Open Access
7 March 2022 Tunable metasurfaces towards versatile metalenses and metaholograms: a review
Author Affiliations +
Abstract

Metasurfaces have attracted great attention due to their ability to manipulate the phase, amplitude, and polarization of light in a compact form. Tunable metasurfaces have been investigated recently through the integration with mechanically moving components and electrically tunable elements. Two interesting applications, in particular, are to vary the focal point of metalenses and to switch between holographic images. We present the recent progress on tunable metasurfaces focused on metalenses and metaholograms, including the basic working principles, advantages, and disadvantages of each working mechanism. We classify the tunable stimuli based on the light source and electrical bias, as well as others such as thermal and mechanical modulation. We conclude by summarizing the recent progress of metalenses and metaholograms, and providing our perspectives for the further development of tunable metasurfaces.

1.

Introduction

Metasurfaces are composed of two-dimensional (2D) periodic arrays of subwavelength-scale artificial elements, called meta-atoms. They have attracted great attention due to their ability to manipulate the properties of electromagnetic waves.18 Several design methods have been proposed with various shapes and compositions of meta-atoms.8,9 The Pancharatnam–Berry (PB) phase, also called geometric phase, has been investigated using rectangular-shaped meta-atoms that impart a phase delay proportional to their rotation angle. Propagation phase has been investigated by exploiting an effective refractive index to manipulate the retardation phase by changing the volume ratio, aspect ratio, and height of meta-atoms.10,11 Resonant effects such as plasmonic resonance,1215 Mie resonance,16,17 and Fabry–Pérot resonance18,19 have also been exploited. By exploiting these resources, optical elements can be highly miniaturized and various optical applications have been implemented, such as beam splitters,2022 absorbers,2329 metalenses,30,31 metaholograms,3240 selective thermal emitters,4143 detecting devices,4446 and structural color.4752

The functionality and efficiency of metasurfaces have been continuously increased by improving the methods to design meta-atoms, and the development of their material composition. Achromatic metalenses have been fabricated using complex geometric-structured meta-atoms that have a wide phase-dispersion set, and therefore enable achromatic focusing with single-layered metasurfaces.53 Furthermore, complex-amplitude metaholograms have been proposed by varying the conversion efficiency of meta-atoms to enable three-dimensional (3D) images.54,55 In terms of the materials, the use of resin with embedded nanoparticles30,31,56 has been proposed as a method to achieve mass production of dielectric metasurfaces through single-step direct nanoimprinting. Low-loss hydrogenated amorphous silicon57 has been proposed for low-cost deposition of visibly transparent thinfilms, and the fabrication cost is much lower and its modulation efficiency is compatible with that of titanium dioxide (TiO2)58 and gallium nitride (GaN)59,60 metasurfaces that work at visible frequencies.

Tunable metasurfaces with multiple functionalities through the active control of electromagnetic waves have been actively studied.61,62 Tunable metasurfaces are made up of meta-atoms that are controlled by external stimuli such as electrical biases or high-intensity light sources. Electrical, thermal, and mechanical stimuli have been used to induce two or more optical responses in single- or double-layered metasurfaces. Also, manipulation of the polarization state of the incident light that changes the output wavefront has been used to provide tunable functionalities for metasurfaces. Tunable metasurfaces provide multiple functionalities, however, generally have limitations in that the efficiency is generally worse than conventional passive metasurfaces, due to inherent problems such as the properties of tunable materials and design principles.1,3,9,63,64

In this review, we define “tunable metasurface” as one that can induce two or more optical responses due to variations in the incident light, or to changes to the meta-atom configuration or relative distances between two adjacent metasurfaces. Additionally, we present recent advances in tunable metasurfaces, in particular, tunable metalenses and metaholograms. Tunable metalenses and metaholograms are important applications of tunable metasurfaces. In the case of tunable metalenses, there is an advantage that it can be applied to an ultrathin zoom lens that can replace bulky optical components required for conventional optical devices.61,65 Meanwhile, owing to the high capacity of tunable metaholograms, it is expected to be one of the fundamental technologies of future metasurface integrated devices, such as holographic memory devices and ultrahigh-density display applications.54,66,67 However, once the fabrication is undertaken, it is difficult to change the period and size of the meta-atoms. In addition, complete modulation principles capable of nanoscale local pixel control have not been established.

In this review, we first briefly introduce the fundamentals of metalenses and metaholograms. Most tunable optical responses are obtained by controlling the light source or through an applied voltage, so we classify tuning methods as (1) controlling the light source, (2) electrical tuning, and (3) non-electrical tuning. Non-electrical tuning includes heat-induced phase change materials (PCMs), mechanical deformation, and changes of a relative position of cascaded metasurfaces. Finally, we summarize the overall contents, and suggest future directions of research on tunable metasurfaces.

2.

Tunable Metalenses

2.1.

Design Principles of Metalenses

Conventional refractive and diffractive lenses have a tradeoff relationship between miniaturization and optical characteristics.68 For example, to achieve achromatic focusing, several diffractive or refractive lenses must be used, but it can be achieved using single-layer metasurfaces.60 To design metalenses, the desired phase profile should be physically constructed using meta-atoms. To focus an incident plane wave at a lens focal point, the target phase Φ at a point P(x,y) on a metalens should satisfy the phase retardation.9

Eq. (1)

Φ(r)=2πλ(r2+f2),
where r=x2+y2 is the radial distance from the center to each point, f is the focal length, and λ is the wavelength. The distance between the focal point and a point on a metasurface varies with r so as to correct for the distance difference, and the meta-atoms should satisfy the phase retardation [Eq. (1)] to achieve constructive interference at the focal point. To design tunable metalenses, meta-atoms must have properties that can change their optical responses so that they satisfy Eq. (1).

To achieve tunable metalenses, some mechanisms such as helicity or spin sensitive geometric phase, liquid crystals (LCs) or graphene-integrated lenses, and PCMs or stretching methods have been used. Active materials such as LCs or graphene can be integrated with metalenses to modulate the phase profile for achieving focus tuning. Their electrically controllable characteristics such as different alignment (LC) and Fermi level and carrier density (graphene) can be used, depending on the external electric field. Moreover, modulating refractive index using PCMs can be used in achieving tunable metalenses. Finally, a method of changing the geometrical parameters of the metalens through stretching flexible substrates can be used.

2.2.

Tunable Metalenses by Light Source

Tunable metalenses can be realized to control properties of light sources such as the polarization state. Spin-decoupled metalenses have been achieved using PB phase to integrate the properties of multiple convex and concave lenses into one metasurface:69 one phase profile focuses left-circularly polarized (LCP) incident light, while the other profile focuses right-circularly polarized (RCP) light. Therefore, the focal point changes when the polarization of incident light changes [Fig. 1(a)]. Additionally, the intensity of multiple focal points can be tuned by controlling the ellipticity of incident light [Fig. 1(b)]. Using only the geometric phase, it has the simplicity of designing a spin decoupled metalens instead of using both the propagation phase and geometric phase because of no need for scanning lots of parameters. However, the proposed metalens has a low efficiency of <50% in theory.

Fig. 1

Tunable metalenses by light source. (a) Schematic illustration of spin-decoupled metalenses.69 (b) Simulated and experimental electric field intensity distributions at line y=1.5  mm for multiple polarization states of incident light (LCP, LECP, LP, RECP, RCP from left to right).69 (c) Schematic of spin selective metalenses that can focus RCP and LCP incident beams to different focal points.70 (d) Schematic view of metalenses using the combination of the PB phase and propagation phase.71 (e) Schematic of step-zoom metalenses in which the focal length is changed according to the linear polarization of the incident light.72 (f) Schematic of metalenses doublet that has different functions depending on the polarization of the incident light.73

AP_4_2_024001_f001.png

Multiple focal points can be generated by controlling the circular polarization state of incident light. Helicity-dependent multifocal metalenses can create multiple focal points in different directions when the polarization of the incident light changes.74 These metalenses are composed of anisotropic rods that have different orientations and can be considered as half waveplates with a high efficiency. Its polarization-conversion efficiency is 97% at 0.64 THz under LCP illumination.

Furthermore, a spin-selected metalens that has a 0.98 numerical aperture (NA) value (simulated data) can focus incident light at two focal points depending on the spin state of the incident light.70 It is composed of a unit structure of silicon nanobricks, and the desired phase profile is implemented by the PB phase. Two silicon nanobricks (red and blue) on the metalens act as a convex lens or a concave lens when the spin state of the incident light changes [Fig. 1(c)]. This spin-selected metalens is useful for applying detecting techniques and spin controlled photonics.

The focal point can be adjusted by combining the PB phase and the propagation phase.71,75 One spin-multiplexed metalens uses the PB phase and propagation phase of TiO2 nanorods.71 It makes a polarization-independent hyperbolic phase and a polarization-dependent linear phase, depending on the polarization state of the incident light, and therefore has different focal points for LCP and RCP light [Fig. 1(d)]. The diameter and NA of this lens are 1.8 and 0.05 mm, respectively. Although the NA is low, this lens demonstrated diffraction-limited focusing. Furthermore, TiO2 is used to obtain high modulation efficiencies in the visible band by exploiting its low loss extinction coefficient and high refractive index. This metalens has an advantage of high focusing efficiency (maximum of 70%). But it is more complex for designing a metalens; it is a tradeoff relationship between efficiency and designing simplicity.

A step zoom metalens that has dual focal lengths and 0.21 NA value has been demonstrated using double-sided metasurfaces.72 These metasurfaces are composed of an array of silicon nanobricks, and the desired phase is obtained by changing the lengths of their long and short axes. Under x polarized light, the first metasurface operates as a concave lens, and the second one operates as a convex lens. In contrast, under y-polarized light, both metasurfaces operate as convex lenses [Fig. 1(e)]. Consequently, focal lengths vary depending on the linear polarization state of incident light. This double-sided metasurfaces design technique has advantages such as compactness, simplicity, and flexibility; and it has great potential for applications in biomedical sciences, optical communications, and wearable electronics.

Additionally, a metalens doublet that has different functions depending on the polarization of the incident light and the distance between two lenses has been reported.73 The first 0.258 NA metalens is composed of TiO2 nanocylinders with different diameters to implement the propagation phase. It is therefore polarization-independent. The second metalens has an NA of 0.66 and is composed of TiO2 rectangular meta-atoms to implement the PB phase, making it a polarization-dependent lens. The two lenses are used in tandem to implement a three-function lens doublet by varying the circular polarization state of the incident light and the distance between the lenses [Fig. 1(f)]. This metalens doublet has the advantages of making the imaging system simple and compact because no additional optical components are required for the multifunctional system. Therefore, it has great promising perspectives for applications in portable imaging systems.

2.3.

Tunable Metalenses by Electrical Bias

Electrically tunable metalenses can be realized by applying an external voltage bias on active materials, such as LCs7682 and graphene.8389 Transmission-type terahertz metalenses that combine dielectric metasurfaces with photopatterned LCs have achieved tunable chromatic aberration.78 When the voltage bias is applied to the LCs, their geometric phase modulation vanishes, and only the resonant phase in the metalenses remains, so the function of the device changes from achromatic to dispersive focusing [Fig. 2(a)]. By integrating two functions into one metalens, it has great promising perspectives for applications in spectroscopy and imaging systems.

Fig. 2

Tunable metalenses by electrical bias. (a) Schematic view of LC integrated metalenses that change functionality achromatic focusing to dispersive focusing when applying voltage bias to LCs.78 (b) Side view of TN LCs integrated electrically tunable metalenses. It modulates the polarization state of the incident beam depending on the applied voltage bias.79 (c) Schematic of a focus tunable graphene metalens when DC voltage bias is applied to this metalens.85 (d) Variation of focal length and focal spot intensity when the design is a packed pattern (top) and shifted bezel pattern (bottom).85 (e) Schematic of tunable graphene metalenses in which the chemical potential of graphene is controlled by applying a gate voltage.88 (f) Schematic view of electrically tunable metalenses in which the refractive index of BTO antennas is changed by applying voltage bias.90

AP_4_2_024001_f002.png

A varifocal metalens that switches between NA 0.21 and 0.7 (simulated data) has been obtained by putting twisted nematic (TN) LCs under a metalens substrate.79 Depending on the voltage applied to the electrode, the TN LCs convert the polarization state of the incident light, and achieve different focal points for different polarization states of incident light [Fig. 2(b)]. Using the combination of a metalens and TN LCs, it has advantages of high image quality and fast response time (sub-millisecond level). Therefore, it has great potential for applications in biomedical and optical technology.

Recently, graphene has been used to achieve tunability.8389,9195 A graphene-based ultrathin square subpixel lens whose focal length can be controlled by a voltage bias has been demonstrated85 [Fig. 2(c)]. By applying the external bias voltage to multilayer graphene, the Fermi level and carrier concentration according to the location in multilayer graphene are changed, resulting in the change of absorption and transmittance of graphene. These changes cause a width change of arc ribbon-shaped graphene, resulting in the modulation of a Fresnel zone plate topology. Therefore, the focal length of the lens is modulated up to 19.42% (from 190 to 226.9  μm) [Fig. 2(d)]. Furthermore, this ultrathin lens mechanism can be applied to multifunctional autostereoscopic areas such as 3D hologram displays and acoustic applications owing to the advantages of its subpixel scale structure.

Additionally, tunable terahertz metalenses composed of a graphene monolayer and gold (Au) film have been demonstrated.88 The application of a voltage to graphene changes its chemical potential and permittivity. This change shifts the transmittance and phase of the incident light, and yields a tunability of focal length of about 1.25λ [Fig. 2(e)]. This design concept can be applied to active terahertz devices for imaging. Changing the refractive index of nanopillars by applying a voltage is another way to tune the focal length.90 A metalens composed of indium tin oxide (ITO) as a transparent electrode, barium titanate (BTO) nanopillars, and a SiO2 substrate has been proposed.90 The refractive index of BTO is proportional to the induced electric field, so by exploiting the electro-optic effect of the BTO crystals, a phase change can be achieved by controlling the external voltage [Fig. 2(f)]. The refractive index of a particular area on metalenses can be tuned by changing the refractive index of the nanopillars without controlling the entire metasurface. This metalens has advantages such as high-speed modulation, compactness, and flexibility.

2.4.

Tunable Metalenses by Non-Electrical Input

In this section, we introduce metalenses that are tuned non-electrically through mechanical actuation and PCMs. First, metalenses that are tuned using mechanical actuation can be realized by stretching or rotating the substrate. For actuation by stretching the substrate, the meta-atoms are placed on a stretchable substrate such as polydimethylsiloxane. The physical locations and therefore the periodicity of the meta-atoms increase when a uniform strain is applied to the substrate. Therefore, stretching the substrate causes a change of the spatial phase profiles, which is used to vary the focal length [Fig. 3(a)].96,102105 In Fig. 3(b), experimental results including a longitudinal beam profile and intensity distribution according to the stretch ratio and stretch ratio versus focal length graph are shown. According to these results, the focal length gradually increases as the stretch ratio increases.96 The target wavelength of this zoom lens is 632.8 nm, but can be altered by changing the geometry and materials of the meta-atoms.

Fig. 3

Tunable metalenses by non-electrical input. (a) Schematic of metalenses that are tuned mechanically by stretching the substrate to tune the focal length.96 (b) Measured longitudinal beam profiles according to different stretch ratios (top), intensity distributions of transmitted cross polarized light with different stretch ratios (bottom left), measured and calculated stretch ratio-focal length graph (bottom right).96 (c) Schematic illustration of metalenses that are tuned mechanically by rotating the substrate (left) and phase distribution of one of two lenses (right).97 (d) Schematic view of a tunable metalenses system that consists of two cubic metasurfaces.98 (e) Schematic illustration of MEMS tunable metalenses in which the focal length is tuned by controlling the distance between two lenses.99 (f) Schematic illustration of varifocal metalenses in which the focal length is tuned by exploiting the phase change of GSST by furnace annealing.100 (g) Schematic of varifocal metalenses using phase-change material Sb2S3. The phase of Sb2S3 is changed by controlling the temperature, and the change affects the focal length.101

AP_4_2_024001_f003.png

Recently, to realize varifocal metalenses, graphene has been used to achieve a range of focal length tuning.106 Graphene has an advantage of being suitable for designing broadband devices due to its dispersionless characteristics over a broadband wavelength region from the ultraviolet to the terahertz regime due to its lack of a bandgap. The focal length of the graphene oxide metalenses can be adjusted by >20% for a single wavelength (red, green, and blue light) by stretching the metalenses laterally. Furthermore, rotation of a metalenses doublet can tune its focal length.107110 Mutual rotation of doublet metalenses has been shown to tune the focal length using Moiré metalenses that are axially asymmetric doublet lens97 [Fig. 3(c)]. The demonstrated metalens has an NA of 0.5 at a target wavelength of 900 nm. This mechanism has an advantage in terms of the wide tuning range of the focal length from negative to positive, compared to other mechanisms such as Alvarez lenses98,111 and micro-electromechanical systems (MEMS).99,112114

Other mechanical actuation mechanisms such as Alvarez lenses and tunable metalenses that integrate MEMS systems have been demonstrated. Varifocal metalenses using the Alvarez lens design have been fabricated by integrating two cubic metasurfaces; one example in particular has a tunable focal length in connections when it is laterally actuated using a translation stage [Fig. 3(d)].98 This mechanism has a large tuning range due to the inverse proportionality between the focal length and displacement for Alvarez lenses, but it is unsuitable for portable lens platforms owing to use of micrometer translation stage to actuate the metasurfaces laterally. Another tunable metalens has been obtained using MEMS by combining two metasurfaces99; one on a glass substrate and is static (NA0.8); the other on a movable membrane (NA0.8); the two metasurfaces are linked by electrostatic actuation, and the focal length is modulated by controlling the distance between the two [Fig. 3(e)]. The NA value of object space and image space is 0.16 and 0.014, respectively. Using this mechanism, high-speed electrical focusing and scanning of the imaging distance was achieved.

Furthermore, PCMs have been widely used to make tunable metalenses, by varying the phase of the PCM through the application of external stimuli such as optical pulses,115 thermal heating,100 and electrical heating.101,116119 A rewritable device that uses Ge2Sb2Te5 (GST) exploits modulation of the state of GST by an optical pulse.115 When a GST thin-film is stimulated by a femtosecond laser pulse, it causes partial state transition of GST and causes phase patterning. Active metasurfaces using a related optical PCM, Ge2Sb2Se4Te1 (GSST) have been demonstrated to realize a highly varifocal metalens that operates at λ=5.2  μm and has NAs of 0.45 (amorphous) and 0.35 (crystalline).100 The state of GSST can be converted by furnace annealing, and the focal length is tuned depending on the state of GSST [Fig. 3(f)]. This mechanism is simple and compact compared to other mechanical mechanisms due to the absence of moving parts. Moreover, GSST has no critical thickness to fully reversible switching, unlike GST, which has a critical thickness <100  nm. However, the crystallization time of GSST is in the order of microseconds, which is slower than that of GST.

A thermally modulated varifocal metalens with NAs of 0.714 and 0.608 (simulated data) using Sb2S3 has also been proposed [Fig. 3(g)].101 By controlling the temperature of Sb2S3, its phase can be converted. The refractive index of Sb2S3 varies depending on its phase; this change causes the phase shift of incident light, and thereby achieves tunable focal length. The target wavelength of this metalens is 1310 nm, so using Sb2S3 as meta-atoms, high focusing efficiency can be achieved due to low absorption of Sb2S3 in the near-infrared region, unlike GST.

3.

Tunable Metaholograms

3.1.

Design Principle of Metaholograms

Holographic technologies have exploited the characteristics of light, such as amplitude, phase, and polarization, to record and reconstruct the interference patterns of targeted objects. Previous holographic technologies have utilized spatial light modulators (SLMs) to produce 3D images. However, the pixel pitch of SLMs is limited to the micrometer scale, which results in low resolution, small viewing angles, unpredicted high-order diffraction, and sampling problems.120 Therefore, metasurface holograms that have a pixel subwavelength scale have received great attention to overcome the shortcomings of conventional holograms.121,122

To design these metaholograms, there exist two main challenges. First, calculating the phase map is needed for attaining the desired light propagation, and second, obtaining the proper meta-atom design necessary to physically implement the desired phase shift range from 0 to 2π. In calculating the overall phase map, employing the Fourier hologram is the most well-known method, and can be retrieved using techniques such as the Gerchberg–Saxton (GS) algorithm, which has demonstrated holograms with an efficiency over 80%.19,123125 Then, the continuous phase map should be quantized to a discrete phased map to be utilized with the designed meta-atoms. Recently, a topologically protected full 2π phase with a reflective plasmonic metasurface has also been reported.126 However, the current design processes have limitations in that they can only realize quantized phase maps instead of a continuous phase map from the above design methods. Establishing the reliable process that can realize continuous phase maps is a challenging problem.127

Metaholograms can control the phase, amplitude, and even polarization by exploiting light-matter interactions at the subwavelength scale. However, conventional metaholograms, once manufactured, are limited to a single function. Above all, tunable metaholographic technologies focus on storing as many holographic images as possible in a single metasurface by exploiting light source properties and active materials. In this section, we review the achievements of tunable metaholograms by controlling light sources (Sec. 3.2), and using active materials (Sec. 3.3).

3.2.

Tunable Metaholograms by Light Source

The properties of light including the amplitude, phase, and polarization can be modulated using various optical components such as lenses, beam splitters, wave retarders, and polarizers. Metasurfaces can also be designed to respond differently to the properties of incident light. A single metasurface can produce tunable hologram images by controlling the properties of incident light sources that enter the metasurfaces. In this section, we discuss tunable metahologram research that considers regulating the polarization state,128135 orbital angular momentum (OAM),54,136139 and coding incident beam.67,140142 Above these, other methods include modulating the wavelength or angle of the incident beam, which have also been used to realize tunable holographic images. One such metaholographic device composed of heterogeneous meta-atoms, operates at the visible (532 nm) and near-infrared (980 nm) wavelengths.36 Also, angle-multiplexed metaholograms composed of dielectric U-shaped meta-atoms have been used by adjusting the angle of incident light.143

3.2.1.

Tunable metaholograms by polarization state

Metasurfaces consist of artificially tailored meta-atoms that can manipulate the polarization state of an interacting light beam. These meta-atoms can be used to create dynamic holographic images by changing the polarization states of light sources. Using these properties, helicity-multiplexed reflective metaholograms have been reported.129 One helicity-multiplexed metasurface composed of silver nanorods on a Si substrate can reconstruct switchable images for RCP and LCP at a wavelength of 633 nm [Fig. 4(a)]. Clear holographic images are also obtained at other visible wavelengths [Fig. 4(b)]. This approach suggests solutions to several main problems with polarization tunable metaholograms such as image quality, efficiency, and broadband bandwidth.

Fig. 4

Tunable metaholograms by light source. (a) Schematic of tunable metaholograms that generate dual images that can be changed by the polarization state of an incident light beam. (b) Experimentally obtained holograms at incident wavelengths of 24 nm (left) and 475 nm (right).129 (c) Design principle of OAM multiplexing metaholograms and metahologram images with a planar wavefront (l=0). (d) Reconstructed images using four different OAM beams with topological charges l=2, 1, 1, and 2.137 (e) Illustration of a video metahologram using OAM. (f) It changes their images depending on the angular momentum of the incident light.54 (g) Schematic of the reprogrammable metaholograms with the modulated incident beam by SLM. (h) The recorded images with the characterized incident light (left), with uniform laser light (middle), and the incident laser beam itself (right).140

AP_4_2_024001_f004.png

Further, polarization-sensitive color-tunable metasurface holograms have been demonstrated.130 The metasurfaces are composed of three kinds of Si meta-atoms and are developed using PB phase to modulate the wavefronts of a visible hologram. Chiral metaholographic technologies have been extended by the propagation phase and PB phase.128 These metasurfaces are designed to respond to arbitrary orthogonal polarization states. This approach improves on previous metaholograms, which only work under orthogonal linearly or circularly polarized light. Use of planar chiral elements extended metaholographic technologies to planar chirality.131 When illuminated with circularly polarized light, the reflective metasurfaces reconstructed dual images, while absorbing the opposite circularly polarized light. This research enables demonstrated various developments of tunable metasurfaces such as full-color display applications, polarization switchable devices, and spatial separation of polarization information channels beyond holographic imaging.

Despite the endeavors to overcome the limitations of polarization switchable metaholograms, conventional metaholograms generate only two images in response to two orthogonal polarization states. Vectorial holography is an innovative technique that exploits metasurfaces that consist of meta-pixels composed of two orthogonal meta-atoms that can respond to multiple polarization states.132 A broadband reflective vectorial metahologram operated at four polarization states. Vectorial hologram extended the degree of freedom of metaholograms, which can switch only two images.

Polarization multiplexing can extend the manipulation channels and augment encryption capability.133 Single-layer metasurfaces can realize multiple independent phase profiles, which can contain distinct information under illumination of different polarization states of light. Such metasurface holograms are actively used in optical encryption144 and applications such as polarization analyzers. An orthogonally polarized metasurface hologram that uses the PB phase can enable direct detection of polarization states in a one-time measurement.134 Although the polarization-analyzing metasurface has a low efficiency (<0.3%) at λ=650  nm, it works well producing holographic images over a broad range of wavelengths. A proposed new encryption process uses different metaholograms as the keys of imaging encryption.135 The process uses dual-channel Malus metasurfaces to exploit high-quality images, which can generate a matrix during the encoding and decoding processes. Different channels of the metasurface can contain different matrix information to increase the number of keys available to encryption or anti-counterfeiting systems.

Although polarization multiplexed metaholograms have achieved storing multichannel information by adjusting the incident polarization states, it is difficult to react to subtle polarization states due to the sensitivity of meta-atoms.

3.2.2.

Tunable metaholograms by orbital angular momentum

OAM is a fundamental property of light that can be controlled. OAM can be applied as a method to multiplex metahologram images. A vortex beam carrying OAM has a helical wavefront and a spiral phase profile expressed by exp(ilφ), where l is the topological charge number and φ is the azimuth angle in cylindrical coordinates. The topological charge number can be arbitrarily controlled, thus the OAM mode of a vortex beam has the advantage of having infinite degrees of freedom, theoretically.145

Multi-momentum transformation metasurfaces have been demonstrated using OAM and the linear momentum of incident light.136 OAM has also been implemented as an information carrier for metaholographic technology.137 The holographic image is sampled using a 2D Dirac function related to OAM modes to reconstruct an OAM selective hologram. These OAM metasurfaces have been designed with four images in the spatial frequency domain by combining different states of OAM. Four holographic images under light with different topological charges (l=2, 1, 1, and 2) are reconstructed [Fig. 4(d)]. Using similar methods, video-holographic metasurfaces have been demonstrated [Fig. 4(e)].54 In addition, OAM video-holographic experimental images with different topological charges are reconstructed [Fig. 4(f)]. More than two hundred images can be stored by exploiting independent OAM channels. The above methods are expected to be utilized in ultrahigh-density video-holographic devices.

Another new method combines polarization control and OAM selectivity for metasurface holograms.138 Through this method, a single metasurface can reconstruct multiple holographic images; polarization selectivity is controlled using the property of birefringence, and OAM selectivity is modulated by changing the topological charge. OAM-encrypted metasurface holograms that depend on the polarization states have been introduced.139 They consist of two metasurfaces: one to generate multiple OAM beams, and another to generate an OAM-selective hologram, which can be applied to an encryption system.

OAM holograms are emerging technologies that can contain infinite information, theoretically. However, they cannot implement broadband selective metaholograms and ultrafast switching. These challenges could be solved through the combination with RGB selective rules and vortex microlasers.

3.2.3.

Tunable metaholograms by coded incident beam

SLMs or dynamic micro-mirror devices (DMDs) have been traditionally used to create dynamic holographic displays, but these approaches have been limited due to their large pixel sizes, which cause sampling problems, small viewing angles, and multiple-order diffractions.146,147 However, SLM and DMD combined with metasurfaces have been used to greatly complicate the information of the incident light by small-pixel coding of the metasurfaces. This method can increase the complexity of metasurface hologram design.

SLMs have been used to produce reprogrammable metahologram encryption algorithms by manipulating the light sources.140 The correct metaholographic image is only reconstructed when the incident light is modulated to a predesigned wavefront by the SLM, whereas a misleading image is shown for undesigned incident light [Fig. 4(g)]. The experimental image set verifies that the phase matrix of incident light can operate in an optical holographic encryption system [Fig. 4(h)]. Dynamic 3D metasurface hologram can store 28-bit different holographic images, and dynamically reconstruct a holographic image with high frame rate in the visible range.67 The proposed metasurface uses DMD to exploit dynamic space coding of the incident structured laser beam and can display 228 different frames and achieve a high frame rate of up to 9523 frames per second. Code-division multiplexing (CDM) dynamic metasurface holograms have been developed using DMD to generate structural light as the incident beam.141 Birefringent metasurfaces adopt the CDM principle to produce 32 distinct holograms. A new method of secret sharing using cascaded metasurface holograms has been demonstrated.142 The process uses metasurface hologram images as encoding keys in place of an SLM or DMD. Light propagation by the cascaded metasurfaces optically reconstructs the secret images with high fidelity and builds up the phase shift of both holograms. Employing additional components such as SLM or DMD is an excellent means of complicating information through small pixel coding at the metasurface. However, due to inherent limitations of SLM and DMD, it requires a lot of optical components to combine metasurfaces. To overcome these fundamental limitations, realizing SLM with a tunable metasurface has been reported.148

3.3.

Tunable Metaholograms by Active Material

Metaholograms can also be tuned using active materials. The tuning methods generally apply external stimuli to a metasurface composed of active materials. External stimuli can be classified as electrical or non-electrical stimuli. Therefore, this section presents electrically tunable149155 and non-electrically tunable34,156164 methods to tune metaholograms.

3.3.1.

Tunable metaholograms by electrical bias

LCs, conductive oxides, and semiconductors that undergo dramatic changes in response to electricity have been used as electrical tuning methods. Among them, LCs are mainly applied to generate different holographic images. LCs can be easily switched between liquid and solid-crystal states by applying an electrical field. This phase transition of LC can generate great optical birefringence, which can be utilized in tunable metaholograms. LCs can assume nematic, smectic, and isotropic phases [Fig. 5(a)]. The nematic phase has a fixed orientation, the smectic phase has a fixed orientation in well-defined planes, and the isotropic phase has random orientations.62 Therefore, LCs can be used to control the local birefringence, which is used to reconstruct various holographic images by exerting external electric fields.

Fig. 5

Tunable metaholograms by electrical bias. (a) Schematic of three major states of LC (nematic, smectic, isotropic). (b) Design and demonstration of electrically tunable dielectric metasurfaces that use LCs.149 (c) Working principle of the electrically controlled digital metasurface device (DMSD). (d) SEM image of the DMSD and experimental results by independent control of seven electrodes.150 (e) Schematic of the bifunctional vectorial metasurfaces with LC analyzer. (f) SEM image of fabricated metasurfaces, and metaholograms that can be tuned by applying different voltages [scale bars: (left) 50  μm, (right) 1  μm].152

AP_4_2_024001_f005.png

Electrically tunable transparent holographic displays have been achieved by integrating dielectric metasurfaces with LCs.149 The displays can be induced to show resonance shifts twice the size of their line width, by applying a voltage through the metasurface. A switchable metasurface display achieving 53% efficiency at λ=669  nm [Fig. 5(b)] has been proven. An electrically controlled digital metasurface device (DMSD) has been developed for light-projection displays.150 The DMSD has unit pixels composed of an Au nanorod metasurface, LC, and an ITO-coated superstrate to exert electric fields in each cell [Fig. 5(c)]. To further broaden the potential of DMSD, it is used in a numeric display composed of seven electrically manipulated segments [Fig. 5(d)]. Multifunctional polarization-dependent metasurfaces can be integrated with electrically tunable LC in the visible region.151 The proposed metasurfaces could combine the polarization-control ability of metasurfaces with the birefringence properties of LC. These devices provide a pragmatic method for dynamic addressable metasurface applications such as laser imaging detecting and ranging.

Further, a proposed photonic security platform uses dynamic vectorial holographic images of pixelated bifunctional metasurfaces.152 In the white light, the proposed metasurfaces show structural color prints, whereas when laser illumination passes through the metasurfaces, the encoded tunable metaholograms are reconstructed [Fig. 5(e)]. The device shows a reflective QR-code image and transmissive vectorial holograms [Fig. 5(f)]. When the QR code is captured, decipher keys about proper voltage values are transferred and the receiver can decode the real key using the vectorial holographic image. A proposed new optical encryption method exploits an improved computer-generated holography (CGH) algorithm to generate holograms that have quantitative correlation.153 A nematic LC layer realizes the function of dynamic holographic display. One set of electrical modulation patterns acts as encryption keys, and the receiver decrypts the message using both cipher text and a table transferred holographically. Electrically tunable metaholographic technologies have also been studied. An electromagnetic reprogrammable hologram device exploits 1-bit coding of the diodes on the metasurfaces.154 Additionally, a conducting polymer and polyaniline can be used to electrochemically control a metaholographic device.155 The reported work has contributed to realizing practical LC-integrated metaholographic displays with encryption systems and data storage. However, controlling LCs locally with nano pixel units through partial voltages and combining them with the metasurface are still a challenging problem.

3.3.2.

Tunable metaholograms by non-electrical input

Various metaholograms have been accomplished using thermal,156159 chemical,155,160162 and mechanical163,164 methods. Thermally sensitive materials, especially PCMs, such as germanium antimony telluride (GST) and vanadium dioxide (VO2), have been exploited for efficient thermal tuning. By heating the material, the crystallization of GST which has nonvolatile properties can be achieved. Also, VO2, which has volatile properties, can undergo an insulator—metal phase transition at a heating temperature of around 68°C. Plasmonic metadevices that have switchable spin–orbit interactions have been evaluated.156 Metasurfaces that enable the spin Hall effect, vortex beam generation, and hologram are obtained using GST. The response of the metasurface can be changed by a change of the phase of GST between amorphous (ON) and crystalline (OFF). A broadband active metahologram that could operate in dynamic holographic imaging has been achieved by exploiting the temperature-dependent properties of VO2.157 A single metasurface is composed of two sets of resonators: the passive one is simple metallic C-shaped split-ring resonators (SRRs); the active one is VO2 integrated SRRs. This device could reconstruct different images in the terahertz range.

The capability of a GST metadevice has been increased using multiple-state switching of photonic angular momentum coupling.158 The proposed metasurfaces could convert spin angular momentum to an OAM beam, depending on three states of GST [Fig. 6(a)]. They can also encrypt optical information using various hologram images at different crystallization levels of GST under RCP and LCP illumination [Fig. 6(b)]. Hybrid-state engineering of GST may also have applications for optical encryption.159 The GST metasurfaces hologram could provide a novel technique that is only recognizable when amorphous and crystalline states coexist, a semi-crystalline state. More discussion about PCM metasurfaces is listed in the conclusion with electrically tunable PCM metasurfaces.165,166

Fig. 6

Tunable metaholograms by non-electrical input. (a) Changing phase geometry (amorphous, semicrystalline, crystalline) that can construct SAM-OAM conversion by tuning the crystallization level of GST. (b) SEM images of fabricated metasurfaces and two different metaholographic images in response to three crystallization levels (top: RCP, bottom: LCP).158 (c) Schematic of the hydrogenating metasurfaces. (d) Four different holographic images during two hydrogenation and two dehydrogenation processes.160 (e) Illustration of the surface pressure-responsive designer LC; (f) experimental results before (left, LCP) and after (right, RCP) surface pressure by touch of finger.164 (g) Schematic of a gas sensor with LC-integrated metahologram. (h) The gas sensor changes optical images when hazardous gas is detected.34

AP_4_2_024001_f006.png

Chemical tuning methods usually exploit hydrogenation of the metasurfaces. Hydrogenation and dehydrogenation process can transform from metallic to dielectric material properties, which can be utilized to implement dynamic metaholograms. This method has been used to create addressable dynamic metasurface holograms that can use chemical reactions to manipulate subwavelength pixels.160 A metallic metasurface composed of magnesium (Mg) nanorods transforms the dielectric metasurfaces as a result of hydrogenation of Mg [Fig. 6(c)]. The devices show four dynamic metasurface holograms during the hydrogenation and dehydrogenation process [Fig. 6(d)]. A reconfigurable metasurface hologram that reconstructs switchable images by exploiting a quantified phase relation of the Fidoc method has been studied.161 The functionality of metasurfaces that use Mg nanorods has been improved to combine the display and holograms.162 The dynamic dual-function metasurfaces can generate a colorful display under white light and reconstruct holographic images under coherent light at λ=633  nm. However, chemical tunable metaholograms require long phase transition times. To overcome this limitation, a metallic polymer combining with electrical tuning method has been implemented.167

A stimulus-responsive, electric, thermal and mechanical, dynamic metahologram has been obtained using a designer LC.164 The LCs can be modulated by three different methods and operate as a switch that could change the holographic images in real-time [Fig. 6(e)]. LC-integrated metasurfaces could reconstruct different images when subjected to different surface pressures [Fig. 6(f)]. This concept was further investigated to realize gas sensors using LC-integrated metasurfaces [Fig. 6(g)].34 It attaches gas-reactive LC on metaholograms, and it changes optical images when gas is detected [Fig. 6(h)]. These studies will be helpful for future touching controllable metasurfaces. However, it is limited to applying the appropriate pressure, so optimizing the design process is necessary.

4.

Conclusion

Metasurfaces can offer unique phenomena to control light by interacting with subwavelength meta-atoms. However, conventional metasurfaces are limited to single functionality, and this constraint impedes potential applications after fabrication. To overcome this limitation, tunable metasurfaces are being developed.62 Here, we have presented an overview of the recent advances in the design of tunable metasurfaces.

We introduced tuning methods such as controlling the light sources or electrical fields, and using active materials. Of these methods, electrical tuning of active materials has received great attention because of its versatility to be integrated with electronics. The use of LCs, semiconductors, and conductive oxides has revealed novel opportunities, but the design of metasurfaces that are tuned electrically is still in its infancy. Furthermore, nonvolatile PCMs such as GST and GSST are not easily reversed back to their amorphous state from a crystalline state. Conventional GST metasurfaces usually require a melt-quenching process with cooling rates >109  K·s1. However, electrical tuning methods that use well-designed micrometer-scale heaters to change GST phase have been developed.165,166 This method has opened a new class of mixed-mode optoelectronic devices that use GST. Metasurfaces that use PCM with electrical control may give a fresh view to reprogrammable metadevices, such as optoelectrical neuromorphic devices and dynamic holograms.

Table 1 shows a summary of the work presented in this review, focusing on tunable metalenses and metaholograms. The main goals of metalenses are to obtain multiple focal spots, and some research has been conducted to change states of focusing and defocusing. On the other hand, metaholograms have been studied as ways to store multiple images by changing the properties of the incident light and nanostructures. More recently, visualized sensing and optical metaholographic encryption have gained attention. In addition, to realize metasurface holographic technology, the dynamic control of the holograms is essential. Even if new approaches for dynamic holographic displays are proposed using SLM or DMD, fully developed metasurface holographic video displays have not been realized yet. OAM holograms that use complex-amplitude metasurfaces capable of more than two hundred independent channels have been proposed. This result is achieved in momentum space by manipulating complex amplitudes of the light. Then holographic videos are successfully realized at different image planes. This method may provide a method to achieve a metasurface video display.

Table 1

Summary of tunable metalenses and metaholograms.

TypesMetalensesMetaholograms
Research goalsMulti-, varifocal length6989,91118Storing multiple images54,129142
Visualized sensing (thermal, chemical, etc.)155162
Optical encryption140142
Common stimuliElectrically tunable LC7682,149155
Polarization (circular and linear polarization)6975,129135
Mechanical (strain, relative position of two metasurfaces)9699,102114,163,164
Vectorial metasurfaces132,133
Different stimuliElectrically tunable 2D materials8389,9195 (graphene, etc.)Topological charges of OAM136139
Chemically/Thermally tunable holograms155162
Pressure responsive LC34,164
Common problems/challenges of controlling light sourceTradeoff relation between efficiency and design simplicityReacting with subtle polarization state differences
RGB OAM hologram and ultrafast switching
Common problems/challenges of electrical biasHard to control meta-atoms locallyLocal control of nano pixels and integrating them with metasurfaces
Tradeoff between the number of images and their quality (efficiency)
Common problems/challenges of non-electrical inputNeed to mechanically move parts in case of stretching methodsLong phase transition time and optimizing design process
Long phase transition time and limited operating bandwidth in infrared regime in case of using PCMs

The development of a new design method or fabrication methods will expand the degrees of freedom of meta-atom design. Accordingly, we expect that alternative design methods such as inverse design and machine learning can assist researchers in designing metasurfaces that have the desired optical and electrical responses.23,168178 Also, recently developed 3D manufacturing methods of plasmonic and dielectric metasurfaces will enhance the flexibility of possible designs.179182 The development of deep learning will improve human intuition and imagination, and their design will be realized with advanced 3D manufacturing methods. These will provide effective strategies to design complete tunable metasurfaces.

Acknowledgments

J.R. conceived of the idea and initiated the project. J.K. and J.S. mainly wrote the manuscript, while Y.Y. and T.B. were partially involved. Y.Y. and S-W.M. gave advice on the recent progress of tunable metasurfaces and provided the outlook of current research. All authors participated in the discussion. J.R. guided all aspects of this work. This work was financially supported by the POSCO-POSTECH-RIST Convergence Research Center program funded by POSCO, and the National Research Foundation (NRF) grants (Grant Nos. NRF-2019R1A2C3003129, CAMM-2019M3A6B3030637, and NRF-2019R1A5A8080290) funded by the Ministry of Science and ICT, Republic of Korea. Y.Y. acknowledges the Hyundai Motor Chung Mong-Koo fellowship, and the NRF fellowship (Grant No. NRF-2021R1A6A3A13038935) funded by the Ministry of Education, Republic of Korea. The authors declare no competing interests.

References

1. 

J. Sung, G.-Y. Lee and B. Lee, “Progresses in the practical metasurface for holography and lens,” Nanophotonics, 8 (10), 1701 –1718 (2019). https://doi.org/10.1515/nanoph-2019-0203 Google Scholar

2. 

H.-T. Chen, A. J. Taylor and N. Yu, “A review of metasurfaces: physics and applications,” Rep. Prog. Phys., 79 (7), 076401 (2016). https://doi.org/10.1088/0034-4885/79/7/076401 RPPHAG 0034-4885 Google Scholar

3. 

S.-W. Moon et al., “Recent progress on ultrathin metalenses for flat Optics,” iScience, 23 (12), 101877 (2020). https://doi.org/10.1016/j.isci.2020.101877 Google Scholar

4. 

G. Yoon et al., “Recent progress on metasurfaces: applications and fabrication,” J. Phys. Appl. Phys., 54 (38), 383002 (2021). https://doi.org/10.1088/1361-6463/ac0faa Google Scholar

5. 

I. Kim et al., “Outfitting next generation displays with optical metasurfaces,” ACS Photonics, 5 (10), 3876 –3895 (2018). https://doi.org/10.1021/acsphotonics.8b00809 Google Scholar

6. 

D. Lee et al., “Metasurfaces-based imaging and applications: from miniaturized optical components to functional imaging platforms,” Nanoscale Adv., 2 (2), 605 –625 (2020). https://doi.org/10.1039/C9NA00751B Google Scholar

7. 

H. Jeong et al., “Emerging advanced metasurfaces: alternatives to conventional bulk optical devices,” Microelectron. Eng., 220 111146 (2020). https://doi.org/10.1016/j.mee.2019.111146 MIENEF 0167-9317 Google Scholar

8. 

J. Kim et al., “Geometric and physical configurations of meta-atoms for advanced metasurface holography,” InfoMat, 3 (7), 739 –754 (2021). https://doi.org/10.1002/inf2.12191 Google Scholar

9. 

W. T. Chen, A. Y. Zhu and F. Capasso, “Flat optics with dispersion-engineered metasurfaces,” Nat. Rev. Mater., 5 (8), 604 –620 (2020). https://doi.org/10.1038/s41578-020-0203-3 Google Scholar

10. 

N. Mahmood et al., “Polarisation insensitive multifunctional metasurfaces based on all-dielectric nanowaveguides,” Nanoscale, 10 (38), 18323 –18330 (2018). https://doi.org/10.1039/C8NR05633A NANOHL 2040-3364 Google Scholar

11. 

J. Mun et al., “Electromagnetic chirality: from fundamentals to nontraditional chiroptical phenomena,” Light Sci. Appl., 9 139 (2020). https://doi.org/10.1038/s41377-020-00367-8 Google Scholar

12. 

Z. H. Jiang et al., “Broadband and wide field-of-view plasmonic metasurface-enabled waveplates,” Sci. Rep., 4 7511 (2014). https://doi.org/10.1038/srep07511 SRCEC3 2045-2322 Google Scholar

13. 

F. Qin et al., “Hybrid bilayer plasmonic metasurface efficiently manipulates visible light,” Sci. Adv., 2 (2), e1501168 (2016). https://doi.org/10.1126/sciadv.1501168 STAMCV 1468-6996 Google Scholar

14. 

T. Stolt et al., “Backward phase-matched second-harmonic generation from stacked metasurfaces,” Phys. Rev. Lett., 126 (3), 033901 (2021). https://doi.org/10.1103/PhysRevLett.126.033901 PRLTAO 0031-9007 Google Scholar

15. 

Y. Yang et al., “Ultra-sharp circular dichroism induced by twisted layered C4 oligomers,” Adv. Theory Simul., 3 (3), 1900229 (2020). https://doi.org/10.1002/adts.201900229 Google Scholar

16. 

J. Bohn et al., “Active tuning of spontaneous emission by Mie-resonant dielectric metasurfaces,” Nano Lett., 18 (6), 3461 –3465 (2018). https://doi.org/10.1021/acs.nanolett.8b00475 NALEFD 1530-6984 Google Scholar

17. 

K.-T. Lee et al., “Electrically biased silicon metasurfaces with magnetic Mie resonance for tunable harmonic generation of light,” ACS Photonics, 6 (11), 2663 –2670 (2019). https://doi.org/10.1021/acsphotonics.9b01398 Google Scholar

18. 

M. Anzan-Uz-Zaman et al., “A novel approach to Fabry–Pérot-resonance-based lens and demonstrating deep-subwavelength imaging,” Sci. Rep., 10 10769 (2020). https://doi.org/10.1038/s41598-020-67409-4 SRCEC3 2045-2322 Google Scholar

19. 

G. Zheng et al., “Metasurface holograms reaching 80% efficiency,” Nat. Nanotechnol., 10 (4), 308 –312 (2015). https://doi.org/10.1038/nnano.2015.2 NNAABX 1748-3387 Google Scholar

20. 

G. Yoon et al., “Geometric metasurface enabling polarization independent beam splitting,” Sci. Rep., 8 9468 (2018). https://doi.org/10.1038/s41598-018-27876-2 SRCEC3 2045-2322 Google Scholar

21. 

W. S. L. Lee et al., “Broadband terahertz circular-polarization beam splitter,” Adv. Opt. Mater., 6 (3), 1700852 (2018). https://doi.org/10.1002/adom.201700852 2195-1071 Google Scholar

22. 

M. Kim, D. Lee and J. Rho, “Spin hall effect under arbitrarily polarized or unpolarized light,” Laser Photonics Rev., 15 (7), 2100138 (2021). https://doi.org/10.1002/lpor.202100138 Google Scholar

23. 

S. So et al., “On-demand design of spectrally sensitive multiband absorbers using an artificial neural network,” Photonics Res., 9 (4), B153 –B158 (2021). https://doi.org/10.1364/PRJ.415789 Google Scholar

24. 

T. Badloe, I. Kim and J. Rho, “Moth-eye shaped on-demand broadband and switchable perfect absorbers based on vanadium dioxide,” Sci. Rep., 10 4522 (2020). https://doi.org/10.1038/s41598-020-59729-2 SRCEC3 2045-2322 Google Scholar

25. 

D. Lee et al., “Multiple-patterning colloidal lithography-implemented scalable manufacturing of heat-tolerant titanium nitride broadband absorbers in the visible to near-infrared,” Microsyst. Nanoeng., 7 14 (2021). https://doi.org/10.1038/s41378-020-00237-8 Google Scholar

26. 

B. Ko et al., “Employing vanadium dioxide nanoparticles for flexible metasurfaces with switchable absorption properties at near-infrared frequencies,” J. Opt., 22 (11), 114002 (2020). https://doi.org/10.1088/2040-8986/abbc55 Google Scholar

27. 

T. Badloe, I. Kim and J. Rho, “Biomimetic ultra-broadband perfect absorbers optimised with reinforcement learning,” Phys. Chem. Chem. Phys., 22 (4), 2337 –2342 (2020). https://doi.org/10.1039/C9CP05621A PPCPFQ 1463-9076 Google Scholar

28. 

D. Lee et al., “Polarization-sensitive tunable absorber in visible and near-infrared regimes,” Sci. Rep., 8 12393 (2018). https://doi.org/10.1038/s41598-018-30835-6 SRCEC3 2045-2322 Google Scholar

29. 

G. Yoon et al., “Electrically tunable metasurface perfect absorber for infrared frequencies,” Nano Converg., 4 36 (2017). https://doi.org/10.1186/s40580-017-0131-0 Google Scholar

30. 

G. Yoon et al., “Printable nanocomposite metalens for high-contrast near-infrared imaging,” ACS Nano, 15 (2), 698 –706 (2021). https://doi.org/10.1021/acsnano.0c06968 ANCAC3 1936-0851 Google Scholar

31. 

G. Yoon et al., “Single-step manufacturing of hierarchical dielectric metalens in the visible,” Nat. Commun., 11 2268 (2020). https://doi.org/10.1038/s41467-020-16136-5 NCAOBW 2041-1723 Google Scholar

32. 

G. Yoon et al., “Pragmatic metasurface hologram at visible wavelength: the balance between diffraction efficiency and fabrication compatibility,” ACS Photonics, 5 (5), 1643 –1647 (2018). https://doi.org/10.1021/acsphotonics.7b01044 Google Scholar

33. 

M. A. Ansari et al., “A spin-encoded all-dielectric metahologram for visible light,” Laser Photonics Rev., 13 (5), 1900065 (2019). https://doi.org/10.1002/lpor.201900065 Google Scholar

34. 

I. Kim et al., “Holographic metasurface gas sensors for instantaneous visual alarms,” Sci. Adv., 7 (15), eabe9943 (2021). https://doi.org/10.1126/sciadv.abe9943 STAMCV 1468-6996 Google Scholar

35. 

M. A. Naveed et al., “Optical spin-symmetry breaking for high-efficiency directional helicity-multiplexed metaholograms,” Microsyst. Nanoeng., 7 5 (2021). https://doi.org/10.1038/s41378-020-00226-x Google Scholar

36. 

I. Kim et al., “Dual-band operating metaholograms with heterogeneous meta-atoms in the visible and near-infrared,” Adv. Opt. Mater., 9 (19), 2100609 (2021). https://doi.org/10.1002/adom.202100609 2195-1071 Google Scholar

37. 

H. S. Khaliq et al., “Giant chiro-optical responses in multipolar-resonances-based single-layer dielectric metasurfaces,” Photonics Res., 9 (9), 1667 –1674 (2021). https://doi.org/10.1364/PRJ.424477 Google Scholar

38. 

H. S. Khaliq et al., “Manifesting simultaneous optical spin conservation and spin isolation in diatomic metasurfaces,” Adv. Opt. Mater., 9 (8), 2002002 (2021). https://doi.org/10.1002/adom.202002002 2195-1071 Google Scholar

39. 

B. Xiong et al., “Realizing colorful holographic mimicry by metasurfaces,” Adv. Mater., 33 (21), 2005864 (2021). https://doi.org/10.1002/adma.202005864 ADVMEW 0935-9648 Google Scholar

40. 

H. Cai et al., “Polarization-insensitive medium-switchable holographic metasurfaces,” ACS Photonics, 8 (9), 2581 –2589 (2021). https://doi.org/10.1021/acsphotonics.1c00836 Google Scholar

41. 

M. Kim et al., “Visibly transparent radiative cooler under direct sunlight,” Adv. Opt. Mater., 9 (13), 2002226 (2021). https://doi.org/10.1002/adom.202002226 2195-1071 Google Scholar

42. 

S. So et al., “Inverse design of ultra-narrowband selective thermal emitters designed by artificial neural networks,” Opt. Mater. Express, 11 (7), 1863 –1873 (2021). https://doi.org/10.1364/OME.430306 Google Scholar

43. 

M. Kim et al., “Switchable diurnal radiative cooling by doped VO2,” Opto-Electron. Adv., 4 (5), 200006 (2021). https://doi.org/10.29026/oea.2021.200006 Google Scholar

44. 

I. Kim et al., “Nanophotonics for light detection and ranging technology,” Nat. Nanotechnol., 16 (5), 508 –524 (2021). https://doi.org/10.1038/s41565-021-00895-3 NNAABX 1748-3387 Google Scholar

45. 

M. Kim et al., “Spin Hall effect of light with near-unity efficiency in the microwave,” Laser Photonics Rev., 15 (2), 2000393 (2021). https://doi.org/10.1002/lpor.202000393 Google Scholar

46. 

M. Kim et al., “Observation of enhanced optical spin Hall effect in a vertical hyperbolic metamaterial,” ACS Photonics, 6 (10), 2530 –2536 (2019). https://doi.org/10.1021/acsphotonics.9b00904 Google Scholar

47. 

J. Jang et al., “Spectral modulation through the hybridization of Mie-scatterers and quasi-guided mode resonances: realizing full and gradients of structural color,” ACS Nano, 14 (11), 15317 –15326 (2020). https://doi.org/10.1021/acsnano.0c05656 ANCAC3 1936-0851 Google Scholar

48. 

J. Jang et al., “Self-powered humidity sensor using chitosan-based plasmonic metal–hydrogel–metal filters,” Adv. Opt. Mater., 8 (9), 1901932 (2020). https://doi.org/10.1002/adom.201901932 2195-1071 Google Scholar

49. 

C. Jung et al., “Near-zero reflection of all-dielectric structural coloration enabling polarization-sensitive optical encryption with enhanced switchability,” Nanophotonics, 10 (2), 919 –926 (2021). https://doi.org/10.1515/nanoph-2020-0440 Google Scholar

50. 

T. Lee et al., “Nearly perfect transmissive subtractive coloration through the spectral amplification of Mie scattering and lattice resonance,” ACS Appl. Mater. Interfaces, 13 (22), 26299 –26307 (2021). https://doi.org/10.1021/acsami.1c03427 AAMICK 1944-8244 Google Scholar

51. 

D. Lee et al., “Sub-ambient daytime radiative cooling by silica-coated porous anodic aluminum oxide,” Nano Energy, 79 105426 (2021). https://doi.org/10.1016/j.nanoen.2020.105426 Google Scholar

52. 

J. Jang et al., “Full and gradient structural colouration by lattice amplified gallium nitride Mie-resonators,” Nanoscale, 12 (41), 21392 –21400 (2020). https://doi.org/10.1039/D0NR05624C NANOHL 2040-3364 Google Scholar

53. 

W. T. Chen et al., “A broadband achromatic metalens for focusing and imaging in the visible,” Nat. Nanotechnol., 13 (3), 220 –226 (2018). https://doi.org/10.1038/s41565-017-0034-6 NNAABX 1748-3387 Google Scholar

54. 

H. Ren et al., “Complex-amplitude metasurface-based orbital angular momentum holography in momentum space,” Nat. Nanotechnol., 15 (11), 948 –955 (2020). https://doi.org/10.1038/s41565-020-0768-4 NNAABX 1748-3387 Google Scholar

55. 

A. C. Overvig et al., “Dielectric metasurfaces for complete and independent control of the optical amplitude and phase,” Light Sci. Appl., 8 92 (2019). https://doi.org/10.1038/s41377-019-0201-7 Google Scholar

56. 

K. Kim et al., “Facile nanocasting of dielectric metasurfaces with sub-100 nm resolution,” ACS Appl. Mater. Interfaces, 11 (29), 26109 –26115 (2019). https://doi.org/10.1021/acsami.9b07774 AAMICK 1944-8244 Google Scholar

57. 

Y. Yang et al., “Revealing structural disorder in hydrogenated amorphous silicon for a low-loss photonic platform at visible frequencies,” Adv. Mater., 33 (9), 2005893 (2021). https://doi.org/10.1002/adma.202005893 ADVMEW 0935-9648 Google Scholar

58. 

M. Khorasaninejad et al., “Metalenses at visible wavelengths: diffraction-limited focusing and subwavelength resolution imaging,” Science, 352 (6290), 1190 –1194 (2016). https://doi.org/10.1126/science.aaf6644 SCIEAS 0036-8075 Google Scholar

59. 

B. H. Chen et al., “GaN metalens for pixel-level full-color routing at visible light,” Nano Lett., 17 (10), 6345 –6352 (2017). https://doi.org/10.1021/acs.nanolett.7b03135 NALEFD 1530-6984 Google Scholar

60. 

S. Wang et al., “A broadband achromatic metalens in the visible,” Nat. Nanotechnol., 13 (3), 227 –232 (2018). https://doi.org/10.1038/s41565-017-0052-4 NNAABX 1748-3387 Google Scholar

61. 

Q. He, S. Sun and L. Zhou, “Tunable/reconfigurable metasurfaces: physics and applications,” Research, 2019 1849272 (2019). https://doi.org/10.34133/2019/1849272 Google Scholar

62. 

T. Badloe et al., “Tunable metasurfaces: the path to fully active nanophotonics,” Adv. Photonics Res., 2 (9), 2000205 (2021). https://doi.org/10.1002/adpr.202000205 Google Scholar

63. 

H.-H. Hsiao, C. H. Chu and D. P. Tsai, “Fundamentals and applications of metasurfaces,” Small Methods, 1 (4), 1600064 (2017). https://doi.org/10.1002/smtd.201600064 Google Scholar

64. 

Q. He et al., “High-efficiency metasurfaces: principles, realizations, and applications,” Adv. Opt. Mater., 6 (19), 1800415 (2018). https://doi.org/10.1002/adom.201800415 2195-1071 Google Scholar

65. 

M. L. Tseng et al., “Metalenses: advances and applications,” Adv. Opt. Mater., 6 (18), 1800554 (2018). https://doi.org/10.1002/adom.201800554 2195-1071 Google Scholar

66. 

W.-J. Joo et al., “Metasurface-driven OLED displays beyond 10,000 pixels per inch,” Science, 370 (6515), 459 –463 (2020). https://doi.org/10.1126/science.abc8530 SCIEAS 0036-8075 Google Scholar

67. 

H. Gao et al., “Dynamic 3D meta-holography in visible range with large frame number and high frame rate,” Sci. Adv., 6 (28), eaba8595 (2020). https://doi.org/10.1126/sciadv.aba8595 STAMCV 1468-6996 Google Scholar

68. 

G.-Y. Lee, J. Sung and B. Lee, “Metasurface optics for imaging applications,” MRS Bull., 45 (3), 202 –209 (2020). https://doi.org/10.1557/mrs.2020.64 MRSBEA 0883-7694 Google Scholar

69. 

B. Yao et al., “Spin-decoupled metalens with intensity-tunable multiple focal points,” Photonics Res., 9 (6), 1019 –1032 (2021). https://doi.org/10.1364/PRJ.420665 Google Scholar

70. 

W. Wang et al., “Spin-selected and spin-independent dielectric metalenses,” J. Opt., 20 (9), 095102 (2018). https://doi.org/10.1088/2040-8986/aad6fd Google Scholar

71. 

B. Groever et al., “High-efficiency chiral meta-lens,” Sci. Rep., 8 7240 (2018). https://doi.org/10.1038/s41598-018-25675-3 SRCEC3 2045-2322 Google Scholar

72. 

R. Fu et al., “Reconfigurable step-zoom metalens without optical and mechanical compensations,” Opt. Express, 27 (9), 12221 –12230 (2019). https://doi.org/10.1364/OE.27.012221 OPEXFF 1094-4087 Google Scholar

73. 

L. Yu et al., “Spin angular momentum controlled multifunctional all-dielectric metasurface doublet,” Laser Photonics Rev., 14 (6), 1900324 (2020). https://doi.org/10.1002/lpor.201900324 Google Scholar

74. 

T. Zhou et al., “Helicity multiplexed terahertz multi-foci metalens,” Opt. Lett., 45 (2), 463 –466 (2020). https://doi.org/10.1364/OL.381105 OPLEDP 0146-9592 Google Scholar

75. 

J. Zhang et al., “Polarization-enabled tunable focusing by visible-light metalenses with geometric and propagation phase,” J. Opt., 21 (11), 115102 (2019). https://doi.org/10.1088/2040-8986/ab48cf Google Scholar

76. 

Y. A. Zhang et al., “Dual-layer electrode-driven liquid crystal lens with electrically tunable focal length and focal plane,” Opt. Commun., 412 114 –120 (2018). https://doi.org/10.1016/j.optcom.2017.12.008 OPCOB8 0030-4018 Google Scholar

77. 

M. Bosch et al., “Electrically actuated varifocal lens based on liquid-crystal-embedded dielectric metasurfaces,” Nano Lett., 21 (9), 3849 –3856 (2021). https://doi.org/10.1021/acs.nanolett.1c00356 NALEFD 1530-6984 Google Scholar

78. 

Z. Shen et al., “Liquid crystal integrated metalens with tunable chromatic aberration,” Adv. Photonics, 2 (3), 036002 (2020). https://doi.org/10.1117/1.AP.2.3.036002 Google Scholar

79. 

C.-Y. Fan et al., “Electrically modulated varifocal metalens combined with twisted nematic liquid crystals,” Opt. Express, 28 (7), 10609 –10617 (2020). https://doi.org/10.1364/OE.386563 OPEXFF 1094-4087 Google Scholar

80. 

M. Sun et al., “Efficient visible light modulation based on electrically tunable all dielectric metasurfaces embedded in thin-layer nematic liquid crystals,” Sci. Rep., 9 8673 (2019). https://doi.org/10.1038/s41598-019-45091-5 SRCEC3 2045-2322 Google Scholar

81. 

S. Zhou et al., “Liquid crystal integrated metalens with dynamic focusing property,” Opt. Lett., 45 (15), 4324 –4327 (2020). https://doi.org/10.1364/OL.398601 OPLEDP 0146-9592 Google Scholar

82. 

T. Badloe et al., “Electrically tunable bifocal metalens with diffraction-limited focusing and imaging at visible wavelengths,” Adv. Sci., 8 (21), 2102646 (2021). https://doi.org/10.1002/advs.202102646 Google Scholar

83. 

W. Liu et al., “Graphene-enabled electrically controlled terahertz meta-lens,” Photonics Res., 6 (7), 703 –708 (2018). https://doi.org/10.1364/PRJ.6.000703 Google Scholar

84. 

P. Ding et al., “Graphene aperture-based metalens for dynamic focusing of terahertz waves,” Opt. Express, 26 (21), 28038 –28050 (2018). https://doi.org/10.1364/OE.26.028038 OPEXFF 1094-4087 Google Scholar

85. 

S. Park et al., “Electrically focus-tuneable ultrathin lens for high-resolution square subpixels,” Light Sci. Appl., 9 98 (2020). https://doi.org/10.1038/s41377-020-0329-5 Google Scholar

86. 

D. Chen et al., “Continuously tunable metasurfaces controlled by single electrode uniform bias-voltage based on nonuniform periodic rectangular graphene arrays,” Opt. Express, 28 (20), 29306 –29317 (2020). https://doi.org/10.1364/OE.401255 OPEXFF 1094-4087 Google Scholar

87. 

Z. Zhang et al., “Graphene-enabled electrically tunability of metalens in the terahertz range,” Opt. Express, 28 (19), 28101 –28112 (2020). https://doi.org/10.1364/OE.401627 OPEXFF 1094-4087 Google Scholar

88. 

Z. Huang et al., “Dynamical tuning of terahertz meta-lens assisted by graphene,” J. Opt. Soc. Am. B, 34 (9), 1848 –1854 (2017). https://doi.org/10.1364/JOSAB.34.001848 JOBPDE 0740-3224 Google Scholar

89. 

S. Park et al., “Focus-tunable planar lenses by controlled carriers over exciton,” Adv. Opt. Mater., 9 (2), 2001526 (2021). https://doi.org/10.1002/adom.202001526 2195-1071 Google Scholar

90. 

N. Xu et al., “Electrically-driven zoom metalens based on dynamically controlling the phase of barium titanate (BTO) column antennas,” Nanomaterials, 11 (3), 729 (2021). https://doi.org/10.3390/nano11030729 Google Scholar

91. 

C. Huang et al., “Graphene-integrated reconfigurable metasurface for independent manipulation of reflection magnitude and phase,” Adv. Opt. Mater., 9 (7), 2001950 (2021). https://doi.org/10.1002/adom.202001950 2195-1071 Google Scholar

92. 

H. Chen et al., “Microwave programmable graphene metasurface,” ACS Photonics, 7 (6), 1425 –1435 (2020). https://doi.org/10.1021/acsphotonics.9b01807 Google Scholar

93. 

B. Zeng et al., “Hybrid graphene metasurfaces for high-speed mid-infrared light modulation and single-pixel imaging,” Light Sci. Appl., 7 51 (2018). https://doi.org/10.1038/s41377-018-0055-4 Google Scholar

94. 

Z. Su et al., “Complete control of Smith-Purcell radiation by graphene metasurfaces,” ACS Photonics, 6 (8), 1947 –1954 (2019). https://doi.org/10.1021/acsphotonics.9b00251 Google Scholar

95. 

W. Ma et al., “Dual-band light focusing using stacked graphene metasurfaces,” ACS Photonics, 4 (7), 1770 –1775 (2017). https://doi.org/10.1021/acsphotonics.7b00351 Google Scholar

96. 

H.-S. Ee and R. Agarwal, “Tunable metasurface and flat optical zoom lens on a stretchable substrate,” Nano Lett., 16 (4), 2818 –2823 (2016). https://doi.org/10.1021/acs.nanolett.6b00618 NALEFD 1530-6984 Google Scholar

97. 

K. Iwami et al., “Demonstration of focal length tuning by rotational varifocal moiré metalens in an ir-A wavelength,” Opt. Express, 28 (24), 35602 –35614 (2020). https://doi.org/10.1364/OE.411054 OPEXFF 1094-4087 Google Scholar

98. 

S. Colburn, A. Zhan and A. Majumdar, “Varifocal zoom imaging with large area focal length adjustable metalenses,” Optica, 5 (7), 825 –831 (2018). https://doi.org/10.1364/OPTICA.5.000825 Google Scholar

99. 

E. Arbabi et al., ““MEMS-tunable dielectric metasurface lens,” Nat. Commun., 9 812 (2018). https://doi.org/10.1038/s41467-018-03155-6 NCAOBW 2041-1723 Google Scholar

100. 

M. Y. Shalaginov et al., “Reconfigurable all-dielectric metalens with diffraction-limited performance,” Nat. Commun., 12 1225 (2021). https://doi.org/10.1038/s41467-021-21440-9 NCAOBW 2041-1723 Google Scholar

101. 

S. Qin et al., “Near-infrared thermally modulated varifocal metalens based on the phase change material Sb2S3,” Opt. Express, 29 (5), 7925 –7934 (2021). https://doi.org/10.1364/OE.420014 OPEXFF 1094-4087 Google Scholar

102. 

Z. Guanxing et al., “Reconfigurable metasurfaces with mechanical actuations: towards flexible and tunable photonic devices,” J. Opt., 23 (2), 013001 (2020). https://doi.org/10.1088/2040-8986/abcc52 Google Scholar

103. 

P. Gutruf et al., “Mechanically tunable dielectric resonator metasurfaces at visible frequencies,” ACS Nano, 10 (2), 133 –141 (2016). https://doi.org/10.1021/acsnano.5b05954 ANCAC3 1936-0851 Google Scholar

104. 

S. M. Kamali et al., “Highly tunable elastic dielectric metasurface lenses,” Laser Photonics Rev., 10 (6), 1002 –1008 (2016). https://doi.org/10.1002/lpor.201600144 Google Scholar

105. 

F. Cheng et al., “Mechanically tunable focusing metamirror in the visible,” Opt. Express, 27 (11), 15194 –15204 (2019). https://doi.org/10.1364/OE.27.015194 OPEXFF 1094-4087 Google Scholar

106. 

S. Wei et al., “A varifocal graphene metalens for broadband zoom imaging covering the entire visible region,” ACS Nano, 15 (3), 4769 –4776 (2021). https://doi.org/10.1021/acsnano.0c09395 ANCAC3 1936-0851 Google Scholar

107. 

Y. Wei et al., “Compact optical polarization-insensitive zoom metalens doublet,” Adv. Opt. Mater., 8 (13), 2000142 (2020). https://doi.org/10.1002/adom.202000142 2195-1071 Google Scholar

108. 

Y. Guo et al., “Experimental demonstration of a continuous varifocal metalens with large zoom range and high imaging resolution,” Appl. Phys. Lett., 115 (16), 163103 (2019). https://doi.org/10.1063/1.5123367 APPLAB 0003-6951 Google Scholar

109. 

N. Yilmaz et al., “Rotationally tunable polarization-insensitive single and multifocal metasurface,” J. Opt., 21 (4), 045105 (2019). https://doi.org/10.1088/2040-8986/ab0d5f Google Scholar

110. 

Y. Cui et al., “Reconfigurable continuous-zoom metalens in visible band,” Chin. Opt. Lett., 17 111603 (2019). CJOEE3 1671-7694 Google Scholar

111. 

S. Colburn and A. Majumdar, “Simultaneous achromatic and varifocal imaging with quartic metasurfaces in the visible,” ACS Photonics, 7 (2), 120 –127 (2020). https://doi.org/10.1021/acsphotonics.9b01216 Google Scholar

112. 

T. Roy et al., “Dynamic metasurface lens based on MEMS technology,” APL Photonics, 3 (2), 021302 (2018). https://doi.org/10.1063/1.5018865 Google Scholar

113. 

Z. Han et al., “MEMS-actuated metasurface Alvarez lens,” Microsyst. Nanoeng., 6 79 (2020). https://doi.org/10.1038/s41378-020-00190-6 Google Scholar

114. 

C. Meng et al., “Dynamic piezoelectric MEMS-based optical metasurfaces,” Sci. Adv., 7 (26), eabg5639 (2021). https://doi.org/10.1126/sciadv.abg5639 STAMCV 1468-6996 Google Scholar

115. 

Q. Wang et al., “Optically reconfigurable metasurfaces and photonic devices based on phase change materials,” Nat. Photonics, 10 60 –65 (2016). https://doi.org/10.1038/nphoton.2015.247 NPAHBY 1749-4885 Google Scholar

116. 

W. Bai et al., “Actively tunable metalens array based on patterned phase change materials,” Appl. Sci., 9 (22), 4927 (2019). https://doi.org/10.3390/app9224927 Google Scholar

117. 

W. Bai et al., “Tunable duplex metalens based on phase-change materials in communication range,” Nanomaterials, 9 (7), 993 (2019). https://doi.org/10.3390/nano9070993 Google Scholar

118. 

W. Bai et al., “Near-infrared tunable metalens based on phase change material Ge2Sb2Te5,” Sci. Rep., 9 5368 (2019). https://doi.org/10.1038/s41598-019-41859-x SRCEC3 2045-2322 Google Scholar

119. 

F.-Z. Shu et al., “Electrically driven tunable broadband polarization states via active metasurfaces based on Joule-heat-induced phase transition of vanadium dioxide,” Laser Photonics Rev., 15 (10), 2100155 (2021). https://doi.org/10.1002/lpor.202100155 Google Scholar

120. 

P.-A. Blanche et al., “Holographic three-dimensional telepresence using large-area photorefractive polymer,” Nature, 468 (7320), 80 –83 (2010). https://doi.org/10.1038/nature09521 Google Scholar

121. 

P. Genevet and F. Capasso, “Holographic optical metasurfaces: a review of current progress,” Rep. Prog. Phys., 78 (2), 024401 (2015). https://doi.org/10.1088/0034-4885/78/2/024401 RPPHAG 0034-4885 Google Scholar

122. 

N. Yu et al., “Light propagation with phase discontinuities: generalized laws of reflection and refraction,” Science, 334 (6054), 333 –337 (2011). https://doi.org/10.1126/science.1210713 SCIEAS 0036-8075 Google Scholar

123. 

R. W. Gerchberg and W. O. Saxton, “A practical algorithm for the determination of phase from image and diffraction plane pictures,” Optik, 35 (2), 237 –249 (1972). OTIKAJ 0030-4026 Google Scholar

124. 

J. R. Fienup, “Phase retrieval algorithms: a comparison,” Appl. Opt., 21 (15), 2758 –2769 (1982). https://doi.org/10.1364/AO.21.002758 APOPAI 0003-6935 Google Scholar

125. 

Y. Yifat et al., “Highly efficient and broadband wide-angle holography using patch-dipole nanoantenna reflectarrays,” Nano Lett., 14 (5), 2485 –2490 (2014). https://doi.org/10.1021/nl5001696 NALEFD 1530-6984 Google Scholar

126. 

Q. Song et al., “Plasmonic topological metasurface by encircling an exceptional point,” Science, 373 (6559), 1133 –1137 (2021). https://doi.org/10.1126/science.abj3179 SCIEAS 0036-8075 Google Scholar

127. 

J. Scheuer, “Metasurfaces-based holography and beam shaping: engineering the phase profile of light,” Nanophotonics, 6 (2), 137 –152 (2017). https://doi.org/10.1515/nanoph-2016-0109 Google Scholar

128. 

J. P. Balthasar Mueller et al., “Metasurface polarization optics: independent phase control of arbitrary orthogonal states of polarization,” Phys. Rev. Lett., 118 (11), 113901 (2017). https://doi.org/10.1103/PhysRevLett.118.113901 PRLTAO 0031-9007 Google Scholar

129. 

D. Wen et al., “Helicity multiplexed broadband metasurface holograms,” Nat. Commun., 6 (2), 8241 (2015). https://doi.org/10.1038/ncomms9241 NCAOBW 2041-1723 Google Scholar

130. 

B. Wang et al., “Polarization-controlled color-tunable holograms with dielectric metasurfaces,” Optica, 4 (11), 1368 –1371 (2017). https://doi.org/10.1364/OPTICA.4.001368 Google Scholar

131. 

Q. Wang et al., “Reflective chiral meta-holography: multiplexing holograms for circularly polarized waves,” Light Sci. Appl., 7 25 (2018). https://doi.org/10.1038/s41377-018-0019-8 Google Scholar

132. 

Z.-L. Deng et al., “Diatomic metasurface for vectorial holography,” Nano Lett., 18 (5), 2885 –2892 (2018). https://doi.org/10.1021/acs.nanolett.8b00047 NALEFD 1530-6984 Google Scholar

133. 

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

134. 

X. Zhang et al., “Direct polarization measurement using a multiplexed Pancharatnam–Berry metahologram,” Optica, 6 (9), 1190 –1198 (2019). https://doi.org/10.1364/OPTICA.6.001190 Google Scholar

135. 

P. Zheng et al., “Metasurface-based key for computational imaging encryption,” Sci. Adv., 7 (21), eabg0363 (2021). https://doi.org/10.1126/sciadv.abg0363 STAMCV 1468-6996 Google Scholar

136. 

L. Jin et al., “Dielectric multi-momentum meta-transformer in the visible,” Nat. Commun., 10 4789 (2019). https://doi.org/10.1038/s41467-019-12637-0 NCAOBW 2041-1723 Google Scholar

137. 

H. Ren et al., “Metasurface orbital angular momentum holography,” Nat. Commun., 10 2986 (2019). https://doi.org/10.1038/s41467-019-11030-1 NCAOBW 2041-1723 Google Scholar

138. 

H. Zhou et al., “Polarization-encrypted orbital angular momentum multiplexed metasurface holography,” ACS Nano, 14 (5), 5553 –5559 (2020). https://doi.org/10.1021/acsnano.9b09814 ANCAC3 1936-0851 Google Scholar

139. 

Q. Xiao et al., “Orbital‐angular‐momentum‐encrypted holography based on coding information metasurface,” Adv. Opt. Mater., 9 (11), 2002155 (2021). https://doi.org/10.1002/adom.202002155 2195-1071 Google Scholar

140. 

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

141. 

X. Li et al., “Code division multiplexing inspired dynamic metasurface holography,” Adv. Funct. Mater., 31 (35), 2103326 (2021). https://doi.org/10.1002/adfm.202103326 AFMDC6 1616-301X Google Scholar

142. 

P. Georgi et al., “Optical secret sharing with cascaded metasurface holography,” Sci. Adv., 7 (16), eabf9718 (2021). https://doi.org/10.1126/sciadv.abf9718 STAMCV 1468-6996 Google Scholar

143. 

S. M. Kamali et al., “Angle-multiplexed metasurfaces: encoding independent wavefronts in a single metasurface under different illumination angles,” Phys. Rev. X, 7 (4), 041056 (2017). https://doi.org/10.1103/PhysRevX.7.041056 PRXHAE 2160-3308 Google Scholar

144. 

C. Jung et al., “Metasurface-driven optically variable devices,” Chem. Rev., 121 (21), 13013 –13050 (2021). https://doi.org/10.1021/acs.chemrev.1c00294 CHREAY 0009-2665 Google Scholar

145. 

J. Wang et al., “Terabit free-space data transmission employing orbital angular momentum multiplexing,” Nat. Photonics, 6 (7), 488 –496 (2012). https://doi.org/10.1038/nphoton.2012.138 NPAHBY 1749-4885 Google Scholar

146. 

G. Tricoles, “Computer generated holograms: an historical review,” Appl. Opt., 26 (20), 4351 –4360 (1987). https://doi.org/10.1364/AO.26.004351 APOPAI 0003-6935 Google Scholar

147. 

J. Hahn et al., “Wide viewing angle dynamic holographic stereogram with a curved array of spatial light modulators,” Opt. Express, 16 (16), 12372 –12386 (2008). https://doi.org/10.1364/OE.16.012372 OPEXFF 1094-4087 Google Scholar

148. 

S.-Q. Li et al., “Phase-only transmissive spatial light modulator based on tunable dielectric metasurface,” Science, 364 (6445), 1087 –1090 (2019). https://doi.org/10.1126/science.aaw6747 SCIEAS 0036-8075 Google Scholar

149. 

C. Zou et al., “Electrically tunable transparent displays for visible light based on dielectric metasurfaces,” ACS Photonics, 6 (6), 1533 –1540 (2019). https://doi.org/10.1021/acsphotonics.9b00301 Google Scholar

150. 

J. Li et al., “Electrically-controlled digital metasurface device for light projection displays,” Nat. Commun., 11 3574 (2020). https://doi.org/10.1038/s41467-020-17390-3 NCAOBW 2041-1723 Google Scholar

151. 

Y. Hu et al., “Electrically tunable multifunctional polarization-dependent metasurfaces integrated with liquid crystals in the visible region,” Nano Lett., 21 (11), 4554 –4562 (2021). https://doi.org/10.1021/acs.nanolett.1c00104 NALEFD 1530-6984 Google Scholar

152. 

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

153. 

S. Zhu et al., “Liquid crystal integrated metadevice for reconfigurable hologram displays and optical encryption,” Opt. Express, 29 (6), 9553 –9564 (2021). https://doi.org/10.1364/OE.419914 OPEXFF 1094-4087 Google Scholar

154. 

L. Li et al., “Electromagnetic reprogrammable coding-metasurface holograms,” Nat. Commun., 8 197 (2017). https://doi.org/10.1038/s41467-017-00164-9 NCAOBW 2041-1723 Google Scholar

155. 

R. Kaissner et al., “Electrochemically controlled metasurfaces with high-contrast switching at visible frequencies,” Sci. Adv., 7 (19), eabd9450 (2021). https://doi.org/10.1126/sciadv.abd9450 STAMCV 1468-6996 Google Scholar

156. 

M. Zhang et al., “Plasmonic metasurfaces for switchable photonic spin-orbit interactions based on phase change materials,” Adv. Sci., 5 (10), 1800835 (2018). https://doi.org/10.1002/advs.201800835 Google Scholar

157. 

X. Liu et al., “Thermally dependent dynamic meta‐holography using a vanadium dioxide integrated metasurface,” Adv. Opt. Mater., 7 (12), 1900175 (2019). https://doi.org/10.1002/adom.201900175 2195-1071 Google Scholar

158. 

F. Zhang et al., “Multistate switching of photonic angular momentum coupling in phase‐change metadevices,” Adv. Mater., 32 (39), 1908194 (2020). https://doi.org/10.1002/adma.201908194 ADVMEW 0935-9648 Google Scholar

159. 

C. Choi et al., “Hybrid state engineering of phase‐change metasurface for all-optical cryptography,” Adv. Funct. Mater., 31 (4), 2007210 (2021). https://doi.org/10.1002/adfm.202007210 AFMDC6 1616-301X Google Scholar

160. 

J. Li et al., “Addressable metasurfaces for dynamic holography and optical information encryption,” Sci. Adv., 4 (6), eaar6768 (2018). https://doi.org/10.1126/sciadv.aar6768 STAMCV 1468-6996 Google Scholar

161. 

T. Li et al., “Reconfigurable metasurface hologram by utilizing addressable dynamic pixels,” Opt. Express, 27 (15), 21153 –21162 (2019). https://doi.org/10.1364/OE.27.021153 OPEXFF 1094-4087 Google Scholar

162. 

J. Li et al., “Magnesium-based metasurfaces for dual-function switching between dynamic holography and dynamic color display,” ACS Nano, 14 (7), 7892 –7898 (2020). https://doi.org/10.1021/acsnano.0c01469 ANCAC3 1936-0851 Google Scholar

163. 

S. C. Malek, H.-S. Ee and R. Agarwal, “Strain multiplexed metasurface holograms on a stretchable substrate,” Nano Lett., 17 (6), 3641 –3645 (2017). https://doi.org/10.1021/acs.nanolett.7b00807 NALEFD 1530-6984 Google Scholar

164. 

I. Kim et al., “Stimuli‐responsive dynamic metaholographic displays with designer liquid crystal modulators,” Adv. Mater., 32 (50), 2004664 (2020). https://doi.org/10.1002/adma.202004664 ADVMEW 0935-9648 Google Scholar

165. 

Y. Zhang et al., “Electrically reconfigurable non-volatile metasurface using low-loss optical phase-change material,” Nat. Nanotechnol., 16 (6), 661 –666 (2021). https://doi.org/10.1038/s41565-021-00881-9 NNAABX 1748-3387 Google Scholar

166. 

Y. Wang et al., “Electrical tuning of phase-change antennas and metasurfaces,” Nat. Nanotechnol., 16 (6), 667 –672 (2021). https://doi.org/10.1038/s41565-021-00882-8 NNAABX 1748-3387 Google Scholar

167. 

J. Karst et al., “Electrically switchable metallic polymer nanoantennas,” Science, 374 (6567), 612 –616 (2021). https://doi.org/10.1126/science.abj3433 SCIEAS 0036-8075 Google Scholar

168. 

W. Ma, F. Cheng and Y. Liu, “Deep-learning-enabled on-demand design of chiral metamaterials,” ACS Nano, 12 (6), 6326 –6334 (2018). https://doi.org/10.1021/acsnano.8b03569 ANCAC3 1936-0851 Google Scholar

169. 

S. Molesky et al., “Inverse design in nanophotonics,” Nat. Photonics, 12 (11), 659 –670 (2018). https://doi.org/10.1038/s41566-018-0246-9 NPAHBY 1749-4885 Google Scholar

170. 

S. So et al., “Deep learning enabled inverse design in nanophotonics,” Nanophotonics, 9 (5), 1041 –1057 (2020). https://doi.org/10.1515/nanoph-2019-0474 Google Scholar

171. 

J. Noh et al., “Design of a transmissive metasurface antenna using deep neural networks,” Opt. Mater. Express, 11 (7), 2310 –2317 (2021). https://doi.org/10.1364/OME.421990 Google Scholar

172. 

S. So, J. Mun and J. Rho, “Simultaneous inverse design of materials and structures via deep learning: demonstration of dipole resonance engineering using core–shell nanoparticles,” ACS Appl. Mater. Interfaces, 11 (27), 24264 –24268 (2019). https://doi.org/10.1021/acsami.9b05857 AAMICK 1944-8244 Google Scholar

173. 

S. So and J. Rho, “Designing nanophotonic structures using conditional deep convolutional generative adversarial networks,” Nanophotonics, 8 (7), 1255 –1261 (2019). https://doi.org/10.1515/nanoph-2019-0117 Google Scholar

174. 

W. Ma et al., “Deep learning for the design of photonic structures,” Nat. Photonics, 15 (2), 77 –90 (2021). https://doi.org/10.1038/s41566-020-0685-y NPAHBY 1749-4885 Google Scholar

175. 

Y. Xu et al., “Interfacing photonics with artificial intelligence: an innovative design strategy for photonic structures and devices based on artificial neural networks,” Photonics Res., 9 (4), B135 –B152 (2021). https://doi.org/10.1364/PRJ.417693 Google Scholar

176. 

X. An et al., “Broadband achromatic metalens design based on deep neural networks,” Opt. Lett., 46 (16), 3881 –3884 (2021). https://doi.org/10.1364/OL.427221 OPLEDP 0146-9592 Google Scholar

177. 

M. M. R. Elsawy et al., “Multiobjective statistical learning optimization of RGB metalens,” ACS Photonics, 8 (8), 2498 –2508 (2021). https://doi.org/10.1021/acsphotonics.1c00753 Google Scholar

178. 

W. Ma et al., “Probabilistic representation and inverse design of metamaterials based on a deep generative model with semi-supervised learning strategy,” Adv. Mater., 31 (35), 1901111 (2019). https://doi.org/10.1002/adma.201901111 ADVMEW 0935-9648 Google Scholar

179. 

W. Jung et al., “Three-dimensional nanoprinting via charged aerosol jets,” Nature, 592 (7852), 54 –59 (2021). https://doi.org/10.1038/s41586-021-03353-1 Google Scholar

180. 

Y. Hou et al., “Design and fabrication of three-dimensional chiral nanostructures based on stepwise glancing angle deposition technology,” Langmuir, 29 (3), 867 –872 (2013). https://doi.org/10.1021/la304122f LANGD5 0743-7463 Google Scholar

181. 

H. Lee et al., “Three-dimensional assembly of nanoparticles from charged aerosols,” Nano Lett., 11 (2), 119 –124 (2011). https://doi.org/10.1021/nl103787k NALEFD 1530-6984 Google Scholar

182. 

A. G. Mark et al., “Hybrid nanocolloids with programmed three-dimensional shape and material composition,” Nat. Mater., 12 (9), 802 –807 (2013). https://doi.org/10.1038/nmat3685 NMAACR 1476-1122 Google Scholar

Biography

Jaekyung Kim received his BS degree in mechanical engineering at Pohang University of Science and Technology (POSTECH), Republic of Korea (2021). He is currently an MS/PhD student under the guidance of Prof. Junsuk Rho at POSTECH. His research interests focus on nanofabrication and dielectric metasurfaces.

Junhwa Seong received his BS degree in mechanical engineering at POSTECH (2021). He is currently an MS/PhD student under the guidance of Prof. Junsuk Rho at POSTECH. His research interests focus on nanofabrication and dielectric metasurfaces.

Younghwan Yang received his BS degree in mechanical engineering from Ajou University, Republic of Korea (2018). He is currently an MS/PhD student under the guidance of Prof. Junsuk Rho at POSTECH. He is a recipient of the Hyundai Motor Chung Mong-Koo fellowship, and of the NRF doctoral candidate fellowship

Seong-Won Moon received his BS and MS degrees in electronics engineering at Kyungpook National University in 2020. He is currently a PhD student under the guidance of Prof. Junsuk Rho at POSTECH. His research interests focus on metasurfaces with orbital angular momentum.

Trevon Badloe received his MPhys (hons) degree from the University of Sheffield, United Kingdom, in 2012, with a year of study abroad at the National University of Singapore in 2010. After three years of teaching courses in English and classical mechanics as an assistant professor at Yeungjin University, Republic of Korea, he started working toward his PhD in mechanical engineering at POSTECH, Republic of Korea, in 2017. His interests include tunable metamaterials and metasurfaces, and machine learning for the design and optimization of nanophotonic applications.

Junsuk Rho is a Mu-Eun-Jae endowed chair professor with a joint appointment in the Department of Mechanical Engineering and the Department of Chemical Engineering at POSTECH. He received his BS (2007) and MS (2008) degrees in mechanical engineering at Seoul National University and the University of Illinois, Urbana–Champaign, respectively. After getting his PhD (2013) in mechanical engineering and nanoscale science and engineering from University of California Berkeley, he worked as a postdoctoral fellow in the Materials Sciences Division at Lawrence Berkeley National Laboratory and as Ugo Fano fellow in the Nanoscience and Technology Division at Argonne National Laboratory.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Jaekyung Kim, Junhwa Seong, Younghwan Yang, Seong-Won Moon, Trevon Badloe, and Junsuk Rho "Tunable metasurfaces towards versatile metalenses and metaholograms: a review," Advanced Photonics 4(2), 024001 (7 March 2022). https://doi.org/10.1117/1.AP.4.2.024001
Received: 31 August 2021; Accepted: 28 December 2021; Published: 7 March 2022
Lens.org Logo
CITATIONS
Cited by 120 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Holograms

Holography

Polarization

Liquid crystals

Modulation

Dielectric polarization

Graphene

Back to Top