wave and its evolution can be described as a point and its trajectory on the surface of a Poincaré sphere. For the cyclic evolution of polarization displayed in Fig. 1(c), the trajectory of polarization on the Poincaré sphere is a closed curve. In this case, Pancharatnam’s phase equals half the solid angle (Ω) subtended at the origin by the area enclosed by the closed curves (C1 and C2) on the Poincaré sphere. The 1/2 factor corresponds to the well-known “half-speed” rotations of spinors seen in experiments with neutrons, since it is an SU(2) transformation[8,9]. Although Pancharatnam’s phase was proposed before the publication of Berry’s paper, it was not noted for many years. Subsequently, Berry demonstrated the equivalence between Pancharatnam’s geometric phase in polarization and the broader phenomenon of the acquired phase in the adiabatic evolution of a state in quantum mechanics[10]. Therefore, such phases are always referred to as PB phases. During the past few years, the PB phase was further extended to nonlinear PB phase and high-order linear PB phase, which will be discussed in the following sections. A simple roadmap of the development of the geometric phase is displayed in Fig. 2. Note that geometric phase is exploited to provide an additional π phase shift, excepting the circuit resonance induced phase in Ref. [11]. Several excellent reviews on geometric phases are helpful to grasp this research field from different perspectives[12-16].

Fig. 2

Roadmap of the development of the geometric phase in optics.

PI_1_1_R03_f002.png

2.

Geometric Phases in Spatially Rotated Anisotropic Elements

In an earlier experiment, the polarization of a uniformly polarized beam was altered by a series of space-invariant (transversely homogeneous) wave plates and polarizers, and the phase difference was introduced through the evolution of the beam in the time domain. For example, as early as 1947, metallic waveguides featuring birefringent propagation with rotatable half-wave plates were sandwiched between two quarter-wave plates to generate adjustable phase shifts in phased antenna arrays for radar applications[17], as indicated in Fig. 3(a). A similar approach was independently proposed by Pancharatnam in 1955 to construct an achromatic elliptic polarizer[18]. It was not until 2001 that Hasman et al. considered the PB phase in the space domain[19] by using space-variant (transversely inhomogeneous) metal stripe subwavelength gratings. They demonstrated that the conversion of circular polarization was accomplished by a space-variant phase shift, which can be illustrated in a Jones matrix[20]. Specifically, a wave plate with a space-variant fast axis can be described by a coordinate dependent matrix:

Eq. (1)

T(r,ϕ)=cos(δ2)(1001)+isinδ2(cos(2α)sin(2α)sin(2α)cos(2α)),
where δ is the retardation phase of the wave plate, and α is the angle formed between the fast axis and the x axis. For circularly polarized incidence Ein(r,ϕ)=E0(r,ϕ)·(1,iδ)T, the output beam can be expressed as

Eq. (2)

Eout(r,ϕ)=T(r,ϕ)Ein(r,ϕ)=E0cos(δ2)(1iσ)+iE0sin(δ2)exp(i2σα)(1iσ).

Fig. 3

Different elements for generating PB phase. (a) An adjustable waveguide phase shift at the microwave band is composed of a rotatable half-wave plate sandwiched between two quarter-wave plates. Reproduced with permission from Ref. [17]. (b) Actual quantized phase as a function of the desired phase, as well as discrete local grating orientation, with a scanning electron microscope (SEM) image of space-variant gratings displayed in the inset. Reproduced with permission from Ref. [25]. (c) Schematic representation of geometric phase metasurface for generating OAMs via PB phase. Reproduced with permission from Ref. [27]. (d) SEM image of geometric phase metasurface composed of space-variant plasmonic nano-apertures. Scale bar represents 500 nm. Reproduced with permission from Ref. [29].

PI_1_1_R03_f003.png

Equation (2) shows that after passing the space-variant anisotropic wave plates, a circularly polarized beam is primarily scattered into waves of the same polarization as that of the incident beam without a phase change and waves of the opposite circular polarization with a spin-dependent phase change Φ=2σα. Here, σ=±1 represents left- and right-handed circular polarization (LCP and RCP), respectively. Obviously, by controlling the local orientation of the fast axes of the wave plate elements between zero and π, one can realize phase variation that covers the full zero to 2π range while maintaining equal transmission amplitude. Note that the amplitudes of opposite circular polarization depend on the phase retardation δ [i.e., sin(δ/2)], and a wave plate with phase retardation of π is desired to realize 100% conversion efficiency. For artificial wave plates based on subwavelength structures, the operation band is limited due to resonant dispersion. Therefore, dispersion engineering technology has been proposed to construct broadband wave plates[21,22].

Subsequently, a series of PB phase optical elements, composed of space-variant large pixel-sized periodic gratings or annular grooves [Fig. 3(b)], was developed for beam deflection, helical beam generation, and coronagraphs[20,23-25]. As shown in Figs. 3(c) and 3(d), with the development of advanced fabrication technology, e.g., laser direct writing, electron beam lithography, and focused ion beams, the building blocks of space-variant PB phase optical elements were generally dominated by subwavelength plasmonic[26-31] or dielectric elements[32-37], wherein the corresponding metasurfaces were referred to as geometric phase metasurfaces. Owing to the abrupt phase that changes over the scale of the wavelength, the propagation behavior of light can be manipulated arbitrarily. The metasurface-assisted law of refraction and reflection (MLRR)[38], a kind of generalized law of reflection and refraction[11], and broadband sub-diffraction focusing and imaging[26,39] have been demonstrated. Metasurfaces find widespread applications in planar lenses[40-42], the photonic spin Hall effect (PSHE)[43,44], directional coupling[45], holographic imaging[46-53], beam generation[30,54-56], and beam steering antennas[57]. For example, by matching the geometric phases of nano-slots on silver to specific superimpositions of the inward and outward surface plasmon profiles for the two spins, arbitrary spin-dependent orbitals can be generated in a slot-free region[58]. In addition to the traditional design approach, heuristic design methodology has been adopted in nanophotonics to improve the design efficiency and device performances[59].

3.

From Discrete Geometric Phase Metasurface to Continuous Counterparts

Geometric phase metasurfaces are generally composed of multiple discrete meta-atoms, which inevitably degrade the overall performance of metasurfaces, for example, decreasing the diffraction efficiency of metasurfaces and the purity of generated orbit angular momentum (OAM)[60]. During the past few years, quasi-continuous geometric phase metasurfaces composed of catenaries[30], concentric rings[60], sinusoidal metallic strip arrays[61], free-shaped meander lines[62,63], and other shaped structures[32,64-68] have been utilized to overcome the shortfalls above, as indicated in Fig. 4. All these proposals have demonstrated that continuous metasurfaces possess higher efficiency or broader bandwidths than their discrete counterparts owing to the elimination of resonance. Interestingly, the continuous and linear phase shift covering [0, 2π] can be imparted by a single optical catenary[30,43], i.e., the so-called “catenary of equal phase gradient.” Consequently, perfect OAM and even high-order non-diffraction Bessel beams can be generated by arranging the catenaries in various forms, as shown in Fig. 4(a). As a branch of subwavelength optics, catenary optics include not only catenary shaped quasi-continuous structures, but also catenary fields and catenary dispersion. The former is generated by the coupling of evanescent waves, with typical applications in super-resolution imaging and lithography. The latter provides a general guideline for realizing the quick design of functional metasurfaces, such as broadband electromagnetic absorbers and flat antennas. Readers may refer to Ref. [69] for a comprehensive review.

Fig. 4

Quasi-continuous geometric phase. (a) SEM figure of the rotational optical catenary array and measured intensity patterns of OAMs generated through PB phase. Reproduced with permission from Ref. [30]. (b) Optical layout of wide-FOV camera with a single layer quadratic phase metalens. Optical and SEM images of fabricated wide-angle metalens composed of catenary-like structures. Reproduced with permission from Ref. [71]. (c) SEM image of cubic phase metasurface composed of catenary-like structures and measured results of the generated accelerating beams. Reproduced with permission from Ref. [73]. (d) SEM image of holographic metasurface composed of C-shaped dielectric nanoarcs and corresponding holographic images at different wavelengths. Reproduced with permission from Ref. [64].

PI_1_1_R03_f004.png

Subsequently, it has been demonstrated that the metamorphosis of an optical catenary, either metallic or dielectric, can be utilized to generate a quadratic phase profile for wide-field-of-view (FOV) imaging[70,71] and beam steering[57] as well as a cubic phase profile for accelerating beam generation[72,73], as displayed in Figs. 4(b) and 4(c). According to the principle of geometric phase, a novel metasurface design methodology by integrating the predefined phase function of space-variant stripes can be extended for arbitrary phase profiles[62,71]. To cancel the propagation phase modulation, the geometries of dielectric catenaries are tailored elaborately. Subsequently, Wang et al. proposed dielectric nanoarc structures as building blocks that support different electromagnetic resonant modes at localized areas. These dielectric nanoarcs have the distinct advantage of a continuous-phase gradient along the curved trajectory within an individual building block[64]. By combining with the principle of a chiral-detour phase, complex holograms with high efficiency were achieved.

4.

Merging PB Phase and Propagation Phase for Spin-Decoupling Modulation

For a common geometric phase metasurface, geometric phase is generated accompanied by spin conversion and phase modulation conjugating with each other when the spin state of incidence is reversed, which is inconvenient in practical applications. For example, for a focusing metalens composed of a geometric phase metasurface, it can operate only under a specific spin state and will become divergent for other spin states[74]. In addition, holographic images will be centrosymmetric with each other under different spin states[50], which means no additional information is imparted.

To break the conjugate symmetry of a geometric phase metasurface under opposite helicity illumination, Guo et al. proposed to merge it with a polarization-independent propagation phase into a single metasurface[60]. Due to the spin independence of the dynamical propagation phase in the surface plasmon polariton (SPP) wave, the total phases for two spin states are decoupled. The composite phase modulation method has been implemented by dielectric nanopillars with spatially variant orientations and geometric parameters. Two research teams independently demonstrated the spin-decoupled holographic imaging and OAM generation in 2017[75-78], and the concept of spin-decoupled metasurfaces has been further expanded from circular polarization to arbitrary orthogonal polarizations.

Besides phase decoupling modulation, amplitude decoupling modulation[76,79,80] has also been demonstrated via two or multiple meta-atom metasurfaces. By adjusting the orientations and geometric parameters of meta-atoms in the supercell, the LCP and RCP components satisfy different interference conditions. As a consequence, one can make one component be transmitted and the other be reflected. By rotating the super unit cell following the PB phase principle, different functionalities, e.g., deflection, OAM generation, and holographic imaging, have been demonstrated[76]. Following a similar methodology, decoupling amplitude modulation has been expanded to orthogonal linear polarizations and arbitrary orthogonal polarizations[80]. Subsequently, more degrees of freedom (DOFs) are introduced to realize more independent channels. For example, a chirality-assisted phase can decouple two co-polarized outputs, and thus be an alternative solution for designing arbitrary modulated-phase metasurfaces with distinct wavefront manipulation in all four output channels[81]. Recently, by combining four nanoblocks in one pixel, optical elements with five DOFs and printing-hologram functionalities with six DOFs of the Jones matrix have been experimentally demonstrated[82].

Since two orthogonal spin states can be independently manipulated with different phase shifts, vectorial optical beams can be generated more conveniently and compactly[83-87]. Liu et al. have proposed and experimentally demonstrated the broadband generation of perfect Poincaré beams via a single-layer dielectric metasurface for visible light[88]. With the spatial superposition of two perfect optical vortices, different perfect Poincaré beams, whose total angular momenta are described by the hybrid-order Poincaré sphere, can be generated. More recently, a longitudinally varied vector field was experimentally demonstrated by a monolayer metasurface[86], as shown in Figs. 5(a) and 5(b), which are distinct from traditional elements that manipulate polarization in a single transverse plane. The underlying mechanism relies on transforming an incident waveform into an ensemble of pencil-like beams with different polarization states that beat along the optical axis, thereby changing the resulting polarization at will, locally, as light propagates. Subsequently, a simultaneous transversally and longitudinally varied vector optical field (VOF) has been further demonstrated by decoupling two orthogonal polarization states through asymmetric photonic SOIs to obtain custom-tailored phase and amplitude differences between opposite components and then coherently synthesize them in three-dimensional space. As illustrated in Fig. 5, the polarization distribution switches continuously and periodically between radial and azimuthal polarizations[89]. Furthermore, the vector field can be dynamically tuned by rotating the incident polarization state. This work extends polarization optics from two-dimensional space to 3D space, allowing the arbitrary generation and manipulation of 3D VOFs with temporal tunability.

Fig. 5

(a) Schematic of a z-dependent polarizing device (polarizer or retarder) that enables variable polarization operations to be performed at different z planes along the optical path. (b) Measured transverse profiles of the output beam at different planes along the propagation direction for x and y incident polarizations. White arrows depict incident polarization (in the xy plane), and red arrows represent the state of polarization analyzed at each z plane. Reproduced with permission from Ref. [86]. Three-dimensional synthetic VOF generated by a monolayer composite phase metasurface: (c) schematic diagram; (d) measured intensity pattern of orthogonal linear polarizations. Black arrows depict the virtual principal axis orientation of the polarizing element at each z plane. Reproduced with permission from Ref. [89].

PI_1_1_R03_f005.png

Considering the principle of reciprocity, VOF generators may also be used to characterize and detect VOFs through spin–orbit coupling effects[90,91]. It has been demonstrated that a single angular metalens with an azimuthal-quadratic phase profile can simultaneously determine the phase and polarization singularities of a cylindrical vortex vector beam (CVVB) via spin-dependent photonic momentum transformation[92]. As shown in Fig. 6, CVVBs with phase and polarization singularities can be treated as a superimposition of an LCP vortex with topological charge l+m and an RCP vortex with topological charge lm, which are separated into top and bottom half-screens, with their azimuthal angles proportional to the topological charge. Luo’s group has proposed and investigated the concept of asymmetric photonic SOI since 2017, and interesting applications can be found in recent reviews[93,94].

Fig. 6

Single spin-decoupled metasurface for vectorial optical field characterization. (a) Schematic diagram. (b) Measured intensity pattern of CVVB with different intensity patterns. CVVB, cylindrical vortex vector beam. Reproduced with permission from Ref. [92].

PI_1_1_R03_f006.png

5.

PB Phase in the Nonlinear Regime

In 2015, the concept of PB phase optical elements was extended to the nonlinear regime. Three research groups independently reported the nonlinear PB phase in metasurfaces nearly at the same time[95-97]. It was found that under the illumination of a circularly polarized fundamental wave, transmitted nonlinear harmonic waves with the same or opposite helicity will carry phase factors of (n1)σϕ or (n+1)σϕ, respectively, where n is the order of harmonic generation[96]. To satisfy the selection rules for the harmonic generation of circularly polarized fundamental waves, a single nanostructure with m-fold rotational symmetry allows only harmonic orders of n=lm±1, where l is an integer, and the “+” and “−” signs correspond to harmonic generation of the same and opposite circular polarization, respectively. For example, for a nanorod structure with two-fold rotational symmetry (C2), third-harmonic generation (THG) signals with both the same and opposite circular polarizations to that of the fundamental wave can be generated, and the corresponding spin-dependent phases of 2σθ and 4σθ are generated. On the contrary, the THG process is not allowed for a nanostructure with four-fold rotational symmetry (C4) for the same polarization state as the incident polarization, as revealed by Fig. 7(a). Hence, only a single THG signal with a geometric phase of 4σθ for opposite circular polarization is generated. Furthermore, the impact of rotational symmetry on THG for circularly polarized light in nonlinear plasmonic crystals has also been investigated, which may inspire the design of novel spin-dependent nonlinear plasmonic devices[98,99]. Such a nonlinear geometric phase was further confirmed through THG signal deflection via a gradient metasurface, as shown in Figs. 7(b) and 7(c). A similar phenomenon was reported in Ref. [95] via multiquantum-well-based nonlinear plasmonic metasurfaces, as indicated in Figs. 7(d) and 7(e). Readers may refer to a recent review for comprehensive details[100]. Subsequently, Li et al. demonstrated the existence of a nonlinear geometric Berry phase in the four-wave mixing process, which was applied to spin-controlled nonlinear light generation from plasmonic metasurfaces[101]. The polarization state of four-wave mixing from ultrathin metasurfaces, comprising gold meta-atoms with four-fold rotational symmetry, can be controlled by manipulating the spin of the excitation beams. The mutual orientation of the meta-atoms in the metasurface influences the intensity of four-wave mixing via the geometric phase effects.

Fig. 7

Nonlinear geometric phase metasurface. (a) Nonlinear geometric phases are generated through THG of a C2 nanostructure and a C4 nanostructure. (b) Illustration of phase-controlled diffraction of THG signals for RCP light at the fundamental frequency. (c) Measured diffraction pattern of THG signals from C2 and C4 metasurfaces for circular polarization states of the fundamental and THG waves. (a)–(c) Reproduced with permission from Ref. [96]. (d) Sketch of the proposed PB nonlinear metasurface with a phase gradient in x direction. The MQW blocks are sandwiched between U-shaped gold resonators and a metallic ground plane. The incident circularly polarized wave at frequency ω generates simultaneously RCP and LCP nonlinear waves at 2ω. (e) Simulated field distributions and deflection directions of second-harmonic generation (SHG) signals of different circular polarization states. (d), (e) Reproduced with permission from Ref. [95].

PI_1_1_R03_f007.png

By exploiting the concept of the PB phase in the nonlinear optical regime, many interesting functional devices such as nonlinear lenses[102-104], nonlinear encryption[105], and nonlinear imaging[106,107], have been proposed. The multidimensional nonlinear polarization response of nano-material was achieved in a single heterodyne measurement by active manipulation of the polarization states of incident light[108], and a new kind of metasurface-based THz emitter with the polarization and phase of THz waves finely controlled was demonstrated[109]. By exploiting linear and nonlinear geometric phases simultaneously as different channels, spin and wavelength multiplexed nonlinear metasurface holography provides independent, nondispersive, and crosstalk-free post-selective channels for holographic multiplexing and multidimensional optical data storage, anti-counterfeiting, and optical encryption[110].

6.

High-Order Linear PB Phase

In principle, all previous methods to generate a linear PB phase rely on the polarization conversion enabled by anisotropic structures, and subwavelength structures with 3 rotational symmetries are thought to be isotropic and no linear geometric phase can be observed[100,111]. In 2021, Xie et al. demonstrated that subwavelength structures with rotational symmetries 3 can tailor optical anisotropy, and high-order linear geometric phases that manifest as multiple times the rotation angle of the meta-atoms can be achieved[112]. The anisotropy is attributed to the lattice coupling effect when the symmetries of the meta-atom and lattice are incompatible. The high-order geometric phase can be understood by analyzing the rotation dependence of meta-atoms and their effective principal axes. As shown in Fig. 8, for meta-atoms with one-fold (C1) or two-fold (C2) rotational symmetry, the principal axis has a rotation angle equal to the single meta-atom. For meta-atoms with three and higher rotational symmetry, however, there is a multiple and lattice-dependent relationship between the rotation angles of the principal axes and meta-atoms. For instance, for triangular meta-atoms in a square lattice, owing to their three-fold rotational symmetry, the principal axis has a rotation angle of 3ϕ when the meta-atoms are rotated by ϕ. For square meta-atoms arranged in a hexagonal lattice, when the meta-atoms are rotated by ϕ, the principal axis has a rotation angle of 10ϕ. By taking advantage of the multifold relationship, high-order geometric phases equal to multiple times the rotation angle of meta-atoms can be obtained.

Fig. 8

High-order linear PB phase metasurfaces. (a) Rotation dependence of the principal axis and meta-atoms with C1, C2, C3, C6, C5, and C10 rotational symmetry in the square lattice. (b) Rotation dependence of the principal axis and meta-atoms with C1, C2, C4, C8, C5, and C10 rotational symmetry in the hexagonal lattice. Double-headed arrows indicate the orientations of the principal array axis.

PI_1_1_R03_f008.png

After rigorous mathematical derivation, for a square lattice, the geometric phase for meta-atoms with n-fold rotational symmetry can be written as

Eq. (3)

Φ={±2nϕ,nis odd±nϕ,nis even.

For a hexagonal lattice, the geometric phase can be deduced with a bit more complex expression:

Eq. (4)

Φ=2×60°γϕ,
or

Eq. (5)

Φ=2×120°γϕ,
where γ denotes the minimal rotation angle of the meta-atoms to rotate the principal axis by 60° or 120°. The choice of expression should be determined by investigating either the geometric phase or the transmission coefficients along the principal axes at different rotational angles.

The nontrivial phase shifts introduced by meta-atoms in the square or hexagonal lattice are summarized in Table 1. According to these relationships, one can obtain a well-defined geometric phase with the desired gradient in the range of 0 to 2π by choosing the appropriate rotational symmetry of meta-atoms and lattices.

Table 1

Geometric Phases Introduced by Meta-Atoms with C1–C10 Rotational Symmetries in the Square or Hexagonal Lattice under LCP and RCP Illuminations.

LatticePolarizationC1C2C3C4C5C6C7C8C9C10
SquareLCP2ϕ2ϕ6ϕ10ϕ6ϕ14ϕ18ϕ10ϕ
RCP2ϕ2ϕ6ϕ10ϕ6ϕ14ϕ18ϕ10ϕ
HexagonalLCP2ϕ2ϕ8ϕ20ϕ14ϕ8ϕ10ϕ
RCP2ϕ2ϕ8ϕ20ϕ14ϕ8ϕ10ϕ

7.

Conclusion and Outlook

In summary, we have briefly reviewed the history of geometric phase and recent advances in geometric phase metasurfaces. Compared to discrete geometric phase metasurfaces, quasi-continuous and continuous geometric phase metasurfaces exhibit great advantages such as higher efficiency and broader bandwidth. Merging geometric phase and propagation phase can break the conjugate symmetry of traditional PB phase and bring new degrees of freedom for light manipulation. For example, one can simultaneously manipulate the phase and dispersion of a transmitted light wave to realize multiple wavelength or achromatic applications[113]. As two kinds of emerging geometric phases, nonlinear PB phase and high-order linear PB phase have been discussed in detail, providing us with a new understanding of the geometric phase as well as light–matter interaction in nanophotonics. Although this paper focuses mainly on the geometric phase metasurface, the geometric phase principle is extended to liquid crystal planar devices by changing the tangential curves of the molecular orientation pattern[114], owing to the high polarization conversion efficiency and easy fabrication process. For a focusing lens, the tangential curves possess a catenary profile.

In the future, there are some interesting research directions that can be explored, for example, the combination of nonlinear geometric phase and lattice-mismatch induced high-order geometric phase, dynamic and reconfigurable modulation based on high-order geometric phase, and merging high-order PB phase with dynamic phase in a single metasurface. The idea of high-order linear PB phase can be applied to nonlinear optics and will promote new findings about nonlinear geometric phase[115], and provides new approaches for manipulations of light in nonlinear metasurfaces. Furthermore, multiple-dimensional light manipulation including polarization amplitude, polarization phase, complex amplitude, phase-dispersion engineering, and VOF[116-121] can be realized. Recently, progress in geometric phase inside a light beam at a sharp interface has provided very different geometric-phase-enabled novel phenomena distinct from those reviewed in this paper[122,123]. For example, a wave-vector-varying PB phase has been found that arises naturally in the transmission and reflection processes in homogeneous media for paraxial beams with small incident angles[124]. The eigenpolarization states of transmission and reflection processes are determined by the local wave vectors of the incident beam. A small incident angle breaks the rotational symmetry and induces a PB phase that varies linearly with the transverse wave vector, resulting in the PSHE. Additionally, an intriguing phase transition between vortex generation and spin Hall shift triggered by varying the incidence angle is revealed. After reflection/refraction of a spin-polarized light beam at sharp interfaces, the beam contains two components: normal and abnormal modes acquiring spin-redirection Berry phases and PB phases, respectively[125]. Efficiency enhancement by several-thousand times compared to that at a conventional slab was observed at a purposely designed metamaterial slab. Therefore, it is expected that the concept and connotation of generalized geometric phase will be greatly enriched. As a typical application of the geometric phase, its dispersionless property has been utilized to realize super-resolving telescopes by generating broadband super-oscillation[39], where a resolution as high as 0.64 times that of the Rayleigh criterion was observed in an experiment with a white light source. In addition, by exploiting the geometric phase, anisotropic super-resolution photolithography can be achieved via extraordinary Young’s interference (EYI)[126].

Acknowledgments

This work was supported by the National Natural Science Foundation of China (61875253, 62105338, and U20A20217), National Key Research and Development Program of China (2021YFA1401000), Sichuan Science and Technology Program (2021ZYCD001), and Chinese Academy of Sciences Youth Innovation Promotion Association (2019371).

Disclosures

The authors declare no conflicts of interest.

References

1. 

P. K. Aravind, “A simple proof of Pancharatnam’s theorem,” Opt. Commun., 94 191 (1992). https://doi.org/10.1016/0030-4018(92)90012-G OPCOB8 0030-4018 Google Scholar

2. 

M. V. Berry, “Quantal phase factors accompanying adiabatic changes,” Proc. R. Soc. Lond. A, 392 45 (1984). https://doi.org/10.1098/rspa.1984.0023 PRLAAZ 1364-5021 Google Scholar

3. 

Y. Aharonov and D. Bohm, “Significance of electromagnetic potentials in the quantum theory,” Phys. Rev., 115 485 (1959). https://doi.org/10.1103/PhysRev.115.485 PHRVAO 0031-899X Google Scholar

4. 

R. Y. Chiao and Y.-S. Wu, “Manifestations of Berry’s topological phase for the photon,” Phys. Rev. Lett., 57 933 (1986). https://doi.org/10.1103/PhysRevLett.57.933 PRLTAO 0031-9007 Google Scholar

5. 

Y. Aharonov and J. Anandan, “Phase change during a cyclic quantum evolution,” Phys. Rev. Lett., 58 1593 (1987). https://doi.org/10.1103/PhysRevLett.58.1593 PRLTAO 0031-9007 Google Scholar

6. 

J. Zak, “Berry’s phase for energy bands in solids,” Phys. Rev. Lett., 62 2747 (1989). https://doi.org/10.1103/PhysRevLett.62.2747 PRLTAO 0031-9007 Google Scholar

7. 

S. Pancharatnam, “Generalized theory of interference and its applications,” Proc. Indian Acad. Sci. A, 44 398 (1956). https://doi.org/10.1007/BF03046095 PISAA7 0370-0089 Google Scholar

8. 

R. Simon, H. J. Kimble and E. C. G. Sudarshan, “Evolving geometric phase and its dynamical manifestation as a frequency shift: an optical experiment,” Phys. Rev. Lett., 61 19 (1988). https://doi.org/10.1103/PhysRevLett.61.19 PRLTAO 0031-9007 Google Scholar

9. 

R. Bhandari, “SU (2) phase jumps and geometric phases,” Phys. Lett. A, 157 221 (1991). https://doi.org/10.1016/0375-9601(91)90055-D PYLAAG 0375-9601 Google Scholar

10. 

M. V. Berry, “The adiabatic phase and Pancharatnam’s phase for polarized light,” J. Mod. Opt., 34 1401 (1987). https://doi.org/10.1080/09500348714551321 JMOPEW 0950-0340 Google Scholar

11. 

N. Yu et al., “Light propagation with phase discontinuities: generalized laws of reflection and refraction,” Science, 334 333 (2011). https://doi.org/10.1126/science.1210713 SCIEAS 0036-8075 Google Scholar

12. 

C. P. Jisha, S. Nolte and A. Alberucci, “Geometric phase in optics: from wavefront manipulation to waveguiding,” Laser Photonics Rev., 15 2100003 (2021). https://doi.org/10.1002/lpor.202100003 Google Scholar

13. 

J. Anandan, “The geometric phase,” Nature, 360 307 (1992). https://doi.org/10.1038/360307a0 Google Scholar

14. 

M. Berry, “Geometric phase memories,” Nat. Phys., 6 148 (2010). https://doi.org/10.1038/nphys1608 NPAHAX 1745-2473 Google Scholar

15. 

C. A. Mead, “The geometric phase in molecular systems,” Rev. Mod. Phys., 64 51 (1992). https://doi.org/10.1103/RevModPhys.64.51 RMPHAT 0034-6861 Google Scholar

16. 

E. Cohen et al., “Geometric phase from Aharonov–Bohm to Pancharatnam–Berry and beyond,” Nat. Rev. Phys., 1 437 (2019). https://doi.org/10.1038/s42254-019-0071-1 Google Scholar

17. 

A. G. Fox, “An adjustable wave-guide phase changer,” Proc. IRE, 35 1489 (1947). https://doi.org/10.1109/JRPROC.1947.234574 PIREAE 0096-8390 Google Scholar

18. 

S. Pancharatnam, “Achromatic combinations of birefringent plates,” Proc. Indian Acad. Sci. A, 41 137 (1955). https://doi.org/10.1007/BF03047098 PISAA7 0370-0089 Google Scholar

19. 

Z. Bomzon, V. Kleiner and E. Hasman, “Pancharatnam-Berry phase in space-variant polarization-state manipulations with subwavelength gratings,” Opt. Lett., 26 1424 (2001). https://doi.org/10.1364/OL.26.001424 OPLEDP 0146-9592 Google Scholar

20. 

Z. Bomzon et al., “Space-variant Pancharatnam-Berry phase optical elements with computer-generated subwavelength gratings,” Opt. Lett., 27 1141 (2002). https://doi.org/10.1364/OL.27.001141 OPLEDP 0146-9592 Google Scholar

21. 

Y. Guo et al., “Dispersion management of anisotropic metamirror for super-octave bandwidth polarization conversion,” Sci. Rep., 5 8434 (2015). https://doi.org/10.1038/srep08434 SRCEC3 2045-2322 Google Scholar

22. 

M. Pu et al., “Spatially and spectrally engineered spin-orbit interaction for achromatic virtual shaping,” Sci. Rep., 5 9822 (2015). https://doi.org/10.1038/srep09822 SRCEC3 2045-2322 Google Scholar

23. 

D. Mawet et al., “Annular groove phase mask coronagraph,” Astrophys. J., 633 1191 (2005). https://doi.org/10.1086/462409 ASJOAB 0004-637X Google Scholar

24. 

G. Biener et al., “Formation of helical beams by use of Pancharatnam-Berry phase optical elements,” Opt. Lett., 27 1875 (2002). https://doi.org/10.1364/OL.27.001875 OPLEDP 0146-9592 Google Scholar

25. 

E. Hasman et al., “Polarization dependent focusing lens by use of quantized Pancharatnam–Berry phase diffractive optics,” Appl. Phys. Lett., 82 328 (2003). https://doi.org/10.1063/1.1539300 APPLAB 0003-6951 Google Scholar

26. 

D. Tang et al., “Ultrabroadband superoscillatory lens composed by plasmonic metasurfaces for subdiffraction light focusing,” Laser Photonics Rev., 9 713 (2015). https://doi.org/10.1002/lpor.201500182 Google Scholar

27. 

E. Karimi et al., “Generating optical orbital angular momentum at visible wavelengths using a plasmonic metasurface,” Light Sci. Appl., 3 e167 (2014). https://doi.org/10.1038/lsa.2014.48 Google Scholar

28. 

G. Zheng et al., “Metasurface holograms reaching 80% efficiency,” Nat. Nanotechnol., 10 308 (2015). https://doi.org/10.1038/nnano.2015.2 NNAABX 1748-3387 Google Scholar

29. 

X. Ma et al., “A planar chiral meta-surface for optical vortex generation and focusing,” Sci. Rep., 5 10365 (2015). https://doi.org/10.1038/srep10365 SRCEC3 2045-2322 Google Scholar

30. 

M. Pu et al., “Catenary optics for achromatic generation of perfect optical angular momentum,” Sci. Adv., 1 e1500396 (2015). https://doi.org/10.1126/sciadv.1500396 STAMCV 1468-6996 Google Scholar

31. 

F. Zhang et al., “Multistate switching of photonic angular momentum coupling in phase-change metadevices,” Adv. Mater., 32 1908194 (2020). https://doi.org/10.1002/adma.201908194 ADVMEW 0935-9648 Google Scholar

32. 

D. Lin et al., “Dielectric gradient metasurface optical elements,” Science, 345 298 (2014). https://doi.org/10.1126/science.1253213 SCIEAS 0036-8075 Google Scholar

33. 

F. Aieta et al., “Multiwavelength achromatic metasurfaces by dispersive phase compensation,” Science, 347 1342 (2015). https://doi.org/10.1126/science.aaa2494 SCIEAS 0036-8075 Google Scholar

34. 

A. Arbabi et al., “Dielectric metasurfaces for complete control of phase and polarization with subwavelength spatial resolution and high transmission,” Nat. Nanotechnol., 10 937 (2015). https://doi.org/10.1038/nnano.2015.186 NNAABX 1748-3387 Google Scholar

35. 

M. Khorasaninejad et al., “Metalenses at visible wavelengths: diffraction-limited focusing and subwavelength resolution imaging,” Science, 352 1190 (2016). https://doi.org/10.1126/science.aaf6644 SCIEAS 0036-8075 Google Scholar

36. 

F. Ding et al., “Versatile polarization generation and manipulation using dielectric metasurfaces,” Laser Photonics Rev., 14 2000116 (2020). https://doi.org/10.1002/lpor.202000116 Google Scholar

37. 

M. Khorasaninejad et al., “Multispectral chiral imaging with a metalens,” Nano Lett., 16 4595 (2016). https://doi.org/10.1021/acs.nanolett.6b01897 NALEFD 1530-6984 Google Scholar

38. 

X. Luo, “Principles of electromagnetic waves in metasurfaces,” Sci. China Phys. Mech. Astron, 58 594201 (2015). https://doi.org/10.1007/s11433-015-5688-1 Google Scholar

39. 

Z. Li et al., “Achromatic broadband super-resolution imaging by super-oscillatory metasurface,” Laser Photonics Rev., 12 1800064 (2018). https://doi.org/10.1002/lpor.201800064 Google Scholar

40. 

Z. Yue et al., “Terahertz metasurface zone plates with arbitrary polarizations to a fixed polarization conversion,” Opto-Electron. Sci., 1 210014 (2022). https://doi.org/10.29026/oes.2022.210014 Google Scholar

41. 

X. Zang et al., “Metasurfaces for manipulating terahertz waves,” Light Adv. Manuf., 2 148 (2021). https://doi.org/10.37188/lam.2021.010 Google Scholar

42. 

K. Liu et al., “Active tuning of electromagnetically induced transparency from chalcogenide-only metasurface,” Light Adv. Manuf., 2 251 (2021). https://doi.org/10.37188/lam.2021.019 Google Scholar

43. 

X. Luo et al., “Broadband spin Hall effect of light in single nanoapertures,” Light Sci. Appl., 6 e16276 (2017). https://doi.org/10.1038/lsa.2016.276 Google Scholar

44. 

X. Ling et al., “Giant photonic spin Hall effect in momentum space in a structured metamaterial with spatially varying birefringence,” Light Sci. Appl., 4 e290 (2015). https://doi.org/10.1038/lsa.2015.63 Google Scholar

45. 

Y. Meng et al., “Optical meta-waveguides for integrated photonics and beyond,” Light Sci. Appl., 10 235 (2021). https://doi.org/10.1038/s41377-021-00655-x Google Scholar

46. 

L. Huang et al., “Three-dimensional optical holography using a plasmonic metasurface,” Nat. Commun., 4 2808 (2013). https://doi.org/10.1038/ncomms3808 NCAOBW 2041-1723 Google Scholar

47. 

X. Li et al., “Multicolor 3D meta-holography by broadband plasmonic modulation,” Sci. Adv., 2 e1601102 (2016). https://doi.org/10.1126/sciadv.1601102 STAMCV 1468-6996 Google Scholar

48. 

G. Qu et al., “Reprogrammable meta-hologram for optical encryption,” Nat. Commun., 11 5484 (2020). https://doi.org/10.1038/s41467-020-19312-9 NCAOBW 2041-1723 Google Scholar

49. 

Y. Hu et al., “Trichromatic and tripolarization-channel holography with noninterleaved dielectric metasurface,” Nano Lett., 20 994 (2020). https://doi.org/10.1021/acs.nanolett.9b04107 NALEFD 1530-6984 Google Scholar

50. 

D. Wen et al., “Helicity multiplexed broadband metasurface holograms,” Nat. Commun., 6 8241 (2015). https://doi.org/10.1038/ncomms9241 NCAOBW 2041-1723 Google Scholar

51. 

J.-H. Park and B. Lee, “Holographic techniques for augmented reality and virtual reality near-eye displays,” Light Adv. Manuf., 3 1 (2022). https://doi.org/10.37188/lam.2022.009 Google Scholar

52. 

Z.-L. Deng et al., “Multi-freedom metasurface empowered vectorial holography,” Nanophotonics, 11 0622 (2022). https://doi.org/10.1515/nanoph-2021-0662 2192-8614 Google Scholar

53. 

H. Gao et al., “Recent advances in optical dynamic meta-holography,” Opto-Electronic Adv., 4 210030 (2021). https://doi.org/10.29026/oea.2021.210030 Google Scholar

54. 

Y. Ming et al., “Creating composite vortex beams with a single geometric metasurface,” Adv. Mater., 34 2109714 (2022). https://doi.org/10.1002/adma.202109714 ADVMEW 0935-9648 Google Scholar

55. 

S. Zhang et al., “Generation of achromatic auto-focusing Airy beam for visible light by an all-dielectric metasurface,” J. Appl. Phys., 131 043104 (2022). https://doi.org/10.1063/5.0077930 JAPIAU 0021-8979 Google Scholar

56. 

Q. Zhou et al., “Generation of perfect vortex beams by dielectric geometric metasurface for visible light,” Laser Photon. Rev., 15 2100390 (2021). https://doi.org/10.1002/lpor.202100390 Google Scholar

57. 

Y. Guo et al., “High-efficiency and wide-angle beam steering based on catenary optical fields in ultrathin metalens,” Adv. Opt. Mater., 6 1800592 (2018). https://doi.org/10.1002/adom.201800592 2195-1071 Google Scholar

58. 

S. Xiao et al., “Flexible coherent control of plasmonic spin-Hall effect,” Nat. Commun., 6 8360 (2015). https://doi.org/10.1038/ncomms9360 NCAOBW 2041-1723 Google Scholar

59. 

T. Ma et al., “Benchmarking deep learning-based models on nanophotonic inverse design problems,” Opto-Electron. Sci., 1 210012 (2022). https://doi.org/10.29026/oes.2022.210012 Google Scholar

60. 

Y. Guo et al., “Merging geometric phase and plasmon retardation phase in continuously shaped metasurfaces for arbitrary orbital angular momentum generation,” ACS Photonics, 3 2022 (2016). https://doi.org/10.1021/acsphotonics.6b00564 Google Scholar

61. 

Y. Guo et al., “Scattering engineering in continuously shaped metasurface: an approach for electromagnetic illusion,” Sci. Rep., 6 30154 (2016). https://doi.org/10.1038/srep30154 SRCEC3 2045-2322 Google Scholar

62. 

X. Zhang et al., “A quasi-continuous all-dielectric metasurface for broadband and high-efficiency holographic images,” J. Phys. D, 53 465105 (2020). https://doi.org/10.1088/1361-6463/abaa70 JPAPBE 0022-3727 Google Scholar

63. 

X. Zhang et al., “Ultra-broadband metasurface holography via quasi-continuous nano-slits,” J. Phys. D, 53 104002 (2019). https://doi.org/10.1088/1361-6463/ab5e44 JPAPBE 0022-3727 Google Scholar

64. 

D. Wang et al., “Broadband high-efficiency chiral splitters and holograms from dielectric nanoarc metasurfaces,” Small, 15 1900483 (2019). https://doi.org/10.1002/smll.201900483 SMALBC 1613-6810 Google Scholar

65. 

D. Hakobyan et al., “Tailoring orbital angular momentum of light in the visible domain with metallic metasurfaces,” Adv. Opt. Mater., 4 306 (2016). https://doi.org/10.1002/adom.201500494 2195-1071 Google Scholar

66. 

M. Xu et al., “Topology-optimized catenary-like metasurface for wide-angle and high-efficiency deflection: from a discrete to continuous geometric phase,” Opt. Express, 29 10181 (2021). https://doi.org/10.1364/OE.422112 OPEXFF 1094-4087 Google Scholar

67. 

Z. Gong et al., “Broadband efficient vortex beam generation with metallic helix array,” Appl. Phys. Lett., 113 071104 (2018). https://doi.org/10.1063/1.5039804 APPLAB 0003-6951 Google Scholar

68. 

D. Hakobyan et al., “Tailoring orbital angular momentum of light in the visible domain with metallic metasurfaces,” Adv. Opt. Mater., 4 306 (2015). https://doi.org/10.1002/adom.201500494 2195-1071 Google Scholar

69. 

X. Luo et al., “Catenary functions meet electromagnetic waves: opportunities and promises,” Adv. Opt. Mater., 8 2001194 (2020). https://doi.org/10.1002/adom.202001194 2195-1071 Google Scholar

70. 

M. Pu et al., “Nanoapertures with ordered rotations: symmetry transformation and wide-angle flat lensing,” Opt. Express, 25 31471 (2017). https://doi.org/10.1364/OE.25.031471 OPEXFF 1094-4087 Google Scholar

71. 

F. Zhang et al., “Extreme-angle silicon infrared optics enabled by streamlined surfaces,” Adv. Mater., 33 2008157 (2021). https://doi.org/10.1002/adma.202008157 ADVMEW 0935-9648 Google Scholar

72. 

Y. Guo et al., “Polarization-controlled broadband accelerating beams generation by single catenary-shaped metasurface,” Adv. Opt. Mater., 7 1900503 (2019). https://doi.org/10.1002/adom.201900503 2195-1071 Google Scholar

73. 

F. Zhang et al., “Broadband and high-efficiency accelerating beam generation by dielectric catenary metasurfaces,” Nanophotonics, 9 20200057 (2020). https://doi.org/10.1515/nanoph-2020-0057 Google Scholar

74. 

X. Chen et al., “Dual-polarity plasmonic metalens for visible light,” Nat. Commun., 3 1198 (2012). https://doi.org/10.1038/ncomms2207 NCAOBW 2041-1723 Google Scholar

75. 

F. Zhang et al., “Symmetry breaking of photonic spin-orbit interactions in metasurfaces,” Opto-Electron. Eng., 44 319 (2017). https://doi.org/10.3969/j.issn.1003-501X.2017.03.006 Google Scholar

76. 

F. Zhang et al., “All-dielectric metasurfaces for simultaneous giant circular asymmetric transmission and wavefront shaping based on asymmetric photonic spin–orbit interactions,” Adv. Fun. Mater., 27 1704295 (2017). https://doi.org/10.1002/adfm.201704295 AFMDC6 1616-301X Google Scholar

77. 

J. P. B. Mueller et al., “Metasurface polarization optics: independent phase control of arbitrary orthogonal states of polarization,” Phys. Rev. Lett., 118 113901 (2017). https://doi.org/10.1103/PhysRevLett.118.113901 PRLTAO 0031-9007 Google Scholar

78. 

R. C. Devlin et al., “Arbitrary spin-to-orbital angular momentum conversion of light,” Science, 358 896 (2017). https://doi.org/10.1126/science.aao5392 SCIEAS 0036-8075 Google Scholar

79. 

J. Cai et al., “Simultaneous polarization filtering and wavefront shaping enabled by localized polarization-selective interference,” Sci. Rep., 10 14477 (2020). https://doi.org/10.1038/s41598-020-71508-7 SRCEC3 2045-2322 Google Scholar

80. 

Q. Fan et al., “Independent amplitude control of arbitrary orthogonal states of polarization via dielectric metasurfaces,” Phys. Rev. Lett., 125 267402 (2020). https://doi.org/10.1103/PhysRevLett.125.267402 PRLTAO 0031-9007 Google Scholar

81. 

Y. Yuan et al., “Independent phase modulation for quadruplex polarization channels enabled by chirality-assisted geometric-phase metasurfaces,” Nat. Commun., 11 4186 (2020). https://doi.org/10.1038/s41467-020-17773-6 NCAOBW 2041-1723 Google Scholar

82. 

Y. Bao et al., “Toward the capacity limit of 2D planar Jones matrix with a single-layer metasurface,” Sci. Adv., 7 eabh0365 (2021). https://doi.org/10.1126/sciadv.abh0365 STAMCV 1468-6996 Google Scholar

83. 

D. Wang et al., “Efficient generation of complex vectorial optical fields with metasurfaces,” Light Sci. Appl., 10 67 (2021). https://doi.org/10.1038/s41377-021-00504-x Google Scholar

84. 

D. Wang et al., “High-efficiency metadevices for bifunctional generations of vectorial optical fields,” Nanophotonics, 10 685 (2021). https://doi.org/10.1515/nanoph-2020-0465 Google Scholar

85. 

I. Kim et al., “Pixelated bifunctional metasurface-driven dynamic vectorial holographic color prints for photonic security platform,” Nat. Commun., 12 3614 (2021). https://doi.org/10.1038/s41467-021-23814-5 NCAOBW 2041-1723 Google Scholar

86. 

A. H. Dorrah et al., “Metasurface optics for on-demand polarization transformations along the optical path,” Nat. Photonics, 15 287 (2021). https://doi.org/10.1038/s41566-020-00750-2 NPAHBY 1749-4885 Google Scholar

87. 

N. A. Rubin et al., “Jones matrix holography with metasurfaces,” Sci. Adv., 7 eabg7488 (2021). https://doi.org/10.1126/sciadv.abg7488 STAMCV 1468-6996 Google Scholar

88. 

M. Liu et al., “Broadband generation of perfect Poincaré beams via dielectric spin-multiplexed metasurface,” Nat. Commun., 12 2230 (2021). https://doi.org/10.1038/s41467-021-22462-z NCAOBW 2041-1723 Google Scholar

89. 

F. Zhang et al., “Synthetic vector optical fields with spatial and temporal tenability,” Sci. China Phys. Mech. Astron., 65 254211 (2022). https://doi.org/10.1007/s11433-021-1851-0 SCPMCL 1674-7348 Google Scholar

90. 

J. Ni et al., “Multidimensional phase singularities in nanophotonics,” Science, 374 eabj0039 (2021). https://doi.org/10.1126/science.abj0039 SCIEAS 0036-8075 Google Scholar

91. 

S. Zhang et al., “Broadband detection of multiple spin and orbital angular momenta via dielectric metasurface,” Laser Photonics Rev., 14 2000062 (2020). https://doi.org/10.1002/lpor.202000062 Google Scholar

92. 

Y. Guo et al., “Spin-decoupled metasurface for simultaneous detection of spin and orbital angular momenta via momentum transformation,” Light Sci. Appl., 10 63 (2021). https://doi.org/10.1038/s41377-021-00497-7 Google Scholar

93. 

X. Luo et al., “Symmetric and asymmetric photonic spin-orbit interaction in metasurfaces,” Prog. Quant. Electron., 79 100344 (2021). https://doi.org/10.1016/j.pquantelec.2021.100344 PQUEAH 0079-6727 Google Scholar

94. 

Z. Fei et al., “Metasurfaces enabled by asymmetric photonic spin-orbit interactions,” Opto-Electron Eng., 47 200366 (2020). https://doi.org/10.12086/oee.2020.200366 Google Scholar

95. 

M. Tymchenko et al., “Gradient nonlinear Pancharatnam-Berry metasurfaces,” Phys. Rev. Lett., 115 207403 (2015). https://doi.org/10.1103/PhysRevLett.115.207403 PRLTAO 0031-9007 Google Scholar

96. 

G. Li et al., “Continuous control of the nonlinearity phase for harmonic generations,” Nat. Mater., 14 607 (2015). https://doi.org/10.1038/nmat4267 NMAACR 1476-1122 Google Scholar

97. 

N. Segal et al., “Controlling light with metamaterial-based nonlinear photonic crystals,” Nat. Photonics, 9 180 (2015). https://doi.org/10.1038/nphoton.2015.17 NPAHBY 1749-4885 Google Scholar

98. 

S. Chen et al., “Symmetry-selective third-harmonic generation from plasmonic metacrystals,” Phys. Rev. Lett., 113 033901 (2014). https://doi.org/10.1103/PhysRevLett.113.033901 PRLTAO 0031-9007 Google Scholar

99. 

K. Konishi et al., “Polarization-controlled circular second-harmonic generation from metal hole arrays with threefold rotational symmetry,” Phys. Rev. Lett., 112 135502 (2014). https://doi.org/10.1103/PhysRevLett.112.135502 PRLTAO 0031-9007 Google Scholar

100. 

G. Li, S. Zhang and T. Zentgraf, “Nonlinear photonic metasurfaces,” Nat. Rev. Mater., 2 17010 (2017). https://doi.org/10.1038/natrevmats.2017.10 Google Scholar

101. 

G. Li et al., “Spin and geometric phase control four-wave mixing from metasurfaces,” Laser Photon. Rev., 12 1800034 (2018). https://doi.org/10.1002/lpor.201800034 Google Scholar

102. 

M. Ma et al., “Optical information multiplexing with nonlinear coding metasurfaces,” Laser Photon. Rev., 13 1900045 (2019). https://doi.org/10.1002/lpor.201900045 Google Scholar

103. 

Z. Li et al., “Multiplexed nondiffracting nonlinear metasurfaces,” Adv. Funct. Mater., 30 1910744 (2020). https://doi.org/10.1002/adfm.201910744 AFMDC6 1616-301X Google Scholar

104. 

C. Schlickriede et al., “Imaging through nonlinear metalens using second harmonic generation,” Adv. Mater., 30 1703843 (2018). https://doi.org/10.1002/adma.201703843 ADVMEW 0935-9648 Google Scholar

105. 

F. Walter et al., “Ultrathin nonlinear metasurface for optical image encoding,” Nano Lett., 17 3171 (2017). https://doi.org/10.1021/acs.nanolett.7b00676 NALEFD 1530-6984 Google Scholar

106. 

B. Reineke et al., “Silicon metasurfaces for third harmonic geometric phase manipulation and multiplexed holography,” Nano Lett., 19 6585 (2019). https://doi.org/10.1021/acs.nanolett.9b02844 NALEFD 1530-6984 Google Scholar

107. 

W. Zhao et al., “Chirality-selected second-harmonic holography with phase and binary amplitude manipulation,” Nanoscale, 12 13330 (2020). https://doi.org/10.1039/D0NR03431B NANOHL 2040-3364 Google Scholar

108. 

Z. Gao et al., “Reconstruction of multidimensional nonlinear polarization response of Pancharatnam-Berry metasurfaces,” Phys. Rev. B, 104 054303 (2021). https://doi.org/10.1103/PhysRevB.104.054303 Google Scholar

109. 

C. McDonnell et al., “Functional THz emitters based on Pancharatnam-Berry phase nonlinear metasurfaces,” Nat. Commun., 12 30 (2021). https://doi.org/10.1038/s41467-020-20283-0 NCAOBW 2041-1723 Google Scholar

110. 

W. Ye et al., “Spin and wavelength multiplexed nonlinear metasurface holography,” Nat. Commun., 7 11930 (2016). https://doi.org/10.1038/ncomms11930 NCAOBW 2041-1723 Google Scholar

111. 

M. A. Kats et al., “Giant birefringence in optical antenna arrays with widely tailorable optical anisotropy,” Proc. Nat. Acad. Sci., 109 12364 (2012). https://doi.org/10.1073/pnas.1210686109 Google Scholar

112. 

X. Xie et al., “Generalizd Pancharatnam-Berry phase in rotationally symmetric meta-atoms,” Phys. Rev. Lett., 126 3902 (2021). https://doi.org/10.1103/PhysRevLett.126.183902 PRLTAO 0031-9007 Google Scholar

113. 

S. Yijia et al., “Achromatic metalens based on coordinative modulation of propagation phase and geometric phase,” Opto-Electron Eng., 47 200237 (2020). https://doi.org/10.12086/oee.2020.200237 Google Scholar

114. 

N. V. Tabiryan et al., “Advances in transparent planar optics: enabling large aperture, ultrathin lenses,” Adv. Opt. Mater., 9 2001692 (2021). https://doi.org/10.1002/adom.202001692 2195-1071 Google Scholar

115. 

Y. Tang et al., “Nonlinear vectorial metasurface for optical encryption,” Phys. Rev. Appl., 12 024028 (2019). https://doi.org/10.1103/PhysRevApplied.12.024028 PRAHB2 2331-7019 Google Scholar

116. 

Y. Zhang et al., “Multidimensional manipulation of wave fields based on artificial microstructures,” Opto-Electron. Adv., 3 200002 (2020). https://doi.org/10.29026/oea.2020.200002 Google Scholar

117. 

Y. Wang, Q. Fan and T. Xu, “Design of high efficiency achromatic metalens with large operation bandwidth using bilayer architecture,” Opto-Electron. Adv., 4 200008 (2021). https://doi.org/10.29026/oea.2021.200008 Google Scholar

118. 

R. Zhao et al., “Multichannel vectorial holographic display and encryption,” Light Sci. Appl., 7 95 (2018). https://doi.org/10.1038/s41377-018-0091-0 Google Scholar

119. 

E. Arbabi et al., “Vectorial holograms with a dielectric metasurface: ultimate polarization pattern generation,” ACS Photonics, 6 2712 (2019). https://doi.org/10.1021/acsphotonics.9b00678 Google Scholar

120. 

L. Fang et al., “Vectorial doppler metrology,” Nat. Commun., 12 4186 (2021). https://doi.org/10.1038/s41467-021-24406-z NCAOBW 2041-1723 Google Scholar

121. 

E. Wang et al., “Complete control of multichannel, angle-multiplexed, and arbitrary spatially varying polarization fields,” Adv. Opt. Mater., 8 1901674 (2020). https://doi.org/10.1002/adom.201901674 2195-1071 Google Scholar

122. 

J. Wang et al., “Shifting beams at normal incidence via controlling momentum-space geometric phases,” Nat. Commun., 12 6046 (2021). https://doi.org/10.1038/s41467-021-26406-5 NCAOBW 2041-1723 Google Scholar

123. 

Z. Zhang et al., “Enhancing the efficiency of the topological phase transitions in spin–orbit photonics,” Appl. Phys. Lett., 120 181102 (2022). https://doi.org/10.1063/5.0086930 APPLAB 0003-6951 Google Scholar

124. 

W. Zhu et al., “Wave-vector-varying Pancharatnam-Berry phase photonic spin hall effect,” Phys. Rev. Lett., 126 083901 (2021). https://doi.org/10.1103/PhysRevLett.126.083901 PRLTAO 0031-9007 Google Scholar

125. 

X. Ling et al., “Topology-induced phase transitions in spin-orbit photonics,” Laser Photon. Rev., 15 2000492 (2021). https://doi.org/10.1002/lpor.202000492 Google Scholar

126. 

M. Pu et al., “Revisitation of extraordinary Young’s interference: from catenary optical fields to spin-orbit interaction in metasurfaces,” ACS Photonics, 5 3198 (2018). https://doi.org/10.1021/acsphotonics.8b00437 Google Scholar
CC BY: © The Authors. Published by CLP and SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Yinghui Guo, Mingbo Pu, Fei Zhang, Mingfeng Xu, Xiong Li, Xiaoliang Ma, and Xiangang Luo "Classical and generalized geometric phase in electromagnetic metasurfaces," Photonics Insights 1(1), R03 (18 August 2022). https://doi.org/10.3788/PI.2022.R03
Received: 26 April 2022; Accepted: 4 July 2022; Published: 18 August 2022
Lens.org Logo
CITATIONS
Cited by 32 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Polarization

Phase shift keying

Spiral phase plates

Dielectric polarization

Electromagnetism

Geometrical optics

Nonlinear optics

Back to Top