Startseite Compound meta-optics: there is plenty of room at the top
Artikel Open Access

Compound meta-optics: there is plenty of room at the top

  • Ahmed H. Dorrah ORCID logo EMAIL logo
Veröffentlicht/Copyright: 24. April 2025
Veröffentlichen auch Sie bei De Gruyter Brill
Nanophotonics
Aus der Zeitschrift Nanophotonics

Abstract

Metasurfaces have been widely exploited in imaging and sensing, holography, light–matter interaction, and optical communications in free space and on chip, thanks to their CMOS compatibility, versatility and compact form. However, as this technology matured from novelty to performance, stringent requirements on diffraction efficiency, scalability, and complex light control have also emerged. For instance, the limited thickness of single-layer meta-optics poses fundamental constraints on dispersion engineering and lossless transmission over large-scale devices, whereas in-plane symmetry limits the polarization transformations that can be realized. Cascaded and multi-layer flat optics can alleviate these constraints, offering new possibilities for realizing high-efficiency devices, full polarization control, and achromatic response. In this perspective, recent advances in multi-layer metasurfaces including inherent challenges and opportunities will be discussed. Compound meta-optics hold the promise for enabling complex optical systems with enhanced performance and unprecedented functionality for a diverse set of applications in sensing, imaging, high-capacity communications, and beyond.

1 Introduction

Optical metasurfaces represent a wide class of planar nano-optics, patterned with sub-wavelength resolution, to control light’s phase, intensity, and polarization profile, point-by-point, over a compact thickness [1], [2], [3], [4], [5], [6], [7], [8], [9], [10], [11]. With a single lithography step, a rich gamut of meta-atoms can be tailored at the nanoscale, allowing a versatile set of functionalities to be achieved using the same planar form factor [12], [13], [14], [15], [16]. To this end, flat optics has enabled a variety of applications, ranging from focusing [17], [18], [19], [20], [21], [22], [23], computational imaging [24], [25], [26], [27], [28], holography [29], [30], polarization imaging [31], [32], in addition to multifunctional [33], [34], [35], [36] and reconfigurable devices [37], [38], [39], [40], [41], [42], [43], [44], [45], [46], [47]. Thanks to their well-established CMOS-compatible nanolithography, deposition and etching techniques [48], [49], [50], [51], metasurfaces can be deployed in free space or integrated on chip [52], [53]. Although metasurfaces have historically evolved as planar photonic components [54], [55], [56], [57], [58], [59], [60] that alleviate the fabrication complexity of 3D metamaterials [61], [62], [63], [64], [65], [66], while approaching its versatility, it has become clear that their limited thickness poses many fundamental constraints in wavefront shaping, especially at high numerical apertures, as will be detailed next. Consequently, new architectures for cascaded, multi-layer, double-sided, folded, and 3D meta-optics (Figure 1) have gained much attention recently, aided by the rapid advancements in multi-level nanofabrication techniques.

2 Limitations of single-layer metasurfaces

The development of compound meta-optics arose partly because a single light interaction with a flat optical surface is fundamentally restricted to a finite set of achievable wavefront transformations, often posing a compromise between diffraction efficiency, bandwidth, field of view, numerical aperture (NA), and complex amplitude modulation. For instance, it is now well established that an arbitrary wave transformation via an ultrathin metasurface (much thinner than the wavelength), for e.g., achieved with plasmonic nano-antennas, even in the prototypical case of beam steering, inherently requires balanced loss and gain across the surface to achieve unitary efficiency [67], [68], [69], [70]. This stems from local impedance mismatch between the input and output fields across the surface of an infinitesimally thin meta-optic. Such a passive and reciprocal interface only supports tangential electric currents that, due to 2D symmetry of the surface, radiate equally on both sides, hindering simultaneous lossless transmission and complete phase control, and causes undesired reflection and limited polarization conversion [67], [69].

Achieving impedance matching between the input and output fields using ultra-thin gradient metasurfaces requires at least two layers to improve the transmission level [71], [72], and three (or two in reflection) to simultaneously control the underlying phase between 0 − 2π [67], [73], [74], [75]. This requirement can be reduced to two layers in transmission using nonlocal metasurfaces by relying on extended resonances or long-range interaction between the nanopillars [76], [77], [78], [79], using Huygen’s metasurfaces with electric and magnetic polarization [80] or through incorporating bianisotropy [81], [82].

Extending the thickness of single-layer meta-optics, from ultra-thin to the wavelength-scale regime, using high index contrast dielectric meta-atoms, have relaxed the tight constraints on diffraction efficiency and complete phase and polarization control [20], [48], [56], [57], [83], [84], [85], [86], [87], [88], [89], [90]. Nevertheless, lossless complex amplitude modulation using phase-only dielectric metasurfaces still mandates the use of a compound (multi-layer) arrangements [91], [92], [93]. Furthermore, on the device level, several trade-offs still exist between bandwidth, field-of-view and NA. For instance, achieving achromatic response at high NA values, wich require steep phase gradients remains an open challenge in metalens design due to limitations imposed by its thin form factor. In principle, a metalens, being a linear time-invariant system, is subject to the delay-bandwidth product, which implies that processing more frequencies can be achieved on the expense of more travel time (delay), ultimately requiring thicker or multi-layer metasurfaces [94], [95]. Similar arguments can be made for realizing meta-optics with wide field-of-view [96], [97], [98]. In short, as Prof. David Miller has aptly explained, optics require a minimum thickness to perform a wide range of tasks [99], [100].

3 Compound meta-optics

More generally, there are numerous applications that cannot be met with a single optic. For example, aligning two skewed optical beams collinearly with just one mirror is impossible. Similarly, spatial mode multiplexing, which involves shaping the wavefront and redirecting the momentum of incoming light, is a task that demands adiabatic light conversion across multiple planes [101]. In general, distributing the wavefront transformation across multiple layers introduces an additional degree-of-freedom, providing beneficial redundancy in metasurface design. This redundancy can enable multifunctional devices with high diffraction efficiency, broadband performance, wide field of view, and adaptable dispersion control as will be shown. It also provides an additional knob for tunability, for instance, through dynamically changing the relative position of the metalenses within a stack. This has been previously demonstrated in bilayer Moiré varifocal metalenses [102] and Airy beam generators [103] in the visible, as well as variable zoom meta-optic triplets in the NIR range [104]. Rotary varifocal doublets and triplets of this kind have also been demonstrated in the THz regime for possible application in 6G communications [105]. Additionally, vertical integration of metasurface structures on top of non-linear crystals could provide a route for the generation and control of complex quantum states, thus enabling a compact and practical platform for the development of advanced on-chip quantum photonic information processing [106]. As metasurface technologies have matured, the demand for these advanced functionalities has led to the exploration of more complex configurations, including cascaded, double-sided, and multi-layer flat optics, as depicted in Figure 1. Compound meta-optics of this nature bring many new opportunities when it comes to polarization control, dispersion engineering, and wavefront shaping as discussed next.

Figure 1: 
Different configurations of multi-layer metasurfaces include: (a) cascaded metasurfaces, metasurface doublets, cladded multi-layer metasurfaces, folded metasurfaces, and (b) free-standing metasurfaces with high index contrast. An SEM image of a free-standing bilayer metasurface is shown at the bottom of (b), reproduced from [107].
Figure 1:

Different configurations of multi-layer metasurfaces include: (a) cascaded metasurfaces, metasurface doublets, cladded multi-layer metasurfaces, folded metasurfaces, and (b) free-standing metasurfaces with high index contrast. An SEM image of a free-standing bilayer metasurface is shown at the bottom of (b), reproduced from [107].

3.1 Metasurface polarization optics

Metasurface polarization optics have led to an extensive developments in vectorial holography as well as compact Stokes and Mueller imaging systems which can capture the polarization of a scene in single shot without moving parts [108]. Developments of this kind have reached a high technology readiness level (TRL) to be adopted in consumer electronics, including biometrics in handheld devices, industrial quality control, autonomous driving, AR/VR and quantum state topography [34], [109], [110].

From a more fundamental standpoint, shape-birefringent dielectric nanofins function akin to waveplates, with their transmission behavior acquiring the form of a 2-by-2 unitary and symmetric Jones matrix [87], [90], [108], [111]. While such polarization control has enabled technologies as single-shot Stokes [31] and Mueller imaging [32], unconventional polarizers [112], [113], [114], [115], and vectorial holograms [116], [117], [118], it falls short when it comes to more complex functionalities. For instance, a single-layer non-chiral nanofin cannot serve as a circular analyzer through shape-birefringence alone, as this requires the decoupling of the off-diagonal components of the Jones matrix [108]. However, rectangular non-chiral nanofins only provide linear shape-birefringence, cast as a symmetric Jones matrix. Notably, this limitation cannot be lifted with topology optimization or inverse design as it stems from the 2D symmetry of a single-layer metasurface with vertical side walls [113]. To relax this condition, bilayer metasurfaces have been introduced to enable more generalized polarization transformations by decoupling all four elements of the Jones matrix, allowing each meta-atom to impart the most general polarization transformation [119], [120], as illustrated in Figure 2. This is based on the mathematical principle that the product of two symmetric matrices (i.e., two cascaded nanofins) can form an arbitrary matrix, that is not necessarily symmetric [121].

Figure 2: 
Multi-layer metasurfaces for polarization control. (a) Schematic of metasurface nanostructures and the corresponding Jones matrices that can be implemented. The letters a, b, c, and d represent complex numbers. (b) Left: Schematic of bilayer metasurface unit cell composed of 790 nm tall amorphous silicon nano posts. The bottom layer is encapsulated in SU-8. Right: SEM image of the fabricated metasurface. (c) Left: Schematic of a free-standing bilayer meta-atom made of 600 nm titanium dioxide nanofins. Right: SEM images of the fabricated device. Scale bar 500 μm. Panels (a) and (b) adapted with permission from Ref. [119]. © 2022 Wiley-VCH GmbH. Panel (c) adapted from Ref. [107].
Figure 2:

Multi-layer metasurfaces for polarization control. (a) Schematic of metasurface nanostructures and the corresponding Jones matrices that can be implemented. The letters a, b, c, and d represent complex numbers. (b) Left: Schematic of bilayer metasurface unit cell composed of 790 nm tall amorphous silicon nano posts. The bottom layer is encapsulated in SU-8. Right: SEM image of the fabricated metasurface. (c) Left: Schematic of a free-standing bilayer meta-atom made of 600 nm titanium dioxide nanofins. Right: SEM images of the fabricated device. Scale bar 500 μm. Panels (a) and (b) adapted with permission from Ref. [119]. © 2022 Wiley-VCH GmbH. Panel (c) adapted from Ref. [107].

Likewise, single-layer shape-birefringent metasurfaces can impart two different phase transformations on any pair of orthogonal polarizations, provided that the input-output polarization states are coupled [122], hindered by matrix symmetry. Cascaded metasurfaces, on the other hand, can lift this constraint by imparting arbitrary and independent amplitude and phase control on any set of two orthogonal polarizations, while completely decoupling the input and output polarization states [123], [124]. In addition to arbitrary polarization conversion, multi-layer metasurfaces also suggest new ways for wavefront shaping using Pancharatnam–Berry phase [125], [126] by allowing the input polarization to trace arbitrary trajectories in polarization space [107], [127]. This versatile control has led to a new class of geometric phase metasurfaces whose eigen polarization states are not necessarily circular but can also be linear or elliptical [107], [128].

Notably, versatile and complete polarization detection and conversion with multi-layer meta-optics can further enhance imaging systems by improving their contrast and resolution, leading to more precise sensors for industries like healthcare, automotive, and security [129], [130], [131]. In AR and VR, they enable compact devices with better 3D display quality [132], [133]. They also offer new opportunities in optical communications, energy harvesting, and quantum technologies by allowing precise and multidimensional control over light from any polarization state to another [110], [134], [135], [136], [137].

3.2 Dispersion-engineered meta-optics

Unlike bulk materials, whose dispersion properties are dictated by their material composition, metasurfaces offer nearly unlimited flexibility in managing dispersion by carefully designing and arranging the individual meta-atoms geometry [138], [139], [140], [141], [142], [143], [144], [145], [146]. This capability is essential in optical design, as it helps produce high-quality images by reducing chromatic aberrations, which can otherwise cause color distortion and blurriness. While methods for correcting dispersion in refractive optics have existed for a long time [147], [148], they often require combining different types of glasses, resulting in bulkier and more expensive optical systems. In contrast, dispersion-engineered metasurfaces provide an alternative route that allows for achromatic focusing using compact, planar structures, thereby minimizing size, weight, and power (SWaP) in optical systems [149], [150], [151], [152]. This has significant implications for enhancing performance in lenses, filters, and sensors. However, the performance of current metasurface platforms is ultimately limited by their thickness regardless of the sophistication of their underlying meta-atoms [94], [95].

Many adopted dispersion-engineered meta-atom configurations suffer from issues such as low transmission, narrow bandwidth, and limited polarization responses. This can, in part, be alleviated with inverse design methods such as adaptive mesh [153], [154], adjoint optimization [155], [156], [157], [158], [159], [160], [161] or deep learning [162], [163], [164] to yield more versatile free-form meta-atoms [165], [166]. Nevertheless, while it is possible to modify the phase response of meta-atoms by tuning their geometry, the imparted phase shift at the design wavelength cannot be freely controlled without affecting the phase imparted on the adjacent spectral components. In other words, for a given meta-atom height, and in the absence of resonances, it is difficult to fix the effective index which dictates the imparted phase shift (or accumulated phase delay) at a given wavelength while freely changing the effective index in the vicinity of the design wavelength (referred-to as the group delay). The degree to which phase and group delay can be controlled separately is fundamentally bound by the delay-bandwidth product [94], [95] which, at the physical layer, is dictated by the geometry, height, and composition of the meta-atom. This leads to common scenarios in dielectric metalenses where the phase coverage (or phase delay) falls short of 2π for a desired group delay, limiting the performance of dispersion-engineered metasurfaces over large area. Notably, inverse design techniques also become computationally intractable as the device size scales beyond a few tens of wavelengths. Moreover, adjoint methods suffer from local minima of the merit function in the design-response space, requiring multiple trials with varying starting points. Therefore, there is a critical need for new metasurface schemes that intrinsically offer high transmission, broadband, polarization-insensitive responses, without the constraints imposed by geometry.

Figure 3(a) and (b) provides a conceptual visualization of the limitations encountered in single-layer metasurfaces when it comes to dispersion control. The diffractive nature of simple nanofins typically causes normally incident multichromatic light to suffer from chromatic aberrations, as a consequence of the generalized Snell’s law [1], [167], as depicted in Figure 3(a). Meta-atoms of this kind can provide full control over the phase delay, by varying the geometry of each individual nanofin, and in turn its effective refractive index. However, advanced dispersion properties, such as group delay, cannot be independently tuned using this class of devices. This sets a fundamental limitation on achieving achromatic response. To realize broadband devices, more complex meta-atoms made of coupled nanofins have been widely utilized by exploiting the resonant coupling and rotational degree-of-freedom to enable more versatile dispersion control [141], [146], [150], [151], [166], [168], [169], as shown in Figure 3(b). The group index of this meta-atom is engineered through its geometry, which determines light confinement within the nanostructure, akin to how waveguides function [85], [89]. Rotating the meta-atom by an angle α imparts a Pancharatnam–Berry phase of 2α, achieved through polarization conversion [170], [171]. However, the operation of this type of metasurfaces comes with limitations on polarization due to the intrinsic properties of the Pancharatnam–Berry phase.

Figure 3: 
Comparison of three metasurface platforms: (a) a single-layer metasurface made up of meta-atoms with simple geometries, exemplified by a square pillar-shaped meta-atom. (b) A single-layer metasurface featuring rotated meta-atoms with varying geometries, illustrated by a double-fin meta-atom with a rotation angle α. (c) A bilayer metasurface created by stacking heterogeneous freeform meta-atoms. The unit cell diagram shows a freeform meta-atom stacked on top of a cross-shaped meta-atom embedded in the substrate. (d) Scanning electron microscopy (SEM) image of a fabricated heterogeneous metalens. The cross section of the stack is taken by using focused ion beam. (e) A hybrid achromatic metalens combines a phase plate and a metalens to simultaneously correct for chromatic aberration and improve focusing efficiency. SEM images of 40 μm diameter air-spaced (0.32 NA) metalens fabricated with multi-photon lithography on fused silica substrate. (f) 3D-printed multilayer achromatic metalens made from a low-index material. SEM images of the fabricated metalens operating in the visible range of 400–800 nm with 0.5 NA. Panels (a–d) adapted with permission from Ref. [172]. Images courtesy of Federico Capasso. Panel (e) reproduced from Ref. [173]. © 2020 Springer Nature. Panel (f) reproduced from Ref. [174]. © 2023 The Authors, some rights reserved; exclusive licensee AAAS. Distributed under a Creative Commons Attribution License (CC BY 4.0).
Figure 3:

Comparison of three metasurface platforms: (a) a single-layer metasurface made up of meta-atoms with simple geometries, exemplified by a square pillar-shaped meta-atom. (b) A single-layer metasurface featuring rotated meta-atoms with varying geometries, illustrated by a double-fin meta-atom with a rotation angle α. (c) A bilayer metasurface created by stacking heterogeneous freeform meta-atoms. The unit cell diagram shows a freeform meta-atom stacked on top of a cross-shaped meta-atom embedded in the substrate. (d) Scanning electron microscopy (SEM) image of a fabricated heterogeneous metalens. The cross section of the stack is taken by using focused ion beam. (e) A hybrid achromatic metalens combines a phase plate and a metalens to simultaneously correct for chromatic aberration and improve focusing efficiency. SEM images of 40 μm diameter air-spaced (0.32 NA) metalens fabricated with multi-photon lithography on fused silica substrate. (f) 3D-printed multilayer achromatic metalens made from a low-index material. SEM images of the fabricated metalens operating in the visible range of 400–800 nm with 0.5 NA. Panels (a–d) adapted with permission from Ref. [172]. Images courtesy of Federico Capasso. Panel (e) reproduced from Ref. [173]. © 2020 Springer Nature. Panel (f) reproduced from Ref. [174]. © 2023 The Authors, some rights reserved; exclusive licensee AAAS. Distributed under a Creative Commons Attribution License (CC BY 4.0).

As shown in Figure 3(b), a dispersion-engineered metasurface can redirect light achromatically from a left-handed to a right-handed polarization state. However, the main drawback stems from the typically low polarization conversion efficiency, which further decreases when operating away from the design wavelength [141], [146], [151], [168].

Multi-layer meta-optics holds the promise of addressing all these limitations [173], [174], [175], [176], [177]. In a first-order approximation, the phase delay and group delay of a bilayer meta-atom can be considered as the sum of the contributions from each individual layer [172]. Hence, by heterogeneously stacking two meta-atoms with different material compositions (for e.g., titanium dioxide and silica), one can significantly expand the decoupling space (independent control) between the phase and group delays, pixel-by-pixel. Furthermore, by carefully designing the meta-atoms in a freeform manner as shown in Figure 3(c), the decoupling range can be further expanded allowing for the tailored design of more advanced dispersion-engineered metasurfaces, ultimately enabling absolute focusing efficiency of 80 % over a broadband range covering from 420 nm to NIR [172]. Metasurfaces of this type can be fabricated using a two-step nanolithography, atom layer deposition, and reactive ion etching as depicted in Figure 3(d).

Alternatively, hybrid composite metalenses comprising a phase plate and a metalens, designed to correct for chromatic aberration, have been stacked into a single element that is only a few wavelengths thick, for achromatic focusing in the NIR (1,000–1,800 nm) range [173], as shown in Figure 3(e), and more recently in the visible range (400–700 nm) by patterning the meta-atoms on top of a dispersion-matching dielectric [178]. Polarization-insensitive multi-layer achromatic metalenses have also been demonstrated in the visible by fabricating inverse designed structures in low-refractive index materials using two-photon polymerization lithography as shown in Figure 3(f).

In addition to multi-wavelength focusing with metalens doublets [176] and trilayer metasurfaces in the visible [179] and NIR [180], cascading a plano-convex and a plano-concave planar metalenses represent another clever strategy for dispersion compensation [181], elucidating the clear advantage that compound meta-optics can bring for realizing high efficiency, polarization insensitive and achromatic response. Ultra-broadband meta-optics are poised to improve imaging and metrology systems, wireless communication, and may enable innovations in fields like autonomous vehicles and wearable technologies.

3.3 Light routing with 3D meta-optics

Three-dimensional (3D) devices leverage a broader range of optical modes to achieve exceptionally tailored performance across various tasks, but require an efficient gradient-based optimization algorithm driven by full-wave electromagnetic simulations. Inverse design allows searching the high-dimensional space of permittivity profiles to find a local optimum for an electromagnetic merit function [156], [182], [183], [184]. By doing so, quasi-2D on-chip photonic devices have been extensively explored, where patterning occurs in the direction of light propagation within a single fabrication layer [160], [185], [186]. However, fully 3D designs for free-space applications, particularly in the infrared and visible spectra, are still in their infancy due to the increased fabrication complexity of volumetric devices. Early works in this area, which used one- and two-layer processes for optical applications or many-layer microwave prototypes, have demonstrated the potential benefits of moving to thicker devices [165], [187], [188].

More recently, a two-photon polymerization (TPP) lithography process has been optimized to fabricate multi-layer structures at optical wavelengths [189]. This technique has been previously used to create refractive, diffractive, gradient index, and extruded 2D inverse-designed optical components [190], [191], [192]. By exploiting the flexibility of TPP for 3D patterning at sub-wavelength resolution, several inverse-designed, multi-layer photonic devices with applications in advanced imaging in the mid-infrared band (3–6 μm) have been experimentally demonstrated [189]. This includes volumetric meta-optics for color, polarization, and vortex beam splitting using a compact footprint as depicted in Figure 4. Such development becomes particularly useful for compact imaging systems which typically deploy wavelength- and polarization-selective elements to analyze the fundamental properties of wavefronts.

Figure 4: 
Volumetric meta-optics. (a) Schematic of full Stokes polarimetry device that sorts four polarizations; i.e., analyzes for four Jones vectors, and projects the outcome to different quadrants. (b) Spectral filter designed with multi-layer lithography. A three-dimensional scattering element created by stacking multiple two-dimensional layers. Each layer is made up of a series of titanium dioxide mesas, which can be fabricated using standard lithography and material deposition techniques. The posts are filled with silica matrix, creating a flat substrate for the subsequent layers. (c) When assembled, the five-layer stack focus linearly polarized light into distinct sensor regions based on frequency and polarization. Panel (a) adapted from Ref. [189]. © 2023 Springer Nature. Panel (b) adapted from Ref. [188]. © 2020 Optical Society of America. Used with permission.
Figure 4:

Volumetric meta-optics. (a) Schematic of full Stokes polarimetry device that sorts four polarizations; i.e., analyzes for four Jones vectors, and projects the outcome to different quadrants. (b) Spectral filter designed with multi-layer lithography. A three-dimensional scattering element created by stacking multiple two-dimensional layers. Each layer is made up of a series of titanium dioxide mesas, which can be fabricated using standard lithography and material deposition techniques. The posts are filled with silica matrix, creating a flat substrate for the subsequent layers. (c) When assembled, the five-layer stack focus linearly polarized light into distinct sensor regions based on frequency and polarization. Panel (a) adapted from Ref. [189]. © 2023 Springer Nature. Panel (b) adapted from Ref. [188]. © 2020 Optical Society of America. Used with permission.

In consumer cameras, for instance, absorptive filters are placed on pixel arrays to capture three or four spectral bands. The standard configuration, known as the Bayer pattern [193], arranges red, blue, and two green filters in a 2-by-2 grid. However, filtering schemes like this incur a transmission efficiency penalty because they absorb all light outside their passband, resulting in an average transmission rate of around 33 % under uniform spectral illumination. To address this issue, emerging meta-optics architectures as shown in Figure 4 can be deployed to capture light incident on a group of pixels and redirect each wavelength band or polarization to a different pixel with high diffraction efficiency [188], [194], [195], [196].

Besides light routing on chip, the development of advanced fabrication techniques for 3D printing of nano-optics will pave the path for many other inverse designed meta-optics capable of achieving high NA focusing [197], [198], asymmetric transmission [199], [200], [201], [202], [203], [204], trans-reflection [205], analogue optical computing [206], [207], [208], [209], [210], and beyond. These developments will inevitably rely on novel computational methods which enable the full-wave simulation of large scale devices and point-by-point optimization of its topology to realize the target wave transformation.

3.4 Multi-plane light control with folded metasurfaces

The challenges associated with fabricating volumetric meta-optics can partly be alleviated by considering folded metasurfaces as depicted in Figure 5. Compound or multi-plane wavefront shaping of this kind is typically required for correcting aberrations in many optical systems but is also widely adopted for mode division (de)multiplexing where simultaneous rerouting and reshaping of the wavefront is realized using a sequence of lossless (unitary) and adiabatic transformations [211], [212], [213], [214] akin to photonic lanterns [215], [216], [217]. The compact form factor of folded metasurfaces makes them suitable for shrinking imaging systems as well as (de)multiplexers. For instance, the thickness of consumer cameras is currently dictated by the volume of its bulk refractive lens stacks and the space in between as shown in Figure 5(a). Metasurface-based folded optics can address this by guiding light along multiple folded paths within the substrate, resulting in an ultra-thin imaging system with a thickness of ∼0.7 mm, while providing quasi-diffraction-limited imaging quality at a wavelength of 852 nm [218].

Figure 5: 
Folded metasurfaces. (a) The lens system of traditional cameras consists of multiple refractive lenses stacked vertically. The large volume of these lenses, along with the empty space between them, contributes to the overall thickness of the system. In contrast, metasurface folded optics cleverly bends the light path diagonally within the glass substrate. This approach makes efficient use of space, preserving the optical paths while significantly reducing the thickness of the lens system. The bottom right panel shows ray-tracing optimization results of the designed lens system. The colors of the rays represent different incident angles of light at the entrance aperture. (b) Schematic of a six-mode (three spatial modes and two polarization states) multiplexer based on a metasurface cavity: (a) the metasurface consists of a mirror, dielectric nanostructures with cladding, a substrate, and another mirror. Apertures on the mirrors allow light to enter and exit, with light reflecting within the cavity and coherently interfering at the output. The right column shows the measured output modes of the metasurface multiplexer. The bottom row depicts optical microscope image of the metasurface before the deposition of the gold mirror, with a scale bar of 100 μm. Scanning electron microscope images of the device without the cladding and mirror layers are also shown, with scale bar of 5 μm, and 2 μm. Panel (a) adapted from Ref. [218]. © 2024 The Authors, some rights reserved; exclusive licensee AAAS. Distributed under a Creative Commons Attribution License (CC BY 4.0). Panel (b) adapted from Ref. [101]. © 2022 The Authors. Published by American Chemical Society. Distributed under a Creative Commons Attribution License (CC BY 4.0).
Figure 5:

Folded metasurfaces. (a) The lens system of traditional cameras consists of multiple refractive lenses stacked vertically. The large volume of these lenses, along with the empty space between them, contributes to the overall thickness of the system. In contrast, metasurface folded optics cleverly bends the light path diagonally within the glass substrate. This approach makes efficient use of space, preserving the optical paths while significantly reducing the thickness of the lens system. The bottom right panel shows ray-tracing optimization results of the designed lens system. The colors of the rays represent different incident angles of light at the entrance aperture. (b) Schematic of a six-mode (three spatial modes and two polarization states) multiplexer based on a metasurface cavity: (a) the metasurface consists of a mirror, dielectric nanostructures with cladding, a substrate, and another mirror. Apertures on the mirrors allow light to enter and exit, with light reflecting within the cavity and coherently interfering at the output. The right column shows the measured output modes of the metasurface multiplexer. The bottom row depicts optical microscope image of the metasurface before the deposition of the gold mirror, with a scale bar of 100 μm. Scanning electron microscope images of the device without the cladding and mirror layers are also shown, with scale bar of 5 μm, and 2 μm. Panel (a) adapted from Ref. [218]. © 2024 The Authors, some rights reserved; exclusive licensee AAAS. Distributed under a Creative Commons Attribution License (CC BY 4.0). Panel (b) adapted from Ref. [101]. © 2022 The Authors. Published by American Chemical Society. Distributed under a Creative Commons Attribution License (CC BY 4.0).

A similar folded metasurface architecture has also been used in fiber-based mode division multiplexing [101]. In this work, the metasurface mode (de)multiplexer maps the fundamental mode from a fiber array to multiple spatial modes and spatially superimposes the output beam for direct coupling to a few mode fiber as depcicted in Figure 5(b). The device uses a metasurface cavity, where light propagates between a bottom mirror and an amorphous silicon metasurface-based phase mask on top.

As additional modes are added, only the lateral size of the phase mask increases, allowing scalability. This integrated design eliminates the need for free-space optics, performing collimation, focusing, beam deflection, and mode conversion on the metasurface, within the substrate’s thickness, resulting in a compact system without misalignment issues. This metasurface was optimized using adjoint analysis to minimize insertion loss, assuming light behaves as a Fabry–Perot cavity with coherent reflections. Consequently, a six-mode (de)MUX that converts three single-mode fiber (SMF) input channels (two polarization states for each) into the first three LP modes of a few-mode fiber (FMF) was realized in the C-band. This platform can potentially control additional properties like dispersion and polarization for broadband and vectorial mode manipulation as discussed in Sections 3.1 and 3.2.

The list goes on with potential applications which can make use of folded meta-optics including compact spectrometers [219], hyperspectral imagers [220], high NA off-axis illumination optics [221], augmented reality displays [222], and metasurface-stabilized microcavities [223], highlighting the range of opportunities that these meta-optics can bring in terms of miniaturization and multifunctionality.

4 Outlook

Although metasurfaces have been widely regarded as the planar analogue of metamaterials, the rising demand for higher diffraction efficiency, broadband operation and multifunctional response are not only difficult to realize with a single-layer metasurface from a design perspective but in many cases are not even feasible from a fundamental standpoint as highlighted throughout this article and others [67], [68], [69], [70], [94], [95], [97], [99]. Various permutations of single-layer meta-atom configurations have been exhausted over the past decade inevitably leading to a persistent trade-off between absolute efficiency, broadband operation, and large-scale devices. While it is true that inverse design and topology optimization have succeeded in pushing these frontiers by a few percentages, there is more significant advantage to be gained through vertical integration. By decomposing a complex target optical function into multiple layers, one can simplify the design requirements mandated by each layer while allowing the resulting stack to break the fundamental limits imposed on each constituent layer.

A paradigm shift in meta-optics which moves the complexity from the sophistication of precisely crafted free-form single-layer meta-atoms to “more forgiving” multi-layer stacks of more primitive building blocks will provide a more straight forward path to decoupling the otherwise entangled optical properties (for e.g., effective index, phase delay, group delay…etc.). This will also accelerate the adoption of techniques such as nanoimprint lithography for creating multi-layer stacks of large area meta-optics without the stringent requirements on deep sub-wavelength patterning. It also facilitates the hybrid integration of (conformal) meta-optics with refractive optics, thereby achieving the best of both worlds [224], [225], [226], while ensuring backward compatibility with existing optical systems by allowing meta-optics to be integrated in larger optical systems such as complex wafer metrology, ranging, or imaging systems. This will especially be suited for performing a tailored task which otherwise cannot be performed by merely relying on refractive optics; for e.g., dispersion engineering [144] and polarization control [108] are especially two key areas which can benefit from such a hybrid integration.

Furthermore, multi-layer stacking inspires alternative meta-optics schemes which leverage the capabilities of local and nonlocal metasurfaces, simultaneously, by integrating the two [227]. By combining the extended resonances enabled by neighbor-to-neighbor interactions of nonlocal metasurfaces with the local phase and polarization control of gradient metasurfaces, lossless and multifunctional wavefront shaping can ultimately be achieved. Realizing multifunctional meta-optics with less than 3–5 % reflection will mark a new era where metasurfaces can be real contenders to existing refractive optics, especially in microscopy and imaging applications which are highly sensitive to zeroth-order optical distortion.

Additionally, the integration of metasurfaces and nanoscale emitters enable access to both weak and strong coupling regimes, thereby enhancing photoluminescence, nanoscale lasing, controlled quantum emission, and exciton-polariton formation. New metasurface designs, such as surface-functionalized, chemically tunable, and multi-layer hybrid metasurfaces, offer potential applications in photocatalysis, sensing, displays, and quantum information [228]. Compound meta-optics will also allow a more direct path to tunability. By strategically co-designing a passive metasurfaces, with complex phase profile, and a less sophisticated active metasurface, a hybrid stack which achieves efficient, versatile and dynamic wavefront shaping can readily be implemented.

From a more fundamental standpoint, while many configurations of multi-layer metasurfaces can be treated as a stack of decoupled single layers, long-range interaction between the vertically stacked meta-atoms along the propagation direction can also lead to interesting phenomena related to light transport [229], [230] and disorder [231] which represent untapped potential of diffractive optics by unlocking numerous degrees-of-freedom which can ultimately be harnessed for optical computing, neuromorphic computing and beyond [208], [209], [232], [233], [234], [235], [236], [237], [238].

The prospect of realizing multi-layer flat optics to serve all these applications at scale hinges on recent advances in nanofabrication including heterogeneous integration of meta-atoms [172], [239], nanoimprint lithography [240], two-photon polymerization [189], thermal scanning probe lithography [241], [242], and roll-to-roll metasurfaces [243], [244] which are poised to enable the patterning of compound meta-optics on curved substrates for the correction of off-axis aberrations [245]. With the abundance of these tools, optical metamaterials are within reach, the future of flat optics is 3D and there is plenty of room at the top.


Corresponding author: Ahmed H. Dorrah, Department of Applied Physics and Science Education, Eindhoven University of Technology, Eindhoven, The Netherlands, E-mail: 

Dedicated to My mentor Prof. Federico Capasso whose visionary and profound work in nanophotonics, flat optics and laser science continues to inspire.


Funding source: Optica Foundation

Award Identifier / Grant number: Optica Foundation Challenge

Acknowledgment

The author acknowledges insightful discussions with N. A. Rubin. A. Zaidi, A. Palmieri, J.-S. Park, J. Oh, P. Dainese, and J. Gómez-Rivas as well as the anonymous reviewers who collectively helped shape this perspective.

  1. Research funding: AHD acknowledges funding support from the Optica Foundation Challenge program, the Natural Sciences and Engineering Research Council of Canada (NSERC), and the Dutch Research Council (NWO) Gravitation program.

  2. Author contributions: The author confirms the sole responsibility for the presented analysis and manuscript preparation.

  3. Conflict of interest: The author states no conflict of interest.

  4. Data availability: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.

References

[1] N. Yu, et al.., “Light propagation with phase discontinuities: generalized laws of reflection and refraction,” Science, vol. 334, no. 6054, pp. 333–337, 2011. https://doi.org/10.1126/science.1210713.Suche in Google Scholar PubMed

[2] A. V. Kildishev, A. Boltasseva, and V. M. Shalaev, “Planar photonics with metasurfaces,” Science, vol. 339, no. 6125, p. 1232009, 2013. https://doi.org/10.1126/science.1232009.Suche in Google Scholar PubMed

[3] N. Yu and F. Capasso, “Flat optics with designer metasurfaces,” Nat. Mater., vol. 13, no. 2, pp. 139–150, 2014. https://doi.org/10.1038/nmat3839.Suche in Google Scholar PubMed

[4] 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.Suche in Google Scholar PubMed

[5] 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.Suche in Google Scholar PubMed

[6] A. E. Minovich, A. E. Miroshnichenko, A. Y. Bykov, T. V. Murzina, D. N. Neshev, and Y. S. Kivshar, “Functional and nonlinear optical metasurfaces,” Laser Photonics Rev., vol. 9, no. 2, pp. 195–213, 2015. https://doi.org/10.1002/lpor.201400402.Suche in Google Scholar

[7] 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.Suche in Google Scholar

[8] A. Y. Zhu, A. I. Kuznetsov, B. Luk’yanchuk, N. Engheta, and P. Genevet, “Traditional and emerging materials for optical metasurfaces,” Nanophotonics, vol. 6, no. 2, pp. 452–471, 2017.10.1515/nanoph-2016-0032Suche in Google Scholar

[9] G. Li, S. Zhang, and T. Zentgraf, “Nonlinear photonic metasurfaces,” Nat. Rev. Mater., vol. 2, no. 5, p. 17010, 2017. https://doi.org/10.1038/natrevmats.2017.10.Suche in Google Scholar

[10] S. Chang, X. Guo, and X. Ni, “Optical metasurfaces: progress and applications,” Annu. Rev. Mater. Res., vol. 48, no. 1, pp. 279–302, 2018. https://doi.org/10.1146/annurev-matsci-070616-124220.Suche in Google Scholar

[11] A. I. Kuznetsov, et al.., “Roadmap for optical metasurfaces,” ACS Photonics, vol. 11, no. 3, pp. 816–865, 2024. https://doi.org/10.1021/acsphotonics.3c00457.Suche in Google Scholar PubMed PubMed Central

[12] H.-H. Hsiao, C. H. Chu, and D. P. Tsai, “Fundamentals and applications of metasurfaces,” Small Methods, vol. 1, no. 4, p. 1600064, 2017. https://doi.org/10.1002/smtd.201600064.Suche in Google Scholar

[13] X. Yin, et al.., “Beam switching and bifocal zoom lensing using active plasmonic metasurfaces,” Light: Sci. Appl., vol. 6, no. 7, p. e17016, 2017. https://doi.org/10.1038/lsa.2017.16.Suche in Google Scholar PubMed PubMed Central

[14] M. Khorasaninejad and F. Capasso, “Metalenses: versatile multifunctional photonic components,” Science, vol. 358, no. 6367, p. eaam8100, 2017. https://doi.org/10.1126/science.aam8100.Suche in Google Scholar PubMed

[15] P. Lalanne, W. Yan, K. Vynck, C. Sauvan, and J.-P. Hugonin, “Light interaction with photonic and plasmonic resonances,” Laser Photonics Rev., vol. 12, no. 5, p. 1700113, 2018. https://doi.org/10.1002/lpor.201700113.Suche in Google Scholar

[16] K. Shastri and F. Monticone, “Nonlocal flat optics,” Nat. Photonics, vol. 17, no. 1, pp. 36–47, 2023. https://doi.org/10.1038/s41566-022-01098-5.Suche in Google Scholar

[17] A. Pors, M. G. Nielsen, R. L. Eriksen, and S. I. Bozhevolnyi, “Broadband focusing flat mirrors based on plasmonic gradient metasurfaces,” Nano Lett., vol. 13, no. 2, pp. 829–834, 2013. https://doi.org/10.1021/nl304761m.Suche in Google Scholar PubMed

[18] X. Ni, S. Ishii, A. V. Kildishev, and V. M. Shalaev, “Ultra-thin, planar, babinet-inverted plasmonic metalenses,” Light: Sci. Appl., vol. 2, no. 4, p. e72, 2013. https://doi.org/10.1038/lsa.2013.28.Suche in Google Scholar

[19] F. Aieta, M. A. Kats, P. Genevet, and F. Capasso, “Multiwavelength achromatic metasurfaces by dispersive phase compensation,” Science, vol. 347, no. 6228, pp. 1342–1345, 2015. https://doi.org/10.1126/science.aaa2494.Suche in Google Scholar PubMed

[20] M. Khorasaninejad, W. T. Chen, R. C. Devlin, J. Oh, A. Y. Zhu, and F. Capasso, “Metalenses at visible wavelengths: diffraction-limited focusing and subwavelength resolution imaging,” Science, vol. 352, no. 6290, pp. 1190–1194, 2016. https://doi.org/10.1126/science.aaf6644.Suche in Google Scholar PubMed

[21] R. Paniagua-Domínguez, et al.., “A metalens with a near-unity numerical aperture,” Nano Lett., vol. 18, no. 3, pp. 2124–2132, 2018. https://doi.org/10.1021/acs.nanolett.8b00368.Suche in Google Scholar PubMed

[22] M. L. Tseng, et al.., “Metalenses: advances and applications,” Adv. Opt. Mater., vol. 6, no. 18, p. 1800554, 2018. https://doi.org/10.1002/adom.201800554.Suche in Google Scholar

[23] A. Arbabi and A. Faraon, “Advances in optical metalenses,” Nat. Photonics, vol. 17, no. 1, pp. 16–25, 2023. https://doi.org/10.1038/s41566-022-01108-6.Suche in Google Scholar

[24] S. Colburn, A. Zhan, and A. Majumdar, “Metasurface optics for full-color computational imaging,” Sci. Adv., vol. 4, no. 2, p. eaar2114, 2018. https://doi.org/10.1126/sciadv.aar2114.Suche in Google Scholar PubMed PubMed Central

[25] Z.-B. Fan, et al.., “A broadband achromatic metalens array for integral imaging in the visible,” Light: Sci. Appl., vol. 8, no. 1, p. 67, 2019. https://doi.org/10.1038/s41377-019-0178-2.Suche in Google Scholar PubMed PubMed Central

[26] P. Zheng, et al.., “Metasurface-based key for computational imaging encryption,” Sci. Adv., vol. 7, no. 21, p. eabg0363, 2021. https://doi.org/10.1126/sciadv.abg0363.Suche in Google Scholar PubMed PubMed Central

[27] Z. Lin, et al.., “End-to-end metasurface inverse design for single-shot multi-channel imaging,” Opt. Express, vol. 30, no. 16, pp. 28358–28370, 2022. https://doi.org/10.1364/oe.449985.Suche in Google Scholar PubMed

[28] H. Zheng, Q. Liu, I. I. Kravchenko, X. Zhang, Y. Huo, and J. G. Valentine, “Multichannel meta-imagers for accelerating machine vision,” Nat. Nanotechnol., vol. 19, no. 4, pp. 471–478, 2024. https://doi.org/10.1038/s41565-023-01557-2.Suche in Google Scholar PubMed PubMed Central

[29] 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.Suche in Google Scholar

[30] H. Ren, et al.., “Metasurface orbital angular momentum holography,” Nat. Commun., vol. 10, no. 1, p. 2986, 2019. https://doi.org/10.1038/s41467-019-11030-1.Suche in Google Scholar PubMed PubMed Central

[31] N. A. Rubin, G. D’Aversa, P. Chevalier, Z. Shi, W. T. Chen, and F. Capasso, “Matrix Fourier optics enables a compact full-Stokes polarization camera,” Science, vol. 365, no. 6448, p. eaax1839, 2019. https://doi.org/10.1126/science.aax1839.Suche in Google Scholar PubMed

[32] A. Zaidi, et al.., “Metasurface-enabled single-shot and complete Mueller matrix imaging,” Nat. Photonics, vol. 18, no. 7, pp. 704–712, 2024. https://doi.org/10.1038/s41566-024-01426-x.Suche in Google Scholar

[33] S. M. Kamali, E. Arbabi, A. Arbabi, and A. Faraon, “A review of dielectric optical metasurfaces for wavefront control,” Nanophotonics, vol. 7, no. 6, pp. 1041–1068, 2018. https://doi.org/10.1515/nanoph-2017-0129.Suche in Google Scholar

[34] 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.Suche in Google Scholar PubMed

[35] A. H. Dorrah and F. Capasso, “Tunable structured light with flat optics,” Science, vol. 376, no. 6591, p. eabi6860, 2022. https://doi.org/10.1126/science.abi6860.Suche in Google Scholar PubMed

[36] Q. Yuan, et al.., “Recent advanced applications of metasurfaces in multi-dimensions,” Nanophotonics, vol. 12, no. 13, pp. 2295–2315, 2023. https://doi.org/10.1515/nanoph-2022-0803.Suche in Google Scholar PubMed PubMed Central

[37] 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.Suche in Google Scholar PubMed

[38] R. Sokhoyan, C. U. Hail, M. Foley, M. Y. Grajower, and H. A. Atwater, “All-dielectric high-Q dynamically tunable transmissive metasurfaces,” Laser Photonics Rev., vol. 18, no. 6, p. 2300980, 2024. https://doi.org/10.1002/lpor.202300980.Suche in Google Scholar

[39] Y.-W. Huang, et al.., “Gate-tunable conducting oxide metasurfaces,” Nano Lett., vol. 16, no. 9, pp. 5319–5325, 2016. https://doi.org/10.1021/acs.nanolett.6b00555.Suche in Google Scholar PubMed

[40] M. Y. Shalaginov, et al.., “Design for quality: reconfigurable flat optics based on active metasurfaces,” Nanophotonics, vol. 9, no. 11, pp. 3505–3534, 2020. https://doi.org/10.1515/nanoph-2020-0033.Suche in Google Scholar

[41] A. Komar, et al.., “Dynamic beam switching by liquid crystal tunable dielectric metasurfaces,” ACS Photonics, vol. 5, no. 5, pp. 1742–1748, 2018. https://doi.org/10.1021/acsphotonics.7b01343.Suche in Google Scholar

[42] S.-Q. Li, X. Xu, R. M. Veetil, V. Valuckas, R. Paniagua-Domínguez, and A. I. Kuznetsov, “Phase-only transmissive spatial light modulator based on tunable dielectric metasurface,” Science, vol. 364, no. 6445, pp. 1087–1090, 2019. https://doi.org/10.1126/science.aaw6747.Suche in Google Scholar PubMed

[43] T. Gu, H. J. Kim, C. Rivero-Baleine, and J. Hu, “Reconfigurable metasurfaces towards commercial success,” Nat. Photonics, vol. 17, no. 1, pp. 48–58, 2023. https://doi.org/10.1038/s41566-022-01099-4.Suche in Google Scholar

[44] Y. Zhang, et al.., “Electrically reconfigurable non-volatile metasurface using low-loss optical phase-change material,” Nat. Nanotechnol., vol. 16, no. 6, pp. 661–666, 2021. https://doi.org/10.1038/s41565-021-00881-9.Suche in Google Scholar PubMed

[45] E. Klopfer, S. Dagli, D. Barton III, M. Lawrence, and J. A. Dionne, “High-quality-factor silicon-on-lithium niobate metasurfaces for electro-optically reconfigurable wavefront shaping,” Nano Lett., vol. 22, no. 4, pp. 1703–1709, 2022. https://doi.org/10.1021/acs.nanolett.1c04723.Suche in Google Scholar PubMed

[46] E. Klopfer, H. C. Delgado, S. Dagli, M. Lawrence, and J. A. Dionne, “A thermally controlled high-Q metasurface lens,” Appl. Phys. Lett., vol. 122, no. 22, p. 221701, 2023. https://doi.org/10.1063/5.0152535.Suche in Google Scholar

[47] J. Sisler, P. Thureja, M. Y. Grajower, R. Sokhoyan, I. Huang, and H. A. Atwater, “Electrically tunable space–time metasurfaces at optical frequencies,” Nat. Nanotechnol., vol. 19, no. 10, pp. 1491–1498, 2024. https://doi.org/10.1038/s41565-024-01728-9.Suche in Google Scholar PubMed

[48] R. C. Devlin, M. Khorasaninejad, W. T. Chen, J. Oh, and F. Capasso, “Broadband high-efficiency dielectric metasurfaces for the visible spectrum,” Proc. Natl. Acad. Sci. U. S. A., vol. 113, no. 38, pp. 10473–10478, 2016. https://doi.org/10.1073/pnas.1611740113.Suche in Google Scholar PubMed PubMed Central

[49] A. She, S. Zhang, S. Shian, D. R. Clarke, and F. Capasso, “Large area metalenses: design, characterization, and mass manufacturing,” Opt. Express, vol. 26, no. 2, pp. 1573–1585, 2018. https://doi.org/10.1364/oe.26.001573.Suche in Google Scholar PubMed

[50] V.-C. Su, C. H. Chu, G. Sun, and D. P. Tsai, “Advances in optical metasurfaces: fabrication and applications,” Opt. Express, vol. 26, no. 10, pp. 13148–13182, 2018. https://doi.org/10.1364/oe.26.013148.Suche in Google Scholar

[51] G. Yoon, T. Tanaka, T. Zentgraf, and J. Rho, “Recent progress on metasurfaces: applications and fabrication,” J. Phys. D: Appl. Phys., vol. 54, no. 38, p. 383002, 2021. https://doi.org/10.1088/1361-6463/ac0faa.Suche in Google Scholar

[52] A. F. Koenderink, A. Alù, and A. Polman, “Nanophotonics: shrinking light-based technology,” Science, vol. 348, no. 6234, pp. 516–521, 2015. https://doi.org/10.1126/science.1261243.Suche in Google Scholar PubMed

[53] Y. Yang, et al.., “Integrated metasurfaces for re-envisioning a near-future disruptive optical platform,” Light: Sci. Appl., vol. 12, no. 1, p. 152, 2023. https://doi.org/10.1038/s41377-023-01169-4.Suche in Google Scholar PubMed PubMed Central

[54] F. T. Chen and H. G. Craighead, “Diffractive phase elements based on two-dimensional artificial dielectrics,” Opt. Lett., vol. 20, no. 2, pp. 121–123, 1995. https://doi.org/10.1364/ol.20.000121.Suche in Google Scholar PubMed

[55] F. T. Chen and H. G. Craighead, “Diffractive lens fabricated with mostly zeroth-order gratings,” Opt. Lett., vol. 21, no. 3, pp. 177–179, 1996. https://doi.org/10.1364/ol.21.000177.Suche in Google Scholar PubMed

[56] P. Lalanne, S. Astilean, P. Chavel, E. Cambril, and H. Launois, “Blazed binary subwavelength gratings with efficiencies larger than those of conventional échelette gratings,” Opt. Lett., vol. 23, no. 14, pp. 1081–1083, 1998. https://doi.org/10.1364/ol.23.001081.Suche in Google Scholar PubMed

[57] P. Lalanne, S. Astilean, P. Chavel, E. Cambril, and H. Launois, “Design and fabrication of blazed binary diffractive elements with sampling periods smaller than the structural cutoff,” J. Opt. Soc. Am. A, vol. 16, no. 5, pp. 1143–1156, 1999. https://doi.org/10.1364/josaa.16.001143.Suche in Google Scholar

[58] Z. Bomzon, G. Biener, V. Kleiner, and E. Hasman, “Space-variant Pancharatnam–Berry phase optical elements with computer-generated subwavelength gratings,” Opt. Lett., vol. 27, no. 13, pp. 1141–1143, 2002. https://doi.org/10.1364/ol.27.001141.Suche in Google Scholar PubMed

[59] Z. Bomzon, G. Biener, V. Kleiner, and E. Hasman, “Radially and azimuthally polarized beams generated by space-variant dielectric subwavelength gratings,” Opt. Lett., vol. 27, no. 5, pp. 285–287, 2002. https://doi.org/10.1364/ol.27.000285.Suche in Google Scholar PubMed

[60] H. Kikuta, H. Toyota, and W. Yu, “Optical elements with subwavelength structured surfaces,” Opt. Rev., vol. 10, no. 2, pp. 63–73, 2003. https://doi.org/10.1007/s10043-003-0063-2.Suche in Google Scholar

[61] D. R. Smith, J. B. Pendry, and M. C. K. Wiltshire, “Metamaterials and negative refractive index,” Science, vol. 305, no. 5685, pp. 788–792, 2004. https://doi.org/10.1126/science.1096796.Suche in Google Scholar PubMed

[62] R. W. Ziolkowski and N. Engheta, “Introduction, History, and Selected Topics in Fundamental Theories of Metamaterials,” in Metamaterials, Hoboken, NJ, John Wiley Sons, Ltd, 2006, pp. 1–41.10.1002/0471784192.ch1Suche in Google Scholar

[63] N. I. Zheludev and Y. S. Kivshar, “From metamaterials to metadevices,” Nat. Mater., vol. 11, no. 11, pp. 917–924, 2012. https://doi.org/10.1038/nmat3431.Suche in Google Scholar PubMed

[64] A. M. Urbas, et al.., “Roadmap on optical metamaterials,” J. Opt., vol. 18, no. 9, p. 093005, 2016. https://doi.org/10.1088/2040-8978/18/9/093005.Suche in Google Scholar

[65] S. Xiao, T. Wang, T. Liu, C. Zhou, X. Jiang, and J. Zhang, “Active metamaterials and metadevices: a review,” J. Phys. D: Appl. Phys., vol. 53, no. 50, p. 503002, 2020. https://doi.org/10.1088/1361-6463/abaced.Suche in Google Scholar

[66] M. Kadic, G. W. Milton, M. van Hecke, and M. Wegener, “3D metamaterials,” Nat. Rev. Phys., vol. 1, no. 3, pp. 198–210, 2019. https://doi.org/10.1038/s42254-018-0018-y.Suche in Google Scholar

[67] F. Monticone, N. M. Estakhri, and A. Alù, “Full control of nanoscale optical transmission with a composite metascreen,” Phys. Rev. Lett., vol. 110, no. 20, p. 203903, 2013. https://doi.org/10.1103/physrevlett.110.203903.Suche in Google Scholar PubMed

[68] N. Mohammadi Estakhri and A. Alù, “Wave-front transformation with gradient metasurfaces,” Phys. Rev. X, vol. 6, no. 4, p. 041008, 2016. https://doi.org/10.1103/physrevx.6.041008.Suche in Google Scholar

[69] A. Arbabi and A. Faraon, “Fundamental limits of ultrathin metasurfaces,” Sci. Rep., vol. 7, no. 1, p. 43722, 2017. https://doi.org/10.1038/srep43722.Suche in Google Scholar PubMed PubMed Central

[70] A. Epstein and G. V. Eleftheriades, “Synthesis of passive lossless metasurfaces using auxiliary fields for reflectionless beam splitting and perfect reflection,” Phys. Rev. Lett., vol. 117, no. 25, p. 256103, 2016. https://doi.org/10.1103/physrevlett.117.256103.Suche in Google Scholar PubMed

[71] B. Memarzadeh and H. Mosallaei, “Array of planar plasmonic scatterers functioning as light concentrator,” Opt. Lett., vol. 36, no. 13, pp. 2569–2571, 2011. https://doi.org/10.1364/ol.36.002569.Suche in Google Scholar

[72] F. Qin, et al.., “Hybrid bilayer plasmonic metasurface efficiently manipulates visible light,” Sci. Adv., vol. 2, no. 1, p. e1501168, 2016. https://doi.org/10.1126/sciadv.1501168.Suche in Google Scholar PubMed PubMed Central

[73] A. Pors and S. I. Bozhevolnyi, “Plasmonic metasurfaces for efficient phase control in reflection,” Opt. Express, vol. 21, no. 22, pp. 27438–27451, 2013. https://doi.org/10.1364/oe.21.027438.Suche in Google Scholar PubMed

[74] Y. Huang, et al.., “Phase-gradient gap-plasmon metasurface based blazed grating for real time dispersive imaging,” Appl. Phys. Lett., vol. 104, no. 16, p. 161106, 2014. https://doi.org/10.1063/1.4872170.Suche in Google Scholar

[75] G. Zheng, H. Mühlenbernd, M. Kenney, G. Li, T. Zentgraf, and S. Zhang, “Metasurface holograms reaching 80 % efficiency,” Nat. Nanotechnol., vol. 10, no. 4, pp. 308–312, 2015. https://doi.org/10.1038/nnano.2015.2.Suche in Google Scholar PubMed

[76] G. Castaldi, V. Galdi, A. Alù, and N. Engheta, “Nonlocal transformation optics,” Phys. Rev. Lett., vol. 108, no. 6, p. 063902, 2012. https://doi.org/10.1103/physrevlett.108.063902.Suche in Google Scholar PubMed

[77] A. Silva, F. Monticone, G. Castaldi, V. Galdi, A. Alù, and N. Engheta, “Performing mathematical operations with metamaterials,” Science, vol. 343, no. 6167, pp. 160–163, 2014. https://doi.org/10.1126/science.1242818.Suche in Google Scholar PubMed

[78] H. Kwon, D. Sounas, A. Cordaro, A. Polman, and A. Alù, “Nonlocal metasurfaces for optical signal processing,” Phys. Rev. Lett., vol. 121, no. 17, p. 173004, 2018. https://doi.org/10.1103/physrevlett.121.173004.Suche in Google Scholar

[79] S. C. Malek, A. C. Overvig, A. Alù, and N. Yu, “Multifunctional resonant wavefront-shaping meta-optics based on multilayer and multi-perturbation nonlocal metasurfaces,” Light: Sci. Appl., vol. 11, no. 1, p. 246, 2022. https://doi.org/10.1038/s41377-022-00905-6.Suche in Google Scholar PubMed PubMed Central

[80] C. Pfeiffer and A. Grbic, “Metamaterial huygens’ surfaces: tailoring wave fronts with reflectionless sheets,” Phys. Rev. Lett., vol. 110, no. 19, p. 197401, 2013. https://doi.org/10.1103/physrevlett.110.197401.Suche in Google Scholar PubMed

[81] A. H. Dorrah, M. Chen, and G. V. Eleftheriades, “Bianisotropic Huygens’ metasurface for wideband impedance matching between two dielectric media,” IEEE Trans. Antennas Propag., vol. 66, no. 9, pp. 4729–4742, 2018. https://doi.org/10.1109/tap.2018.2851361.Suche in Google Scholar

[82] A. H. Dorrah and G. V. Eleftheriades, “Bianisotropic Huygens’ metasurface pairs for nonlocal power-conserving wave transformations,” IEEE Antennas Wirel. Propag. Lett., vol. 17, no. 10, pp. 1788–1792, 2018. https://doi.org/10.1109/lawp.2018.2866874.Suche in Google Scholar

[83] M. E. Warren, R. E. Smith, G. A. Vawter, and J. R. Wendt, “High-efficiency subwavelength diffractive optical element in GaAs for 975 nm,” Opt. Lett., vol. 20, no. 12, pp. 1441–1443, 1995. https://doi.org/10.1364/ol.20.001441.Suche in Google Scholar PubMed

[84] S. Astilean, P. Lalanne, P. Chavel, E. Cambril, and H. Launois, “High-efficiency subwavelength diffractive element patterned in a high-refractive-index material for 633 nm,” Opt. Lett., vol. 23, no. 7, pp. 552–554, 1998. https://doi.org/10.1364/ol.23.000552.Suche in Google Scholar PubMed

[85] P. Lalanne, “Waveguiding in blazed-binary diffractive elements,” J. Opt. Soc. Am. A, vol. 16, no. 10, pp. 2517–2520, 1999. https://doi.org/10.1364/josaa.16.002517.Suche in Google Scholar

[86] A. Arbabi, Y. Horie, A. J. Ball, M. Bagheri, and A. Faraon, “Subwavelength-thick lenses with high numerical apertures and large efficiency based on high-contrast transmitarrays,” Nat. Commun., vol. 6, no. 1, p. 7069, 2015. https://doi.org/10.1038/ncomms8069.Suche in Google Scholar PubMed

[87] A. Arbabi, Y. Horie, M. Bagheri, and A. Faraon, “Dielectric metasurfaces for complete control of phase and polarization with subwavelength spatial resolution and high transmission,” Nat. Nanotechnol., vol. 10, no. 11, pp. 937–943, 2015. https://doi.org/10.1038/nnano.2015.186.Suche in Google Scholar PubMed

[88] M. Decker, 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.Suche in Google Scholar

[89] M. Khorasaninejad and F. Capasso, “Broadband multifunctional efficient meta-gratings based on dielectric waveguide phase shifters,” Nano Lett., vol. 15, no. 10, pp. 6709–6715, 2015. https://doi.org/10.1021/acs.nanolett.5b02524.Suche in Google Scholar PubMed

[90] J. P. Balthasar Mueller, N. A. Rubin, R. C. Devlin, B. Groever, and F. Capasso, “Metasurface polarization optics: independent phase control of arbitrary orthogonal states of polarization,” Phys. Rev. Lett., vol. 118, no. 11, p. 113901, 2017. https://doi.org/10.1103/physrevlett.118.113901.Suche in Google Scholar

[91] H. Zheng, M. He, Y. Zhou, I. I. Kravchenko, J. D. Caldwell, and J. G. Valentine, “Compound meta-optics for complete and loss-less field control,” ACS Nano, vol. 16, no. 9, pp. 15100–15107, 2022. https://doi.org/10.1021/acsnano.2c06248.Suche in Google Scholar PubMed

[92] B. O. Raeker, H. Zheng, Y. Zhou, I. I. Kravchenko, J. Valentine, and A. Grbic, “All-dielectric meta-optics for high-efficiency independent amplitude and phase manipulation,” Adv. Photonics Res., vol. 3, no. 3, p. 2100285, 2022. https://doi.org/10.1002/adpr.202100285.Suche in Google Scholar

[93] B. Mirzapourbeinekalaye, A. McClung, and A. Arbabi, “General lossless polarization and phase transformation using bilayer metasurfaces,” Adv. Opt. Mater., vol. 10, no. 11, p. 2102591, 2022. https://doi.org/10.1002/adom.202102591.Suche in Google Scholar

[94] F. Presutti and F. Monticone, “Focusing on bandwidth: achromatic metalens limits,” Optica, vol. 7, no. 6, pp. 624–631, 2020. https://doi.org/10.1364/optica.389404.Suche in Google Scholar

[95] J. Engelberg and U. Levy, “Achromatic flat lens performance limits,” Optica, vol. 8, no. 6, pp. 834–845, 2021. https://doi.org/10.1364/optica.422843.Suche in Google Scholar

[96] F. Yang, et al.., “Wide field-of-view metalens: a tutorial,” Adv. Photonics, vol. 5, no. 3, p. 033001, 2023. https://doi.org/10.1117/1.ap.5.3.033001.Suche in Google Scholar

[97] K. Shastri and F. Monticone, “Bandwidth bounds for wide-field-of-view dispersion-engineered achromatic metalenses,” EPJ Appl. Metamater., vol. 9, no. 16, pp. 1–7, 2022. https://doi.org/10.1051/epjam/2022012.Suche in Google Scholar

[98] A. Wirth-Singh, et al.., “Wide field of view large aperture meta-doublet eyepiece,” Light: Sci. Appl., vol. 14, no. 1, p. 17, 2025. https://doi.org/10.1038/s41377-024-01674-0.Suche in Google Scholar PubMed PubMed Central

[99] D. A. B. Miller, “Why optics needs thickness,” Science, vol. 379, no. 6627, pp. 41–45, 2023. https://doi.org/10.1126/science.ade3395.Suche in Google Scholar PubMed

[100] D. A. B. Miller, “Fundamental limit to linear one-dimensional slow light structures,” Phys. Rev. Lett., vol. 99, no. 20, p. 203903, 2007. https://doi.org/10.1103/physrevlett.99.203903.Suche in Google Scholar

[101] J. Oh, et al.., “Adjoint-optimized metasurfaces for compact mode-division multiplexing,” ACS Photonics, vol. 9, no. 3, pp. 929–937, 2022. https://doi.org/10.1021/acsphotonics.1c01744.Suche in Google Scholar PubMed PubMed Central

[102] Y. Luo, et al.., “Varifocal metalens for optical sectioning fluorescence microscopy,” Nano Lett., vol. 21, no. 12, pp. 5133–5142, 2021. https://doi.org/10.1021/acs.nanolett.1c01114.Suche in Google Scholar PubMed

[103] J. C. Zhang, et al.., “Miniature tunable Airy beam optical meta-device,” Opto-Electron. Adv., vol. 7, no. 2, p. 230171, 2024. https://doi.org/10.29026/oea.2024.230171.Suche in Google Scholar

[104] A. Wirth-Singh, et al.., “Meta-optics triplet for zoom imaging at mid-wave infrared,” Appl. Phys. Lett., vol. 125, no. 21, p. 211705, 2024. https://doi.org/10.1063/5.0227368.Suche in Google Scholar

[105] J. C. Zhang, et al.., “A 6G meta-device for 3D varifocal,” Sci. Adv., vol. 9, no. 4, p. eadf8478, 2023. https://doi.org/10.1126/sciadv.adf8478.Suche in Google Scholar PubMed PubMed Central

[106] L. Li, et al.., “Metalens-array–based high-dimensional and multiphoton quantum source,” Science, vol. 368, no. 6498, pp. 1487–1490, 2020. https://doi.org/10.1126/science.aba9779.Suche in Google Scholar PubMed

[107] A. H. Dorrah, J.-S. Park, A. Palmieri, and F. Capasso, “Free-standing bilayer metasurfaces in the visible,” Nat Commun., vol. 16 , no. 1, p. 3126, 2025. https://doi.org/10.1038/s41467-025-58205-7Suche in Google Scholar PubMed PubMed Central

[108] N. A. Rubin, Z. Shi, and F. Capasso, “Polarization in diffractive optics and metasurfaces,” Adv. Opt. Photonics, vol. 13, no. 4, pp. 836–970, 2021. https://doi.org/10.1364/aop.439986.Suche in Google Scholar

[109] W. T. Chen and F. Capasso, “Will flat optics appear in everyday life anytime soon?” Appl. Phys. Lett., vol. 118, no. 10, p. 100503, 2021. https://doi.org/10.1063/5.0039885.Suche in Google Scholar

[110] M. L. Brongersma, et al.., “The second optical metasurface revolution: moving from science to technology,” Nat. Rev. Electr. Eng., vol. 2, no. 2, pp. 125–143, 2025. https://doi.org/10.1038/s44287-024-00136-4.Suche in Google Scholar

[111] R. C. Jones, “A new calculus for the treatment of optical systemsi. description and discussion of the calculus,” J. Opt. Soc. Am., vol. 31, no. 7, pp. 488–493, 1941. https://doi.org/10.1364/josa.31.000488.Suche in Google Scholar

[112] E. Arbabi, S. M. Kamali, A. Arbabi, and A. Faraon, “Full-Stokes imaging polarimetry using dielectric metasurfaces,” ACS Photonics, vol. 5, no. 8, pp. 3132–3140, 2018. https://doi.org/10.1021/acsphotonics.8b00362.Suche in Google Scholar

[113] Z. Shi, et al.., “Continuous angle-tunable birefringence with freeform metasurfaces for arbitrary polarization conversion,” Sci. Adv., vol. 6, no. 23, p. eaba3367, 2020. https://doi.org/10.1126/sciadv.aba3367.Suche in Google Scholar PubMed PubMed Central

[114] A. H. Dorrah, N. A. Rubin, A. Zaidi, M. Tamagnone, and F. Capasso, “Metasurface optics for on-demand polarization transformations along the optical path,” Nat. Photonics, vol. 15, no. 4, pp. 287–296, 2021. https://doi.org/10.1038/s41566-020-00750-2.Suche in Google Scholar

[115] A. Zaidi, N. A. Rubin, A. H. Dorrah, J.-S. Park, and F. Capasso, “Generalized polarization transformations with metasurfaces,” Opt. Express, vol. 29, no. 24, pp. 39065–39078, 2021. https://doi.org/10.1364/oe.442844.Suche in Google Scholar PubMed

[116] N. A. Rubin, A. Zaidi, A. H. Dorrah, Z. Shi, and F. Capasso, “Jones matrix holography with metasurfaces,” Sci. Adv., vol. 7, no. 33, p. eabg7488, 2021. https://doi.org/10.1126/sciadv.abg7488.Suche in Google Scholar PubMed PubMed Central

[117] E. Arbabi, S. M. Kamali, A. Arbabi, and A. Faraon, “Vectorial holograms with a dielectric metasurface: ultimate polarization pattern generation,” ACS Photonics, vol. 6, no. 11, pp. 2712–2718, 2019. https://doi.org/10.1021/acsphotonics.9b00678.Suche in Google Scholar

[118] Q. Song, X. Liu, C.-W. Qiu, and P. Genevet, “Vectorial metasurface holography,” Appl. Phys. Rev., vol. 9, no. 1, p. 011311, 2022. https://doi.org/10.1063/5.0078610.Suche in Google Scholar

[119] T. Chang, et al.., “Universal metasurfaces for complete linear control of coherent light transmission,” Adv. Mater., vol. 34, no. 44, p. 2204085, 2022. https://doi.org/10.1002/adma.202204085.Suche in Google Scholar PubMed

[120] Z. Shi, N. A. Rubin, J.-S. Park, and F. Capasso, “Nonseparable polarization wavefront transformation,” Phys. Rev. Lett., vol. 129, no. 16, p. 167403, 2022. https://doi.org/10.1103/physrevlett.129.167403.Suche in Google Scholar PubMed

[121] A. Palmieri, A. H. Dorrah, J. Yang, J. Oh, P. Dainese, and F. Capasso, “Do dielectric bilayer metasurfaces behave as a stack of decoupled single-layer metasurfaces?,” Opt. Express, vol. 32, no. 5, pp. 8146–8159, 2024. https://doi.org/10.1364/oe.505401.Suche in Google Scholar PubMed

[122] Q. Fan, et al.., “Independent amplitude control of arbitrary orthogonal states of polarization via dielectric metasurfaces,” Phys. Rev. Lett., vol. 125, no. 26, p. 267402, 2020. https://doi.org/10.1103/physrevlett.125.267402.Suche in Google Scholar

[123] Y.-W. Huang, et al.., “Versatile total angular momentum generation using cascaded J-plates,” Opt. Express, vol. 27, no. 5, pp. 7469–7484, 2019. https://doi.org/10.1364/oe.27.007469.Suche in Google Scholar

[124] Y. Bao, F. Nan, J. Yan, X. Yang, C.-W. Qiu, and B. Li, “Observation of full-parameter Jones matrix in bilayer metasurface,” Nat. Commun., vol. 13, no. 1, p. 7550, 2022. https://doi.org/10.1038/s41467-022-35313-2.Suche in Google Scholar PubMed PubMed Central

[125] E. Cohen, H. Larocque, F. Bouchard, F. Nejadsattari, Y. Gefen, and E. Karimi, “Geometric phase from Aharonov–Bohm to Pancharatnam–Berry and beyond,” Nat. Rev. Phys., vol. 1, no. 7, pp. 437–449, 2019. https://doi.org/10.1038/s42254-019-0071-1.Suche in Google Scholar

[126] A. H. Dorrah, M. Tamagnone, N. A. Rubin, A. Zaidi, and F. Capasso, “Introducing Berry phase gradients along the optical path via propagation-dependent polarization transformations,” Nanophotonics, vol. 11, no. 4, pp. 713–725, 2022. https://doi.org/10.1515/nanoph-2021-0560.Suche in Google Scholar PubMed PubMed Central

[127] E. W. Wang, T. Phan, S.-J. Yu, S. Dhuey, and J. A. Fan, “Dynamic circular birefringence response with fractured geometric phase metasurface systems,” Proc. Natl. Acad. Sci. U. S. A., vol. 119, no. 12, p. e2122085119, 2022. https://doi.org/10.1073/pnas.2122085119.Suche in Google Scholar PubMed PubMed Central

[128] J. R. Nolen, A. C. Overvig, M. Cotrufo, and A. Alù, “Local control of polarization and geometric phase in thermal metasurfaces,” Nat. Nanotechnol., vol. 19, no. 11, pp. 1627–1634, 2024. https://doi.org/10.1038/s41565-024-01763-6.Suche in Google Scholar PubMed

[129] C. He, H. He, J. Chang, B. Chen, H. Ma, and M. J. Booth, “Polarisation optics for biomedical and clinical applications: a review,” Light: Sci. Appl., vol. 10, no. 1, p. 194, 2021. https://doi.org/10.1038/s41377-021-00639-x.Suche in Google Scholar PubMed PubMed Central

[130] F. Snik, et al.., “An overview of polarimetric sensing techniques and technology with applications to different research fields,” in Polarization: Measurement, Analysis, and Remote Sensing XI, vol. 9099, D. B. Chenault, and D. H. Goldstein, Eds., International Society for Optics and Photonics, SPIE, 2014, p. 90990B.10.1117/12.2053245Suche in Google Scholar

[131] J.-S. Lee and E. Pottier, Polarimetric Radar Imaging: From Basics to Applications, 1st ed. Boca Raton, CRC Press, 2009.10.1201/9781420054989-1Suche in Google Scholar

[132] T. Zhan, K. Yin, J. Xiong, Z. He, and S.-T. Wu, “Augmented reality and virtual reality displays: perspectives and challenges,” iScience, vol. 23, no. 8, p. 101397, 2020. https://doi.org/10.1016/j.isci.2020.101397.Suche in Google Scholar PubMed PubMed Central

[133] J. Xiong, E.-L. Hsiang, Z. He, T. Zhan, and S.-T. Wu, “Augmented reality and virtual reality displays: emerging technologies and future perspectives,” Light: Sci. Appl., vol. 10, no. 1, p. 216, 2021. https://doi.org/10.1038/s41377-021-00658-8.Suche in Google Scholar PubMed PubMed Central

[134] A. Z. Goldberg, et al.., “Quantum concepts in optical polarization,” Adv. Opt. Photonics, vol. 13, no. 1, pp. 1–73, 2021. https://doi.org/10.1364/aop.404175.Suche in Google Scholar

[135] A. Luis, “Degree of polarization in quantum optics,” Phys. Rev. A, vol. 66, no. 1, p. 013806, 2002. https://doi.org/10.1103/physreva.66.013806.Suche in Google Scholar

[136] Z. Wu, L. Li, Y. Li, and X. Chen, “Metasurface superstrate antenna with wideband circular polarization for satellite communication application,” IEEE Antennas Wirel. Propag. Lett., vol. 15, pp. 374–377, 2016. https://doi.org/10.1109/lawp.2015.2446505.Suche in Google Scholar

[137] A. Z. Ashoor and O. M. Ramahi, “Polarization-independent cross-dipole energy harvesting surface,” IEEE Trans. Microwave Theory Tech., vol. 67, no. 3, pp. 1130–1137, 2019. https://doi.org/10.1109/tmtt.2018.2885754.Suche in Google Scholar

[138] M. Khorasaninejad, et al.., “Achromatic metalens over 60 nm bandwidth in the visible and metalens with reverse chromatic dispersion,” Nano Lett., vol. 17, no. 3, pp. 1819–1824, 2017. https://doi.org/10.1021/acs.nanolett.6b05137.Suche in Google Scholar PubMed

[139] S. Wang, et al.., “Broadband achromatic optical metasurface devices,” Nat. Commun., vol. 8, no. 1, p. 187, 2017. https://doi.org/10.1038/s41467-017-00166-7.Suche in Google Scholar PubMed PubMed Central

[140] X. Li, M. Pu, X. Ma, Y. Guo, P. Gao, and X. Luo, “Dispersion engineering in metamaterials and metasurfaces,” J. Phys. D: Appl. Phys., vol. 51, no. 5, p. 054002, 2018. https://doi.org/10.1088/1361-6463/aaa233.Suche in Google Scholar

[141] S. Shrestha, A. C. Overvig, M. Lu, A. Stein, and N. Yu, “Broadband achromatic dielectric metalenses,” Light: Sci. Appl., vol. 7, no. 1, p. 85, 2018. https://doi.org/10.1038/s41377-018-0078-x.Suche in Google Scholar PubMed PubMed Central

[142] W. T. Chen, A. Y. Zhu, J. Sisler, Z. Bharwani, and F. Capasso, “A broadband achromatic polarization-insensitive metalens consisting of anisotropic nanostructures,” Nat. Commun., vol. 10, no. 1, p. 355, 2019. https://doi.org/10.1038/s41467-019-08305-y.Suche in Google Scholar PubMed PubMed Central

[143] A. Ndao, L. Hsu, J. Ha, J.-H. Park, C. Chang-Hasnain, and B. Kanté, “Octave bandwidth photonic fishnet-achromatic-metalens,” Nat. Commun., vol. 11, no. 1, p. 3205, 2020. https://doi.org/10.1038/s41467-020-17015-9.Suche in Google Scholar PubMed PubMed Central

[144] W. T. Chen, A. Y. Zhu, and F. Capasso, “Flat optics with dispersion-engineered metasurfaces,” Nat. Rev. Mater., vol. 5, no. 8, pp. 604–620, 2020. https://doi.org/10.1038/s41578-020-0203-3.Suche in Google Scholar

[145] Y. Wang, et al.., “High-efficiency broadband achromatic metalens for near-IR biological imaging window,” Nat. Commun., vol. 12, no. 1, p. 5560, 2021. https://doi.org/10.1038/s41467-021-25797-9.Suche in Google Scholar PubMed PubMed Central

[146] Z. Li, et al.., “Meta-optics achieves RGB-achromatic focusing for virtual reality,” Sci. Adv., vol. 7, no. 5, p. eabe4458, 2021. https://doi.org/10.1126/sciadv.abe4458.Suche in Google Scholar PubMed PubMed Central

[147] M. J. Booth, “Adaptive optics in microscopy,” Philos. Trans. R. Soc., A, vol. 365, no. 1861, pp. 2829–2843, 2007. https://doi.org/10.1098/rsta.2007.0013.Suche in Google Scholar PubMed

[148] T. Stone and N. George, “Hybrid diffractive-refractive lenses and achromats,” Appl. Opt., vol. 27, no. 14, pp. 2960–2971, 1988. https://doi.org/10.1364/ao.27.002960.Suche in Google Scholar

[149] S. D. Campbell, D. E. Brocker, J. Nagar, and D. H. Werner, “SWaP reduction regimes in achromatic grin singlets,” Appl. Opt., vol. 55, no. 13, pp. 3594–3598, 2016. https://doi.org/10.1364/ao.55.003594.Suche in Google Scholar PubMed

[150] S. Wang, et al.., “A broadband achromatic metalens in the visible,” Nat. Nanotechnol., vol. 13, no. 3, pp. 227–232, 2018. https://doi.org/10.1038/s41565-017-0052-4.Suche in Google Scholar PubMed

[151] W. T. Chen, et al.., “A broadband achromatic metalens for focusing and imaging in the visible,” Nat. Nanotechnol., vol. 13, no. 3, pp. 220–226, 2018. https://doi.org/10.1038/s41565-017-0034-6.Suche in Google Scholar PubMed

[152] W. T. Chen, et al.., “Broadband achromatic metasurface-refractive optics,” Nano Lett., vol. 18, no. 12, pp. 7801–7808, 2018. https://doi.org/10.1021/acs.nanolett.8b03567.Suche in Google Scholar PubMed

[153] S. J. Byrnes, A. Lenef, F. Aieta, and F. Capasso, “Designing large, high-efficiency, high-numerical-aperture, transmissive meta-lenses for visible light,” Opt. Express, vol. 24, no. 5, pp. 5110–5124, 2016. https://doi.org/10.1364/oe.24.005110.Suche in Google Scholar PubMed

[154] Y. Zhou, C. Mao, E. Gershnabel, M. Chen, and J. A. Fan, “Large-area, high-numerical-aperture, freeform metasurfaces,” Laser Photonics Rev., vol. 18, no. 6, p. 2300988, 2024. https://doi.org/10.1002/lpor.202300988.Suche in Google Scholar

[155] P. Dainese, et al.., “Shape optimization for high efficiency metasurfaces: theory and implementation,” Light: Sci. Appl., vol. 13, no. 1, p. 300, 2024. https://doi.org/10.1038/s41377-024-01629-5.Suche in Google Scholar PubMed PubMed Central

[156] J. Jensen and O. Sigmund, “Topology optimization for nano-photonics,” Laser Photonics Rev., vol. 5, no. 2, pp. 308–321, 2011. https://doi.org/10.1002/lpor.201000014.Suche in Google Scholar

[157] O. Yesilyurt, Z. A. Kudyshev, A. Boltasseva, V. M. Shalaev, and A. V. Kildishev, “Efficient topology-optimized couplers for on-chip single-photon sources,” ACS Photonics, vol. 8, no. 10, pp. 3061–3068, 2021. https://doi.org/10.1021/acsphotonics.1c01070.Suche in Google Scholar

[158] T. Phan, et al.., “High-efficiency, large-area, topology-optimized metasurfaces,” Light: Sci. Appl., vol. 8, no. 1, p. 48, 2019. https://doi.org/10.1038/s41377-019-0159-5.Suche in Google Scholar PubMed PubMed Central

[159] S. D. Campbell, D. Sell, R. P. Jenkins, E. B. Whiting, J. A. Fan, and D. H. Werner, “Review of numerical optimization techniques for meta-device design,” Opt. Mater. Express, vol. 9, no. 4, pp. 1842–1863, 2019. https://doi.org/10.1364/ome.9.001842.Suche in Google Scholar

[160] C. M. Lalau-Keraly, S. Bhargava, O. D. Miller, and E. Yablonovitch, “Adjoint shape optimization applied to electromagnetic design,” Opt. Express, vol. 21, no. 18, pp. 21693–21701, 2013. https://doi.org/10.1364/oe.21.021693.Suche in Google Scholar PubMed

[161] Z. Li, R. Pestourie, Z. Lin, S. G. Johnson, and F. Capasso, “Empowering metasurfaces with inverse design: principles and applications,” ACS Photonics, vol. 9, no. 7, pp. 2178–2192, 2022. https://doi.org/10.1021/acsphotonics.1c01850.Suche in Google Scholar

[162] R. P. Jenkins, S. D. Campbell, and D. H. Werner, “Establishing exhaustive metasurface robustness against fabrication uncertainties through deep learning,” Nanophotonics, vol. 10, no. 18, pp. 4497–4509, 2021. https://doi.org/10.1515/nanoph-2021-0428.Suche in Google Scholar

[163] Z. A. Kudyshev, A. V. Kildishev, V. M. Shalaev, and A. Boltasseva, “Machine-learning-assisted metasurface design for high-efficiency thermal emitter optimization,” Appl. Phys. Rev., vol. 7, no. 2, p. 021407, 2020. https://doi.org/10.1063/1.5134792.Suche in Google Scholar

[164] M. K. Chen, X. Liu, Y. Sun, and D. P. Tsai, “Artificial intelligence in meta-optics,” Chem. Rev., vol. 122, no. 19, pp. 15356–15413, 2022. https://doi.org/10.1021/acs.chemrev.2c00012.Suche in Google Scholar PubMed PubMed Central

[165] D. Sell, J. Yang, S. Doshay, R. Yang, and J. A. Fan, “Large-angle, multifunctional metagratings based on freeform multimode geometries,” Nano Lett., vol. 17, no. 6, pp. 3752–3757, 2017. https://doi.org/10.1021/acs.nanolett.7b01082.Suche in Google Scholar PubMed

[166] W. T. Chen, J.-S. Park, J. Marchioni, S. Millay, K. M. A. Yousef, and F. Capasso, “Dispersion-engineered metasurfaces reaching broadband 90 % relative diffraction efficiency,” Nat. Commun., vol. 14, no. 1, p. 2544, 2023. https://doi.org/10.1038/s41467-023-38185-2.Suche in Google Scholar PubMed PubMed Central

[167] X. Ni, N. K. Emani, A. V. Kildishev, A. Boltasseva, and V. M. Shalaev, “Broadband light bending with plasmonic nanoantennas,” Science, vol. 335, no. 6067, p. 427, 2012. https://doi.org/10.1126/science.1214686.Suche in Google Scholar PubMed

[168] H.-H. Hsiao, et al.., “Integrated resonant unit of metasurfaces for broadband efficiency and phase manipulation,” Adv. Opt. Mater., vol. 6, no. 12, p. 1800031, 2018. https://doi.org/10.1002/adom.201800031.Suche in Google Scholar

[169] J. Yao, R. Lin, M. K. Chen, and D. P. Tsai, “Integrated-resonant metadevices: a review,” Adv. Photonics, vol. 5, no. 2, p. 024001, 2023. https://doi.org/10.1117/1.ap.5.2.024001.Suche in Google Scholar

[170] S. Pancharatnam, “Generalized theory of interference, and its applications,” Proc. India Acad. Sci. Sect. A, vol. 44, no. 5, pp. 247–262, 1956. https://doi.org/10.1007/bf03046050.Suche in Google Scholar

[171] M. Berry, “The adiabatic phase and pancharatnam’s phase for polarized light,” J. Mod. Opt., vol. 34, no. 11, pp. 1401–1407, 1987. https://doi.org/10.1080/09500348714551321.Suche in Google Scholar

[172] Z. Li, et al.., “Heterogeneous freeform metasurfaces: a platform for advanced broadband dispersion engineering,” arXiv:2412.12028, 2024. https://doi.org/10.48550/arXiv.2412.12028.Suche in Google Scholar

[173] F. Balli, M. Sultan, S. K. Lami, and J. T. Hastings, “A hybrid achromatic metalens,” Nat. Commun., vol. 11, no. 1, p. 3892, 2020. https://doi.org/10.1038/s41467-020-17646-y.Suche in Google Scholar PubMed PubMed Central

[174] C.-F. Pan, et al.., “3D-printed multilayer structures for high–numerical aperture achromatic metalenses,” Sci. Adv., vol. 9, no. 51, p. eadj9262, 2023. https://doi.org/10.1126/sciadv.adj9262.Suche in Google Scholar PubMed PubMed Central

[175] Z. Lin and S. G. Johnson, “Overlapping domains for topology optimization of large-area metasurfaces,” Opt. Express, vol. 27, no. 22, pp. 32445–32453, 2019. https://doi.org/10.1364/oe.27.032445.Suche in Google Scholar

[176] W. Feng, et al.., “RGB achromatic metalens doublet for digital imaging,” Nano Lett., vol. 22, no. 10, pp. 3969–3975, 2022. https://doi.org/10.1021/acs.nanolett.2c00486.Suche in Google Scholar PubMed

[177] C. Kim, S.-J. Kim, and B. Lee, “Doublet metalens design for high numerical aperture and simultaneous correction of chromatic and monochromatic aberrations,” Opt. Express, vol. 28, no. 12, pp. 18059–18076, 2020. https://doi.org/10.1364/oe.387794.Suche in Google Scholar

[178] S. Chang, L. Zhang, Y. Duan, M. T. Rahman, A. Islam, and X. Ni, “Achromatic metalenses for full visible spectrum with extended group delay control via dispersion-matched layers,” Nat. Commun., vol. 15, no. 1, p. 9627, 2024. https://doi.org/10.1038/s41467-024-53701-8.Suche in Google Scholar PubMed PubMed Central

[179] O. Avayu, E. Almeida, Y. Prior, and T. Ellenbogen, “Composite functional metasurfaces for multispectral achromatic optics,” Nat. Commun., vol. 8, no. 1, p. 14992, 2017. https://doi.org/10.1038/ncomms14992.Suche in Google Scholar PubMed PubMed Central

[180] Y. Zhou, I. I. Kravchenko, H. Wang, J. R. Nolen, G. Gu, and J. Valentine, “Multilayer noninteracting dielectric metasurfaces for multiwavelength metaoptics,” Nano Lett., vol. 18, no. 12, pp. 7529–7537, 2018. https://doi.org/10.1021/acs.nanolett.8b03017.Suche in Google Scholar PubMed

[181] M. Li, S. Li, L. K. Chin, Y. Yu, D. P. Tsai, and R. Chen, “Dual-layer achromatic metalens design with an effective Abbe number,” Opt. Express, vol. 28, no. 18, pp. 26041–26055, 2020. https://doi.org/10.1364/oe.402478.Suche in Google Scholar PubMed

[182] O. D. Miller, “Photonic design: from fundamental solar cell physics to computational inverse design,” arXiv:1308.0212, 2013. https://doi.org/10.48550/arXiv.1308.0212.Suche in Google Scholar

[183] S. Molesky, Z. Lin, A. Y. Piggott, W. Jin, J. Vucković, and A. W. Rodriguez, “Inverse design in nanophotonics,” Nat. Photonics, vol. 12, no. 11, pp. 659–670, 2018. https://doi.org/10.1038/s41566-018-0246-9.Suche in Google Scholar

[184] M. P. Bendsøe and N. Kikuchi, “Generating optimal topologies in structural design using a homogenization method,” Comput. Methods Appl. Mech. Eng., vol. 71, no. 2, pp. 197–224, 1988. https://doi.org/10.1016/0045-7825(88)90086-2.Suche in Google Scholar

[185] L. Su, A. Y. Piggott, N. V. Sapra, J. Petykiewicz, and J. Vučković, “Inverse design and demonstration of a compact on-chip narrowband three-channelc wavelength demultiplexer,” ACS Photonics, vol. 5, no. 2, pp. 301–305, 2018. https://doi.org/10.1021/acsphotonics.7b00987.Suche in Google Scholar

[186] P. I. Borel, et al.., “Topology optimization and fabrication of photonic crystal structures,” Opt. Express, vol. 12, no. 9, pp. 1996–2001, 2004. https://doi.org/10.1364/opex.12.001996.Suche in Google Scholar PubMed

[187] M. Mansouree, H. Kwon, E. Arbabi, A. McClung, A. Faraon, and A. Arbabi, “Multifunctional 2.5D metastructures enabled by adjoint optimization,” Optica, vol. 7, no. 1, pp. 77–84, 2020. https://doi.org/10.1364/optica.374787.Suche in Google Scholar

[188] P. Camayd-Muñoz, C. Ballew, G. Roberts, and A. Faraon, “Multifunctional volumetric meta-optics for color and polarization image sensors,” Optica, vol. 7, no. 4, pp. 280–283, 2020. https://doi.org/10.1364/optica.384228.Suche in Google Scholar

[189] G. Roberts, et al.., “3D-patterned inverse-designed mid-infrared metaoptics,” Nat. Commun., vol. 14, no. 1, p. 2768, 2023. https://doi.org/10.1038/s41467-023-38258-2.Suche in Google Scholar PubMed PubMed Central

[190] T. Gissibl, S. Thiele, A. Herkommer, and H. Giessen, “Two-photon direct laser writing of ultracompact multi-lens objectives,” Nat. Photonics, vol. 10, no. 8, pp. 554–560, 2016. https://doi.org/10.1038/nphoton.2016.121.Suche in Google Scholar

[191] S. Thiele, C. Pruss, A. M. Herkommer, and H. Giessen, “3D printed stacked diffractive microlenses,” Opt. Express, vol. 27, no. 24, pp. 35621–35630, 2019. https://doi.org/10.1364/oe.27.035621.Suche in Google Scholar

[192] C. R. Ocier, et al.., “Direct laser writing of volumetric gradient index lenses and waveguides,” Light: Sci. Appl., vol. 9, no. 1, p. 196, 2020. https://doi.org/10.1038/s41377-020-00431-3.Suche in Google Scholar PubMed PubMed Central

[193] B. E. Bayer, Color imaging array, U.S. Patent No. 3971065, 1976.Suche in Google Scholar

[194] M. Miyata, N. Nemoto, K. Shikama, F. Kobayashi, and T. Hashimoto, “Full-color-sorting metalenses for high-sensitivity image sensors,” Optica, vol. 8, no. 12, pp. 1596–1604, 2021. https://doi.org/10.1364/optica.444255.Suche in Google Scholar

[195] S. Nishiwaki, T. Nakamura, M. Hiramoto, T. Fujii, and M.-a. Suzuki, “Efficient colour splitters for high-pixel-density image sensors,” Nat. Photonics, vol. 7, no. 3, pp. 240–246, 2013. https://doi.org/10.1038/nphoton.2012.345.Suche in Google Scholar

[196] N. Zhao, P. B. Catrysse, and S. Fan, “Perfect RGB-IR color routers for sub-wavelength size CMOS image sensor pixels,” Adv. Photonics Res., vol. 2, no. 3, p. 2000048, 2021. https://doi.org/10.1002/adpr.202000048.Suche in Google Scholar

[197] Z. Lin, V. Liu, R. Pestourie, and S. G. Johnson, “Topology optimization of freeform large-area metasurfaces,” Opt. Express, vol. 27, no. 11, pp. 15765–15775, 2019. https://doi.org/10.1364/oe.27.015765.Suche in Google Scholar

[198] H. Chung and O. D. Miller, “High-NA achromatic metalenses by inverse design,” Opt. Express, vol. 28, no. 5, pp. 6945–6965, 2020. https://doi.org/10.1364/oe.385440.Suche in Google Scholar

[199] B. Yao, et al.., “Dual-layered metasurfaces for asymmetric focusing,” Photonics Res., vol. 8, no. 6, pp. 830–843, 2020. https://doi.org/10.1364/prj.387672.Suche in Google Scholar

[200] Y. Ren, C. Jiang, and B. Tang, “Asymmetric transmission in bilayer chiral metasurfaces for both linearly and circularly polarized waves,” J. Opt. Soc. Am. B, vol. 37, no. 11, pp. 3379–3385, 2020. https://doi.org/10.1364/josab.401783.Suche in Google Scholar

[201] H. Kim, J. Jung, and J. Shin, “Bidirectional vectorial holography using bi-layer metasurfaces and its application to optical encryption,” Adv. Mater., vol. 36, no. 44, p. 2406717, 2024. https://doi.org/10.1002/adma.202406717.Suche in Google Scholar PubMed

[202] S. S. Kruk, et al.., “Asymmetric parametric generation of images with nonlinear dielectric metasurfaces,” Nat. Photonics, vol. 16, no. 8, pp. 561–565, 2022. https://doi.org/10.1038/s41566-022-01018-7.Suche in Google Scholar

[203] K. Chen, et al.., “Directional Janus metasurface,” Adv. Mater., vol. 32, no. 2, p. 1906352, 2020. https://doi.org/10.1002/adma.201906352.Suche in Google Scholar PubMed

[204] M. Cotrufo, A. Cordaro, D. L. Sounas, A. Polman, and A. Alù, “Passive bias-free non-reciprocal metasurfaces based on thermally nonlinear quasi-bound states in the continuum,” Nat. Photonics, vol. 18, no. 1, pp. 81–90, 2024. https://doi.org/10.1038/s41566-023-01333-7.Suche in Google Scholar

[205] L. Deng, et al.., “Bilayer-metasurface design, fabrication, and functionalization for full-space light manipulation,” Adv. Opt. Mater., vol. 10, no. 7, p. 2102179, 2022. https://doi.org/10.1002/adom.202102179.Suche in Google Scholar

[206] D. Xu, S. Wen, and H. Luo, “Metasurface-based optical analog computing: from fundamentals to applications,” Adv. Devices Instrum., vol. 2022, p. 0002, 2022. https://doi.org/10.34133/adi.0002.Suche in Google Scholar

[207] S. Abdollahramezani, O. Hemmatyar, and A. Adibi, “Meta-optics for spatial optical analog computing,” Nanophotonics, vol. 9, no. 13, pp. 4075–4095, 2020. https://doi.org/10.1515/nanoph-2020-0285.Suche in Google Scholar

[208] H. Zhou, C. Zhao, C. He, L. Huang, T. Man, and Y. Wan, “Optical computing metasurfaces: applications and advances,” Nanophotonics, vol. 13, no. 4, pp. 419–441, 2024. https://doi.org/10.1515/nanoph-2023-0871.Suche in Google Scholar PubMed PubMed Central

[209] F. Zangeneh-Nejad, D. L. Sounas, A. Alù, and R. Fleury, “Analogue computing with metamaterials,” Nat. Rev. Mater., vol. 6, no. 3, pp. 207–225, 2021. https://doi.org/10.1038/s41578-020-00243-2.Suche in Google Scholar

[210] V. Nikkhah, A. Pirmoradi, F. Ashtiani, B. Edwards, F. Aflatouni, and N. Engheta, “Inverse-designed low-index-contrast structures on a silicon photonics platform for vector–matrix multiplication,” Nat. Photonics, vol. 18, no. 5, pp. 501–508, 2024. https://doi.org/10.1038/s41566-024-01394-2.Suche in Google Scholar

[211] G. Labroille, B. Denolle, P. Jian, P. Genevaux, N. Treps, and J.-F. Morizur, “Efficient and mode selective spatial mode multiplexer based on multi-plane light conversion,” Opt. Express, vol. 22, no. 13, pp. 15599–15607, 2014. https://doi.org/10.1364/oe.22.015599.Suche in Google Scholar PubMed

[212] Y. Zhang and N. K. Fontaine, “Multi-plane light conversion: a practical tutorial,” arXiv:2304.11323, 2023. https://doi.org/10.48550/arXiv.2304.11323.Suche in Google Scholar

[213] N. K. Fontaine, et al.., “Photonic lanterns, 3-D waveguides, multiplane light conversion, and other components that enable space-division multiplexing,” Proc. IEEE, vol. 110, no. 11, pp. 1821–1834, 2022. https://doi.org/10.1109/jproc.2022.3207046.Suche in Google Scholar

[214] R. Tanomura, R. Tang, T. Umezaki, G. Soma, T. Tanemura, and Y. Nakano, “Scalable and robust photonic integrated unitary converter based on multiplane light conversion,” Phys. Rev. Appl., vol. 17, no. 2, p. 024071, 2022. https://doi.org/10.1103/physrevapplied.17.024071.Suche in Google Scholar

[215] S. G. Leon-Saval, A. Argyros, and J. Bland-Hawthorn, “Photonic lanterns: a study of light propagation in multimode to single-mode converters,” Opt. Express, vol. 18, no. 8, pp. 8430–8439, 2010. https://doi.org/10.1364/oe.18.008430.Suche in Google Scholar

[216] D. Noordegraaf, P. M. W. Skovgaard, M. D. Nielsen, and J. Bland-Hawthorn, “Efficient multi-mode to single-mode coupling in a photonic lantern,” Opt. Express, vol. 17, no. 3, pp. 1988–1994, 2009. https://doi.org/10.1364/oe.17.001988.Suche in Google Scholar PubMed

[217] T. A. Birks, I. Gris-Sánchez, S. Yerolatsitis, S. G. Leon-Saval, and R. R. Thomson, “The photonic lantern,” Adv. Opt. Photonics, vol. 7, no. 2, pp. 107–167, 2015. https://doi.org/10.1364/aop.7.000107.Suche in Google Scholar

[218] Y. Kim, et al.., “Metasurface folded lens system for ultrathin cameras,” Sci. Adv., vol. 10, no. 44, p. eadr2319, 2024. https://doi.org/10.1126/sciadv.adr2319.Suche in Google Scholar PubMed PubMed Central

[219] M. Faraji-Dana, E. Arbabi, A. Arbabi, S. M. Kamali, H. Kwon, and A. Faraon, “Compact folded metasurface spectrometer,” Nat. Commun., vol. 9, no. 1, p. 4196, 2018. https://doi.org/10.1038/s41467-018-06495-5.Suche in Google Scholar PubMed PubMed Central

[220] M. Faraji-Dana, et al.., “Hyperspectral imager with folded metasurface optics,” ACS Photonics, vol. 6, no. 8, pp. 2161–2167, 2019. https://doi.org/10.1021/acsphotonics.9b00744.Suche in Google Scholar

[221] B. Born, S.-H. Lee, J.-H. Song, J. Y. Lee, W. Ko, and M. L. Brongersma, “Off-axis metasurfaces for folded flat optics,” Nat. Commun., vol. 14, no. 1, p. 5602, 2023. https://doi.org/10.1038/s41467-023-41123-x.Suche in Google Scholar PubMed PubMed Central

[222] J. Tang, S. Wan, Y. Shi, C. Wan, Z. Wang, and Z. Li, “Dynamic augmented reality display by layer-folded metasurface via electrical-driven liquid crystal,” Adv. Opt. Mater., vol. 10, no. 12, p. 2200418, 2022. https://doi.org/10.1002/adom.202200418.Suche in Google Scholar

[223] M. Ossiander, et al.., “Metasurface-stabilized optical microcavities,” Nat. Commun., vol. 14, no. 1, p. 1114, 2023. https://doi.org/10.1038/s41467-023-36873-7.Suche in Google Scholar PubMed PubMed Central

[224] C. Vo, O. Anderson, A. Wirth-Singh, R. Johnson, A. Majumdar, and Z. Coppens, “Broadband long-range thermal imaging via meta-correctors,” arXiv:2410.17072, 2024. https://doi.org/10.48550/arXiv.2410.17072.Suche in Google Scholar

[225] S. Banerji, M. Meem, A. Majumder, F. G. Vasquez, B. Sensale-Rodriguez, and R. Menon, “Imaging with flat optics: metalenses or diffractive lenses?,” Optica, vol. 6, no. 6, pp. 805–810, 2019. https://doi.org/10.1364/optica.6.000805.Suche in Google Scholar

[226] J. Engelberg and U. Levy, “The advantages of metalenses over diffractive lenses,” Nat. Commun., vol. 11, no. 1, p. 1991, 2020. https://doi.org/10.1038/s41467-020-15972-9.Suche in Google Scholar PubMed PubMed Central

[227] A. C. Overvig, S. A. Mann, and A. Alù, “Thermal metasurfaces: complete emission control by combining local and nonlocal light-matter interactions,” Phys. Rev. X, vol. 11, no. 2, p. 021050, 2021. https://doi.org/10.1103/physrevx.11.021050.Suche in Google Scholar

[228] J. Guan, J.-E. Park, S. Deng, M. J. H. Tan, J. Hu, and T. W. Odom, “Light–matter interactions in hybrid material metasurfaces,” Chem. Rev., vol. 122, no. 19, pp. 15177–15203, 2022. https://doi.org/10.1021/acs.chemrev.2c00011.Suche in Google Scholar PubMed

[229] K. Vynck, et al.., “Light in correlated disordered media,” Rev. Mod. Phys., vol. 95, no. 4, p. 045003, 2023. https://doi.org/10.1103/revmodphys.95.045003.Suche in Google Scholar

[230] S. Rotter and S. Gigan, “Light fields in complex media: mesoscopic scattering meets wave control,” Rev. Mod. Phys., vol. 89, no. 1, p. 015005, 2017. https://doi.org/10.1103/revmodphys.89.015005.Suche in Google Scholar

[231] P. Lalanne, M. Chen, C. Rockstuhl, A. Sprafke, A. Dmitriev, and K. Vynck, “Disordered optical metasurfaces: basics, properties, and applications,” Adv. Opt. Photonics, vol. 17, no. 1, pp. 45–112, 2025. https://doi.org/10.1364/aop.537175.Suche in Google Scholar

[232] D. Marković, A. Mizrahi, D. Querlioz, and J. Grollier, “Physics for neuromorphic computing,” Nat. Rev. Phys., vol. 2, no. 9, pp. 499–510, 2020.10.1038/s42254-020-0208-2Suche in Google Scholar

[233] A. Argyris, “Photonic neuromorphic technologies in optical communications,” Nanophotonics, vol. 11, no. 5, pp. 897–916, 2022. https://doi.org/10.1515/nanoph-2021-0578.Suche in Google Scholar PubMed PubMed Central

[234] B. J. Shastri, et al.., “Photonics for artificial intelligence and neuromorphic computing,” Nat. Photonics, vol. 15, no. 2, pp. 102–114, 2021. https://doi.org/10.1038/s41566-020-00754-y.Suche in Google Scholar

[235] Y. Zuo, et al.., “All-optical neural network with nonlinear activation functions,” Optica, vol. 6, no. 9, pp. 1132–1137, 2019. https://doi.org/10.1364/optica.6.001132.Suche in Google Scholar

[236] J. Bueno, et al.., “Reinforcement learning in a large-scale photonic recurrent neural network,” Optica, vol. 5, no. 6, pp. 756–760, 2018. https://doi.org/10.1364/optica.5.000756.Suche in Google Scholar

[237] M. I. Shalaev, W. Walasik, A. Tsukernik, Y. Xu, and N. M. Litchinitser, “Robust topologically protected transport in photonic crystals at telecommunication wavelengths,” Nat. Nanotechnol., vol. 14, no. 1, pp. 31–34, 2019. https://doi.org/10.1038/s41565-018-0297-6.Suche in Google Scholar PubMed

[238] G.-J. Tang, X.-T. He, F.-L. Shi, J.-W. Liu, X.-D. Chen, and J.-W. Dong, “Topological photonic crystals: physics, designs, and applications,” Laser Photonics Rev., vol. 16, no. 4, p. 2100300, 2022. https://doi.org/10.1002/lpor.202100300.Suche in Google Scholar

[239] M. Miyata, I. Hiraoka, Y. Yamada, F. Nakajima, and T. Hashimoto, “Anti-reflection coating for all-semiconductor metasurface optical elements,” Opt. Express, vol. 32, no. 27, pp. 48943–48956, 2024. https://doi.org/10.1364/oe.547876.Suche in Google Scholar PubMed

[240] V. J. Einck, et al.., “Scalable nanoimprint lithography process for manufacturing visible metasurfaces composed of high aspect ratio TiO2 meta-atoms,” ACS Photonics, vol. 8, no. 8, pp. 2400–2409, 2021. https://doi.org/10.1021/acsphotonics.1c00609.Suche in Google Scholar

[241] B. Erbas, et al.., “Combining thermal scanning probe lithography and dry etching for grayscale nanopattern amplification,” Microsyst. Nanoeng., vol. 10, no. 1, p. 28, 2024. https://doi.org/10.1038/s41378-024-00655-y.Suche in Google Scholar PubMed PubMed Central

[242] N. Lassaline, “Generating smooth potential landscapes with thermal scanning-probe lithography,” J. Phys.: Mater., vol. 7, no. 1, p. 015008, 2023. https://doi.org/10.1088/2515-7639/ad0f31.Suche in Google Scholar

[243] X. Yin, et al.., “Roll-to-roll dielectric metasurfaces,” in Metamaterials, Metadevices, and Metasystems 2020, vol. 11460, N. Engheta, M. A. Noginov, and N. I. Zheludev, Eds., International Society for Optics and Photonics, SPIE, 2020, p. 114600S.10.1117/12.2576210Suche in Google Scholar

[244] Y. Ma, “Roll-to-roll printing trench-like metasurface film for radiative cooling,” Light: Sci. Appl., vol. 12, no. 1, p. 277, 2023. https://doi.org/10.1038/s41377-023-01317-w.Suche in Google Scholar PubMed PubMed Central

[245] F. Aieta, P. Genevet, M. Kats, and F. Capasso, “Aberrations of flat lenses and aplanatic metasurfaces,” Opt. Express, vol. 21, no. 25, pp. 31530–31539, 2013. https://doi.org/10.1364/oe.21.031530.Suche in Google Scholar

Received: 2024-12-27
Accepted: 2025-04-07
Published Online: 2025-04-24

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Heruntergeladen am 24.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2024-0772/html
Button zum nach oben scrollen