Abstract
The rapid development of optical technologies, such as optical manipulation, data processing, sensing, microscopy, and communications, necessitates new degrees of freedom to sculpt optical beams in space and time beyond conventionally used spatially homogenous amplitude, phase, and polarization. Structuring light in space and time has been indeed shown to open new opportunities for both applied and fundamental science of light. Rapid progress in nanophotonics has opened up new ways of “engineering” ultra-compact, versatile optical nanostructures, such as optical two-dimensional metasurfaces or three-dimensional metamaterials that facilitate new ways of optical beam shaping and manipulation. Here, we review recent progress in the field of structured light–matter interactions with a focus on all-dielectric nanostructures. First, we introduce the concept of singular optics and then discuss several other families of spatially and temporally structured light beams. Next, we summarize recent progress in the design and optimization of photonic platforms, and then we outline some new phenomena enabled by the synergy of structured light and structured materials. Finally, we outline promising directions for applications of structured light beams and their interactions with engineered nanostructures.
1 Introduction
Despite the long history of optics, it remains a dynamic research field with an expanding range of applications. It is noteworthy that until the 1990s, optics primarily dealt with plane waves, Gaussian beams, and smooth wavefronts [1]. This situation drastically changed thanks to the pioneering works of Marat Soskin of Ukraine [2–4] and Les Allen of the UK [5–7] who introduced the concepts of experimental “singular optics” and “orbital angular momentum of light”, respectively. Since then, the concepts of singular beams, optical vortices, structured light (also known as sculpted, custom, customized, tailored, and complex light beams) permanently entered the domain of modern optics. Nowadays, it is well-established that a beam of light can carry both spin and orbital angular momenta (SAM and OAM) parallel to the beam axis [8]. In particular, the orbital angular momentum is related to the helical phase front of the light beam, while the spin angular momentum is associated with its polarization [5, 9]. Today, the term “structured light” describes a variety of optical waveforms with the spatial inhomogeneity of one or more physical parameters in two- or three-dimensional space and time, including beams carrying SAM and OAM, Bessel–Gaussian beams [10, 11], radially and azimuthally polarized vector beams [12], optical links and knots [13–15], spatio-temporal optical vortices (STOVs) [16–18], and flying donut pulses [19–21]. Some of these beams already find applications in particle manipulation, optical communications, quantum information processing, sensing, and microscopy [22–25], while others remain the subjects of scientific curiosity primarily because of the practical challenges associated with their experimental realization. It is worth mentioning that several intricate optical structures, such optical knots and skyrmions, have only been achieved in recent years, reflecting the rapid progress and ongoing innovation within this field of research [26–28].
Rapid progress in nanofabrication and computational electromagnetics has opened up new prospects for the realization of versatile optical nanostructures, such as optical two-dimensional metasurfaces [29–31] or three-dimensional metamaterials [32–34] that facilitate unprecedented opportunities for tailoring optical waveforms. For instance, revolutionary developments in computational optics enabled powerful Maxwell’s equation solvers that facilitated the generalization of classical Mie theory first introduced more than a century ago for spherical particles [35] arbitrary shaped complex particles. In parallel, recent progress in electron beam lithography and focused ion beam lithography has enabled the experimental realization of arbitrary shaped particles at nanoscale. These scatterers can be arranged in isolated, two- and three-dimensional arrangements, which are referred to as meta-atoms, metasurfaces, and metamaterials, respectively. It should be remarked that the prefix “meta” means “beyond” in Greek, reflecting that these engineered nanostructures enable optical properties beyond those realizable using conventional (non-structured) materials. Owing to the success and various breakthroughs enabled by such engineered (dielectric) materials, nowadays the unique name of Mie-tronics has been dedicated to this field of research [36–39] (see the top row in Figure 1 for the main categories of Mie-tronics). Consequently, Mie-tronics offers exceptional opportunities to design ultra-compact, multifunctional, and reconfigurable components that facilitate structured light generation, detection, shaping, steering, and multiplexing, among various other applications [38–47]. While to date, the research in the field of Mie-tronics was largely limited to conventional Gaussian light beams, in this paper, we aim at reviewing the new phenomena and potential applications that emerge from the synergy of these two branches of optical science – structured light and Mie-tronics. We will briefly review several families of structured light beams and discuss their amplitude, phase, and polarization properties. Next, we review the basic results of Mie theory and discuss both the exact and long wavelength approximation (LWA) multipole decompositions. In the third section, we discuss recent progress in the field of conventional (Gaussian illumination-based) Mie-tronics. Afterward, we provide an overview of recent developments in the field of structured light-structured matter interactions and discuss the potential and unique opportunities enabled by such a synergy. Finally, we provide an outlook for the outstanding opportunities in this rapidly developing field.
![Figure 1:
(Top row) The bird’s eye view of main categories of Mie-tronics, ranging from passive structures including single resonator (meta-atom), one dimensional (meta-chain), and two dimensional (meta-surface) periodic resonators as well as possible tuning mechanisms for generating active and reconfigurable platforms. (Bottom row) The zoology of structured lights comprises of various types of light beams ranging from LG beams to flying donut pulses. The combination of these two seemingly unrelated fields leads to unique phenomena such as selective excitation of induced multipolar moments, and nonlinear harmonic generation. The STOV and FD are adapted from Ref [17, 23].](/document/doi/10.1515/nanoph-2023-0030/asset/graphic/j_nanoph-2023-0030_fig_001.jpg)
(Top row) The bird’s eye view of main categories of Mie-tronics, ranging from passive structures including single resonator (meta-atom), one dimensional (meta-chain), and two dimensional (meta-surface) periodic resonators as well as possible tuning mechanisms for generating active and reconfigurable platforms. (Bottom row) The zoology of structured lights comprises of various types of light beams ranging from LG beams to flying donut pulses. The combination of these two seemingly unrelated fields leads to unique phenomena such as selective excitation of induced multipolar moments, and nonlinear harmonic generation. The STOV and FD are adapted from Ref [17, 23].
2 Structured light beams
Contrary to conventional light beams such as plane waves or Gaussian beams, structured light beams are complex optical fields with unique spatial and temporal distributions of amplitude, phase, and/or polarizations [48–54]. Several experimental techniques including a pair of cylindrical lenses [5], spiral phase plates [55, 56] q-plates [57–59], spatial light modulators (SLM) [60–62], and optical metasurfaces [63–69] have been developed to tailor the amplitude, wavefront, and polarization of light for numerous applications such as optical communications, quantum information processing, optical imaging, metrology, microscopy, and optical manipulation [22–25]. Here, we briefly review the mathematical description of structured light beams and highlight their unique features. More details on this subject can be found in [12, 23, 24, 49, 50, 64].
2.1 Family of paraxial wave solutions
As the starting point, we consider the paraxial solution of the Helmholtz equation which can be expressed as
where
where
where n
LG and m
LG represent the radial and azimuthal mode indices of the LG beam, and

Laguerre–Gaussian and Hermite–Gaussian mode intensity and phase profiles for various combinations of order indices. Apart from the fundamental mode, with
The relation between the LG and HG modes are obtained based expressing the
By inserting Eq. (5) into Eq. (4), the LG of even and odd modes can be represented in terms of HG light beam as
As can be seen from Eq. (6), the LG light beam of even and odd modes can be represented as the superposition of HG beams with various mode indices. Specifically, even modes can be expressed using a combination of
It is noteworthy that while Poincare sphere is commonly used to visualize the polarization of light, it can also be a useful tool for understanding complex polarization states, and other degrees of freedom such as OAM [75]. Similar to its conventional counterpart, such a representation is referred to as higher order Poincare sphere throughout the literature and can provide a more complete understanding of the properties and behavior of electromagnetic waves [76]. In addition, the so-called Ince–Gaussian (IG) beams are known to be the exact analytical solutions of the paraxial wave equation in an elliptic cylindrical coordinates system [77]. The general expression of the IG can be decomposed into even and odd solutions, each written as
where ɛ indicates the ellipticity parameter,

Transverse field distributions of amplitude and phase of (a) even and (b) odd Inc-Gaussian light beams with ɛ = 2 and for various combinations of
As can be seen from Figure 3, the beam widths of the
2.2 Cylindrical vector beams
While traditionally the polarization state of light was assumed to be spatially uniform (linear, circular, and elliptical), light beams with spatially varying polarizations are expected to open new avenues to novel phenomena in optical systems. Indeed, the vectorial nature of light beams, their generation, as well as their interaction with matter gained significant attention in the last decade [12]. In particular, cylindrical vector beams (CVBs) are considered to be one special class of such vectorial light beams having cylindrically symmetric inhomogeneous polarizations, which its simplest form (1st-order CVB) can be expressed in terms of the superposition of two
where the subscripts of APB and RPB denote the azimuthally and radially polarized beams shown in Figure 4(a) and (b), respectively. We note that the superposition of RPB and APB leads to the so-called hybrid CVB shown in Figure 4(c).

The field distributions of 1st-order CVBs obtained based on the linear superposition of
2.3 Non-separable states of light
So far, we discussed time-independent solutions of Maxwell’s equations. However, a completely different class of exact solutions to Maxwell’s equations exists, known as non-separable space-time solutions, where the spatial and temporal features of the optical fields are coupled and cannot be separated [48]. The first three-dimensional, non-dispersive, source-free solutions to homogeneous Maxwell’s equations have been introduced by Brittingham and named the focused wave mode (FWM) solutions [79]. However, these first mathematical FWM solutions carried infinite energies. Later, in 1985, Ziolkowski solved the paradox of infinite energy by superposing focus wave modes with carefully chosen weighting functions. These new solutions were called electromagnetic directed energy pulse trains (EDEPTs) [80, 81]. The family of EDEPTs includes non-diffracting pulses with azimuthal dependence [82], focused pancake pulses [83, 84], and Flying Donut (FD) pulses [19, 85]. In particular, FD pulses are characterized by exotic donut-like topological structures and have been shown to provide nontrivial light–matter interaction on account of the similarity between its structural features and the radiation pattern of toroidal dipolar moments [86]. Nevertheless, due to their spatiotemporal structural complexity, the studies of FD pulses have been limited to theory and numerical simulations. However, recently Zheludev et al. proposed and realized for the first time an experimental approach for the generation and propagation of FD pulses using photonic metasurfaces [21]. Another class of structured light beams is the spatiotemporal optical vortex whose phase and energy circulate in both planes of space and time [16, 17]. Thanks to the recent experimental demonstration of STOVs by Milchberg et al. [16, 18, 87], these exotic light beams have gained the attention of the scientific community due to their importance in both theoretical and practical studies such as in describing the particle collisions, optics of moving media and quantum communications [17].
3 Light-matter interaction at nanoscale
3.1 Mie scattering of nanoparticles
Light scattering by subwavelength spherical particles can be qualitatively described by Mie theory [35, 88, 89]. According to this theory, the extinction, and scattering efficiencies that are defined as the ratios between their corresponding cross-sections to the geometrical cross-section are given by
where R is the radius of the spherical particle, l is the orbital mode order (
where β = 2πR/λ is the normalized size parameter, n is the refractive index of the particle, and

Optical response of a subwavelength scatterer. (a) The scattering response of a dielectric nanosphere derived from Mie theory up to the octupolar terms (a 3 and b 3). (b) The optical behavior of the subwavelength scatterers when its refractive index changes from n 1 = 3.5 to n 2 = 4. (c) The scattering spectrum of the same particle when the host medium refractive index is altered from air to glass.
Figure 5(c) demonstrates the scenario wherein the radius and refractive index of the spherical particle are the same as that of Figure 5(a), yet the optical properties of the host medium change from the air (n
air = 1) to glass (
3.2 Multipole expansion for arbitrary shape meta-atoms: exact moments and long-wavelength approximation
The optical response of nonmagnetic meta-atoms is characterized by the electric current density
J
in and polarization
P
in induced by external electromagnetic fields. In the case of monochromatic fields, these values are connected by
J
in = −iω
P
in, wherein ω represents the optical angular frequency [68]. Upon the interaction of light with the meta-atom, the induced polarization is related to the field distributions within the particle via
where
n
=
r
/r is the unit vector directed from the particle’s center towards an observation point, k
d
is the wavenumber in the surrounding medium, μ
0 is the permeability of free space, and I is a 3 × 3 unitary matrix. The multipole decomposition of scattered waves can be obtained based on three approaches: (a) Taylor expansion of
wherein
where
where
D
corresponds to the exact total electric dipole (TED),
m
is the MD moment, and
where the second term in
D
LWA is known as the electric toroidal dipole (TD) moment that can interfere with its conventional basic counterpart (first term) either constructively or destructively, leading to a wide spectrum of exotic phenomena such as the formation of anapole state [86]. Using these notations, the far-field scattered power can be readily related to the scattered fields of Eq. (13) using a time-averaged Poynting vector defined as
wherein

The calculated scattering cross-sections and contribution of exact and LWA moments for (a) cubic and (b) cylindrical subwavelength meta-atoms. The dashed lines correspond to the exact multipolar moments, while the circles represent the LWA moments.
4 Mie-tronics in a Nutshell
While the interactions between the conventional Gaussian light beams and nanostructured materials, or metamaterials, have been extensively studied in the last decade [36], [37], [38, 95], [96], [97], in this section, we will briefly review this rapidly developing field research and highlight some unique regimes of such light–matter interactions, and outline their potential applications. While early studies in this field of research focused on the interaction of light with plasmonic nanoantennas [98–100], high refractive index engineered dielectric meta-atoms, have been shown to provide an alternative route to manipulate light through the excitation of their resonant modes, or Mie resonances discussed in the previous section. Such all-dielectric meta-atoms offer the significant advantage of being virtually lossless in the visible and/or near-infrared wavelength ranges, where their plasmonic counterparts suffer from ohmic losses. Moreover, dielectric meta-atoms support both electric and magnetic resonances determined by their topology and material properties (i.e., dielectric constant or refractive index) [101]. In particular, magnetic dipole resonance was proposed and demonstrated in 2012 by Kuznetsov et al. [102] in the visible regime as shown in Figure 7(a). Following this initial demonstration of magnetic resonances in a nonmagnetic, dielectric material, significant effort has been devoted to the studies of the effect of geometries and material properties on such resonances [103–114].
![Figure 7:
An overview of Mie-tronics covering the traditional light interaction with isolated optical scatterers. (a) Kuznetsov et al. experimentally demonstrated that a spherical silicon nanoparticle can support a strong magnetic dipole resonance at various wavelengths depending on its relative size [102]. (b) Fu et al. experimentally demonstrated that by tuning the size of a spherical silicon nanoparticle, its corresponding far-field radiation pattern can be manipulated such that directional light scattering with a high forward-to-backward scattering ratio is obtained [115]. (c) Chong et al. experimentally demonstrated light scattering by all-dielectric oligomers composed of silicon nanoparticles and the possibility of the excitation of Fano resonances within these configurations [116]. (d) Miroshnichenko et al. demonstrated the existence of non-radiating anapole moments in optics [117]. (e) Grinblat et al. studied the third-harmonic generation from low-loss subwavelength germanium particles supporting anapole moment [118]. (f) Zenin et al. experimentally demonstrated the efficient excitation of higher-order anapole moments within all-dielectric silicon particles that can exceed the first-order anapole state [119]. (g) Cihan et al. experimentally demonstrated that by utilizing a silicon nanowire supporting Mie resonances, one can vary the directionality, polarization state, and spectral emission of two-dimensional materials [120]. (h) Verre et al. showed that nanostructures made of transition metal dichalcogenides can support various multipolar moments such as anapole states [121]. (i) Zograf et al. demonstrated the stimulated Raman scattering for isolated crystalline silicon nanoparticles and experimentally observed a transition from spontaneous to stimulated scattering [122]. (j) Kepic et al. demonstrated that vanadium dioxide nanoantennas can be utilized for designing tunable metasurfaces in the visible range thanks to the phase transition properties of VO2 [123]. (k) Asadchy et al. described light scattering by a spherical particle whose permittivity is modulated in time and presented a route to obtain directional light amplification [124].](/document/doi/10.1515/nanoph-2023-0030/asset/graphic/j_nanoph-2023-0030_fig_007.jpg)
An overview of Mie-tronics covering the traditional light interaction with isolated optical scatterers. (a) Kuznetsov et al. experimentally demonstrated that a spherical silicon nanoparticle can support a strong magnetic dipole resonance at various wavelengths depending on its relative size [102]. (b) Fu et al. experimentally demonstrated that by tuning the size of a spherical silicon nanoparticle, its corresponding far-field radiation pattern can be manipulated such that directional light scattering with a high forward-to-backward scattering ratio is obtained [115]. (c) Chong et al. experimentally demonstrated light scattering by all-dielectric oligomers composed of silicon nanoparticles and the possibility of the excitation of Fano resonances within these configurations [116]. (d) Miroshnichenko et al. demonstrated the existence of non-radiating anapole moments in optics [117]. (e) Grinblat et al. studied the third-harmonic generation from low-loss subwavelength germanium particles supporting anapole moment [118]. (f) Zenin et al. experimentally demonstrated the efficient excitation of higher-order anapole moments within all-dielectric silicon particles that can exceed the first-order anapole state [119]. (g) Cihan et al. experimentally demonstrated that by utilizing a silicon nanowire supporting Mie resonances, one can vary the directionality, polarization state, and spectral emission of two-dimensional materials [120]. (h) Verre et al. showed that nanostructures made of transition metal dichalcogenides can support various multipolar moments such as anapole states [121]. (i) Zograf et al. demonstrated the stimulated Raman scattering for isolated crystalline silicon nanoparticles and experimentally observed a transition from spontaneous to stimulated scattering [122]. (j) Kepic et al. demonstrated that vanadium dioxide nanoantennas can be utilized for designing tunable metasurfaces in the visible range thanks to the phase transition properties of VO2 [123]. (k) Asadchy et al. described light scattering by a spherical particle whose permittivity is modulated in time and presented a route to obtain directional light amplification [124].
In addition to the excitation of various electric and magnetic type resonances, the interference and interplay between different Mie-type resonances have been shown to enable many remarkable spectral properties of all-dielectric structures [115, 125–129]. In particular, in 2013, Fu et al. have experimentally shown that silicon spherical particles can support strong unidirectional radiation patterns that rely on the interference of electric and magnetic resonances excited within the nanoparticles as shown in Figure 7(b) [115]. Apart from single scatterers, clusters of subwavelength particles of different geometries including dimers, trimers, quadrumers, and hexamers, have been shown to support new regimes of light–matter interactions [116, 130–141]. For instance, in 2014, Chong et al. have both theoretically and experimentally demonstrated the excitation of Fano resonances within an all-dielectric cylindrical heptamer that form due to the interference of the Mie-type magnetic mode of the central meta-atom with the collective resonant mode of the other meta-atoms within the configuration as shown in Figure 7(c) [116]. While a majority of the studies in the field of Mie-tronics before 2013 have been based on the assumption that any alternating current distribution should radiate electromagnetic energy to the far field, significant efforts have been devoted to finding non-radiating sources, which were theoretically predicted by Afanasiev and Stepanovsky in 1995 as the destructive interference between electric and toroid dipole moments [142]. Moreover, Zheludev et al. have experimentally shown such a peculiar nonradiating state for the first time in the microwave regime, which was then termed as anapole (from Greek “ana”, “without”, thus meaning “without poles”) [143]. Followed by such a breakthrough, in 2015, Miroshnichenko et al. demonstrated that anapoles can be also realized in the visible region based on the destructive interference between the basic and toroidal contribution of the electric dipole in the far field as shown in Figure 7(d) [117]. Noteworthy that such anapole states are characterized by significant confinement of energy within the optical scatterer, making it a promising platform for enhanced light–matter interactions on the nanoscale that may be particularly important for nonlinear optics applications such as harmonics generation and wavelength conversion as shown in Figure 7(e) [118]. More details on the topic of anapole state and its higher-order counterparts (shown in Figure 7(f)) can be found in [119, 144–162].
Recently, a variety of high-index materials including germanium (Ge), silicon (Si), and titanium dioxide (TiO2) has been used to explore various phenomena in the field of Mie-tronics [163]. However, besides these conventional dielectrics, there is a plethora of other materials, such as transition metal dichalcogenides (TMDCs) that exhibit attractive characteristics in the visible and infrared (IR) spectral regions [164–166]. In particular, TMDC crystals are two-dimensional (2D) layers configured in stacked arrangements and are weakly adhered by van de Waals interactions [167]. Each consisting layer comprises three atomic layers, represented as MX2, wherein X denotes a chalcogen atom (e.g., X = S, Se, Te) and M indicates a transition metal atom (e.g., M = Mo, W) and can be synthesis by various methods such as molecular beam epitaxy (MBE) [168] and chemical vapor deposition (CVD) [169]. Aside from the numerous applications enabled by these 2D materials, their integration with structured matters supporting Mie-type resonances leads to hybrid entities, which can potentially open a plethora of applications in nano-photonic such as the manipulation of excitonic emission in near- and far-field regimes, routing valley polarized chiral emission, and establishing strong coupling regime of interaction [170, 171], to name just a few. For instance, Cihan et al. experimentally demonstrated that the integration of a silicon nanowire supporting Mie resonances with a MoS2-TMDC can change the directionality, polarization state, and spectral emission of the integrated 2D material on demand as it is shown in Figure 7(g) [120]. Recently, it has been revealed that bulk TMDCs can provide an unusually high refractive index in both near-infrared and visible regimes [121, 170, 172, 173], making them favorable candidates for optical applications. Moreover, due to rapid advancements in nanofabrication techniques, it is now possible to create desired patterns directly from bulk TMDCs. In this context, Verre et al. experimentally demonstrated that such TMDC-based nanoparticles can support Mie-type resonances, and the induced resonant modes can interfere with one another to excite nonradiating anapole states as shown in Figure 7(h) [121].
Besides the mentioned applications of Mie-tronics, in recent years several studies have been dedicated to tailoring spontaneous Raman scattering with Mie-like resonances provided by subwavelength particles of various shapes and materials [174]. In this perspective, Zograf et al. [122] observed Stimulated Raman scattering from isolated subwavelength crystalline silicon (c-Si) nanoparticles, for the first time as shown in Figure 7(i). In particular, by optimizing the corresponding dimensions of the nanoparticle and substrate, they have been able to enhance the stimulated emission without overheating, which can provide an additional degree of freedom for obtaining Raman scattering from nanoscale structures. However, despite the wide spectrum of applications and great potential enabled by subwavelength scatterers, a grand challenge facing these platforms is that their response is set in stone after they have been designed and cannot be changed afterward. Therefore, an immense effort has been made to overcome this limitation of passive isolated meta-atoms by exploiting mechanical, thermal, and electrical tuning mechanisms such that the scattering response of the meta-atoms can be controlled in real-time by control over external stimuli (such as strain, temperature or voltage) rather than the change in the geometry [123, 175–189]. An example of such efforts is illustrated in Figure 7(j), where Kepic et al. [123] experimentally demonstrated that vanadium dioxide (VO2) meta-atoms can be tuned using temperature and operate as optical switches changing their properties from dielectric to plasmonics. In addition, more recently, time-modulated platforms have emerged as a new class of active devices in which the external stimuli controlling the optical properties are varying periodically in time [190–201]. Such temporally varying structures have been shown to enable a wide range of novel physical phenomena including nonreciprocity [202–208], and signal amplification [124, 209, 210], as shown in Figure 7(k). One of the most remarkable properties of such temporally modulated structures is the possibility of frequency mixing enabling the generation of higher-order frequency harmonics without the need for high-intensity pump beams [190–192].
Finally, in addition to the studies of resonant excitations of individual meta-atoms, one- and two-dimensional periodic arrangements of these meta-atoms, or meta-chain and metasurfaces, have been also shown to facilitate beam steering [211], holography [212], nonlinear harmonic generation [213–216], Kerker, anti-Kerker, and transverse Kerker effects [112, 217, 218], invisibility [219–221], absorber [222–225], optical force manipulation [226–227], and topological waveguiding [230] to name a few. However, as this review aims to address the interaction of structured lights with structured matters and elucidate the potential avenues opened by their synergy, we limit our discussion to the case of single scatterers.
5 Structured-light- structured-matter interaction
In recent years, a significant effort has been put to generate and implement structured light for a wide range of applications including optical trapping, metrology, probing, and data processing [49, 50, 64], while in parallel, the field of Mie-tronics continued to develop rapidly, opening new avenues for shaping and manipulation of light [36–47]. The interaction of structured light with structured matter is expected to give rise to a wide range of light–matter interactions that are not accessible using conventional Gaussian beams (or plane waves) or nonresonant Rayleigh scatterers. In this section, we focus on structured light-structured matter interactions and their experimental studies and outline their potential applications.
5.1 Interactions of SAM- and OAM-carrying light beams with matter
In most general case, light can possess both spin and orbital angular momentum. The spin angular momentum (SAM) is related to the polarization state of the light. In this context, each photon carries an angular momentum of ±ℏ, where the ± sign indicates whether the polarization is right-handed or left-handed. In addition to the SAM, light beams with an azimuthal phase dependence of
![Figure 8:
The role of angular momentum of light in scattering from subwavelength meta-atoms. (a) The experimental setup used by Garbin et al. [231] to distinguish the topological charge of an optical vortex from the measured Mie scattering spectra. The inset demonstrates the scattered intensity distributions of a displaced off-axis spherical particle for Gaussian (m = 0) and LG (m = 1 and m = 2) illuminations. Upon illumination of a Gaussian beam, the measured intensity distribution moves parallel to the direction of translation, while for an LG illumination, the fringes move along a diagonal plane. (b) The experimental investigation carried out by Petrov et al. [237] to characterize the geometrical properties of spherical particles from their interactions with beams carrying angular momentum. (b-1) The dependence of the optical response of the scatterers under the illumination with the beams with two different topological charges of m = 0 and m = 1 as a function of the spherical meta-atom position. The first and second columns correspond to a particle whose radius is 2 μm and 1 μm, respectively. (c) The calculated scattering efficiency of a spherical particle as a function of wavelength for various mode indices of LG beams,
n
LG
,
m
LG
$\left({n}_{\text{LG}},{m}_{\text{LG}}\right)$
, shown on top of each plot [238]. (d) The experimental setup used by Zambrana-Puyalto et al. [239] aimed at exploring the role of SAM and OAM on the scattering spectra of spherical meta-atoms. The back-scattered light beams measured upon the interaction of LG beams with the SAM of (d-1) +ℏ and (d-2) −ℏ and topological charges −2 < m < +2 as a function of wavelength.](/document/doi/10.1515/nanoph-2023-0030/asset/graphic/j_nanoph-2023-0030_fig_008.jpg)
The role of angular momentum of light in scattering from subwavelength meta-atoms. (a) The experimental setup used by Garbin et al. [231] to distinguish the topological charge of an optical vortex from the measured Mie scattering spectra. The inset demonstrates the scattered intensity distributions of a displaced off-axis spherical particle for Gaussian (m = 0) and LG (m = 1 and m = 2) illuminations. Upon illumination of a Gaussian beam, the measured intensity distribution moves parallel to the direction of translation, while for an LG illumination, the fringes move along a diagonal plane. (b) The experimental investigation carried out by Petrov et al. [237] to characterize the geometrical properties of spherical particles from their interactions with beams carrying angular momentum. (b-1) The dependence of the optical response of the scatterers under the illumination with the beams with two different topological charges of m = 0 and m = 1 as a function of the spherical meta-atom position. The first and second columns correspond to a particle whose radius is 2 μm and 1 μm, respectively. (c) The calculated scattering efficiency of a spherical particle as a function of wavelength for various mode indices of LG beams,
Besides the studies of structured light beams’ interactions with structured matters shown in Figure 8, many other remarkable results have been reported in recent years [240–254]. For instance, the interaction of nanoholes with a vortex beam carrying both SAM and OAM is studied in [240] demonstrating that in contrast to the total angular momentum, the polarization handedness of light may not be preserved in the process of spin–orbit coupling. Moreover, a significant effort has also focused on the studies of the selective excitation of multipolar resonances [247] and optical manipulations [242] of all-dielectric nanostructures interacting with various types of structured light beams. We also note that besides the dielectric platform, the interactions between structured light and plasmonic resonators have attracted considerable interest in recent years due to their potential applications in various fields, such as sensing, and imaging [255]. For instance, recently, it has been demonstrated that plasmonic nanoantennas can be used as an alternative mechanism to directly measure the information encoded in twisted lights via converting the OAM into spectral information using bright and dark modes [253].
5.2 Interaction of cylindrical vector beams with structured media
In addition to the interaction of OAM and SAM carrying light beams with structured media, several studies were dedicated to the investigation of vector beams’ interactions with optical nanostructures. Importantly, it was shown that the light beams themselves can be engineered to excite new resonances in nanoscale particles. For example, Wozniak et al. [256] demonstrated that by tailoring the spatial structure of the incident light beam, particularly using a tightly focused APB and RPB with a high NA = 0.9 objective lens, individual multipole resonances can be selectively excited while other multipoles are simultaneously suppressed, as it is schematically shown in Figure 9(a). To experimentally demonstrate such new regimes of light–matter interactions, well-separated individual silicon nanospheres were fabricated on top of a glass substrate and investigated using the setup shown in Figure 9(a-1). The reflectance and transmittance are shown in Figure 9(a-2) and (a-3), for APB and RPB, respectively. As can be seen from these two panels, upon changing the polarization of the incident beams, the scattering spectra of the same particle vary significantly, and new resonances emerge. For instance, upon the illumination with the APB, both MD and MQ resonances emerge within the spectrum of the particle, while when the polarization is changed to the RPB, the previously excited resonances (MD and MQ) disappear, and new ED resonance emerges. Therefore, by choosing a tightly focused CVB, the desired multipolar moment can be selectively excited and manipulated, which allows for the identification of resonances in terms of their type and orientation.
![Figure 9:
The potential avenues opened by CVB interaction with meta-atoms. (a) The schematic illustration of selective excitation of multipolar moments within the subwavelength dielectric particle. (a-1) The experimental setup utilized by Wozniak et al. for studying the interaction of a CVB with an isolated nanosphere [256]. The measured transmittance and reflectance of the isolated particle under (a-2) azimuthally and (a-3) radially polarized light beams. Upon the change in the structure of the incoming light beam, the induced Mie-type resonant mode within the subwavelength scatterer varies. (b) The artistic demonstration of selective excitation of Mie resonances for nonlinear enhancement application. The first meta-atom supports EQ under RP excitation and generates TH, whereas efficient TH is generated in the second meta-atom due to the excitation of MQ under APB [257]. (b-1) The setup used for experimental investigation carried out by Melik-Gaykazyan et al. to measure the TH efficiency under (b-2) RPB and (b-3) APB. (c) The schematic demonstration of the excitation setup used in [258] to investigate the origin of QBIC in terms of Fano resonances. (c-1) The measured reflectance spectra of the subwavelength scatterer as a function of its diameter for azimuthal (first column) and linear (second column) polarizations, respectively. (c-2) The calculated results of the Q-factor extracted from the measured reflectance of the panel (c-1). (d) The graphic demonstration of the SHG in a cylindrical subwavelength meta-atom under APB excitation [259]. (d-1) The three-dimensional measurement result of the SH intensity under APB illumination as a function of the nanodisk dimension and pump wavelength. (d-2) The intensity enhancement map measured for three scenarios of APB (first column), RPB (second column), and linear (third column) excitations.](/document/doi/10.1515/nanoph-2023-0030/asset/graphic/j_nanoph-2023-0030_fig_009.jpg)
The potential avenues opened by CVB interaction with meta-atoms. (a) The schematic illustration of selective excitation of multipolar moments within the subwavelength dielectric particle. (a-1) The experimental setup utilized by Wozniak et al. for studying the interaction of a CVB with an isolated nanosphere [256]. The measured transmittance and reflectance of the isolated particle under (a-2) azimuthally and (a-3) radially polarized light beams. Upon the change in the structure of the incoming light beam, the induced Mie-type resonant mode within the subwavelength scatterer varies. (b) The artistic demonstration of selective excitation of Mie resonances for nonlinear enhancement application. The first meta-atom supports EQ under RP excitation and generates TH, whereas efficient TH is generated in the second meta-atom due to the excitation of MQ under APB [257]. (b-1) The setup used for experimental investigation carried out by Melik-Gaykazyan et al. to measure the TH efficiency under (b-2) RPB and (b-3) APB. (c) The schematic demonstration of the excitation setup used in [258] to investigate the origin of QBIC in terms of Fano resonances. (c-1) The measured reflectance spectra of the subwavelength scatterer as a function of its diameter for azimuthal (first column) and linear (second column) polarizations, respectively. (c-2) The calculated results of the Q-factor extracted from the measured reflectance of the panel (c-1). (d) The graphic demonstration of the SHG in a cylindrical subwavelength meta-atom under APB excitation [259]. (d-1) The three-dimensional measurement result of the SH intensity under APB illumination as a function of the nanodisk dimension and pump wavelength. (d-2) The intensity enhancement map measured for three scenarios of APB (first column), RPB (second column), and linear (third column) excitations.
While an optical response of a nanoresonator is largely determined by its geometry and material, the efficiency of light coupling to a particular resonant mode strongly depends on the structure of the excitation beam. Recently, structured light-induced Mie resonances have been exploited in the context of nonlinear optics as shown in Figure 9(b). In particular, Melik-Gaykazyan et al. [257] utilized APB and RPB for the excitation of the resonances of the silicon nanocylinder. The meta-atom dimensions were chosen such that once the structure of the incident light beam changes from radially polarized to azimuthally polarized beam, the optical response of the particles switches from EQ to MQ as shown in panel (b). The optical response of the cylindrical meta-atom with a height fixed to 700 nm, and a radius varying between 270 and 500 nm was investigated under RPB and APB excitation at the pump wavelength of λ p = 1550 nm as shown in Figure 9(b-1). The linear optical response of the meta-atom indicates that for the case of RPB excitation, the nanocylinder with the radius of 310 nm (first meta-atom) possesses a strong EQ response while when the incoming light beam changes to APB, this resonant peak disappears. On the other hand, under the APB excitation of the particle with a radius of 460 nm (second meta-atom), the optical response is dominated by the MQ resonant mode. To experimentally demonstrate such a capability of CVB interaction with structured matter, an experimental setup of Figure 9(b-1) was utilized. In a different approach from previous works, Kruk et al. [65] obtained radially polarized and azimuthally polarized light beams using a fabricated metasurface. The laser beam at the fundamental frequency of 193.55 THz illuminated the sample and its third harmonics (TH) counterpart was collected as shown in Figure 9(b-2) and (b-3) for RPB and APB, respectively. Both the simulation and experimental results indicate that the APB light beam enhances the TH in the close vicinity of the MQ resonance, while for the RPB excitation, dominant TH power is observed near the EQ resonance. These results demonstrate the possibility of using structured light as another degree of freedom in tailoring the optical response of subwavelength dielectric nanoparticles in nonlinear applications.
Up to this point, we focused on the bright modes that efficiently radiate into free space. In contrast, dark modes, such as the bound state in the continuum (BIC), are confined to the structure with infinite lifetime and zero radiation. In general, BICs are a unique class of resonant states in which the state is perfectly discrete and spatially confined in space but can exist at the same energy as a continuum of states propagating to the infinity [260]. Subsequently, BIC resonant modes do not couple to the continuum of radiation modes outside the structure and are “trapped” in the structure. In quantum mechanics, momentum and position are related through the uncertainty principle, ΔxΔp ≥ ℏ/2, indicating that a particle’s momentum and position cannot be simultaneously known with complete precision. In this context, the confinement of the bound state to a finite region in real space results in the delocalization of the system’s properties in momentum space. In other words, from a momentum-space perspective, a BIC resonant mode can be seen as anomalies where the BIC energy is equal to the energy of the continuum radiation modes. In contrast, other resonant states, such as Fabry–Perot modes or surface plasmon polaritons, do not exhibit such anomalies in momentum space. These modes have finite lifetimes and can couple to the radiation continuum outside the structure, leading to a broadening of the resonance peak in momentum space. While ideal BICs are inaccessible from external excitation, a quasi-BIC (QBIC) can potentially be accessed from the free space either under the oblique incidence or upon breaking the in-plane symmetry under normal incidence (symmetry-protected BIC), providing a platform for achieving high-quality factor (Q-factor) resonances and boosting light–matter interaction at the nanoscale [261]. It is noteworthy that while most of the studies of QBIC resonances were limited to plane wave interactions, recently, several studies focused on the potential applications of azimuthally polarized light beams in the efficient coupling of light to the QBIC resonant modes [258, 259, 262, 263]. Very similar to BIC resonant modes, Fano resonances represent a universal wave phenomenon having two essential features of asymmetric and ultrasharp spectral line shapes, which are attributed to the interference of a continuum of bright modes with the dark modes [264]. In 2021, Melik-Gaykazyan et al. [258] experimentally uncovered the underlying physical mechanism linking the physics of QBIC and the asymmetric features of Fano resonances in the isolated aluminum gallium arsenide (AlGaAs) nanodisks of various diameters and explained how the transition between these two resonant modes can be obtained, as shown in Figure 9(c). The measurement results for various size nanocylinders in the spectral range of 1450 nm–1700 nm for two types of polarizations, that is azimuthal (first column) and linear (second column) are shown in Figure 9(c-1). As can be seen from the first column, an asymmetric resonant line shape emerged in the scattering spectrum of the nanocylinder, whereas for the linear polarization (second column), the induced resonant peak is suppressed regardless of the nanocylinder size. Importantly, for the azimuthal polarization, depending on the size of the meta-atoms, the sign of the Fano parameter changes from negative (red curve for 980 nm diameter) to positive (purple curve for 890 nm diameter). In this process, the Fano resonance transitions to the QBIC resonant mode at the meta-atom diameter of 930 nm (orange curve) where the asymmetrical line shape changes to a symmetrical one as shown in Figure 9(c-2). The obtained results of panel (c) highlight the role of the pump beam structuring to enable selective excitation of high-Q modes that may find applications for the enhancement of nonlinear conversion efficiency.
As we discussed earlier, azimuthally polarized light beams provide unique opportunities to excite QBIC resonant modes within subwavelength particles with high Q-factors. Recently, the interaction of APB with the isolated AlGaAs nanocylinders was studied by Koshelev et al. [259], who demonstrated the possibility to boost the second-harmonic generation (SHG) via inducing QBIC resonant modes with a Q-factor of 190 (Figure 9(d)). In particular, the proposed configuration is a three-layer heterostructure consisting of a silica substrate, indium tin oxide (ITO), and an AlGaAs resonator supporting a Mie resonance at the second harmonic (SH) wavelength simultaneously with the fundamental frequency. Two orders of magnitude enhancement of the SH intensity in the close vicinity of the QBIC resonance under the illumination of APB is shown in Figure 9(d-1). For comparison, the SHG intensity maps for other types of illuminations (RPB and linear polarization) are provided in Figure 9(d-2), clearly revealing the difference in the enhancement of the SHG between the APB excitation and other kinds of beams. It should be noted that the advances of structured light–matter interactions in the context of nonlinear optics are not limited to the aforementioned works and many other remarkable results have been reported, including the possibility of selective excitation of magnetic multipolar moments in isolated cylinders for boosting the SHG or the enhancement of the THG from all-dielectric oligomers [265–273].
5.3 From radiative to nonradiative states with structured lights
So far, we focused on the discussion of isolated multipolar moments. Here we show that the interference between induced multipolar moments leads to a plethora of fascinating light–matter interactions at the nanoscale. In particular, the theory of electromagnetic multipole expansion, including charge-current spherical and Cartesian decompositions, is important for the theoretical description of these interactions [88–94]. For instance, it allows us to describe a toroidal moments family, which can be represented as the poloidal currents flowing along the meridians of a torus for the electric toroidal dipole (ETD) moment and more complex current distributions for higher-order toroidal moments, also known as mean square radii [274]. In particular, an electric dipole anapole (EDA) state forms upon the destructive interference between an electric dipole (ED) and its toroidal counterpart (i.e., ETD), which subsequently leads to vanishing scattering accompanied by strong energy confinement within the subwavelength scatterer [161]. The studies of the anapole states in engineered meta-atoms have received considerable attention due to their potential applications in strong exciton coupling, second and third-harmonic generation, Raman scattering, photocatalysis, guiding energy, and lasing [160–162]. However, despite the fruitful progress in this field of research, the complete suppression of scattering from the EDA state is prohibited, due to the simultaneous excitation of MQ modes within the presented meta-atoms. To overcome such a problem, Wei et al. [275] carried out a theoretical investigation to excite an ideal radiationless anapole state within an isotropic high-index dielectric nanosphere via a “4π configuration” comprising of two counter-propagating radially polarized light beams with the same amplitude and a π phase difference, as shown in Figure 10(a). In particular, owing to the presence of longitudinally polarized electric field
where τ = z − ct, σ = z + ct, and f
0 represent an arbitrary normalization constant, while q
1 and q
2 denote the effective wavelength of the pulse and the focal region depth, respectively. The results of numerical simulations shown in the first row of Figure 10(b-1) suggest that while for the particular spectral position of ν = 0.46c/q
1 (marked by black dashed line), the ED (blue line) and TD (red line) possess resonant peaks, the EQ moment has a minimum. However, despite the presence of such electrical resonant modes, the total scattering intensity spectrum given in the second row of the panel (b-1) closely resembles the EQ moment. This similarity is attributed to the large value of the interference term between electric and toroidal dipoles, given by
![Figure 10:
Formation of non-radiating anapole state under different types of structured light incidences. (a) The 4π illumination setup utilized by Wei et al. in Ref [275] to suppress the radiation to the far field (left side) and the results of numerical simulations corresponding to the normalized internal energy and scattering power spectrum (right side). (a-1) The calculated incident (left column), total (middle column), and scattered (right column) electric field distributions at the operating wavelength corresponding to the anapole excitation. (b) The artistic illustration of a TM FD pulse impinging on a dielectric spherical meta-atom [276]. (b-1) The numerical calculation of LWA moments excited within the nanoresonator of radius R = q
1 (top row) and its corresponding scattering intensity. (c) The measurement results of the scattering spectra of isolated silicon-based nanospheres (the SEM image shown in the inset) under the illumination of linear (black curve), radial (red curve), and azimuthal (blue curve) polarization light beams [277]. (d) The diagram of a Si nanodisk fabricated on a glass substrate and its SEM images utilized in Ref [278] to experimentally investigate the formation of two distinct radiative and non-radiative resonant modes, under the excitation with CVB. (d-1) The optical setup utilized by Lu et al. for measuring the backscattering spectra of the fabricated sample and its (d-2) measured spectra for APB (blue color graph) and RPB (red color curve), respectively.](/document/doi/10.1515/nanoph-2023-0030/asset/graphic/j_nanoph-2023-0030_fig_010.jpg)
Formation of non-radiating anapole state under different types of structured light incidences. (a) The 4π illumination setup utilized by Wei et al. in Ref [275] to suppress the radiation to the far field (left side) and the results of numerical simulations corresponding to the normalized internal energy and scattering power spectrum (right side). (a-1) The calculated incident (left column), total (middle column), and scattered (right column) electric field distributions at the operating wavelength corresponding to the anapole excitation. (b) The artistic illustration of a TM FD pulse impinging on a dielectric spherical meta-atom [276]. (b-1) The numerical calculation of LWA moments excited within the nanoresonator of radius R = q 1 (top row) and its corresponding scattering intensity. (c) The measurement results of the scattering spectra of isolated silicon-based nanospheres (the SEM image shown in the inset) under the illumination of linear (black curve), radial (red curve), and azimuthal (blue curve) polarization light beams [277]. (d) The diagram of a Si nanodisk fabricated on a glass substrate and its SEM images utilized in Ref [278] to experimentally investigate the formation of two distinct radiative and non-radiative resonant modes, under the excitation with CVB. (d-1) The optical setup utilized by Lu et al. for measuring the backscattering spectra of the fabricated sample and its (d-2) measured spectra for APB (blue color graph) and RPB (red color curve), respectively.
Despite the recent theoretical and experimental advancements in the field of anapole state with plane wave illuminations, the experimental demonstration of these nonradiating resonant modes with structured lights is still in its infancy. To this end, Parker et al. [277] have recently illustrated the excitation of a non-scattering anapole state in all-dielectric silicon-based spherical particles with various sizes, using tightly focused radially and azimuthally polarized CVBs as shown in Figure 10(c). In particular, the experimental studies of particles with a diameter of 160 nm, indicate that upon RPB illumination, the scattering spectra of the meta-atom possess pronounced minima in contrast with the APB and linear polarization excitation. This result can be explained by the destructive interference of electrical and toroidal dipole moments. Followed by such an experimental demonstration, and the ability of CVBs to selectively excite Mie resonant modes, very recently, Lu et al. [278] have experimentally demonstrated the transition from nonradiating anapole state to radiative MQ resonant mode within a cylindrical meta-atom (the top and side views of scanning electron microscopy (SEM) images of the fabricated samples are shown in Figure 10(d)) using radially and azimuthally light beams, respectively, as illustrated in Figure 10(d-1).
The schematic of the experimental setup is shown in Figure 10 (d-1). Figure 10 (d-2) shows a resonant peak at 735 nm under APB illumination, which is attributed to MQ resonant mode (blue curve), while when the polarization was changed to RPB (red curve), the scattering spectrum possesses a dip at the same spectral position corresponding to the excitation of an anapole state. These results offer new opportunities for switching between the maximum and minimum scattering at the same wavelength by utilizing tightly focused AP and RP beams.
5.4 Structured light-induced chiroptical response
Chiral objects, including the DNA and proteins, are structures whose mirror images (enantiomers) are not superimposable with that of their original topologies. They can possess different chiroptical responses, such as optical rotatory dispersion (ORD), and circular dichroism (CD) [279, 280]. Circular dichroism is defined as the differential absorption of left and right circular polarization (LCP or RCP). While traditionally the CD is associated with chiral geometries, Zambrana-Puyalto et al. [281] have experimentally shown the possibility of inducing chiroptical responses from achiral objects using light beams carrying an OAM as shown in Figure 11(a). The experimental setup and the process of generating OAM carrying light beam are the same as that of Figure 8 with the high NA = 1.1 objective lens, yet the achiral samples are gold-based circular apertures with various diameters (the SEM image is shown in the inset) that are centered with respect to the incident beam. The CD of such a sample is measured in a two-step process, such that first the transmitted intensity of the LCP light beam carrying the topological charge of m (i.e.,
![Figure 11:
The potential applications of structured lights in chiral light–matter interactions. (a) The experimental setup utilized by Zambrana-Puyalto in Ref [281] for investigating the role of angular momentum of the incident light in the CD of a non-chiral nanostructure shown in the inset. (a-1) The measurements result of the CD(%), calculated from
C
D
m
%
=
I
m
L
−
I
m
R
/
I
m
L
+
I
m
R
×
100
$\text{C}{\text{D}}_{m}\left(\%\right)=\left[\left({I}_{m}^{L}-{I}_{m}^{R}\right)/\left({I}_{m}^{L}+{I}_{m}^{R}\right)\right]\times 100$
, for three different TC of
m
=
−
1,0
,
+
1
$m=\left[-1,0,+1\right]$
as a function of the nano-holes diameters. While the green curve is associated with the TC of m = 0, the black and pink color graphs correspond to m = 1 and m = −1, respectively. (b) The artistic illustration of the chiral structure and its SEM image used by Wozniak et al. in Ref [282] for probing the vorticity of the incident beam and its (b-1) measured spectra for various incident topological charges as a function of the operating wavelength. (c) The experimental results associated with the scattering intensity of the left-handed (top left), right-handed (top right), and achiral (bottom left) microstructures illuminated by a linearly polarized light beam carrying an orbital angular momentum of mℏ [283]. The corresponding VD of the measured results are also shown in the bottom right figure with red (for left-handed meta-atom), blue (for right-handed scatterer), and orange (achiral structure) color curves. (d) and (d-1) The conceptual sketch of tunable VD based on the twisted stereometamaterials proposed by Liu et al. in Ref [284] and its (d-2) corresponding experimental results as a function of the incident beam topological charge and the twisted angle of θ.](/document/doi/10.1515/nanoph-2023-0030/asset/graphic/j_nanoph-2023-0030_fig_011.jpg)
The potential applications of structured lights in chiral light–matter interactions. (a) The experimental setup utilized by Zambrana-Puyalto in Ref [281] for investigating the role of angular momentum of the incident light in the CD of a non-chiral nanostructure shown in the inset. (a-1) The measurements result of the CD(%), calculated from
Ni et al. [283] have both theoretically and experimentally illustrated a significant chiroptical response arising from the interaction of a non-paraxial OAM-carrying light beam (focused with NA of 0.9) with intrinsically chiral microstructures as it is illustrated in Figure 11(c). As can be seen from the reported results of this work, for intrinsically chiral structures possessing both left- and right-handiness (see panels (c-1) and (c-2)), the scattering spectra exhibit different chiroptical responses as a function of the topological charge of the incoming beam, while for the achiral sample (panel (c-2)) no prominent change is observed. By defining the VD in a similar fashion as that of Zambrana-Puyalto’s work, (i.e.,
Finally, in 2022 Liu et al. [284] proposed a twisted stereo-metamaterial, shown in Figure 11(d), that can lead to a tunable VD ranging from −97% to +98%. In particular, it was shown that regardless of the sign and value of the topological charge of the incoming wave, when the twisted angle (θ) is zero (achiral structure), the handiness of the incident beam does not affect the chiroptical response of the meta-atom (Figure 11(d)), whereas different behavior can be observed once the twisted angle is changed to nonzero values (chiral geometry shown in Figure 11(d-1)), providing an opportunity to achieve tailorable response upon engineering the rotational angle. Experimentally measured VD spectra with respect to the rotational angle and the value of the topological charge at the fixed wavelength of 800 nm are shown in Figure 11 (d-2). As can be seen from this figure, for the achiral structure (θ = 0°), the measured VD is zero regardless of the value of TC, whereas once the level of geometrical achirality increases, the VD spectrum reaches the maximum value of +98% for θ = 22.5° and then drops to the minimum value of −97% for θ = 67.5°, confirming the tailorable chiroptical response resulting from the structured light–matter interaction. As the final remark, it is worth noting that in the context of structured light, both paraxial and non-paraxial solutions have been demonstrated to enable chiral light–matter interactions [279–284]. In the paraxial approximation, the transverse components of the wave vector are significantly smaller than the longitudinal component, resulting in light propagating primarily along the optical axis. On the other hand, non-paraxial light beams do not adhere to these assumptions, giving rise to more intricate and diverse phenomena, such as vortical dichroism.
6 Summary and outlook
In this paper, we have outlined the recent advances in the field of structured light–matter interactions and elucidated the role of the spatial structure of light beams in manipulating the scattering spectra of dielectric meta-atoms and their assemblies. We overviewed various theoretical and experimental approaches to study the synergy of structured light and structured matter in linear, nonlinear, and chiral optics. In this outlook section, we attempt to highlight the remaining open questions, potential trends, and future directions in this fascinating branch of modern optics, which we named “singular optics empowered by engineered optical materials”.
Despite significant progress in the field of structured light interaction with resonant Mie particles, the majority of the studies focused on spherically or cylindrically symmetric particles, while the synergy of structured light beams and complex-shaped meta-atoms has not been fully explored. One possible research direction would be the study of the interaction of an OAM-carrying light beam, with nonspherical and randomly tilted meta-atoms with varying aspect ratios. For instance, owing to the particular structure of the LG beam and its OAM, such an interaction with arbitrary shape meta-atoms can lead to entirely new scattering characteristics and degrees of freedom for selective excitation, suppression, and manipulation of individual resonant modes as compared to those obtained with a conventional Gaussian beam and spherically/cylindrically symmetric particles, which in turn opens new prospects for applications such as remote sensing and communications in scattering media.
Moreover, electric anapole state, which arises due to the destructive interferences of electric and toroidal electric dipoles, has been introduced as the fundamental class of non-radiating sources and has enabled a plethora of applications ranging from nonlinear optics to thermodynamics. Nevertheless, from the standpoint of physics and duality of Maxwell equations, the existence of other types of anapole states such as magnetic anapole is also possible and can potentially provide enhanced radiation suppression and confinement of electromagnetic energy, if they are spectrally overlapped with their electrical counterparts. While the existence of magnetic anapole with structured light has been theoretically investigated in [285], the possibility of exciting hybrid anapoles is yet to be addressed. Therefore, another research direction would be the study and the design of meta-atoms that support these new non-radiative states, called hybrid anapoles, which are formed due to the destructive interference between electric and magnetic basic and toroidal multipoles of various types.
Another interesting scattering phenomenon not discussed here in detail is super-scattering, which is the phenomenon opposite of a non-radiating state. In particular, energy conservation enforces a limit on the portion of the energy scattered into one scattering channel. As a result, the maximum contribution of scattering cross-section (SCR) from any given multipole moment is limited to
Finally, as we discussed in this review, optical metasurfaces offer several exciting opportunities for the generation and manipulation of structured light including the OAM-carrying beams, CVBs, and FD pulses. However, engineered nanostructures may revolutionize many other branches of singular optics, including the generation of STOVs, higher-dimensional structured light (such optical links, knots, etc.), and linear and nonlinear quantum optics with structured light for atom trapping, communications, and computing. Therefore, we expect that the combination of singular optics with Mie-tronics will continue broadening the horizons of nanophotonics and offer new perspectives for meta-optics, leading to novel applications beyond our current imagination.
Funding source: Army Research Office
Award Identifier / Grant number: W911NF1810348
Funding source: Office of Naval Research
Award Identifier / Grant number: N00014-20-1-2558
-
Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: This paper was supported in part by the Office of Naval Research (ONR) (Grant No. N00014-20-1-2558), and Army Research Office Award (Grant No. W911NF1810348).
-
Conflict of interest statement: The authors declare no conflict of interest.
References
[1] O. V. Angelsky, A. Y. Bekshaev, S. G. Hanson, C. Y. Zenkova, I. I. Mokhun, and Z. Jun, “Structured light: ideas and concepts,” Front. Phys., vol. 8, p. 114, 2020. https://doi.org/10.3389/fphy.2020.00114.Search in Google Scholar
[2] M. S. Soskin, M. V. Vasnetsov, and I. V. Basistiy, “Optical wavefront dislocations,” in International Conf. on Holography and Correlation Optics, vol. 2647, Chernivsti, Ukraine, SPIE, 1995.10.1117/12.226741Search in Google Scholar
[3] M. S. Soskin and M. V. Vasnetsov, Eds., Proc. Int. Conf. on Singular Optics (Partenit, October, 1997); Proc. SPIE 3487, 1998.Search in Google Scholar
[4] M. Soskin, S. V. Boriskina, Y. Chong, M. R. Dennis, and A. Desyatnikov, “Singular optics and topological photonics,” J. Opt., vol. 19, no. 1, p. 010401, 2016. https://doi.org/10.1088/2040-8986/19/1/010401.Search in Google Scholar
[5] L. Allen, M. W. Beijersbergen, R. J. C. Spreeuw, and J. P. Woerdman, “Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes,” Phys. Rev. A, vol. 45, no. 11, p. 8185, 1992. https://doi.org/10.1103/physreva.45.8185.Search in Google Scholar PubMed
[6] M. W. Beijersbergen, L. Allen, H. van der Veen, and J. Woerdman, “Astigmatic laser mode converters and transfer of orbital angular momentum,” Opt. Commun., vol. 96, nos. 1–3, pp. 123–132, 1993. https://doi.org/10.1016/0030-4018(93)90535-d.Search in Google Scholar
[7] S. Franke‐Arnold, L. Allen, and M. Padgett, “Advances in optical angular momentum,” Laser Photon. Rev., vol. 2, no. 4, pp. 299–313, 2008. https://doi.org/10.1002/lpor.200810007.Search in Google Scholar
[8] R. P. Cameron, S. M. Barnett, and A. M. Yao, “Optical helicity, optical spin and related quantities in electromagnetic theory,” New J. Phys., vol. 14, no. 5, p. 053050, 2012. https://doi.org/10.1088/1367-2630/14/5/053050.Search in Google Scholar
[9] L. Allen, S. M. Barnett, and M. J. Padgett, Optical Angular Momentum, Boca Raton, Florida, CRC Press, 2016.10.1201/9781482269017Search in Google Scholar
[10] X. Chu, Q. Sun, J. Wang, P. Lu, W. Xie, and X. Xu, “Generating a Bessel-Gaussian beam for the application in optical engineering,” Sci. Rep., vol. 5, no. 1, pp. 1–8, 2015. https://doi.org/10.1038/srep18665.Search in Google Scholar PubMed PubMed Central
[11] J. K. Miller, D. Tsvetkov, P. Terekhov, et al.., “Spatio-temporal controlled filamentation using higher order Bessel-Gaussian beams integrated in time,” Opt. Express, vol. 29, no. 13, pp. 19362–19372, 2021. https://doi.org/10.1364/oe.428742.Search in Google Scholar PubMed
[12] Q. Zhan, “Cylindrical vector beams: from mathematical concepts to applications,” Adv. Opt. Photonics, vol. 1, no. 1, pp. 1–57, 2009. https://doi.org/10.1364/aop.1.000001.Search in Google Scholar
[13] J. Leach, M. R. Dennis, J. Courtial, and M. J. Padgett, “Knotted threads of darkness,” Nature, vol. 432, no. 7014, p. 165, 2004. https://doi.org/10.1038/432165a.Search in Google Scholar PubMed
[14] M. R. Dennis, R. P. King, B. Jack, K. O’Holleran, and M. J. Padgett, “Isolated optical vortex knots,” Nat. Phys., vol. 6, no. 2, pp. 118–121, 2010. https://doi.org/10.1038/nphys1504.Search in Google Scholar
[15] M. R. Dennis, K. O’holleran, and M. J. Padgett, “Singular optics: optical vortices and polarization singularities,” in Progress in Optics, vol. 53, Amsterdam, Elsevier, 2009, pp. 293–363.10.1016/S0079-6638(08)00205-9Search in Google Scholar
[16] N. Jhajj, I. Larkin, E. Rosenthal, S. Zahedpour, J. Wahlstrand, and H. Milchberg, “Spatiotemporal optical vortices,” Phys. Rev. X, vol. 6, no. 3, p. 031037, 2016. https://doi.org/10.1103/physrevx.6.031037.Search in Google Scholar
[17] A. Chong, C. Wan, J. Chen, and Q. Zhan, “Generation of spatiotemporal optical vortices with controllable transverse orbital angular momentum,” Nat. Photonics, vol. 14, no. 6, pp. 350–354, 2020. https://doi.org/10.1038/s41566-020-0587-z.Search in Google Scholar
[18] S. W. Hancock, S. Zahedpour, and H. M. Milchberg, “Second-harmonic generation of spatiotemporal optical vortices and conservation of orbital angular momentum,” Optica, vol. 8, no. 5, pp. 594–597, 2021. https://doi.org/10.1364/optica.422743.Search in Google Scholar
[19] R. W. Hellwarth and P. Nouchi, “Focused one-cycle electromagnetic pulses,” Phys. Rev. E, vol. 54, no. 1, p. 889, 1996. https://doi.org/10.1103/physreve.54.889.Search in Google Scholar PubMed
[20] A. Zdagkas, N. Papasimakis, V. Savinov, M. R. Dennis, and N. I. Zheludev, “Singularities in the flying electromagnetic doughnuts,” Nanophotonics, vol. 8, no. 8, pp. 1379–1385, 2019. https://doi.org/10.1515/nanoph-2019-0101.Search in Google Scholar
[21] A. Zdagkas, C. McDonnell, J. Deng, et al.., “Observation of toroidal pulses of light,” Nat. Photonics, vol. 16, no. 7, pp. 523–528, 2022. https://doi.org/10.1038/s41566-022-01028-5.Search in Google Scholar
[22] N. M. Litchinitser, “Structured light meets structured matter,” Science, vol. 337, no. 6098, pp. 1054–1055, 2012. https://doi.org/10.1126/science.1226204.Search in Google Scholar PubMed
[23] C. He, Y. Shen, and A. Forbes, “Towards higher-dimensional structured light,” Light Sci. Appl., vol. 11, no. 1, pp. 1–17, 2022. https://doi.org/10.1038/s41377-022-00897-3.Search in Google Scholar PubMed PubMed Central
[24] A. Forbes, M. de Oliveira, and M. R. Dennis, “Structured light,” Nat. Photonics, vol. 15, no. 4, pp. 253–262, 2021. https://doi.org/10.1038/s41566-021-00780-4.Search in Google Scholar
[25] H. Rubinsztein-Dunlop, A. Forbes, M. V. Berry, et al.., “Roadmap on structured light,” J. Opt., vol. 19, no. 1, p. 013001, 2016. https://doi.org/10.1088/2040-8978/19/1/013001.Search in Google Scholar
[26] S. Tsesses, E. Ostrovsky, K. Cohen, B. Gjonaj, N. H. Lindner, and G. Bartal, “Optical skyrmion lattice in evanescent electromagnetic fields,” Science, vol. 361, no. 6406, pp. 993–996, 2018. https://doi.org/10.1126/science.aau0227.Search in Google Scholar PubMed
[27] L. Du, A. Yang, A. V. Zayats, and X. Yuan, “Deep-subwavelength features of photonic skyrmions in a confined electromagnetic field with orbital angular momentum,” Nat. Phys., vol. 15, no. 7, pp. 650–654, 2019. https://doi.org/10.1038/s41567-019-0487-7.Search in Google Scholar
[28] Y. Shen, E. C. Martínez, and C. Rosales-Guzmán, “Generation of optical skyrmions with tunable topological textures,” ACS Photonics, vol. 9, no. 1, pp. 296–303, 2022. https://doi.org/10.1021/acsphotonics.1c01703.Search in Google Scholar
[29] H.-T. Chen, A. J. Taylor, and N. Yu, “A review of metasurfaces: physics and applications,” Rep. Prog. Phys., vol. 79, no. 7, p. 076401, 2016. https://doi.org/10.1088/0034-4885/79/7/076401.Search in Google Scholar PubMed
[30] F. Ding, A. Pors, and S. I. Bozhevolnyi, “Gradient metasurfaces: a review of fundamentals and applications,” Rep. Prog. Phys., vol. 81, no. 2, p. 026401, 2017. https://doi.org/10.1088/1361-6633/aa8732.Search in Google Scholar PubMed
[31] C.-W. Qiu, T. Zhang, G. Hu, and Y. Kivshar, “Quo vadis, metasurfaces?” Nano Lett., vol. 21, no. 13, pp. 5461–5474, 2021. https://doi.org/10.1021/acs.nanolett.1c00828.Search in Google Scholar PubMed
[32] N. Engheta and R. W. Ziolkowski, Eds., Metamaterials: Physics and Engineering Explorations, Hoboken, New Jersey, John Wiley & Sons, 2006.10.1002/0471784192Search in Google Scholar
[33] I. Staude and J. Schilling, “Metamaterial-inspired silicon nanophotonics,” Nat. Photonics, vol. 11, no. 5, pp. 274–284, 2017. https://doi.org/10.1038/nphoton.2017.39.Search in Google Scholar
[34] Y. Liu and X. Zhang, “Metamaterials: a new Frontier of science and technology,” Chem. Soc. Rev., vol. 40, no. 5, pp. 2494–2507, 2011. https://doi.org/10.1039/c0cs00184h.Search in Google Scholar PubMed
[35] G. Mie, “Beiträge zur Optik trüber Medien, speziell kolloidaler Metallösungen,” Ann. Phys., vol. 330, no. 3, pp. 377–445, 1908. https://doi.org/10.1002/andp.19083300302.Search in Google Scholar
[36] Y. Kivshar, “The rise of Mie-tronics,” Nano Lett., vol. 22, no. 9, pp. 3513–3515, 2022. https://doi.org/10.1021/acs.nanolett.2c00548.Search in Google Scholar PubMed
[37] T. Liu, R. Xu, P. Yu, Z. Wang, and J. Takahara, “Multipole and multimode engineering in Mie resonance-based metastructures,” Nanophotonics, vol. 9, no. 5, pp. 1115–1137, 2020. https://doi.org/10.1515/nanoph-2019-0505.Search in Google Scholar
[38] K. Koshelev and Y. Kivshar, “Dielectric resonant metaphotonics,” ACS Photonics, vol. 8, no. 1, pp. 102–112, 2020. https://doi.org/10.1021/acsphotonics.0c01315.Search in Google Scholar
[39] D. Smirnova and Y. S. Kivshar, “Multipolar nonlinear nanophotonics,” Optica, vol. 3, no. 11, pp. 1241–1255, 2016. https://doi.org/10.1364/optica.3.001241.Search in Google Scholar
[40] L. Cao, P. Fan, E. S. Barnard, A. M. Brown, and M. L. Brongersma, “Tuning the color of silicon nanostructures,” Nano Lett., vol. 10, no. 7, pp. 2649–2654, 2010. https://doi.org/10.1021/nl1013794.Search in Google Scholar PubMed
[41] J. A. Schuller, R. Zia, T. Taubner, and M. L. Brongersma, “Dielectric metamaterials based on electric and magnetic resonances of silicon carbide particles,” Phys. Rev. Lett., vol. 99, no. 10, p. 107401, 2007. https://doi.org/10.1103/physrevlett.99.107401.Search in Google Scholar
[42] C. Gigli, Q. Li, P. Chavel, G. Leo, M. L. Brongersma, and P. Lalanne, “Fundamental limitations of Huygens’ metasurfaces for optical beam shaping,” Laser Photon. Rev., vol. 15, no. 8, p. 2000448, 2021. https://doi.org/10.1002/lpor.202000448.Search in Google Scholar
[43] A. Cordaro, J. van de Groep, S. Raza, E. F. Pecora, F. Priolo, and M. L. Brongersma, “Antireflection high-index metasurfaces combining Mie and Fabry-Pérot resonances,” ACS Photonics, vol. 6, no. 2, pp. 453–459, 2019. https://doi.org/10.1021/acsphotonics.8b01406.Search in Google Scholar
[44] J. van de Groep and M. L. Brongersma, “Metasurface mirrors for external control of Mie resonances,” Nano Lett., vol. 18, no. 6, pp. 3857–3864, 2018. https://doi.org/10.1021/acs.nanolett.8b01148.Search in Google Scholar PubMed
[45] S. J. Kim, I. Thomann, J. Park, J. H. Kang, A. P. Vasudev, and M. L. Brongersma, “Light trapping for solar fuel generation with Mie resonances,” Nano Lett., vol. 14, no. 3, pp. 1446–1452, 2014. https://doi.org/10.1021/nl404575e.Search in Google Scholar PubMed
[46] Q. Zhao, J. Zhou, F. Zhang, and D. Lippens, “Mie resonance-based dielectric metamaterials,” Mater. Today, vol. 12, no. 12, pp. 60–69, 2009. https://doi.org/10.1016/s1369-7021(09)70318-9.Search in Google Scholar
[47] A. I. Kuznetsov, A. E. Miroshnichenko, M. L. Brongersma, Y. S. Kivshar, and B. Luk’yanchuk, “Optically resonant dielectric nanostructures,” Science, vol. 354, no. 6314, p. aag2472, 2016. https://doi.org/10.1126/science.aag2472.Search in Google Scholar PubMed
[48] Y. Shen and C. Rosales‐Guzmán, “Nonseparable states of light: from quantum to classical,” Laser Photon. Rev., vol. 16, p. 2100533, 2022. https://doi.org/10.1002/lpor.202100533.Search in Google Scholar
[49] C. Rosales-Guzmán, B. Ndagano, and A. Forbes, “A review of complex vector light fields and their applications,” J. Opt., vol. 20, no. 12, p. 123001, 2018. https://doi.org/10.1088/2040-8986/aaeb7d.Search in Google Scholar
[50] A. Forbes, “Structured light from lasers,” Laser Photon. Rev., vol. 13, no. 11, p. 1900140, 2019. https://doi.org/10.1002/lpor.201900140.Search in Google Scholar
[51] Y. Shen, X. Wang, Z. Xie, et al.., “Optical vortices 30 years on: OAM manipulation from topological charge to multiple singularities,” Light Sci. Appl., vol. 8, no. 1, p. 90, 2019. https://doi.org/10.1038/s41377-019-0194-2.Search in Google Scholar PubMed PubMed Central
[52] D. Lin, P. Fan, E. Hasman, and M. L. Brongersma, “Dielectric gradient metasurface optical elements,” Science, vol. 345, no. 6194, pp. 298–302, 2014. https://doi.org/10.1126/science.1253213.Search in Google Scholar PubMed
[53] N. Radwell, R. F. Offer, A. Selyem, and S. Franke-Arnold, “Optimisation of arbitrary light beam generation with spatial light modulators,” J. Opt., vol. 19, no. 9, p. 095605, 2017. https://doi.org/10.1088/2040-8986/aa7f50.Search in Google Scholar
[54] A. M. Weiner, “Ultrafast optical pulse shaping: a tutorial review,” Opt. Commun., vol. 284, no. 15, pp. 3669–3692, 2011. https://doi.org/10.1016/j.optcom.2011.03.084.Search in Google Scholar
[55] S. N. Khonina, A. V. Ustinov, V. I. Logachev, and A. P. Porfirev, “Properties of vortex light fields generated by generalized spiral phase plates,” Phys. Rev. A, vol. 101, no. 4, p. 043829, 2020. https://doi.org/10.1103/physreva.101.043829.Search in Google Scholar
[56] M. W. Beijersbergen, R. Coerwinkel, M. Kristensen, and J. Woerdman, “Helical-wavefront laser beams produced with a spiral phaseplate,” Opt. Commun., vol. 112, nos. 5–6, pp. 321–327, 1994. https://doi.org/10.1016/0030-4018(94)90638-6.Search in Google Scholar
[57] L. Marrucci, C. Manzo, and D. Paparo, “Optical spin-to-orbital angular momentum conversion in inhomogeneous anisotropic media,” Phys. Rev. Lett., vol. 96, no. 16, p. 163905, 2006. https://doi.org/10.1103/physrevlett.96.163905.Search in Google Scholar PubMed
[58] L. Marrucci, C. Manzo, and D. Paparo, “Pancharatnam-Berry phase optical elements for wave front shaping in the visible domain: switchable helical mode generation,” Appl. Phys. Lett., vol. 88, no. 22, p. 221102, 2006. https://doi.org/10.1063/1.2207993.Search in Google Scholar
[59] L. Marrucci, “The q-plate and its future,” J. Nanophotonics, vol. 7, no. 1, p. 078598, 2013. https://doi.org/10.1117/1.jnp.7.078598.Search in Google Scholar
[60] C. Maurer, A. Jesacher, S. Bernet, and M. Ritsch-Marte, “What spatial light modulators can do for optical microscopy,” Laser Photon. Rev., vol. 5, no. 1, pp. 81–101, 2011. https://doi.org/10.1002/lpor.200900047.Search in Google Scholar
[61] C. Rosales-Guzmán and A. Forbes, How to Shape Light with Spatial Light Modulators, Bellingham, Washington, Society of Photo-Optical Instrumentation Engineers (SPIE), 2017.10.1117/3.2281295Search in Google Scholar
[62] J. Pinnell, I. Nape, B. Sephton, M. A. Cox, V. Rodriguez-Fajardo, and A. Forbes, “Modal analysis of structured light with spatial light modulators: a practical tutorial,” J. Opt. Soc. Am. A, vol. 37, no. 11, pp. C146–C160, 2020. https://doi.org/10.1364/josaa.398712.Search in Google Scholar PubMed
[63] J. Ni, C. Huang, L. M. Zhou, et al.., “Multidimensional phase singularities in nanophotonics,” Science, vol. 374, no. 6566, p. eabj0039, 2021. https://doi.org/10.1126/science.abj0039.Search in Google Scholar PubMed
[64] A. H. Dorrah and C. Federico, “Tunable structured light with flat optics,” Science, vol. 376, no. 6591, p. eabi6860, 2022. https://doi.org/10.1126/science.abi6860.Search in Google Scholar PubMed
[65] S. Kruk, B. Hopkins, I. I. Kravchenko, A. Miroshnichenko, D. N. Neshev, and Y. S. Kivshar, “Invited Article: broadband highly efficient dielectric metadevices for polarization control,” APL Photonics, vol. 1, no. 3, p. 030801, 2016. https://doi.org/10.1063/1.4949007.Search in Google Scholar
[66] X. Bai, F. Kong, Y. Sun, et al.., “High‐efficiency transmissive programmable metasurface for multimode OAM generation,” Adv. Opt. Mater., vol. 8, no. 17, p. 2000570, 2020. https://doi.org/10.1002/adom.202000570.Search in Google Scholar
[67] R. C. Devlin, A. Ambrosio, N. A. Rubin, J. P. B. Mueller, and F. Capasso, “Arbitrary spin-to–orbital angular momentum conversion of light,” Science, vol. 358, no. 6365, pp. 896–901, 2017. https://doi.org/10.1126/science.aao5392.Search in Google Scholar PubMed
[68] K. E. Chong, I. Staude, A. James, et al.., “Polarization-independent silicon metadevices for efficient optical wavefront control,” Nano Lett., vol. 15, no. 8, pp. 5369–5374, 2015. https://doi.org/10.1021/acs.nanolett.5b01752.Search in Google Scholar PubMed
[69] M. I. Shalaev, J. Sun, A. Tsukernik, A. Pandey, K. Nikolskiy, and N. M. Litchinitser, “High-efficiency all-dielectric metasurfaces for ultracompact beam manipulation in transmission mode,” Nano Lett., vol. 15, no. 9, pp. 6261–6266, 2015. https://doi.org/10.1021/acs.nanolett.5b02926.Search in Google Scholar PubMed
[70] G. J. Gbur, Singular Optics, Boca Raton, Florida, CRC Press, 2016.10.1201/9781315374260Search in Google Scholar
[71] B. E. A. Saleh and M. Carl Teich, Fundamentals of Photonics, Hoboken, New Jersey, John Wiley & Sons, 2019.Search in Google Scholar
[72] G. J. Gbur, Mathematical Methods for Optical Physics and Engineering, Cambridge, UK, Cambridge University Press, 2011.10.1017/CBO9780511777677Search in Google Scholar
[73] I. Kimel and L. R. Elias, “Relations between Hermite and Laguerre Gaussian modes,” IEEE J. Quant. Electron., vol. 29, no. 9, pp. 2562–2567, 1993. https://doi.org/10.1109/3.247715.Search in Google Scholar
[74] M. A. Bandres and J. C. Gutiérrez-Vega, “Ince–Gaussian modes of the paraxial wave equation and stable resonators,” J. Opt. Soc. Am. A, vol. 21, no. 5, pp. 873–880, 2004. https://doi.org/10.1364/josaa.21.000873.Search in Google Scholar PubMed
[75] Y. Shen, Z. Wang, X. Fu, D. Naidoo, and A. Forbes, “SU (2) Poincaré sphere: a generalized representation for multidimensional structured light,” Phys. Rev. A, vol. 102, no. 3, p. 031501, 2020. https://doi.org/10.1103/physreva.102.031501.Search in Google Scholar
[76] P. Senthilkumaran, “Polarization singularities and intensity degeneracies,” Front. Phys., vol. 8, p. 140, 2020.10.3389/fphy.2020.00140Search in Google Scholar
[77] M. A. Bandres and J. C. Gutiérrez-Vega, “Ince–Gaussian beams,” Opt. Lett., vol. 29, no. 2, pp. 144–146, 2004. https://doi.org/10.1364/ol.29.000144.Search in Google Scholar PubMed
[78] J. A. Davis, J. B. Bentley, M. A. Bandres, and J. C. Gutiérrez-Vega, “Generation of helical Ince-Gaussian beams: beam-shaping with a liquid crystal display,” in Laser Beam Shaping VII, vol. 6290, San Diego, California, SPIE, 2006.10.1117/12.679533Search in Google Scholar
[79] J. N. Brittingham, “Focus waves modes in homogeneous Maxwell’s equations: transverse electric mode,” J. Appl. Phys., vol. 54, no. 3, pp. 1179–1189, 1983. https://doi.org/10.1063/1.332196.Search in Google Scholar
[80] R. W. Ziolkowski, “Exact solutions of the wave equation with complex source locations,” J. Math. Phys., vol. 26, no. 4, pp. 861–863, 1985. https://doi.org/10.1063/1.526579.Search in Google Scholar
[81] R. W. Ziolkowski, “Localized transmission of electromagnetic energy,” Phys. Rev. A, vol. 39, no. 4, p. 2005, 1989. https://doi.org/10.1103/physreva.39.2005.Search in Google Scholar PubMed
[82] J. Lekner, “Helical light pulses,” J. Opt. A: Pure Appl. Opt., vol. 6, no. 10, p. L29, 2004. https://doi.org/10.1088/1464-4258/6/10/l01.Search in Google Scholar
[83] S. Feng, H. G. Winful, and R. W. Hellwarth, “Gouy shift and temporal reshaping of focused single-cycle electromagnetic pulses,” Opt. Lett., vol. 23, no. 5, pp. 385–387, 1998. https://doi.org/10.1364/ol.23.000385.Search in Google Scholar PubMed
[84] S. Feng, H. G. Winful, and R. W. Hellwarth, “Spatiotemporal evolution of focused single-cycle electromagnetic pulses,” Phys. Rev. E, vol. 59, no. 4, p. 4630, 1999. https://doi.org/10.1103/physreve.59.4630.Search in Google Scholar
[85] N. Papasimakis, T. Raybould, V. A. Fedotov, D. P. Tsai, I. Youngs, and N. I. Zheludev, “Pulse generation scheme for flying electromagnetic doughnuts,” Phys. Rev. B, vol. 97, no. 20, p. 201409, 2018. https://doi.org/10.1103/physrevb.97.201409.Search in Google Scholar
[86] V. Savinov, N. Papasimakis, D. P. Tsai, and N. I. Zheludev, “Optical anapoles,” Commun. Phys., vol. 2, no. 1, pp. 1–4, 2019. https://doi.org/10.1038/s42005-019-0167-z.Search in Google Scholar
[87] S. W. Hancock, S. Zahedpour, A. Goffin, and H. M. Milchberg, “Free-space propagation of spatiotemporal optical vortices,” Optica, vol. 6, no. 12, pp. 1547–1553, 2019. https://doi.org/10.1364/optica.6.001547.Search in Google Scholar
[88] C. F. Bohren, D. R. Huffman, and Z. Kam, “Book-review-absorption and scattering of light by small particles,” Nature, vol. 306, no. 5943, p. 625, 1983.Search in Google Scholar
[89] J. D. Jackson, Classical Electrodynamics, New York, John Wiley & Sons, 1999, p. 13.Search in Google Scholar
[90] A. B. Evlyukhin and B. N. Chichkov, “Multipole decompositions for directional light scattering,” Phys. Rev. B, vol. 100, no. 12, p. 125415, 2019. https://doi.org/10.1103/physrevb.100.125415.Search in Google Scholar
[91] A. B. Evlyukhin, C. Reinhardt, E. Evlyukhin, and B. N. Chichkov, “Multipole analysis of light scattering by arbitrary-shaped nanoparticles on a plane surface,” J. Opt. Soc. Am. B, vol. 30, no. 10, pp. 2589–2598, 2013. https://doi.org/10.1364/josab.30.002589.Search in Google Scholar
[92] A. B. Evlyukhin, T. Fischer, C. Reinhardt, and B. N. Chichkov, “Optical theorem and multipole scattering of light by arbitrarily shaped nanoparticles,” Phys. Rev. B, vol. 94, no. 20, p. 205434, 2016. https://doi.org/10.1103/physrevb.94.205434.Search in Google Scholar
[93] R. Alaee, C. Rockstuhl, and I. Fernandez-Corbaton, “An electromagnetic multipole expansion beyond the long-wavelength approximation,” Opt. Commun., vol. 407, pp. 17–21, 2018. https://doi.org/10.1016/j.optcom.2017.08.064.Search in Google Scholar
[94] R. Alaee, C. Rockstuhl, and I. Fernandez‐Corbaton, “Exact multipolar decompositions with applications in nanophotonics,” Adv. Opt. Mater., vol. 7, no. 1, p. 1800783, 2019. https://doi.org/10.1002/adom.201800783.Search in Google Scholar
[95] J. Xu, Y. Wu, P. Zhang, et al.., “Resonant scattering manipulation of dielectric nanoparticles,” Adv. Opt. Mater., vol. 9, no. 15, p. 2100112, 2021. https://doi.org/10.1002/adom.202100112.Search in Google Scholar
[96] P. Tonkaev and Y. Kivshar, “All-dielectric resonant metaphotonics: opinion,” Opt. Mater. Express, vol. 12, no. 7, pp. 2879–2885, 2022. https://doi.org/10.1364/ome.467655.Search in Google Scholar
[97] P. Tonkaev and Y. Kivshar, “High-Q dielectric mie-resonant nanostructures (brief review),” JETP Lett., vol. 112, no. 10, pp. 615–622, 2020. https://doi.org/10.1134/s0021364020220038.Search in Google Scholar
[98] N. Meinzer, W. L. Barnes, and I. R. Hooper, “Plasmonic meta-atoms and metasurfaces,” Nat. Photonics, vol. 8, no. 12, pp. 889–898, 2014. https://doi.org/10.1038/nphoton.2014.247.Search in Google Scholar
[99] K. Yao and Y. Liu, “Plasmonic metamaterials,” Nanotechnol. Rev., vol. 3, no. 2, pp. 177–210, 2014. https://doi.org/10.1515/ntrev-2012-0071.Search in Google Scholar
[100] P. Genevet, F. Capasso, F. Aieta, M. Khorasaninejad, and R. Devlin, “Recent advances in planar optics: from plasmonic to dielectric metasurfaces,” Optica, vol. 4, no. 1, pp. 139–152, 2017. https://doi.org/10.1364/optica.4.000139.Search in Google Scholar
[101] S. Jahani and Z. Jacob, “All-dielectric metamaterials,” Nat. Nanotechnol., vol. 11, no. 1, pp. 23–36, 2016. https://doi.org/10.1038/nnano.2015.304.Search in Google Scholar PubMed
[102] A. I. Kuznetsov, A. E. Miroshnichenko, Y. H. Fu, J. Zhang, and B. Luk’yanchuk, “Magnetic light,” Sci. Rep., vol. 2, no. 1, pp. 1–6, 2012. https://doi.org/10.1038/srep00492.Search in Google Scholar PubMed PubMed Central
[103] I. Staude, A. E. Miroshnichenko, M. Decker, et al.., “Tailoring directional scattering through magnetic and electric resonances in subwavelength silicon nanodisks,” ACS Nano, vol. 7, no. 9, pp. 7824–7832, 2013. https://doi.org/10.1021/nn402736f.Search in Google Scholar PubMed
[104] Y. F. Yu, A. Y. Zhu, R. Paniagua-Dominguez, Y. H. Fu, B. Luk’yanchuk, and A. I. Kuznetsov, “High‐transmission dielectric metasurface with 2π phase control at visible wavelengths,” Laser Photon. Rev., vol. 9, no. 4, pp. 412–418, 2015. https://doi.org/10.1002/lpor.201500041.Search in Google Scholar
[105] A. E. Krasnok, A. E. Miroshnichenko, P. A. Belov, and Y. S. Kivshar, “All-dielectric optical nanoantennas,” Opt. Express, vol. 20, no. 18, pp. 20599–20604, 2012. https://doi.org/10.1364/oe.20.020599.Search in Google Scholar
[106] A. B. Evlyukhin, S. M. Novikov, U. Zywietz, et al.., “Demonstration of magnetic dipole resonances of dielectric nanospheres in the visible region,” Nano Lett., vol. 12, no. 7, pp. 3749–3755, 2012. https://doi.org/10.1021/nl301594s.Search in Google Scholar PubMed
[107] L. Shi, T. U. Tuzer, R. Fenollosa, and F. Meseguer, “A new dielectric metamaterial building block with a strong magnetic response in the sub‐1.5‐micrometer region: silicon colloid nanocavities,” Adv. Mater., vol. 24, no. 44, pp. 5934–5938, 2012. https://doi.org/10.1002/adma.201201987.Search in Google Scholar PubMed
[108] J.-M. Geffrin, B. Garcia-Camara, R. Gomez-Medina, et al.., “Magnetic and electric coherence in forward-and back-scattered electromagnetic waves by a single dielectric subwavelength sphere,” Nat. Commun., vol. 3, no. 1, pp. 1–8, 2012. https://doi.org/10.1038/ncomms2167.Search in Google Scholar PubMed
[109] S. Person, M. Jain, Z. Lapin, J. J. Saenz, G. Wicks, and L. Novotny, “Demonstration of zero optical backscattering from single nanoparticles,” Nano Lett., vol. 13, no. 4, pp. 1806–1809, 2013. https://doi.org/10.1021/nl4005018.Search in Google Scholar PubMed
[110] S. T. Ha, Y. H. Fu, N. K. Emani, et al.., “Directional lasing in resonant semiconductor nanoantenna arrays,” Nat. Nanotechnol., vol. 13, no. 11, pp. 1042–1047, 2018. https://doi.org/10.1038/s41565-018-0245-5.Search in Google Scholar PubMed
[111] R. M. Bakker, D. Permyakov, Y. F. Yu, et al.., “Magnetic and electric hotspots with silicon nanodimers,” Nano Lett., vol. 15, no. 3, pp. 2137–2142, 2015. https://doi.org/10.1021/acs.nanolett.5b00128.Search in Google Scholar PubMed
[112] W. Liu and Y. S. Kivshar, “Generalized Kerker effects in nanophotonics and meta-optics,” Opt. Express, vol. 26, no. 10, pp. 13085–13105, 2018. https://doi.org/10.1364/oe.26.013085.Search in Google Scholar PubMed
[113] P. Albella, M. A. Poyli, M. K. Schmidt, et al.., “Low-loss electric and magnetic field-enhanced spectroscopy with subwavelength silicon dimers,” J. Phys. Chem. C, vol. 117, no. 26, pp. 13573–13584, 2013. https://doi.org/10.1021/jp4027018.Search in Google Scholar
[114] M. Hentschel, K. Koshelev, F. Sterl, et al.., “Dielectric Mie voids: confining light in air,” Light Sci. Appl., vol. 12, no. 1, pp. 1–12, 2023. https://doi.org/10.1038/s41377-022-01015-z.Search in Google Scholar PubMed PubMed Central
[115] Y. H. Fu, A. I. Kuznetsov, A. E. Miroshnichenko, Y. F. Yu, and B. Luk’yanchuk, “Directional visible light scattering by silicon nanoparticles,” Nat. Commun., vol. 4, no. 1, pp. 1–6, 2013. https://doi.org/10.1038/ncomms2538.Search in Google Scholar PubMed
[116] K. E. Chong, B. Hopkins, I. Staude, et al.., “Observation of Fano resonances in all‐dielectric nanoparticle oligomers,” Small, vol. 10, no. 10, pp. 1985–1990, 2014. https://doi.org/10.1002/smll.201303612.Search in Google Scholar PubMed
[117] A. E. Miroshnichenko, A. B. Evlyukhin, Y. F. Yu, et al.., “Nonradiating anapole modes in dielectric nanoparticles,” Nat. Commun., vol. 6, no. 1, pp. 1–8, 2015. https://doi.org/10.1038/ncomms9069.Search in Google Scholar PubMed PubMed Central
[118] G. Grinblat, Y. Li, M. P. Nielsen, R. F. Oulton, and S. A. Maier, “Enhanced third harmonic generation in single germanium nanodisks excited at the anapole mode,” Nano Lett., vol. 16, no. 7, pp. 4635–4640, 2016. https://doi.org/10.1021/acs.nanolett.6b01958.Search in Google Scholar PubMed
[119] V. A. Zenin, A. B. Evlyukhin, S. M. Novikov, et al.., “Direct amplitude-phase near-field observation of higher-order anapole states,” Nano Lett., vol. 17, no. 11, pp. 7152–7159, 2017. https://doi.org/10.1021/acs.nanolett.7b04200.Search in Google Scholar PubMed
[120] A. F. Cihan, S. Raza, P. G. Kik, and M. L. Brongersma, “Silicon Mie resonators for highly directional light emission from monolayer MoS2,” Nat. Photonics, vol. 12, no. 5, pp. 284–290, 2018. https://doi.org/10.1038/s41566-018-0155-y.Search in Google Scholar
[121] R. Verre, D. G. Baranov, B. Munkhbat, J. Cuadra, M. Kall, and T. Shegai, “Transition metal dichalcogenide nanodisks as high-index dielectric Mie nanoresonators,” Nat. Nanotechnol., vol. 14, no. 7, pp. 679–683, 2019. https://doi.org/10.1038/s41565-019-0442-x.Search in Google Scholar PubMed
[122] G. P. Zograf, D. Ryabov, V. Rutckaia, et al.., “Stimulated Raman scattering from Mie-resonant subwavelength nanoparticles,” Nano Lett., vol. 20, no. 8, pp. 5786–5791, 2020. https://doi.org/10.1021/acs.nanolett.0c01646.Search in Google Scholar PubMed
[123] P. Kepic, F. Ligmajer, M. Hrton, et al.., “Optically tunable Mie resonance VO2 nanoantennas for metasurfaces in the visible,” ACS Photonics, vol. 8, no. 4, pp. 1048–1057, 2021. https://doi.org/10.1021/acsphotonics.1c00222.Search in Google Scholar
[124] V. Asadchy, A. Lamprianidis, G. Ptitcyn, et al.., “Parametric mie resonances and directional amplification in time-modulated scatterers,” Phys. Rev. Appl., vol. 18, no. 5, p. 054065, 2022. https://doi.org/10.1103/physrevapplied.18.054065.Search in Google Scholar
[125] M. Decker, I. Staude, M. Falkner, et al.., “High‐efficiency dielectric Huygens’ surfaces,” Adv. Opt. Mater., vol. 3, no. 6, pp. 813–820, 2015. https://doi.org/10.1002/adom.201400584.Search in Google Scholar
[126] S. Kruk and Y. Kivshar, “Functional meta-optics and nanophotonics governed by Mie resonances,” ACS Photonics, vol. 4, no. 11, pp. 2638–2649, 2017. https://doi.org/10.1021/acsphotonics.7b01038.Search in Google Scholar
[127] J. Van de Groep and A. Polman, “Designing dielectric resonators on substrates: combining magnetic and electric resonances,” Opt. Express, vol. 21, no. 22, pp. 26285–26302, 2013. https://doi.org/10.1364/oe.21.026285.Search in Google Scholar
[128] C. Liu, L. Chen, T. Wu, et al.., “Characteristics of electric quadrupole and magnetic quadrupole coupling in a symmetric silicon structure,” New J. Phys., vol. 22, no. 2, p. 023018, 2020. https://doi.org/10.1088/1367-2630/ab6cde.Search in Google Scholar
[129] K. Baryshnikova, D. Gets, T. Liashenko, et al.., “Broadband antireflection with halide perovskite metasurfaces,” Laser Photon. Rev., vol. 14, no. 12, p. 2000338, 2020. https://doi.org/10.1002/lpor.202000338.Search in Google Scholar
[130] J. Yan, P. Liu, Z. Lin, et al.., “Directional Fano resonance in a silicon nanosphere dimer,” ACS Nano, vol. 9, no. 3, pp. 2968–2980, 2015. https://doi.org/10.1021/nn507148z.Search in Google Scholar PubMed
[131] B. Hopkins, A. N. Poddubny, A. E. Miroshnichenko, and Y. S. Kivshar, “Revisiting the physics of Fano resonances for nanoparticle oligomers,” Phys. Rev. A, vol. 88, no. 5, p. 053819, 2013. https://doi.org/10.1103/physreva.88.053819.Search in Google Scholar
[132] A. E. Miroshnichenko and Y. S. Kivshar, “Fano resonances in all-dielectric oligomers,” Nano Lett., vol. 12, no. 12, pp. 6459–6463, 2012. https://doi.org/10.1021/nl303927q.Search in Google Scholar PubMed
[133] H.-J. Hu, F. W. Zhang, J. Y. Chen, Q. Li, and L. J. Wu, “Fano resonances with a high figure of merit in silver oligomer systems,” Photon. Res., vol. 6, no. 3, pp. 204–213, 2018. https://doi.org/10.1364/prj.6.000204.Search in Google Scholar
[134] B. Hopkins, D. S. Filonov, A. E. Miroshnichenko, F. Monticone, A. Alu, and Y. S. Kivshar, “Interplay of magnetic responses in all-dielectric oligomers to realize magnetic Fano resonances,” ACS Photonics, vol. 2, no. 6, pp. 724–729, 2015. https://doi.org/10.1021/acsphotonics.5b00082.Search in Google Scholar
[135] D. S. Filonov, A. P. Slobozhanyuk, A. E. Krasnok, et al.., “Near-field mapping of Fano resonances in all-dielectric oligomers,” Appl. Phys. Lett., vol. 104, no. 2, p. 021104, 2014. https://doi.org/10.1063/1.4858969.Search in Google Scholar
[136] G. Zhang, C. Lan, R. Gao, Y. Wen, and J. Zhou, “Toroidal dipole resonances in all‐dielectric oligomer metasurfaces,” Adv. Theory Simul., vol. 2, no. 10, p. 1900123, 2019. https://doi.org/10.1002/adts.201900123.Search in Google Scholar
[137] Z.-J. Yang, Y. H. Deng, Y. Yu, and J. He, “Magnetic toroidal dipole response in individual all-dielectric nanodisk clusters,” Nanoscale, vol. 12, no. 19, pp. 10639–10646, 2020. https://doi.org/10.1039/d0nr01440k.Search in Google Scholar PubMed
[138] A. A. Basharin, M. Kafesaki, E. N. Economou, et al.., “Dielectric metamaterials with toroidal dipolar response,” Phys. Rev. X, vol. 5, no. 1, p. 011036, 2015. https://doi.org/10.1103/physrevx.5.011036.Search in Google Scholar
[139] V. R. Tuz, V. Dmitriev, and A. B. Evlyukhin, “Antitoroidic and toroidic orders in all-dielectric metasurfaces for optical near-field manipulation,” ACS Appl. Nano Mater., vol. 3, no. 11, pp. 11315–11325, 2020. https://doi.org/10.1021/acsanm.0c02421.Search in Google Scholar
[140] S. Xu, A. Sayanskiy, A. S. Kupriianov, et al.., “Experimental observation of toroidal dipole modes in all‐dielectric metasurfaces,” Adv. Opt. Mater., vol. 7, no. 4, p. 1801166, 2019. https://doi.org/10.1002/adom.201801166.Search in Google Scholar
[141] X. Liu, J. Li, Q. Zhang, and Y. Wang, “Dual-toroidal dipole excitation on permittivity-asymmetric dielectric metasurfaces,” Opt. Lett., vol. 45, no. 10, pp. 2826–2829, 2020. https://doi.org/10.1364/ol.387872.Search in Google Scholar
[142] G. N. Afanasiev and Y. P. Stepanovsky, “The electromagnetic field of elementary time-dependent toroidal sources,” J. Phys. A: Math. Gen., vol. 28, no. 16, p. 4565, 1995. https://doi.org/10.1088/0305-4470/28/16/014.Search in Google Scholar
[143] V. A. Fedotov, A. V. Rogacheva, V. Savinov, D. P. Tsai, and N. I. Zheludev, “Resonant transparency and non-trivial non-radiating excitations in toroidal metamaterials,” Sci. Rep., vol. 3, no. 1, pp. 1–5, 2013. https://doi.org/10.1038/srep02967.Search in Google Scholar PubMed PubMed Central
[144] D. G. Baranov, R. Verre, P. Karpinski, and M. Kall, “Anapole-enhanced intrinsic Raman scattering from silicon nanodisks,” ACS Photonics, vol. 5, no. 7, pp. 2730–2736, 2018. https://doi.org/10.1021/acsphotonics.8b00480.Search in Google Scholar
[145] Y. Yang, V. A. Zenin, and S. I. Bozhevolnyi, “Anapole-assisted strong field enhancement in individual all-dielectric nanostructures,” ACS Photonics, vol. 5, no. 5, pp. 1960–1966, 2018. https://doi.org/10.1021/acsphotonics.7b01440.Search in Google Scholar
[146] A. A. Basharin, V. Chuguevsky, N. Volsky, M. Kafesaki, and E. N. Economou, “Extremely high Q-factor metamaterials due to anapole excitation,” Phys. Rev. B, vol. 95, no. 3, p. 035104, 2017. https://doi.org/10.1103/physrevb.95.035104.Search in Google Scholar
[147] P. C. Wu, C. Y. Liao, V. Savinov, et al.., “Optical anapole metamaterial,” ACS Nano, vol. 12, no. 2, pp. 1920–1927, 2018. https://doi.org/10.1021/acsnano.7b08828.Search in Google Scholar PubMed
[148] A. K. Ospanova, I. V. Stenishchev, and A. A. Basharin, “Anapole mode sustaining silicon metamaterials in visible spectral range,” Laser Photon. Rev., vol. 12, no. 7, p. 1800005, 2018. https://doi.org/10.1002/lpor.201800005.Search in Google Scholar
[149] G. Grinblat, Y. Li, M. P. Nielsen, R. F. Oulton, and S. A. Maier, “Efficient third harmonic generation and nonlinear subwavelength imaging at a higher-order anapole mode in a single germanium nanodisk,” ACS Nano, vol. 11, no. 1, pp. 953–960, 2017. https://doi.org/10.1021/acsnano.6b07568.Search in Google Scholar PubMed
[150] T. Shibanuma, G. Grinblat, P. Albella, and S. A. Maier, “Efficient third harmonic generation from metal–dielectric hybrid nanoantennas,” Nano Lett., vol. 17, no. 4, pp. 2647–2651, 2017. https://doi.org/10.1021/acs.nanolett.7b00462.Search in Google Scholar PubMed
[151] L. Xu, M. Rahmani, K. Zangeneh Kamali, et al.., “Boosting third-harmonic generation by a mirror-enhanced anapole resonator,” Light Sci. Appl., vol. 7, no. 1, pp. 1–8, 2018. https://doi.org/10.1038/s41377-018-0051-8.Search in Google Scholar PubMed PubMed Central
[152] W. Liu, B. Lei, J. Shi, H. Hu, and A. E. Miroshnichenko, “Elusive pure anapole excitation in homogenous spherical nanoparticles with radial anisotropy,” J. Nanomater., vol. 2015, p. 672957, 2015. https://doi.org/10.1155/2015/672957.Search in Google Scholar
[153] B. Luk’yanchuk, R. Paniagua-Dominguez, A. I. Kuznetsov, A. E. Miroshnichenko, and Y. S. Kivshar, “Hybrid anapole modes of high-index dielectric nanoparticles,” Phys. Rev. A, vol. 95, no. 6, p. 063820, 2017. https://doi.org/10.1103/physreva.95.063820.Search in Google Scholar
[154] S.-Q. Li and K. B. Crozier, “Origin of the anapole condition as revealed by a simple expansion beyond the toroidal multipole,” Phys. Rev. B, vol. 97, no. 24, p. 245423, 2018. https://doi.org/10.1103/physrevb.97.245423.Search in Google Scholar
[155] J. S. Totero Gongora, A. E. Miroshnichenko, Y. S. Kivshar, and A. Fratalocchi, “Anapole nanolasers for mode-locking and ultrafast pulse generation,” Nat. Commun., vol. 8, no. 1, pp. 1–9, 2017.10.1038/ncomms15535Search in Google Scholar PubMed PubMed Central
[156] B. Luk’yanchuk, R. Paniagua-Dominguez, A. I. Kuznetsov, A. E. Miroshnichenko, and Y. S. Kivshar, “Suppression of scattering for small dielectric particles: anapole mode and invisibility,” Philos. Trans. R. Soc., A, vol. 375, no. 2090, p. 20160069, 2017. https://doi.org/10.1098/rsta.2016.0069.Search in Google Scholar PubMed PubMed Central
[157] J. Tian, H. Luo, Y. Yang, et al.., “Active control of anapole states by structuring the phase-change alloy Ge2Sb2Te5,” Nat. Commun., vol. 10, no. 1, pp. 1–9, 2019. https://doi.org/10.1038/s41467-018-08057-1.Search in Google Scholar PubMed PubMed Central
[158] M. Timofeeva, L. Lang, F. Timpu, et al.., “Anapoles in free-standing III–V nanodisks enhancing second-harmonic generation,” Nano Lett., vol. 18, no. 6, pp. 3695–3702, 2018. https://doi.org/10.1021/acs.nanolett.8b00830.Search in Google Scholar PubMed
[159] E. Zanganeh, A. Evlyukhin, A. Miroshnichenko, M. Song, E. Nenasheva, and P. Kapitanova, “Anapole meta-atoms: nonradiating electric and magnetic sources,” Phys. Rev. Lett., vol. 127, no. 9, p. 096804, 2021. https://doi.org/10.1103/physrevlett.127.096804.Search in Google Scholar PubMed
[160] K. V. Baryshnikova, D. A. Smirnova, B. S. Luk’yanchuk, and Y. S. Kivshar, “Optical anapoles: concepts and applications,” Adv. Opt. Mater., vol. 7, no. 14, p. 1801350, 2019. https://doi.org/10.1002/adom.201801350.Search in Google Scholar
[161] Y. Yang and S. I. Bozhevolnyi, “Nonradiating anapole states in nanophotonics: from fundamentals to applications,” Nanotechnology, vol. 30, no. 20, p. 204001, 2019. https://doi.org/10.1088/1361-6528/ab02b0.Search in Google Scholar PubMed
[162] R. M. Saadabad, L. Huang, A. B. Evlyukhin, and A. E. Miroshnichenko, “Multifaceted anapole: from physics to applications,” Opt. Mater. Express, vol. 12, no. 5, pp. 1817–1837, 2022.10.1364/OME.456070Search in Google Scholar
[163] D. G. Baranov, D. A. Zuev, S. I. Lepeshov, et al.., “All-dielectric nanophotonics: the quest for better materials and fabrication techniques,” Optica, vol. 4, no. 7, pp. 814–825, 2017. https://doi.org/10.1364/optica.4.000814.Search in Google Scholar
[164] S. Manzeli, D. Ovchinnikov, D. Pasquier, O. V. Yazyev, and A. Kis, “2D transition metal dichalcogenides,” Nat. Rev. Mater., vol. 2, no. 8, pp. 1–15, 2017. https://doi.org/10.1038/natrevmats.2017.33.Search in Google Scholar
[165] W. Choi, N. Choudhary, G. H. Han, J. Park, D. Akinwande, and Y. H. Lee, “Recent development of two-dimensional transition metal dichalcogenides and their applications,” Mater. Today, vol. 20, no. 3, pp. 116–130, 2017. https://doi.org/10.1016/j.mattod.2016.10.002.Search in Google Scholar
[166] K. F. Mak and J. Shan, “Photonics and optoelectronics of 2D semiconductor transition metal dichalcogenides,” Nat. Photonics, vol. 10, no. 4, pp. 216–226, 2016. https://doi.org/10.1038/nphoton.2015.282.Search in Google Scholar
[167] A. K. Geim and I. V. Grigorieva, “Van der Waals heterostructures,” Nature, vol. 499, no. 7459, pp. 419–425, 2013. https://doi.org/10.1038/nature12385.Search in Google Scholar PubMed
[168] Y. Zhang, T. R. Chang, B. Zhou, et al.., “Direct observation of the transition from indirect to direct bandgap in atomically thin epitaxial MoSe2,” Nat. Nanotechnol., vol. 9, no. 2, pp. 111–115, 2014. https://doi.org/10.1038/nnano.2013.277.Search in Google Scholar PubMed
[169] Y.‐H. Lee, X. Q. Zhang, W. Zhang, et al.., “Synthesis of large‐area MoS2 atomic layers with chemical vapor deposition,” Adv. Mater., vol. 24, no. 17, pp. 2320–2325, 2012. https://doi.org/10.1002/adma.201104798.Search in Google Scholar PubMed
[170] R. Mupparapu, T. Bucher, and I. Staude, “Integration of two-dimensional transition metal dichalcogenides with Mie-resonant dielectric nanostructures,” Adv. Phys. X, vol. 5, no. 1, p. 1734083, 2020. https://doi.org/10.1080/23746149.2020.1734083.Search in Google Scholar
[171] M. L. Brongersma, “The road to atomically thin metasurface optics,” Nanophotonics, vol. 10, no. 1, pp. 643–654, 2021. https://doi.org/10.1515/nanoph-2020-0444.Search in Google Scholar
[172] F. Hu, Y. Luan, M. E. Scott, et al.., “Imaging exciton–polariton transport in MoSe2 waveguides,” Nat. Photonics, vol. 11, no. 6, pp. 356–360, 2017. https://doi.org/10.1038/nphoton.2017.65.Search in Google Scholar
[173] J. A. Wilson and A. D. Yoffe, “The transition metal dichalcogenides discussion and interpretation of the observed optical, electrical and structural properties,” Adv. Phys., vol. 18, no. 73, pp. 193–335, 1969. https://doi.org/10.1080/00018736900101307.Search in Google Scholar
[174] I. Alessandri and J. R. Lombardi, “Enhanced Raman scattering with dielectrics,” Chem. Rev., vol. 116, no. 24, pp. 14921–14981, 2016. https://doi.org/10.1021/acs.chemrev.6b00365.Search in Google Scholar PubMed
[175] T. Cui, B. Bai, and H.‐B. Sun, “Tunable metasurfaces based on active materials,” Adv. Funct. Mater., vol. 29, no. 10, p. 1806692, 2019. https://doi.org/10.1002/adfm.201806692.Search in Google Scholar
[176] L. Kang, R. P. Jenkins, and D. H. Werner, “Recent progress in active optical metasurfaces,” Adv. Opt. Mater., vol. 7, no. 14, p. 1801813, 2019. https://doi.org/10.1002/adom.201801813.Search in Google Scholar
[177] Q. He, S. Sun, and L. Zhou, “Tunable/reconfigurable metasurfaces: physics and applications,” Research, vol. 2019, p. 1849272, 2019. https://doi.org/10.34133/2019/1849272.Search in Google Scholar PubMed PubMed Central
[178] A. M. Shaltout, V. M. Shalaev, and M. L. Brongersma, “Spatiotemporal light control with active metasurfaces,” Science, vol. 364, no. 6441, p. eaat3100, 2019. https://doi.org/10.1126/science.aat3100.Search in Google Scholar PubMed
[179] G. P. Zograf, M. I. Petrov, S. V. Makarov, and Y. S. Kivshar, “All-dielectric thermonanophotonics,” Adv. Opt. Photonics, vol. 13, no. 3, pp. 643–702, 2021. https://doi.org/10.1364/aop.426047.Search in Google Scholar
[180] T. Lewi, N. A. Butakov, H. A. Evans, et al.., “Thermally reconfigurable meta-optics,” IEEE Photon. J., vol. 11, no. 2, pp. 1–16, 2019. https://doi.org/10.1109/jphot.2019.2916161.Search in Google Scholar
[181] T. Brunet, K. Zimny, B. Mascaro, et al.., “Tuning Mie scattering resonances in soft materials with magnetic fields,” Phys. Rev. Lett., vol. 111, no. 26, p. 264301, 2013. https://doi.org/10.1103/physrevlett.111.264301.Search in Google Scholar
[182] F. Laible, A. Horneber, and M. Fleischer, “Mechanically tunable nanogap antennas: single‐structure effects and multi‐structure applications,” Adv. Opt. Mater., vol. 9, no. 20, p. 2100326, 2021. https://doi.org/10.1002/adom.202100326.Search in Google Scholar
[183] J. Gao, M. A. Vincenti, J. Frantz, et al.., “All-optical tunable wavelength conversion in opaque nonlinear nanostructures,” Nanophotonics, vol. 11, pp. 4027–4035, 2022. https://doi.org/10.1515/nanoph-2022-0078.Search in Google Scholar
[184] Z. Zhang, X. Qiao, B. Midya, et al.., “Tunable topological charge vortex microlaser,” Science, vol. 368, no. 6492, pp. 760–763, 2020. https://doi.org/10.1126/science.aba8996.Search in Google Scholar PubMed
[185] M. I. Shalaev, W. Walasik, and N. M. Litchinitser, “Optically tunable topological photonic crystal,” Optica, vol. 6, no. 7, pp. 839–844, 2019. https://doi.org/10.1364/optica.6.000839.Search in Google Scholar
[186] Y. Xu, J. Sun, J. Frantz, et al.., “Nonlinear metasurface for structured light with tunable orbital angular momentum,” Appl. Sci., vol. 9, no. 5, p. 958, 2019. https://doi.org/10.3390/app9050958.Search in Google Scholar
[187] T. Kang, B. Fan, J. Qin, et al.., “Mid-infrared active metasurface based on Si/VO 2 hybrid meta-atoms,” Photon. Res., vol. 10, no. 2, pp. 373–380, 2022. https://doi.org/10.1364/prj.445571.Search in Google Scholar
[188] M. M. Salary and H. Mosallaei, “Tunable all-dielectric metasurfaces for phase-only modulation of transmitted light based on quasi-bound states in the continuum,” ACS Photonics, vol. 7, no. 7, pp. 1813–1829, 2020. https://doi.org/10.1021/acsphotonics.0c00554.Search in Google Scholar
[189] S. Jafar‐Zanjani, M. M. Salary, D. Huynh, E. Elhamifar, and H. Mosallaei, “TCO‐based active dielectric metasurfaces design by conditional generative adversarial networks,” Adv. Theory Simul., vol. 4, no. 2, p. 2000196, 2021. https://doi.org/10.1002/adts.202000196.Search in Google Scholar
[190] C. Caloz and Z.-L. Deck-Léger, “Spacetime metamaterials—part I: general concepts,” IEEE Trans. Antenn. Propag., vol. 68, no. 3, pp. 1569–1582, 2019. https://doi.org/10.1109/tap.2019.2944225.Search in Google Scholar
[191] C. Caloz and Z.-L. Deck-Leger, “Spacetime metamaterials—part II: theory and applications,” IEEE Trans. Antenn. Propag., vol. 68, no. 3, pp. 1583–1598, 2019. https://doi.org/10.1109/tap.2019.2944216.Search in Google Scholar
[192] C. Caloz, Z. L. Deck-Léger, A. Bahrami, O. S. Vicente, and Z. Li, “Generalized Space-Time Engineered Modulation (GSTEM) Metamaterials: A global and extended perspective,” in IEEE Antennas and Propagation Magazine, 2022, https://doi.org/10.1109/MAP.2022.3216773.10.1109/MAP.2022.3216773Search in Google Scholar
[193] E. Galiffi, R. Tirole, S. Yin, et al.., “Photonics of time-varying media,” Adv. Photon., vol. 4, no. 1, p. 014002, 2022. https://doi.org/10.1117/1.ap.4.1.014002.Search in Google Scholar
[194] S. Taravati and G. V. Eleftheriades, “Microwave space-time-modulated metasurfaces,” ACS Photonics, vol. 9, no. 2, pp. 305–318, 2022. https://doi.org/10.1021/acsphotonics.1c01041.Search in Google Scholar
[195] S. Taravati and A. A. Kishk, “Space-time modulation: principles and applications,” IEEE Microw. Mag., vol. 21, no. 4, pp. 30–56, 2020. https://doi.org/10.1109/mmm.2019.2963606.Search in Google Scholar
[196] G. Ptitcyn, M. S. Mirmoosa, and S. A. Tretyakov, “Time-modulated meta-atoms,” Phys. Rev. Res., vol. 1, no. 2, p. 023014, 2019. https://doi.org/10.1103/physrevresearch.1.023014.Search in Google Scholar
[197] V. S. Asadchy, M. S. Mirmoosa, A. Diaz-Rubio, S. Fan, and S. A. Tretyakov, “Tutorial on electromagnetic nonreciprocity and its origins,” Proc. IEEE, vol. 108, no. 10, pp. 1684–1727, 2020. https://doi.org/10.1109/jproc.2020.3012381.Search in Google Scholar
[198] M. S. Mirmoosa, T. T. Koutserimpas, G. A. Ptitcyn, S. A. Tretyakov, and R. Fleury, “Dipole polarizability of time-varying particles,” New J. Phys., vol. 24, p. 063004, 2022. https://doi.org/10.1088/1367-2630/ac6b4c.Search in Google Scholar
[199] H. B. Sedeh, M. M. Salary, and H. Mosallaei, “Time-varying optical vortices enabled by time-modulated metasurfaces,” Nanophotonics, vol. 9, no. 9, pp. 2957–2976, 2020. https://doi.org/10.1515/nanoph-2020-0202.Search in Google Scholar
[200] H. Barati Sedeh, M. M. Salary, and H. Mosallaei, “Topological space‐time photonic transitions in angular‐momentum‐biased metasurfaces,” Adv. Opt. Mater., vol. 8, no. 11, p. 2000075, 2020. https://doi.org/10.1002/adom.202000075.Search in Google Scholar
[201] V. Pacheco-Peña and N. Engheta, “Spatiotemporal isotropic-to-anisotropic meta-atoms,” New J. Phys., vol. 23, no. 9, p. 095006, 2021. https://doi.org/10.1088/1367-2630/ac21df.Search in Google Scholar
[202] A. Shaltout, A. Kildishev, and V. Shalaev, “Time-varying metasurfaces and Lorentz non-reciprocity,” Opt. Mater. Express, vol. 5, no. 11, pp. 2459–2467, 2015. https://doi.org/10.1364/ome.5.002459.Search in Google Scholar
[203] X. Guo, Y. Ding, Y. Duan, and X. Ni, “Nonreciprocal metasurface with space–time phase modulation,” Light Sci. Appl., vol. 8, no. 1, pp. 1–9, 2019. https://doi.org/10.1038/s41377-019-0225-z.Search in Google Scholar PubMed PubMed Central
[204] Y. Hadad, D. L. Sounas, and A. Alu, “Space-time gradient metasurfaces,” Phys. Rev. B, vol. 92, no. 10, p. 100304, 2015. https://doi.org/10.1103/physrevb.92.100304.Search in Google Scholar
[205] H. B. Sedeh, H. M. Dinani, and H. Mosallaei, “Optical nonreciprocity via transmissive time-modulated metasurfaces,” Nanophotonics, vol. 11, no. 17, pp. 4135–4148, 2022. https://doi.org/10.1515/nanoph-2022-0373.Search in Google Scholar
[206] M. M. Salary, S. Jafar-Zanjani, and H. Mosallaei, “Nonreciprocal optical links based on time-modulated nanoantenna arrays: full-duplex communication,” Phys. Rev. B, vol. 99, no. 4, p. 045416, 2019. https://doi.org/10.1103/physrevb.99.045416.Search in Google Scholar
[207] Y. Shi, S. Han, and S. Fan, “Optical circulation and isolation based on indirect photonic transitions of guided resonance modes,” ACS Photonics, vol. 4, no. 7, pp. 1639–1645, 2017. https://doi.org/10.1021/acsphotonics.7b00420.Search in Google Scholar
[208] S. Taravati and G. V. Eleftheriades, “Full-duplex nonreciprocal beam steering by time-modulated phase-gradient metasurfaces,” Phys. Rev. Appl., vol. 14, no. 1, p. 014027, 2020. https://doi.org/10.1103/physrevapplied.14.014027.Search in Google Scholar
[209] E. Galiffi, P. A. Huidobro, and J. B. Pendry, “Broadband nonreciprocal amplification in luminal metamaterials,” Phys. Rev. Lett., vol. 123, no. 20, p. 206101, 2019. https://doi.org/10.1103/physrevlett.123.206101.Search in Google Scholar PubMed
[210] H. Barati Sedeh, M. M. Salary, and H. Mosallaei, “Optical pulse compression assisted by high‐Q time‐modulated transmissive metasurfaces,” Laser Photon. Rev., vol. 16, p. 2100449, 2022. https://doi.org/10.1002/lpor.202100449.Search in Google Scholar
[211] P. Berini, “Optical beam steering using tunable metasurfaces,” ACS Photonics, vol. 9, no. 7, pp. 2204–2218, 2022. https://doi.org/10.1021/acsphotonics.2c00439.Search in Google Scholar
[212] L. Huang, S. Zhang, and T. Zentgraf, “Metasurface holography: from fundamentals to applications,” Nanophotonics, vol. 7, no. 6, pp. 1169–1190, 2018. https://doi.org/10.1515/nanoph-2017-0118.Search in Google Scholar
[213] G. Li, S. Zhang, and T. Zentgraf, “Nonlinear photonic metasurfaces,” Nat. Rev. Mater., vol. 2, no. 5, pp. 1–14, 2017. https://doi.org/10.1038/natrevmats.2017.10.Search in Google Scholar
[214] A. Krasnok, M. Tymchenko, and A. Alù, “Nonlinear metasurfaces: a paradigm shift in nonlinear optics,” Mater. Today, vol. 21, no. 1, pp. 8–21, 2018. https://doi.org/10.1016/j.mattod.2017.06.007.Search in Google Scholar
[215] M. A. Vincenti, J. Gao, D. de Ceglia, J. A. Frantz, M. Scalora, and N. M. Litchinitser, “Stacked chalcogenide metasurfaces for third harmonic generation in the UV range,” New J. Phys., vol. 24, no. 3, p. 035005, 2022. https://doi.org/10.1088/1367-2630/ac599c.Search in Google Scholar
[216] J. Gao, M. A. Vincenti, J. Frantz, et al.., “Near-infrared to ultra-violet frequency conversion in chalcogenide metasurfaces,” Nat. Commun., vol. 12, no. 1, pp. 1–5, 2021. https://doi.org/10.1038/s41467-021-26094-1.Search in Google Scholar PubMed PubMed Central
[217] H. K. Shamkhi, K. V. Baryshnikova, A. Sayanskiy, et al.., “Transverse scattering and generalized kerker effects in all-dielectric Mie-resonant metaoptics,” Phys. Rev. Lett., vol. 122, no. 19, p. 193905, 2019. https://doi.org/10.1103/physrevlett.122.193905.Search in Google Scholar PubMed
[218] H. K. Shamkhi, A. Sayanskiy, A. C. Valero, et al.., “Transparency and perfect absorption of all-dielectric resonant metasurfaces governed by the transverse Kerker effect,” Phys. Rev. Mater., vol. 3, no. 8, p. 085201, 2019. https://doi.org/10.1103/physrevmaterials.3.085201.Search in Google Scholar
[219] X. Ni, Z. J. Wong, M. Mrejen, Y. Wang, and X. Zhang, “An ultrathin invisibility skin cloak for visible light,” Science, vol. 349, no. 6254, pp. 1310–1314, 2015. https://doi.org/10.1126/science.aac9411.Search in Google Scholar PubMed
[220] V. A. Bushuev, D. M. Tsvetkov, V. V. Konotop, and B. I. Mantsyzov, “Unidirectional invisibility and enhanced reflection of short pulses in quasi-PT-symmetric media,” Opt. Lett., vol. 44, no. 23, pp. 5667–5670, 2019. https://doi.org/10.1364/ol.44.005667.Search in Google Scholar PubMed
[221] H. Chu, Q. Li, B. Liu, et al.., “A hybrid invisibility cloak based on integration of transparent metasurfaces and zero-index materials,” Light Sci. Appl., vol. 7, no. 1, pp. 1–8, 2018. https://doi.org/10.1038/s41377-018-0052-7.Search in Google Scholar PubMed PubMed Central
[222] M. M. Sadafi, H. Karami, and M. Hosseini, “A tunable hybrid graphene-metal metamaterial absorber for sensing in the THz regime,” Curr. Appl. Phys., vol. 31, pp. 132–140, 2021. https://doi.org/10.1016/j.cap.2021.07.020.Search in Google Scholar
[223] J. Gao, C. Lan, Q. Zhao, B. Li, and J. Zhou, “Experimental realization of Mie-resonance terahertz absorber by self-assembly method,” Opt. Express, vol. 26, no. 10, pp. 13001–13011, 2018. https://doi.org/10.1364/oe.26.013001.Search in Google Scholar
[224] J. Gao, C. Lan, Q. Zhao, B. Li, and J. Zhou, “Electrically controlled Mie-resonance absorber,” Opt. Express, vol. 25, no. 19, pp. 22658–22666, 2017. https://doi.org/10.1364/oe.25.022658.Search in Google Scholar PubMed
[225] C. Lan, D. Zhu, J. Gao, B. Li, and Z. Gao, “Flexible and tunable terahertz all-dielectric metasurface composed of ceramic spheres embedded in ferroelectric/elastomer composite,” Opt. Express, vol. 26, no. 9, pp. 11633–11638, 2018. https://doi.org/10.1364/oe.26.011633.Search in Google Scholar PubMed
[226] M. Taghavi and H. Mosallaei, “Increasing the stability margins using multi-pattern metasails and multi-modal laser beams,” Sci. Rep., vol. 12, no. 1, p. 20034, 2022. https://doi.org/10.1038/s41598-022-24681-w.Search in Google Scholar PubMed PubMed Central
[227] D. A. Kislov, E. A. Gurvitz, V. Bobrovs, et al.., “Multipole engineering of Attractive− repulsive and bending optical forces,” Adv. Photon. Res., vol. 2, no. 9, p. 2100082, 2021. https://doi.org/10.1002/adpr.202100082.Search in Google Scholar
[228] H. A. Atwater, A. R. Davoyan, O. Ilic, et al.., “Materials challenges for the Starshot lightsail,” Nat. Mater., vol. 17, no. 10, pp. 861–867, 2018. https://doi.org/10.1038/s41563-018-0075-8.Search in Google Scholar PubMed
[229] M. Taghavi, M. M. Salary, and H. Mosallaei, “Multifunctional metasails for self-stabilized beam-riding and optical communication,” Nanoscale Adv., vol. 4, no. 7, pp. 1727–1740, 2022. https://doi.org/10.1039/d1na00747e.Search in Google Scholar PubMed PubMed Central
[230] D. V. Zhirihin and Y. S. Kivshar, “Topological photonics on a small scale,” Small Sci., vol. 1, no. 12, p. 2100065, 2021. https://doi.org/10.1002/smsc.202170032.Search in Google Scholar
[231] V. Garbin, G. Volpe, E. Ferrari, M. Versluis, D. Cojoc, and D. Petrov, “Mie scattering distinguishes the topological charge of an optical vortex: a homage to Gustav Mie,” New J. Phys., vol. 11, no. 1, p. 013046, 2009. https://doi.org/10.1088/1367-2630/11/1/013046.Search in Google Scholar
[232] A. Belmonte, C. Rosales-Guzmán, and J. P. Torres, “Measurement of flow vorticity with helical beams of light,” Optica, vol. 2, no. 11, pp. 1002–1005, 2015. https://doi.org/10.1364/optica.2.001002.Search in Google Scholar
[233] S. Berg-Johansen, F. Toppel, B. Stiller, et al.., “Classically entangled optical beams for high-speed kinematic sensing,” Optica, vol. 2, no. 10, pp. 864–868, 2015. https://doi.org/10.1364/optica.2.000864.Search in Google Scholar
[234] Y. Zhao, J. S. Edgar, G. D. M. Jeffries, D. McGloin, and D. T. Chiu, “Spin-to-orbital angular momentum conversion in a strongly focused optical beam,” Phys. Rev. Lett., vol. 99, no. 7, p. 073901, 2007. https://doi.org/10.1103/physrevlett.99.073901.Search in Google Scholar PubMed
[235] Y. Iketaki, T. Watanabe, N. Bokor, and M. Fujii, “Investigation of the center intensity of first-and second-order Laguerre-Gaussian beams with linear and circular polarization,” Opt. Lett., vol. 32, no. 16, pp. 2357–2359, 2007. https://doi.org/10.1364/ol.32.002357.Search in Google Scholar PubMed
[236] N. M. Mojarad, V. Sandoghdar, and M. Agio, “Plasmon spectra of nanospheres under a tightly focused beam,” J. Opt. Soc. Am. B, vol. 25, no. 4, pp. 651–658, 2008. https://doi.org/10.1364/josab.25.000651.Search in Google Scholar
[237] D. Petrov, N. Rahuel, G. Molina-Terriza, and L. Torner, “Characterization of dielectric spheres by spiral imaging,” Opt. Lett., vol. 37, no. 5, pp. 869–871, 2012. https://doi.org/10.1364/ol.37.000869.Search in Google Scholar PubMed
[238] X. Zambrana-Puyalto and G. Molina-Terriza, “The role of the angular momentum of light in Mie scattering. Excitation of dielectric spheres with Laguerre–Gaussian modes,” J. Quant. Spectrosc. Radiat. Transf., vol. 126, pp. 50–55, 2013. https://doi.org/10.1016/j.jqsrt.2012.10.010.Search in Google Scholar
[239] X. Zambrana-Puyalto, X. Vidal, P. Wozniak, P. Banzer, and G. Molina-Terriza, “Tailoring multipolar Mie scattering with helicity and angular momentum,” ACS Photonics, vol. 5, no. 7, pp. 2936–2944, 2018. https://doi.org/10.1021/acsphotonics.8b00268.Search in Google Scholar
[240] X. Zambrana-Puyalto, X. Vidal, I. Fernandez-Corbaton, and G. Molina-Terriza, “Far-field measurements of vortex beams interacting with nanoholes,” Sci. Rep., vol. 6, no. 1, pp. 1–10, 2016. https://doi.org/10.1038/srep22185.Search in Google Scholar PubMed PubMed Central
[241] N. Tischler, X. Zambrana-Puyalto, and G. Molina-Terriza, “The role of angular momentum in the construction of electromagnetic multipolar fields,” Eur. J. Phys., vol. 33, no. 5, p. 1099, 2012. https://doi.org/10.1088/0143-0807/33/5/1099.Search in Google Scholar
[242] A. D. Kiselev and D. O. Plutenko, “Optical trapping by Laguerre-Gaussian beams: far-field matching, equilibria, and dynamics,” Phys. Rev. A, vol. 94, no. 1, p. 013804, 2016. https://doi.org/10.1103/physreva.94.013804.Search in Google Scholar
[243] A. D. Kiselev and D. O. Plutenko, “Mie scattering of Laguerre-Gaussian beams: photonic nanojets and near-field optical vortices,” Phys. Rev. A, vol. 89, no. 4, p. 043803, 2014. https://doi.org/10.1103/physreva.89.043803.Search in Google Scholar
[244] A. S. Rury and R. Freeling, “Mie scattering of purely azimuthal Laguerre-Gauss beams: angular-momentum-induced transparency,” Phys. Rev. A, vol. 86, no. 5, p. 053830, 2012. https://doi.org/10.1103/physreva.86.053830.Search in Google Scholar
[245] L.-M. Zhou, K. W. Xiao, Z. Q. Yin, J. Chen, and N. Zhao, “Sensitivity of displacement detection for a particle levitated in the doughnut beam,” Opt. Lett., vol. 43, no. 19, pp. 4582–4585, 2018. https://doi.org/10.1364/ol.43.004582.Search in Google Scholar
[246] X. Zambrana-Puyalto, X. Vidal, M. L. Juan, and G. Molina-Terriza, “Dual and anti-dual modes in dielectric spheres,” Opt. Express, vol. 21, no. 15, pp. 17520–17530, 2013. https://doi.org/10.1364/oe.21.017520.Search in Google Scholar PubMed
[247] X. Zambrana-Puyalto, X. Vidal, and G. Molina-Terriza, “Excitation of single multipolar modes with engineered cylindrically symmetric fields,” Opt. Express, vol. 20, no. 22, pp. 24536–24544, 2012. https://doi.org/10.1364/oe.20.024536.Search in Google Scholar
[248] L. Guo, Q. Huang, M. Cheng, J. Li, and X. Yan, “Remote sensing for aerosol particles in marine atmosphere using scattering of optical vortex,” in High-Performance Computing in Geoscience and Remote Sensing VII, vol. 10430, Warsaw, Poland, SPIE, 2017.10.1117/12.2279428Search in Google Scholar
[249] L. F. Votto, A. Chafiq, A. Belafhal, G. Gouesbet, and L. A. Ambrosio, “Hermite–Gaussian beams in the generalized Lorenz–Mie theory through finite–series Laguerre–Gaussian beam shape coefficients,” J. Opt. Soc. Am. B, vol. 39, no. 4, pp. 1027–1032, 2022. https://doi.org/10.1364/josab.445314.Search in Google Scholar
[250] T. Qu, Z. S. Wu, Q. C. Shang, and Z. J. Li, “Light scattering of a Laguerre–Gaussian vortex beam by a chiral sphere,” J. Opt. Soc. Am. A, vol. 33, no. 4, pp. 475–482, 2016. https://doi.org/10.1364/josaa.33.000475.Search in Google Scholar
[251] Y. Jiang, Y. Shao, X. Qu, and H. Hua, “Scattering of a focused Laguerre–Gaussian beam by a spheroidal particle,” J. Opt., vol. 14, no. 12, p. 125709, 2012. https://doi.org/10.1088/2040-8978/14/12/125709.Search in Google Scholar
[252] P. Banzer, U. Peschel, S. Quabis, and G. Leuchs, “On the experimental investigation of the electric and magnetic response of a single nano-structure,” Opt. Express, vol. 18, no. 10, pp. 10905–10923, 2010. https://doi.org/10.1364/oe.18.010905.Search in Google Scholar
[253] R. M. Kerber, J. M. Fitzgerald, D. E. Reiter, S. S. Oh, and O. Hess, “Reading the orbital angular momentum of light using plasmonic nanoantennas,” ACS Photonics, vol. 4, no. 4, pp. 891–896, 2017. https://doi.org/10.1021/acsphotonics.6b00980.Search in Google Scholar
[254] H. Barati Sedeh, D. G. Pires, N. Chandra, et al.., “Manipulation of scattering spectra with topology of light and matter,” Laser Photon. Rev., vol. 17, p. 2200472, 2023. https://doi.org/10.1002/lpor.202200472.Search in Google Scholar
[255] S. Yoo and Q.-H. Park, “Metamaterials and chiral sensing: a review of fundamentals and applications,” Nanophotonics, vol. 8, no. 2, pp. 249–261, 2019. https://doi.org/10.1515/nanoph-2018-0167.Search in Google Scholar
[256] P. Woźniak, P. Banzer, and G. Leuchs, “Selective switching of individual multipole resonances in single dielectric nanoparticles,” Laser Photon. Rev., vol. 9, no. 2, pp. 231–240, 2015. https://doi.org/10.1002/lpor.201400188.Search in Google Scholar
[257] E. V. Melik-Gaykazyan, S. S. Kruk, R. Camacho-Morales, et al.., “Selective third-harmonic generation by structured light in Mie-resonant nanoparticles,” ACS Photonics, vol. 5, no. 3, pp. 728–733, 2017. https://doi.org/10.1021/acsphotonics.7b01277.Search in Google Scholar
[258] E. Melik-Gaykazyan, K. Koshelev, J. H. Choi, et al.., “From Fano to quasi-BIC resonances in individual dielectric nanoantennas,” Nano Lett., vol. 21, no. 4, pp. 1765–1771, 2021. https://doi.org/10.1021/acs.nanolett.0c04660.Search in Google Scholar PubMed
[259] K. Koshelev, S. Kruk, E. Melik-Gaykazyan, et al.., “Subwavelength dielectric resonators for nonlinear nanophotonics,” Science, vol. 367, no. 6475, pp. 288–292, 2020. https://doi.org/10.1126/science.aaz3985.Search in Google Scholar PubMed
[260] C. W. Hsu, B. Zhen, A. D. Stone, J. D. Joannopoulos, and M. Soljacic, “Bound states in the continuum,” Nat. Rev. Mater., vol. 1, no. 9, pp. 1–13, 2016. https://doi.org/10.1038/natrevmats.2016.48.Search in Google Scholar
[261] K. Koshelev, A. Bogdanov, and Y. Kivshar, “Meta-optics and bound states in the continuum,” Sci. Bull., vol. 64, no. 12, pp. 836–842, 2019. https://doi.org/10.1016/j.scib.2018.12.003.Search in Google Scholar PubMed
[262] L. Carletti, K. Koshelev, C. De Angelis, and Y. Kivshar, “Giant nonlinear response at the nanoscale driven by bound states in the continuum,” Phys. Rev. Lett., vol. 121, no. 3, p. 033903, 2018. https://doi.org/10.1103/physrevlett.121.033903.Search in Google Scholar
[263] I. Volkovskaya, L. Xu, L. Huang, A. I. Smirnov, A. E. Miroshnichenko, and D. Smirnova, “Multipolar second-harmonic generation from high-Q quasi-BIC states in subwavelength resonators,” Nanophotonics, vol. 9, no. 12, pp. 3953–3963, 2020. https://doi.org/10.1515/nanoph-2020-0156.Search in Google Scholar
[264] M. F. Limonov, M. V. Rybin, A. N. Poddubny, and Y. S. Kivshar, “Fano resonances in photonics,” Nat. Photonics, vol. 11, no. 9, pp. 543–554, 2017. https://doi.org/10.1038/nphoton.2017.142.Search in Google Scholar
[265] R. Camacho-Morales, G. Bautista, X. Zang, et al.., “Resonant harmonic generation in AlGaAs nanoantennas probed by cylindrical vector beams,” Nanoscale, vol. 11, no. 4, pp. 1745–1753, 2019. https://doi.org/10.1039/c8nr08034h.Search in Google Scholar PubMed
[266] J. Zeng, M. Darvishzadeh-Varcheie, M. Albooyeh, et al.., “Exclusive magnetic excitation enabled by structured light illumination in a nanoscale mie resonator,” ACS Nano, vol. 12, no. 12, pp. 12159–12168, 2018. https://doi.org/10.1021/acsnano.8b05778.Search in Google Scholar PubMed
[267] G. Bautista, C. Dreser, X. Zang, D. P. Kern, M. Kauranen, and M. Fleischer, “Collective effects in second-harmonic generation from plasmonic oligomers,” Nano Lett., vol. 18, no. 4, pp. 2571–2580, 2018. https://doi.org/10.1021/acs.nanolett.8b00308.Search in Google Scholar PubMed PubMed Central
[268] M. K. Kroychuk, A. S. Shorokhov, D. F. Yagudin, et al.., “Enhanced nonlinear light generation in oligomers of silicon nanoparticles under vector beam illumination,” Nano Lett., vol. 20, no. 5, pp. 3471–3477, 2020. https://doi.org/10.1021/acs.nanolett.0c00393.Search in Google Scholar PubMed
[269] M. Montagnac, G. Agez, A. Patoux, A. Arbouet, and V. Paillard, “Engineered near-and far-field optical response of dielectric nanostructures using focused cylindrical vector beams,” J. Appl. Phys., vol. 131, no. 13, p. 133101, 2022. https://doi.org/10.1063/5.0085940.Search in Google Scholar
[270] U. Manna, H. Sugimoto, D. Eggena, et al.., “Selective excitation and enhancement of multipolar resonances in dielectric nanospheres using cylindrical vector beams,” J. Appl. Phys., vol. 127, no. 3, p. 033101, 2020. https://doi.org/10.1063/1.5132791.Search in Google Scholar
[271] E. V. Melik-Gaykazyan, K. L. Koshelev, J. H. Choi, et al.., “Enhanced second-harmonic generation with structured light in AlGaAs nanoparticles governed by magnetic response,” JETP Lett., vol. 109, no. 2, pp. 131–135, 2019. https://doi.org/10.1134/s0021364019020036.Search in Google Scholar
[272] T. Das, P. P. Iyer, R. A. DeCrescent, and J. A. Schuller, “Beam engineering for selective and enhanced coupling to multipolar resonances,” Phys. Rev. B, vol. 92, no. 24, p. 241110, 2015. https://doi.org/10.1103/physrevb.92.241110.Search in Google Scholar
[273] J. Sancho-Parramon and S. Bosch, “Dark modes and Fano resonances in plasmonic clusters excited by cylindrical vector beams,” ACS Nano, vol. 6, no. 9, pp. 8415–8423, 2012. https://doi.org/10.1021/nn303243p.Search in Google Scholar PubMed
[274] A. Canós Valero, E. A. Gurvitz, F. A. Benimetskiy, et al.., “Theory, observation, and ultrafast response of the hybrid anapole regime in light scattering,” Laser Photon. Rev., vol. 15, no. 10, p. 2100114, 2021. https://doi.org/10.1002/lpor.202100114.Search in Google Scholar
[275] L. Wei, Z. Xi, N. Bhattacharya, and H. P. Urbach, “Excitation of the radiationless anapole mode,” Optica, vol. 3, no. 8, pp. 799–802, 2016. https://doi.org/10.1364/optica.3.000799.Search in Google Scholar
[276] T. Raybould, V. A. Fedotov, N. Papasimakis, I. Youngs, and N. I. Zheludev, “Exciting dynamic anapoles with electromagnetic doughnut pulses,” Appl. Phys. Lett., vol. 111, no. 8, p. 081104, 2017. https://doi.org/10.1063/1.4999368.Search in Google Scholar
[277] J. A. Parker, H. Sugimoto, B. Coe, et al.., “Excitation of nonradiating anapoles in dielectric nanospheres,” Phys. Rev. Lett., vol. 124, no. 9, p. 097402, 2020. https://doi.org/10.1103/physrevlett.124.097402.Search in Google Scholar
[278] Y. D. Lu, Y. Xu, X. Ouyang, et al.., “Cylindrical vector beams reveal radiationless anapole condition in a resonant state,” Opto-Electron. Adv., vol. 5, p. 210014, 2022. https://doi.org/10.29026/oea.2022.210014.Search in Google Scholar
[279] J. Mun, M. Kim, Y. Yang, et al.., “Electromagnetic chirality: from fundamentals to nontraditional chiroptical phenomena,” Light Sci. Appl., vol. 9, no. 1, pp. 1–18, 2020. https://doi.org/10.1038/s41377-020-00367-8.Search in Google Scholar PubMed PubMed Central
[280] A. Lininger, G. Palermo, A. Guglielmelli, et al.., “Chirality in light–matter interaction,” Adv. Mater., p. 2107325, 2022. https://doi.org/10.1002/adma.202107325.Search in Google Scholar PubMed
[281] X. Zambrana-Puyalto, X. Vidal, and G. Molina-Terriza, “Angular momentum-induced circular dichroism in non-chiral nanostructures,” Nat. Commun., vol. 5, no. 1, pp. 1–7, 2014. https://doi.org/10.1038/ncomms5922.Search in Google Scholar PubMed
[282] P. Woźniak, I. De Leon, K. Hoflich, G. Leuchs, and P. Banzer, “Interaction of light carrying orbital angular momentum with a chiral dipolar scatterer,” Optica, vol. 6, no. 8, pp. 961–965, 2019. https://doi.org/10.1364/optica.6.000961.Search in Google Scholar
[283] J. Ni, S. Liu, D. Wu, et al.., “Gigantic vortical differential scattering as a monochromatic probe for multiscale chiral structures,” Proc. Natl. Acad. Sci., vol. 118, no. 2, p. e2020055118, 2021. https://doi.org/10.1073/pnas.2020055118.Search in Google Scholar PubMed PubMed Central
[284] S. Liu, J. Ni, C. Zhang, et al.., “Tailoring optical vortical dichroism with stereometamaterials,” Laser Photon. Rev., vol. 16, no. 2, p. 2100518, 2022. https://doi.org/10.1002/lpor.202100518.Search in Google Scholar
[285] A. G. Lamprianidis and A. E. Miroshnichenko, “Excitation of nonradiating magnetic anapole states with azimuthally polarized vector beams,” Beilstein J. Nanotechnol., vol. 9, no. 1, pp. 1478–1490, 2018. https://doi.org/10.3762/bjnano.9.139.Search in Google Scholar PubMed PubMed Central
[286] Z. Ruan and S. Fan, “Superscattering of light from subwavelength nanostructures,” Phys. Rev. Lett., vol. 105, no. 1, p. 013901, 2010. https://doi.org/10.1103/physrevlett.105.013901.Search in Google Scholar
[287] C. Qian, X. Lin, Y. Yang, et al.., “Experimental observation of superscattering,” Phys. Rev. Lett., vol. 122, no. 6, p. 063901, 2019. https://doi.org/10.1103/physrevlett.122.063901.Search in Google Scholar
© 2023 the author(s), 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
- Editorial
- Nanophotonics in support of Ukrainian Scientists
- Reviews
- Asymmetric transmission in nanophotonics
- Integrated circuits based on broadband pixel-array metasurfaces for generating data-carrying optical and THz orbital angular momentum beams
- Singular optics empowered by engineered optical materials
- Electrochemical photonics: a pathway towards electrovariable optical metamaterials
- Sustainable chemistry with plasmonic photocatalysts
- Perspectives
- Ukraine and singular optics
- Machine learning to optimize additive manufacturing for visible photonics
- Through thick and thin: how optical cavities control spin
- Research Articles
- Spin–orbit coupling induced by ascorbic acid crystals
- Broadband transfer of binary images via optically long wire media
- Counting and mapping of subwavelength nanoparticles from a single shot scattering pattern
- Controlling surface waves with temporal discontinuities of metasurfaces
- On the relation between electrical and electro-optical properties of tunnelling injection quantum dot lasers
- On-chip multivariant COVID 19 photonic sensor based on silicon nitride double-microring resonators
- Nano-infrared imaging of metal insulator transition in few-layer 1T-TaS2
- Electrical generation of surface phonon polaritons
- Dynamic beam control based on electrically switchable nanogratings from conducting polymers
- Tilting light’s polarization plane to spatially separate the ultrafast nonlinear response of chiral molecules
- Spin-dependent phenomena at chiral temporal interfaces
- Spin-controlled photonics via temporal anisotropy
- Coherent control of symmetry breaking in transverse-field Ising chains using few-cycle pulses
- Field enhancement of epsilon-near-zero modes in realistic ultrathin absorbing films
- Controlled compression, amplification and frequency up-conversion of optical pulses by media with time-dependent refractive index
- Tailored thermal emission in bulk calcite through optic axis reorientation
- Tip-enhanced photoluminescence of monolayer MoS2 increased and spectrally shifted by injection of electrons
- Quantum-enhanced interferometer using Kerr squeezing
- Nonlocal electro-optic metasurfaces for free-space light modulation
- Dispersion braiding and band knots in plasmonic arrays with broken symmetries
- Dual-mode hyperbolicity, supercanalization, and leakage in self-complementary metasurfaces
- Monocular depth sensing using metalens
- Multimode hybrid gold-silicon nanoantennas for tailored nanoscale optical confinement
- Replicating physical motion with Minkowskian isorefractive spacetime crystals
- Reconfigurable nonlinear optical element using tunable couplers and inverse-designed structure
Articles in the same Issue
- Frontmatter
- Editorial
- Nanophotonics in support of Ukrainian Scientists
- Reviews
- Asymmetric transmission in nanophotonics
- Integrated circuits based on broadband pixel-array metasurfaces for generating data-carrying optical and THz orbital angular momentum beams
- Singular optics empowered by engineered optical materials
- Electrochemical photonics: a pathway towards electrovariable optical metamaterials
- Sustainable chemistry with plasmonic photocatalysts
- Perspectives
- Ukraine and singular optics
- Machine learning to optimize additive manufacturing for visible photonics
- Through thick and thin: how optical cavities control spin
- Research Articles
- Spin–orbit coupling induced by ascorbic acid crystals
- Broadband transfer of binary images via optically long wire media
- Counting and mapping of subwavelength nanoparticles from a single shot scattering pattern
- Controlling surface waves with temporal discontinuities of metasurfaces
- On the relation between electrical and electro-optical properties of tunnelling injection quantum dot lasers
- On-chip multivariant COVID 19 photonic sensor based on silicon nitride double-microring resonators
- Nano-infrared imaging of metal insulator transition in few-layer 1T-TaS2
- Electrical generation of surface phonon polaritons
- Dynamic beam control based on electrically switchable nanogratings from conducting polymers
- Tilting light’s polarization plane to spatially separate the ultrafast nonlinear response of chiral molecules
- Spin-dependent phenomena at chiral temporal interfaces
- Spin-controlled photonics via temporal anisotropy
- Coherent control of symmetry breaking in transverse-field Ising chains using few-cycle pulses
- Field enhancement of epsilon-near-zero modes in realistic ultrathin absorbing films
- Controlled compression, amplification and frequency up-conversion of optical pulses by media with time-dependent refractive index
- Tailored thermal emission in bulk calcite through optic axis reorientation
- Tip-enhanced photoluminescence of monolayer MoS2 increased and spectrally shifted by injection of electrons
- Quantum-enhanced interferometer using Kerr squeezing
- Nonlocal electro-optic metasurfaces for free-space light modulation
- Dispersion braiding and band knots in plasmonic arrays with broken symmetries
- Dual-mode hyperbolicity, supercanalization, and leakage in self-complementary metasurfaces
- Monocular depth sensing using metalens
- Multimode hybrid gold-silicon nanoantennas for tailored nanoscale optical confinement
- Replicating physical motion with Minkowskian isorefractive spacetime crystals
- Reconfigurable nonlinear optical element using tunable couplers and inverse-designed structure