Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter February 6, 2017

Traditional and emerging materials for optical metasurfaces

  • Alexander Y. Zhu , Arseniy I. Kuznetsov , Boris Luk’yanchuk , Nader Engheta and Patrice Genevet EMAIL logo
From the journal Nanophotonics

Abstract

One of the most promising and vibrant research areas in nanotechnology has been the field of metasurfaces. These are two dimensional representations of metaatoms, or artificial interfaces designed to possess specialized electromagnetic properties which do not occur in nature, for specific applications. In this article, we present a brief review of metasurfaces from a materials perspective, and examine how the choice of different materials impact functionalities ranging from operating bandwidth to efficiencies. We place particular emphasis on emerging and non-traditional materials for metasurfaces such as high index dielectrics, topological insulators and digital metamaterials, and the potentially transformative role they could play in shaping further advances in the field.

1 General Introduction to optical metasurfaces

Unexpected and unusual effects have been recently reported in various fields of research, such as but not limited to acoustics, seismology, thermal physics, and electromagnetism. At the heart of these developments is a progressive understanding of the wave-matter interaction and our ability to artificially manipulate it, particularly at small length scales. This has in turn been largely driven by the discovery and engineering of materials at extreme limit. These “metamaterials” possess exotic properties that go beyond conventional or naturally occurring materials. Encompassing many new research directions, the field of metamaterials is rapidly expanding, and therefore, writing a complete review on this subject is a formidable task; here instead, we present a comprehensive review in which we discuss the progress and the emerging materials for metasurfaces, i.e. artificially designed ultrathin two dimensional optical metamaterials with customizable functionalities to produce designer outputs.

Metasurfaces are often considered as the two dimensional versions of 3D metamaterials. The former, also known as frequency selective surfaces, or reflect- and transmit-arrays in the microwave community [16], which has been recently shown to produce a variety of spectacular optical effects ranging from negative refraction [7], super/hyper-lensing [810], optical mirages to cloaking [1119], are comprised of artificial materials engineered at the subwavelength scale to guide light along any arbitrary direction. The functionalities of individual optical components, which were previously essentially determined by the optical properties of materials, can now be specified at will. Effective material optical parameters can be rigorously derived via homogenization techniques by averaging the electromagnetic fields over a subwavelength volume encompassing each individual element of the periodic ensemble of resonant inclusions, with arbitrary geometry [12, 1519]. Thus although metamaterials can be tailored for exotic optical functions, the basic mechanism responsible for shaping the optical wavefronts remains the same as that in conventional refractive optics: it is based on the propagation of light in the composite materials. Nevertheless the idea of manipulating light deep into the subwavelength regime can be further exploited to the extreme case where metamaterials become infinitely – or at least sufficiently-thin with respect to the wavelength of light to create metasurfaces (Fig. 1a).

Figure 1 Figure 1 (a) schematic of a metasurface. It consists of an ensemble of subwavelength optical resonators which have been disposed along 2-dimensional planes to scatter light in a controlled manner. Metasurfaces can be designed to control almost any light characteristics including phase, amplitude, polarization and dispersion. (b) Generalized reflection and refraction of light. (c, d) Substituting the individual scatterers by their effective electric and magnetic surface susceptibilities, the generalized sheet boundary conditions are powerful theoretical tools to predict and model the rapid changes in the electromagnetic fields across a metasurface.
Figure 1

Figure 1 (a) schematic of a metasurface. It consists of an ensemble of subwavelength optical resonators which have been disposed along 2-dimensional planes to scatter light in a controlled manner. Metasurfaces can be designed to control almost any light characteristics including phase, amplitude, polarization and dispersion. (b) Generalized reflection and refraction of light. (c, d) Substituting the individual scatterers by their effective electric and magnetic surface susceptibilities, the generalized sheet boundary conditions are powerful theoretical tools to predict and model the rapid changes in the electromagnetic fields across a metasurface.

Breaking away from our reliance on gradual phase accumulation, one can instead consider functional interfaces covered with a collection of subwavelength nanostructures/slits or structured optical thin-films (i.e. almost 2D) to introduce abrupt modifications on the wavefront of light over the scale of the wavelength. For example by imposing gradients of optical phase discontinuities at interfaces [2029], it became possible to bend light propagation in any desired directions in perfect agreement with the generalized laws of reflection and refraction (Fig. 1b)[24]:

ntsinθtnisinθi=λ02πϕx(1)

where θj and θt denote respectively the incident and the transmitted angles, ni, nt are the refractive indices of the incident and the transmitted medium and λo is the vacuum wavelength. Equation 1 assumes that the phase gradient ϕx is a continuous function of position; however in reality, the interface is comprised of discrete resonators. Similarly, although the basic mechanism for controlling light with an array of subwavelength scatterers can be intuited by considering a subwavelength distribution of discrete phase delayed light sources, its theoretical description is rather more complex. The effective medium theory, which is well suited to derive the bulk-parameters of 3D metamaterials, may not always apply for media with reduced dimensionality such as interfaces. This point has been discussed in detail by C.L. Holloway et al. [36]. In these papers, it is shown that the scattering of light by nanostructures of sub-wavelength thickness is best characterized by a generalized version of the boundary conditions, known as generalized sheet transition conditions (GSTCs) [3]:

z^×ΔH=jωP||z^×||Mz
ΔE×z^=jωμM||||(Pzϵ)×z^z^ΔD=P||z^ΔB=μM||(2)

where Δ represents the difference between the fields in the two side of the interface, and P and M are the surface electric and magnetic polarization densities derived from the local electric and magnetic induced dipole moments in the plane z=0. Considerable efforts have been made to express the GSTCs using macroscopic quantities and we can now rewrite the GSTCs using the equivalent six-by-six surface susceptibility tensor[1] to relate the fields and the induced currents on the metasurfaces by using the relation [30]:

P=ϵχ=eeEav+χ=emμϵHav
M=χ=mmHav+χ=me=ϵμEav(3)

where the average fields are defined by: Eu,av=Eut+(Eui+Eur)2,u=xyz

Unlike 3D metamaterials where both permittivity and/or permeability are engineered for the bulk, metasurfaces act directly upon the boundary conditions of Maxwell’s equations, and for this reason the surface susceptibility (or equivalently the admittance) becomes the most appropriate quantity characterizing the optical response [46]. The large number of degrees of freedom in choosing the components of the tensor in Eq. (3) is such that solving the system in Eq. (2) poses a complex theoretical challenge, particularly in terms of design. The characterization of a metasurface at the microscopic level given arbitrary electromagnetic fields at either side represents a tremendous task, and we believe that this may provide potentials for even greater theoretical progress in this direction.

From the above discussion, it is clear that the manipulation of light with metasurfaces goes beyond the phase; it is possible to achieve simultaneous control of the amplitude, phase (full 2π coverage), as well as polarization and the dispersion of light at interfaces. These techniques have already been exploited to create numerous new and intrinsically flat devices such as phase plates [31], collimators [32], waveplates [33] and flat lenses [34, 35] for diverse applications ranging from holography [3638], subdiffraction imaging [3941] and nano-circuitry [4244] to cloaking [13, 4548].

In addition, from the manufacturing point of view metasurfaces possess a number of advantages compared to traditional metamaterials. Since light is not propagating in the bulk, metasurfaces are less absorbing and are particularly easier to fabricate than their 3D counterparts. While metasurfaces are created using planar micro/nanofabrication technologies, 3D metamaterials are fabricated using mature but still expensive processes that involve carving, polishing and surface coating of different sort of bulk materials. Traditional manufacturing optics technology requires a large amount of raw materials and qualified engineers that have to deal with difficult and precise tasks. Fabrication of refractive devices consumes a lot of materials, coating the surfaces is performed by depositing multilayers of various dielectrics, and, the characterization of the end products is generally performed manually one-by-one, device after device. Instead, 2D planar metasurfaces can be created using advanced manufacturing processes that are now already well-established in the semiconductor integrated circuit industry. Thus one could in principle arbitrarily engineer the phase, polarization and amplitude profile with a single device fabricated using in a single lithographic step, subject to the complexities introduced by near-field coupling between individual elements. In the following, we highlight various materials and design techniques developed in the recent years and comment on the progress in utilizing them for novel metasurface applications, with particular emphasis on non-traditional and emerging materials.

2 Loss

One key concern in almost any device application is loss. Despite significantly mitigating propagation losses due to their reduced spatial footprint compared to bulk metamaterials, metasurfaces can have considerable intrinsic absorption. To fully understand the origin of the problem, one should bear in mind that the individual building blocks of metasurfaces may in principle be optically resonant elements. Traditional optical resonators have consisted of Fabry-Perot cavities, whispering gallery structures, or photonic crystals and various forms of dielectric waveguides [49, 50]. While these generally low loss structures have proved to be incredibly useful for applications ranging from lasers to sensitive biochemical sensors and even frequency generation devices [5156], they present an overly large physical footprint. Rigorously the electromagnetic mode volume either exceeds or is on the order of (λ/n)3, where n is the refractive index of the material. It was only with the development of plasmonics [57, 58] that deeply subwavelength confinement, or manipulation of light beyond the Rayleigh diffraction limit, became a real and enticing possibility. Plasmonics involves the study of electromagnetic surface waves (plasmons) which exist at the interface between media whose dielectric permittivities have opposite signs; in turn, electromagnetic energy is confined to mode volumes significantly smaller than λ3 at resonance [56]. This combination of possessing a strong resonance at a deeply subwavelength size scale lends itself readily to metasurface applications, and for enhanced light-matter interactions in general [59, 60]. As a result, the discussion of traditional metasurface materials has been until recently the discussion of plasmonic materials.

Figure 2 illustrates various plasmonic materials, both traditional and emerging, consolidated on a common plot of quality factor (Q) of their localized surface plasmon resonance (within scaling constants) against wavelength. The Q factor is rigorously defined as the ratio of stored electromagnetic energy to the energy dissipated (or absorbed) per cycle in a resonator; it is an indication of photon lifetime in a cavity and thus loosely corresponds to how “good” a resonance is. This is a meaningful metric to assess the relative strengths of plasmonic materials beyond simply ensuring that they satisfy the plasmon dispersion condition (negative real permittivity or that they have minimal imaginary permittivity), although numerous other figures-of-merit exist and might be more useful depending on the area of application, such as normalized surface plasmon propagation lengths [61]. Here we consider only the contributions arising from material properties, i.e. neglecting the influence of shape, size and structure in influencing resonance behaviour. Although important for the practical implementations of efficient metasurfaces, we do not include scattering losses of the resonators in order to make a fair comparison based on intrinsic material properties alone. This intrinsic Q factor can be shown [62, 63] to scale as the ratio of the real and imaginary permittivities −ε′/ε″ (its propagating surface plasmon counterpart scales as ε2/ε″) in the quasistatic limit, and with low losses (e.g. finite ε″). In terms of the standard Drude model the former expression becomes QmaxLSP(ωp2γ2)3/2/ωp2γ (correspondingly QmaxSPPωp2/γ2)[62]: intuitively we seek materials with a large ratio of plasma frequency to damping, as well as intrinsically small damping. It would be inadequate to base the discussion on either one of these quantities alone.

Figure 2 Figure 2 Localized surface plasmon resonance quality factors determined from material properties alone, plotted as a function of wavelength. In each case experimental data of permittivities were directly obtained or derived from fitting based on experimental results (data from Ref [64], Ag data from Ref [65]) with the exception of graphene (abbreviated Gr), whose dielectric function was computed using the Kubo formula in the local random phase approximation (RPA) limit, and assuming a relatively modest carrier lifetime of 0.4ps [67]. Backscattering due to extrinsic impurities or substrate induced effects are not included. Voltage values for ITO corresponds to the gating geometry shown in Ref [66], and translates to carrier densities of 1 × 1011/cm3,1.65 × 1022/cm3, and 2.77 ×1022/cm3 respectively. The figure is meant as a general guide for fair comparison of quality of resonances in plasmonic materials, and their effective spectral ranges.
Figure 2

Figure 2 Localized surface plasmon resonance quality factors determined from material properties alone, plotted as a function of wavelength. In each case experimental data of permittivities were directly obtained or derived from fitting based on experimental results (data from Ref [64], Ag data from Ref [65]) with the exception of graphene (abbreviated Gr), whose dielectric function was computed using the Kubo formula in the local random phase approximation (RPA) limit, and assuming a relatively modest carrier lifetime of 0.4ps [67]. Backscattering due to extrinsic impurities or substrate induced effects are not included. Voltage values for ITO corresponds to the gating geometry shown in Ref [66], and translates to carrier densities of 1 × 1011/cm3,1.65 × 1022/cm3, and 2.77 ×1022/cm3 respectively. The figure is meant as a general guide for fair comparison of quality of resonances in plasmonic materials, and their effective spectral ranges.

From Fig. 2, it is clear that while there are an abundance of plasmonic materials in general, ranging from noble metals like gold and silver which perform well at optical to near-infrared wavelengths [64, 65], to transparent conducting oxides (ITO) [66], as well as silicides [67], graphene [68, 69] and refractory plasmonic materials like titanium nitride [70, 71] whose resonances span the near to mid-infrared region, their material quality factors are generally in the same order of magnitude (between 10 to 100). This value is low, particularly in comparison to, for example, values approaching 108 that can now be routinely achieved in whispering gallery mode dielectric resonators [72, 73]. This implies that traditional plasmonic materials discovered so far are unavoidably lossy at optical and infrared wavelengths, due to the finite skin depth of the mode in a lossy medium. In turn, metasurface efficiencies have been significantly affected by these losses alone [74]. This strongly motivates the search for better materials (not necessarily plasmonic) that still satisfy the criteria of supporting optical resonances with subwavelength confinement [7577], as well as alternative design schemes and strategies that can mitigate all sources of loss (e.g. both absorption and reflection) and achieve high efficiencies via interplay between multiple resonant channels.

One such method to fully and efficiently engineer the phase of a light beam consists of utilizing an anisotropic nanostructure designed to have two orthogonal polarization states with a π phase difference [78]. This element can be intuitively visualized as a tiny half wave plate whose principal axis is determined by its orientation with respect to the incident polarization. By employing a series of the same anisotropic scatterers that are geometrically rotated with respect one to another, resulting in the acquisition of a geometric Pancharatnam-Berry phase [7982], one can achieve holographic projection of an image with a record reflection efficiencies close to unity [78, 83]. A recent example is illustrated in Fig. 3a. Instead of using two distinct modes of resonance whose transmission is largely fixed by the desired phase profile, these metasurfaces have an additional degree of freedom: the final imparted phase originates only from the geometric phase acquired as a circularly polarized light beam is taken along a closed cycle in the polarization space due to interacting with geometrically rotated scatterers. This leaves room to separately design the transmission amplitude of the metasurface by varying geometrical parameters (besides orientation, which is fixed). This is in contrast to regular dynamic phase acquired via propagation or from resonator response. In this way, the basic structure of each individual scatterer can be optimized to provide results with better efficiency, for example, by suppressing transmission channel and improving reflection via constructive interference with cavity resonances (Fig. 3a), although it does not overcome intrinsic losses and operates with the caveat of being in cross polarization.

Figure 3 Left-to-right: (a) From [78]: Unit cell of a nanorod supporting two resonances under orthogonal polarizations with π phase separation. By introducing a space variant change in orientation, a geometric phase ranging from 0 to In can be imparted to incident light. This frees up the design of the antenna resonances to maximize reflection. The current scheme utilizes a mirror to achieve ~90% reflection efficiency in the visible. (b) From [84]: A similar application of the Berry phase principle for the manipulation and focusing of surface plasmon beams via rectangular nano-slits. However due to lack of helicity of the plasmons, the phase change that can be achieved with orientation changes is only π. Reprinted with permission from Refs [78] and [84]. Copyright 2015 Nature Publishing Group, 2015 Optical Society of America.
Figure 3

Left-to-right: (a) From [78]: Unit cell of a nanorod supporting two resonances under orthogonal polarizations with π phase separation. By introducing a space variant change in orientation, a geometric phase ranging from 0 to In can be imparted to incident light. This frees up the design of the antenna resonances to maximize reflection. The current scheme utilizes a mirror to achieve ~90% reflection efficiency in the visible. (b) From [84]: A similar application of the Berry phase principle for the manipulation and focusing of surface plasmon beams via rectangular nano-slits. However due to lack of helicity of the plasmons, the phase change that can be achieved with orientation changes is only π. Reprinted with permission from Refs [78] and [84]. Copyright 2015 Nature Publishing Group, 2015 Optical Society of America.

We also note here an added subtlety that while this technique applies to incident light beams, the same is generally not true for bound electromagnetic waves such as surface plasmons, particularly in planar geometries such as interfaces. The reason is the lack of helicity (σ) in the latter which limits overall phase acquired to π, instead of the whole range of 2π. This situation is illustrated in the panels on the right in Fig. 3a and 3b[84]. For the same rotation angle of 90° in the rectangular resonator element, it is seen that the accompanying phase changes for the cross polarized and surface plasmon beams are π and π/ 2 respectively (Fig. 3b), in contrast to circularly polarized free space light beams that possess helicities ±1 (Fig. 3a). It is only in certain geometries (e.g. cylindrical) where multiple surface plasmon modes can be superposed with a defined phase relationship, that surface plasmons can possess helicity and chirality [85].

An alternative approach to achieving high performance efficiencies is via engineered interfaces (without the use of geometric phase), particularly in a cavity configuration utilizing regular reflectors or multiple metasurface layers. This has been employed in the design of highly efficient reflectors and absorbers by overlaying patterned metallic structures over opaque metal ground plates separated by a dielectric spacer [8688]. By considering the interplay between the radiated fields of the induced image dipole in the ground plane upon resonant excitation of the top metasurface [86], or equivalently the formation of “gap surface plasmon” (GSP) modes in the entire composite structure [88, 89], any incident power in a spectrally broad range can be almost entirely absorbed or reflected. Importantly such designs utilize an optically thin spacer, such that the overall device is compact and subwavelength in all dimensions. For transmission mode devices, a similar principle applies, although neither reflection nor absorption can be trivially suppressed. Intuitively a single interface can only support tangential electric currents that radiate equally on both sides, thereby giving rise to unavoidable reflection loss. By utilizing the notion of optical metatronics [42, 90] and employing a series of cascaded metasurfaces with prescribed transverse inhomogeneity, one can realize sufficient degrees of freedom to achieve full control of the transmission phase, while at the same time to minimize reflection via impedance matching [91]. A proposed design from [91] is illustrated in Figure 4a. It utilizes a basic building block of fixed geometry with varying fractions of a plasmonic material and dielectric. By varying the fill factor, both inductive reactances are obtained which span a broad range of values [91].

Figure 4 Left-to-right: (a) From [91]: Schematic showing a basic building block of a multilayer, impedance matched metasurface consisting of two materials with real permittivities of opposite sign (i.e. plasmonic and dielectric). As an example aluminium doped zinc oxide (AZO) and silicon were chosen respectively at a working wavelength of 3 µm. Here l=250 nm and h=250 nm. By varying the width of the block a wide range of reactance (per unit length) can be obtained. Color maps demonstrate the transmission amplitude and phase as a function of width of the block. High transmission efficiencies (>−3 dB) and full control of the phase can be seen (enclosed by white contour lines). (b) From [93]: Schematic and SEM images showing a cascaded series of Ti/Au metasurfaces that together function as an asymmetric circular polarizer, due to magnetic response generated via interactions between the sheets, in addition to each individual tangential electric field response. Note that overall device thickness is still subwavelength and can be modelled as a single bianisotropic metasurface. Adapted with permission from Refs [91, 93]. Copyright 2013, 2014, American Physical Society.
Figure 4

Left-to-right: (a) From [91]: Schematic showing a basic building block of a multilayer, impedance matched metasurface consisting of two materials with real permittivities of opposite sign (i.e. plasmonic and dielectric). As an example aluminium doped zinc oxide (AZO) and silicon were chosen respectively at a working wavelength of 3 µm. Here l=250 nm and h=250 nm. By varying the width of the block a wide range of reactance (per unit length) can be obtained. Color maps demonstrate the transmission amplitude and phase as a function of width of the block. High transmission efficiencies (>−3 dB) and full control of the phase can be seen (enclosed by white contour lines). (b) From [93]: Schematic and SEM images showing a cascaded series of Ti/Au metasurfaces that together function as an asymmetric circular polarizer, due to magnetic response generated via interactions between the sheets, in addition to each individual tangential electric field response. Note that overall device thickness is still subwavelength and can be modelled as a single bianisotropic metasurface. Adapted with permission from Refs [91, 93]. Copyright 2013, 2014, American Physical Society.

A variant of this technique utilizes magnetic responses in addition to electrical ones in traditional metasurfaces, also via a cascaded metasurface geometry [9294]. A representative example of such schemes is shown in Fig. 4b. The overlay of multiple interfaces generates a magnetic response due to longitudinal circulating electrical currents, so that ideally (absent material losses) each unit cell radiates unidirectionally due to interference. By optimizing the net anisotropic sheet admittance of the layers, it is shown that a minimum of three surfaces is sufficient to achieve maximum transmission. It is worth noting that despite requiring proper design of interlayer spacing in order to tailor the magnetic response, these cascaded sheets are still subwavelength in thickness and can be modelled as a single, bianisotropic metasurface capable of exhibiting electric, magnetic and chiral responses [93].

The aforementioned techniques demonstrate how overall losses in a metasurface can be minimized by exploiting device architecture; however net efficiencies are still constrained by intrinsic material parameters (which in the language of the previous examples, manifest as shunt resistances and complex sheet admittances with non-zero real components). From a materials perspective, the recent investigation of strong resonances in subwavelength high index dielectric particles thus opens up tremendous opportunities for the further development of nanoscale optics in general. In particular, the observation that the strong electrical displacement currents in such particles support an orthogonal magnetic dipole, in addition to the electric dipole caused by the driving field of incident light, makes these structures especially relevant for metasurfaces [9598]: the aforementioned criteria of strong resonances at subwavelength scales while maintaining full phase control has been satisfied.

3 Metasurfaces based on resonant dielectric elements

Due to their low intrinsic absorption at visible wavelengths, transparent dielectrics have always been considered as major materials of choice for constructing optical components. Apart from conventional bulk optics various quasi-flat devices, such as binary optics, have been extensively developed [99]. Flat elements have obvious advantage of compactness, which now becomes more and more important in the context of wearable and mobile photonic devices. Some of the earlier approaches, such as binary blazed gratings [100], directly relied on subwavelength patterning of the dielectric surface to achieve variations of effective refractive index profile in 2D, which allows for transmitted or reflected phase engineering. This approach typically requires densely-packed micrometertall nanostructures, which are challenging to fabricate and implement into a robust device. High refractive index materials such as silicon allows reduction of the thickness while still keeping it on the order of half micron or above and maintaining a high aspect ratio (typically >1) [101, 102]. Another approach for phase control with structured dielectric surfaces relies on rotating grating elements and Pancharatnam-Berry (PB) phase, as previously introduced. Initially proposed for metallic gratings [20] it was quickly adapted to dielectric materials [80], resulting in e.g. demonstration of metasurface-based holography [103]. In recent studies a similar approach was applied to realize silicon gratings, lenses and axicons with only 100 nm thickness on quartz substrate [28]. Figure 5a shows an example of an axicon constructed from sub-diffractive arrays of silicon nanorods [28].

Figure 5 (a) From [28]: SEM image of an axicon lens constructed from silicon nanorods based on nanorod resonances and Pancharatnam-Berry phase concept. (b) From [118]: Huygens gradient metasurface for beam deflection constructed from silicon nanodisks with interfering electric and magnetic dipoles. Right: SEM image; left: back focal plane image of transmitted beam, demonstrating deflection of beam at 715 nm. (c) From [121]: Dielectric metasurface comprised of elliptical cylindrical posts, demonstrating complex polarization and phase control. Radially and azimuthally polarized light beams are simultaneously generated and focused, from incident x (blue arrow) and y (green arrow) polarized light respectively. Right: SEM and optical image of the device; left: simulations and experimental measurements of the generated beams. (d) From [126]: Generalized electro-magnetic Brewster effect demonstrated at a metasurface constructed of resonant silicon nanodisks. Left: (top) schematic of suppression of angular reflection from a sub-diffractive array of nanoparticles with both electric and magnetic dipoles excited, and (bottom) SEM image of the fabricated silicon nanoparticles. Right: experimental (dots) and theoretical (solid lines) demonstration of wavelength-dependent generalized Brewster effect. Scale bar: 500 nm. Reprinted with permissions from Refs [28], [118], [121] and [126] respectively. Copyright 2014 American Association for the Advancement of Science, 2015 Wiley.
Figure 5

(a) From [28]: SEM image of an axicon lens constructed from silicon nanorods based on nanorod resonances and Pancharatnam-Berry phase concept. (b) From [118]: Huygens gradient metasurface for beam deflection constructed from silicon nanodisks with interfering electric and magnetic dipoles. Right: SEM image; left: back focal plane image of transmitted beam, demonstrating deflection of beam at 715 nm. (c) From [121]: Dielectric metasurface comprised of elliptical cylindrical posts, demonstrating complex polarization and phase control. Radially and azimuthally polarized light beams are simultaneously generated and focused, from incident x (blue arrow) and y (green arrow) polarized light respectively. Right: SEM and optical image of the device; left: simulations and experimental measurements of the generated beams. (d) From [126]: Generalized electro-magnetic Brewster effect demonstrated at a metasurface constructed of resonant silicon nanodisks. Left: (top) schematic of suppression of angular reflection from a sub-diffractive array of nanoparticles with both electric and magnetic dipoles excited, and (bottom) SEM image of the fabricated silicon nanoparticles. Right: experimental (dots) and theoretical (solid lines) demonstration of wavelength-dependent generalized Brewster effect. Scale bar: 500 nm. Reprinted with permissions from Refs [28], [118], [121] and [126] respectively. Copyright 2014 American Association for the Advancement of Science, 2015 Wiley.

The studies mentioned above considered dielectric structures, which rely on additional mechanisms to achieve the full phase control (e.g. PB phase). Recently the field of resonant dielectric nanophotonics has witnessed a rapid development, providing completely new solutions for both isolated dielectric nanoantennas as well as metasurfaces. Although resonances in small dielectric spherical particles with high refractive index have attracted the attention of numerous theorists over the last century – a case completely solved by the Mie theory [104] – it is only very recently these resonances were experimentally realized in the visible and near-IR spectrum. We also invite the reader to consider several major examples treated in references [105107]. Coming back to recent work, the first example of controlled first-order resonance at visible frequencies was performed with germanium [108] and silicon [109] nanorods and later with silicon [97, 98] and gallium arsenide [110] nanoparticles. It was shown [97, 98, 111] that such nanoparticles may have strong electric and magnetic dipole resonances at visible and IR frequencies. Both types of dipoles can be excited simultaneously within simple nanoparticle geometries (such as spheres or disks) which provide numerous possibilities for directivity control of scattering. In particular, as was first predicted by Kerker [106] and recently demonstrated at microwaves [112] and in the visible spectrum [110, 113], the two dipoles may interfere and strongly suppress the particle radiation pattern in backward direction, making it analogous to Huygens’ source with only forward scattering. Tuning the particle shape (e.g. by changing aspect ratio of disks [114, 115] or spheroids [116]) enables one to control spectral position of the electric and magnetic dipole resonances maxima and have them coincide in amplitude and in phase at a single wavelength, which makes the effect of directional forward scattering broadband.

Combining these directional dielectric scatterers with only forward scattering into an array allows constructing a Huygens’ metasurface with enhanced transmission at resonant wavelengths [9294, 115]. This allows for the possibility of full 2π control of transmission phase (π phase shift for each of the electric and magnetic dipole resonances) [117]. The operation principle of such a metasurface is similar to multi-layer plasmonic Huygens’ metasurfaces described above [93]. The obvious advantages compared to metals are lower intrinsic losses, and compatibility with CMOS processes, which makes them more likely to have an impact in real-world applications. Very recently high efficiency resonant transmission (>85%) and transmitted beam deflection (efficiency >45%) with these dielectric metasurfaces with only 130 nm thickness has been experimentally demonstrated [118](Fig. 5b). A similar principle has also been applied to vortex beam generation [119, 120]. Additionally, efficiencies exceeding 80% for beam focusing and forming has been achieved with high-aspect ratio pillars (see example of vortex beam generation in Fig. 5c) [121, 122]. In this case however, due to the high-aspect ratio of the nanostructures fabrication and handling of such a device can become more challenging.

Apart from transmitted phase engineering, resonant dielectric metasurfaces may provide multiple other functionalities related to the interplay between electric and magnetic resonances of their constituent elements. Some examples are extra-high reflectivity at a single metasurface layer [123], which can also be combined with wavefront engineering [83], or ultra-narrow electromagnetically induced transparency window obtained with dielectric nanoantenna array [124]. Recently analogy to magnetic mirror behavior [125] and demonstration of generalized electro-magnetic Brewster effect [126] with dielectric metasurfaces have been reported. The latter case is depicted in Fig. 5d and shows that a single-layer dielectric metasurface can produces an equivalent effect of a conventional bulk material with both electric and magnetic properties at visible wavelengths (ε ≠ 1, μ ≠ 1).

These effects which have been demonstrated in the visible and near infrared spectral region are not restricted to a particular material or a specific spectral range. Similar exciting optical properties can be observed at any frequency by scaling the size of the nanostructures (e.g. for mid-IR frequencies [111, 127]), providing that the dielectric material possesses a relatively high refractive index (>2) and low losses in the required frequency range. In general, strong electric and magnetic resonant behavior, low losses, advanced scattering directivity control, high melting temperature and CMOS compatibility make resonant dielectric nanoantennas and metasurfaces very promising for real-world applications.

4 Active/2D materials

A cornerstone of device operation is active behaviour where output changes in response to a control signal. For example, spatial light modulators comprising of liquid crystal displays are capable of modulating amplitude, phase and polarization of light in both space and time, and are integral components of many systems such as imaging and projection devices in consumer electronics as well as advanced microscopy (optical tweezing, stimulated emission depletion (STED) microscopy) setups. It is thus apparent that in order for the concept of photonics, and by extension metasurfaces, to truly realize the vision of superseding traditional optical components and even optoelectronics, a robust system that enables active behaviour and preserves key advantages of these structures must be established. To this end, significant work has been done in the research and synthesis of new functional materials, and their integration with traditional metamaterial building blocks [128]. Examples include the use of microelectromechanical (MEMs) systems [129131], charge carrier injection [132136], phase change materials [137145] and liquid crystals [146148] for dynamic reconfigurability.

Here we highlight in particular the emerging and rapidly developing field of two dimensional (2D) materials, as exemplified by the earliest isolated carbon monolayer graphene. While it is now well-known that graphene for example has significant potential in various electronic applications due to its 2D nature that results in effectively massless electrons capable of ballistic transport at room temperature [149151], significantly less attention has been paid to its use in various photonic applications. The latter is capable of leveraging upon most of graphene’s advantages for electronic devices, but without the corresponding drawbacks (most notably, the lack of a bandgap which severely hinders electrical switching devices). Intuitively and as a starting point, we consider its use as a functional, passivation layer whose dielectric permittivity can be dynamically altered by a wide range. The low intrinsic carrier concentration of graphene and its unique band structure implies that the former can be tuned by a much larger percentage than traditional bulk materials via external doping, and is typically only limited by gate breakdown voltages. This is the basic principle behind graphene’s use as an electrical transducer, due to its gate-tunable Fermi energy level and hence conductivity [152154]. Optically, this directly translates to a highly tunable dielectric permittivity [155162]. Examples of planar photonic devices that make use of this principle abound (Fig. 6), ranging from standard plasmonic nanoantenna [163, 164], to exotic Fano resonant asymmetric structures [165] and even various biosensors [166]. The key principle underlying the performance is the mode overlap and the interaction volume of the graphene monolayer with the resonantly enhanced electromagnetic fields: for waveguide-like structures where the fields are concentrated in very small gaps for extended distances (30 nm and approximately 500 nm respectively), a tuning range of 20% of the central resonant frequency can be achieved [164](Fig. 6).

Figure 6 Left-to-right: (a) From [157]: Nanoribbons in graphene, demonstrating the observation of graphene localized surface plasmon resonances in the THz regime. Resonances peaks (normalized to transmission through an ungated graphene layer at charge-neutrality point) are seen to tune with gate voltage. (b) From [164]: SEM image of gold waveguide antennae, with a slot size of 30 nm. Electromagnetic fields on resonance are strongly confined within the gaps, and the geometry allows for relatively significant tuning of antenna resonances (approximately 20% of center resonance frequency) using a back-gated graphene layer introduced over the antennae. (c) From [170]: Extinction spectra of nanoslit arrays in a graphene monolayer at small size scales, enabling scaling of the plasmonic response into the mid-infrared (inset: SEM image). Relatively high quality factor resonances are observed, which tune by several times their FWHM under gating. Additional resonance peaks due to plasmon-phonon coupling with surface optical phonons in the substrate are also observed, as evidence of strong light-matter interactions in the graphene system due to the very small mode volumes of its plasmonic response. Reprinted with permissions from Refs [157], [164] and [170] respectively. Copyright 2011 Nature Publishing Group, 2013 American Chemical Society.
Figure 6

Left-to-right: (a) From [157]: Nanoribbons in graphene, demonstrating the observation of graphene localized surface plasmon resonances in the THz regime. Resonances peaks (normalized to transmission through an ungated graphene layer at charge-neutrality point) are seen to tune with gate voltage. (b) From [164]: SEM image of gold waveguide antennae, with a slot size of 30 nm. Electromagnetic fields on resonance are strongly confined within the gaps, and the geometry allows for relatively significant tuning of antenna resonances (approximately 20% of center resonance frequency) using a back-gated graphene layer introduced over the antennae. (c) From [170]: Extinction spectra of nanoslit arrays in a graphene monolayer at small size scales, enabling scaling of the plasmonic response into the mid-infrared (inset: SEM image). Relatively high quality factor resonances are observed, which tune by several times their FWHM under gating. Additional resonance peaks due to plasmon-phonon coupling with surface optical phonons in the substrate are also observed, as evidence of strong light-matter interactions in the graphene system due to the very small mode volumes of its plasmonic response. Reprinted with permissions from Refs [157], [164] and [170] respectively. Copyright 2011 Nature Publishing Group, 2013 American Chemical Society.

Thus far the physical mechanisms involved in the active control of metasurfaces are similar to the charge injection or phase change materials. Graphene truly comes into its own when we consider it to be the active photonic element, since it is capable of supporting plasmonic resonances at mid-infrared frequencies and beyond with relatively high quality factors comparable to noble metals [157, 167, 168] (with variations dependent on doping, see curves denoted by Gr on Fig. 2). Graphene plasmons possess an extremely small mode volume, and consequently exhibit very high field confinement and light matter interaction strengths [168173]. Intuitively the mode volume of localized plasmonic resonances scales proportionally to the particle size; as a result, graphene represents the ultimate limit in scaling, and patterned graphene structures are in many ways the ideal planar metasurface; e.g. in terms of transverse spatial localization, overall device footprint and field enhancement strengths. These structures, ranging from nanorods, to disks and inverse hole arrays [157, 170, 174] (Fig. 6) have been shown to exhibit resonances that are in some cases tunable by several times of their FWHM, enabling broadband functional behavior, which is in turn made possible because of the linear band-structure around their Dirac point.

Moreover, it ought to be emphasized these aforementioned properties are general features of two-dimensional electron gas (2DEG) systems, of which graphene is the most well-known example due to the relative ease of isolation and room temperature operation. Detailed studies into alternative plasmonic 2DEG systems, such as GaAs/AlGaAs at cryogenic temperatures indicate that their properties are robust to standard industrial fabrication processes [175]. As a result the entire spectrum of optoelectronic components, ranging from gated plasmonic circuits operating at GHz frequencies to optical band filters, interferometers, and negative-index screens in the optical regime can be realized [175, 176]. Similar to their graphene counterparts, the intrinsic low dimensionality of such devices enables them to possess unprecedented miniaturization. Examples of such components are illustrated in Figure 7. As such the complete manipulation of amplitude, phase and polarization of electromagnetic waves and their associated charge carriers is shown to be possible at a truly planar level at a wide range of operating frequencies.

Figure 7 Examples of plasmonic components utilizing shaped 2DEG in GaAs/AlGaAs heterostructures at cryogenic temepratures. (a)
From [175]: Optical image (top) and cross-section schematic (bottom) of a GHz-regime plasmonic bandgap crystal comprising of shaped 2D plasmon waveguides and a grounded metallic top gate (labelled “G”). This creates localized input and output terminals (labelled “S”). The plasmonic waveguide can be directly (nonresonantly) excited over a continuous frequency range by altering the density and hence kinetic inductance of the 2DEG via the gate contact. Effective wavelengths of plasmons were shown to be approximately λ/660, corresponding to an effective area up to 440 000 times smaller than typical electrical components. (b) From [176]: Negative index metamaterial in AlGaAs strips on a GaAs substrate operating at GHz frequencies. Incident light on a signal port excites 2D plasmons propagating lengthwise along the strips. This response capacitatively couples to neighbouring strips resulting in an effective wave propagating transversely to the structures; due to the significant kinetic inductance in this 2DEG system this transverse wave experiences a negative index as large as −700. Similar dynamic control can be achieved by varying the carrier properties of the 2DEG via external electrical bias. Reprinted with permissions from Ref [175], [176]. Copyright 2012 Nature Publishing Group, American Chemical Society.
Figure 7

Examples of plasmonic components utilizing shaped 2DEG in GaAs/AlGaAs heterostructures at cryogenic temepratures. (a) From [175]: Optical image (top) and cross-section schematic (bottom) of a GHz-regime plasmonic bandgap crystal comprising of shaped 2D plasmon waveguides and a grounded metallic top gate (labelled “G”). This creates localized input and output terminals (labelled “S”). The plasmonic waveguide can be directly (nonresonantly) excited over a continuous frequency range by altering the density and hence kinetic inductance of the 2DEG via the gate contact. Effective wavelengths of plasmons were shown to be approximately λ/660, corresponding to an effective area up to 440 000 times smaller than typical electrical components. (b) From [176]: Negative index metamaterial in AlGaAs strips on a GaAs substrate operating at GHz frequencies. Incident light on a signal port excites 2D plasmons propagating lengthwise along the strips. This response capacitatively couples to neighbouring strips resulting in an effective wave propagating transversely to the structures; due to the significant kinetic inductance in this 2DEG system this transverse wave experiences a negative index as large as −700. Similar dynamic control can be achieved by varying the carrier properties of the 2DEG via external electrical bias. Reprinted with permissions from Ref [175], [176]. Copyright 2012 Nature Publishing Group, American Chemical Society.

5 Broadband and achromatic operation

Beyond active tunability of devices, the bandwidth of operation is also a key figure of merit in many applications. Numerous examples of broadband anomalous behavior [177] based on metasurfaces have already been reported. For example, a tunable metasurface comprised of graphene nanostructures is termed broadband because the resonances can be dynamically varied over a range equivalent to several FWHM [170, 178]. Similar examples abound, e.g. 80% efficiency reflect-metasurface which has high transmission over an approximate 300–500 nm range in the visible – near infrared regime [78]. Alternatively, one can make use of deeply subwavelength nanostructures which interact with broadband white light to achieve high fidelity colors at various design wavelengths for applications ranging from imaging to printing [179181].

We highlight here an additional subtlety: broadband does not equate to achromaticity. A standard example is that of a metasurface lens: it cannot be truly broadband if its functionality (e.g. 2π phase accumulation) is inherently monochromatic. In other words, the spatial phase profile, which is key to focusing light, is still dispersive, even though the lens might possess high efficiency values across a wide range of wavelengths (i.e. suffer from chromatic aberration). An ideal “broadband” metasurface lens should retain an achromatic phase response for different wavelengths within its operating bandwidth. Clearly this is a significant challenge that fundamentally departs from the design philosophy and methodology of the vast majority of current devices.

It was recently demonstrated [182, 183] (Fig. 8) that such achromatic devices, where the spatial phase profiles are constant at various wavelengths, are indeed possible with only a single interface by designing the overall phase response to take into account both the dispersive phase accumulation in free space for several different wavelengths and the intrinsic metasurface phase response. In other words, the general design principle takes the form φp + φm = φtot = constant, where φp,m refer to phase accumulation from free space propagation and metasurface respectively. This in turn implies that φm=2πλl(r), where l (r) is the physical distance between the interface at some position r, and contains information corresponding to the device function. By introducing multiple coupled resonators [184], one can in principle obtain sufficient degrees of freedom to design the desired phase mask for an arbitrary number of discrete wavelengths. This approach is completely generalizable to almost arbitrary wavelengths, and is limited only by intrinsic material properties such as absorption losses; most importantly this design strategy retains the single-step fabrication of metasurfaces even for complex, multiple wavelength operation.

Figure 8 From [182]: An achromatic beam deflector preserves its functionality (i.e. focusing) for different wavelengths, i.e. multiple distinct wavelengths are anomalously refracted at the same angle. (a) Schematic and SEM of the achromatic metasurface design, comprising of near-field coupled resonators of different thicknesses. Here s=1000 nm, t=400 nm, w1 =300 nm, w2 =100 nm and g=175 nm. (b) Measured far field intensities as a function of incidence angle (from normal) for light at three discrete wavelengths of 1300 nm, 1550 nm and 1800 nm (color coded blue-green-red respectively). The beams coincide at the desired order (here approximately −17 degrees). From [183]: (c) False color SEM and (d) experimentally measured intensity profiles across the focal plane of an achromatic lens for three wavelengths of 1300,1550, and 1800 nm. Scale bar: 400 nm. Reprinted with permission from Ref. [182], [183]. Copyright 2015 American Association for the Advancement of Science, 2015 American Chemical Society.
Figure 8

From [182]: An achromatic beam deflector preserves its functionality (i.e. focusing) for different wavelengths, i.e. multiple distinct wavelengths are anomalously refracted at the same angle. (a) Schematic and SEM of the achromatic metasurface design, comprising of near-field coupled resonators of different thicknesses. Here s=1000 nm, t=400 nm, w1 =300 nm, w2 =100 nm and g=175 nm. (b) Measured far field intensities as a function of incidence angle (from normal) for light at three discrete wavelengths of 1300 nm, 1550 nm and 1800 nm (color coded blue-green-red respectively). The beams coincide at the desired order (here approximately −17 degrees). From [183]: (c) False color SEM and (d) experimentally measured intensity profiles across the focal plane of an achromatic lens for three wavelengths of 1300,1550, and 1800 nm. Scale bar: 400 nm. Reprinted with permission from Ref. [182], [183]. Copyright 2015 American Association for the Advancement of Science, 2015 American Chemical Society.

6 Emerging Materials

The previous discussions share the underlying theme of how metasurface functionality can be improved, by a combination of appropriate device level design together with judicious choice of materials. At this juncture it is appropriate to highlight some emerging materials which push the limits of the traditional optics paradigm and promise to extend functionalities of metasurfaces into areas previously thought impossible.

One intriguing example is that of electronic topological insulators, a new class of materials which exhibit strong spin-orbit coupling in charge carriers, resulting in an insulating bulk state and conducting surface states that are protected under time-reversal symmetry [185, 186]. In particular, the charge carriers in these surface states exhibit spin-momentum locking (the spin being always directed in the surface plane, but orthogonal to the momentum) which makes it possible to support a new class of collective charge carrier oscillations with welldefined spin states (i.e. spin polarization), known as “spinplasmons” [173, 187189]. It is instructive to note that by definition of the spin-orbit interaction, these spin excitations are separate from and always transverse to the usual charge oscillations that comprise the plasmon response in the form sk × n, which refer to the directions of the spin, plasmon propagation and surface normal vectors respectively. This does not equate to the surface plasmons or charge excitations picking up non-zero angular momentum (optical spin): their excitation and polarization conditions remain the same as with traditional systems, and they therefore retain a helicity of zero. However, due to different screening effects in the host material these charge and spin excitations generally possess different amplitudes and occur at slightly different frequencies. This gives rise to the possibility of observing and manipulating their response via helicity-dependent optical sources, opening up an additional degree of freedom in the study of nanoscale light-matter interactions [190192]. This is not unlike the rapidly evolving field of spintronics [193, 194], and could serve to further increase the information/bit density for optics-based communication purposes. In addition, due to topologically protected states, plasmon excitations in such systems are immune to non-magnetic scatterers/ defects, and hence possess even greater propagation lengths than their 2DEG counterparts [188].

Of more immediate interest, the family of bismuth based topological insulator compounds have been studied and shown to possess strong plasmonic responses in the ultra-violet spectrum [195, 196], far beyond the generally expected terahertz regime for such materials [197]. This could potentially bridge a critical gap in the general functionality of nanophotonic devices: most plasmonic materials, ranging from noble metals to transition metal oxides and graphene, have their responses centered in the visible to IR spectral regime (Fig. 2). Interband transitions prevent any plasmonic response in the noble metals at short wavelengths, while plasmonic frequencies in the NIR represent the limit of realistically achievable doping levels in graphene and conducting oxides. Similarly, the requirement of a high refractive index in dielectrics at ultra-violet wavelengths in order to sustain strong electric and magnetic resonances, as well as low losses, imposes strict constraints on the range of available materials. Currently the only semi-viable choice is aluminium, which suffers from persistent native oxidation.

Ref. [196](Fig. 9) shows evidence in favour of plasmonic resonances observed in nano-slit patterned Bi1.5Sb0.5Te1.8Se1.2, which is known to be a stable topological insulator at room temperature. The data is found to agree well with a theoretical model of the material as a thin-film Drude metal overlaid on a bulk insulating substrate, analogous to the distribution of charge carrier states in a topological insulator. Similarly, Ref. [195] presents evidence in favour of plasmonic enhancement of photoluminescence in ZnO structures using a topological insulator flake from the same family (Bi2Te3). These are intriguing results and warrant significantly more research into these materials. Of particular interest is how the native screening of the surface state excitations that usually result in plasmonic resonances occurring in the far-infrared to terahertz regime could lead to responses for the demonstrated structures in the ultraviolet.

Figure 9 (a) From [197]: False color SEM and (b) corresponding measured extinction spectrum of patterned Bi2Se3 nanoribbons of varying widths. In (b) measured (circles) data are fitted with an analytical model (black lines), enabling the extraction of plasmon (red lines) and phonon (green lines) contributions. The plasmonic resonances occur in the THz regime and are found to scale quadratically with ribbon width, as expected for a 2DEG on the surface of a topological insulator. (c) From [196]: SEM image and measured spectrum of a patterned Bi1.5Sb0.5Te1.8Se1.2 material at significantly smaller size scales. Evidence pointing to plasmonic response in the UV (<400 nm) is observed. (d) From [195]: SEM image of a Bi2Te3 film on a ZnO substrate, which exhibits photoluminescence near the bandgap energy in the UV spectrum. The boxed are indicates the region where the photoluminescence measurements are collected. Enhanced intensities are observed in the spatial region corresponding to the flake location. This is attributed to plasmonic excitations within the film as a result of being overlaid on the rough substrate. Reprinted with permission from Refs [197], [196] respectively. Copyright 2013, 2014 Nature Publishing Group. Material from Ref. [195] is reproduced under the Creative Commons License.
Figure 9

(a) From [197]: False color SEM and (b) corresponding measured extinction spectrum of patterned Bi2Se3 nanoribbons of varying widths. In (b) measured (circles) data are fitted with an analytical model (black lines), enabling the extraction of plasmon (red lines) and phonon (green lines) contributions. The plasmonic resonances occur in the THz regime and are found to scale quadratically with ribbon width, as expected for a 2DEG on the surface of a topological insulator. (c) From [196]: SEM image and measured spectrum of a patterned Bi1.5Sb0.5Te1.8Se1.2 material at significantly smaller size scales. Evidence pointing to plasmonic response in the UV (<400 nm) is observed. (d) From [195]: SEM image of a Bi2Te3 film on a ZnO substrate, which exhibits photoluminescence near the bandgap energy in the UV spectrum. The boxed are indicates the region where the photoluminescence measurements are collected. Enhanced intensities are observed in the spatial region corresponding to the flake location. This is attributed to plasmonic excitations within the film as a result of being overlaid on the rough substrate. Reprinted with permission from Refs [197], [196] respectively. Copyright 2013, 2014 Nature Publishing Group. Material from Ref. [195] is reproduced under the Creative Commons License.

Finally, we highlight the concept of artificial engineered materials [198], which invites a re-think of the entire paradigm of material choice to fit application requirements. Analogous to the concept of modern computing, which can formulate and solve tasks to almost arbitrary complexity despite working with only binary valued bits, it is interesting to consider if one can make redundant the entire previous portion of this article on material properties by considering some fundamental building blocks of materials that, when appropriately combined, could yield any desired, arbitrary permittivity at a specific wavelength. This is frequently introduced as the concept of “digital metamaterials”. We also briefly highlight works that extend the idea of “digital building blocks” to metasurface design, by using only a few different unit cell elements in various sequences to achieve tailored near and far-field radiation patterns.

7 ‘Digital’ Materials Building Blocks

One of the interesting features of the concept of metamaterials and metasurfaces is the possibility to engineer composite materials with desired values for material parameters, such as permittivity and permeability for bulk three-dimensional (3D) volumetric metamaterials, or surface impedance for two-dimensional (2D) cases of metasurfaces. Depending on different scenarios, sometimes one is required to have access to a relatively wide range of values for one or more of these material parameters. For example, if one wants to design a converging lens that is geometrically flat (and thin), the permittivity along the transverse dimensions of the lens needs to be inhomogeneous with higher value in the middle and lower values at points away from the center. It would be hard, if not impossible, to make structures with varying values of permittivity along such structures using different materials (or different doping concentrations in semiconductors). However, another way to achieve such a goal is to exploit the notion of “digital metamaterials” [198], which is briefly outlined here: First, for the case of binary structures one needs to select two elemental materials, which we can call metamaterial ‘bits’ (Fig. 10a). From the electromagnetic point of view, for such bits one needs to choose two materials whose permittivity functions have oppositely-signed real parts at the operating frequency of interest. Other considerations, e.g., thermal properties, mechanical properties, etc., may be taken into accounts in selecting these two building blocks. Then one needs to combine these two bits, each of which is deeply subwavelength, in order to construct another cell, bigger but still subwavelength, which we can call metamaterial ‘byte’ (Fig. 10b). With a suitable combination, the effective permittivity of the byte as viewed by an outside observer may exhibit values vastly different from the permittivity values of the bits. This is consistent with the notion of ‘internal homogenization’. Intuitively speaking, since the two bits have permittivity values whose real parts have opposite signs, their combination into the form of a ‘byte’ may lead to plasmonic resonance whose polarizability may be higher than that of each bit, and therefore, the effective permittivity of such a byte can be outside the range between the two permittivity values of the original constituents. The reader is referred to Ref. [198] for more detail. It is interesting to note that analogous to the notion of binary numbers, where the positions of “1’s” and “0’s” may affect the final value of the multibit binary numbers, here also the location of bits within a byte influences the value of the effective permittivity of the byte. This is symbolically shown in Fig. 10b. Consider the subwavelength sized spherical ‘byte’ comprised of 3 layers made of the two bits 1 and 2 (where Re1 and Re2have opposite signs). Two examples are shown in the Figure, differing in the locations (i.e., the arrangement order) of the bits in the spherical structure. One is symbolically shown as ‘1 0 1” representing three layers of bits 2, 1 and 2. The other is formed by three layers of 1, 2 and 1 effectively “0 1 0”. The effective polarizability values of such spherical blocks are different when they are illuminated with electromagnetic waves, thus highlighting the analogy with the binary systems and the importance of spatial orders of the elemental material blocks in constructing structures with a range of values of effective permittivity. These bytes themselves can be used as new building blocks to construct larger structures with the desired functionalities such as the one shown in Fig. 10c, showing how alternate layers of collections of such elements can lead to macroscopic layers with alternating effective permittivity values. This method can be exploited to construct a variety of 3D metameterials and 2D metasurfaces using the combinations of only two elemental materials judiciously chosen in the wavelength of interest. It is also worth noting that while here we discuss the case of binary digital metamaterials, in which two materials can be used as the “bits”, the concept can be generalized to ternary or N-ary composites in which 3 or N different materials can be used as the digital building blocks.

Figure 10 Sketch of notion of “digital’ metamaterials: (a) Selection of two elemental building material blocks as ‘bits’ with the following condition: Real parts of the permittivity values of the two bits, ε1 and ε2, should have opposite signs. (b) By combiningthe bits one can construct different material ‘bytes’ with differing effective permittivity values as viewed by an outside observer. (c) Material ‘bytes’ can be used to construct larger structures, e.g., layers of core-shell bytes with different orders of ‘bits’, i.e., “1 0 1” bits and “0 1 0” bits.
Figure 10

Sketch of notion of “digital’ metamaterials: (a) Selection of two elemental building material blocks as ‘bits’ with the following condition: Real parts of the permittivity values of the two bits, ε1 and ε2, should have opposite signs. (b) By combiningthe bits one can construct different material ‘bytes’ with differing effective permittivity values as viewed by an outside observer. (c) Material ‘bytes’ can be used to construct larger structures, e.g., layers of core-shell bytes with different orders of ‘bits’, i.e., “1 0 1” bits and “0 1 0” bits.

A related but conceptually distinct demonstration of tailoring electromagnetic device functionality by the use of binary valued building blocks is shown in Fig. 11[199]. By using periodic sequences comprised of alternating bytes in arbitrarily sized lattices, e.g. “010101. . . /101010. . . /. . . ”, large scale metasurfaces that alter the directivity of reflection and strongly modify scattering properties can be realized [199]. In the example illustrated in Fig. 11, the constituent bytes are comprised of metallic structures with two different dimensions overlaid on dielectric substrates, to realize a null or π phase response (the “0” and “1” bytes respectively) at the targeted wavelength. Note the subtle distinction between the “digital metamaterial” concept introduced earlier, which centers on the realization of any arbitrary (effective) permittivity through the use of two fixed permittivities (of opposite sign) and these current examples, where different combinations and sequences of two basic metasurface elements are used to generate the farfield response. In this case the phase and amplitude response of each unit cell arises from geometric scales and structural design. This latter concept can be extended to comprise of multi-resonance elements in order to achieve multi-bit coding and to broaden the bandwidth of coding metasurfaces [200]. In addition, the concept of digital metamaterials can be naturally extended to encompass active behaviour, i.e. flipping between individual bytes in response to external stimuli, which paves the way for more complicated operations that are in line with the general, overarching concept of computational metasurfaces.

Figure 11 From [199]: (a) Simulated and (b) experimental metasurface with optimized 1 bit encoding for manipulating directivity of reflection under normal incidence and lowering the electromagnetic radar cross section (RCS). (c) Simulated and measured values of RCS reduction over a wide frequency range of 7.5 to 13 GHZ. (d-f) Simulated scattering patterns at 8, 10 and 11.5 GHz respectively. The metasurface redirects incident light to a much broader range of angles, resulting in significant reductions (more than 10 dB difference on average) to RCS. Reprinted with permission from Ref. [199]. Copyright 2014 Nature Publishing Group.
Figure 11

From [199]: (a) Simulated and (b) experimental metasurface with optimized 1 bit encoding for manipulating directivity of reflection under normal incidence and lowering the electromagnetic radar cross section (RCS). (c) Simulated and measured values of RCS reduction over a wide frequency range of 7.5 to 13 GHZ. (d-f) Simulated scattering patterns at 8, 10 and 11.5 GHz respectively. The metasurface redirects incident light to a much broader range of angles, resulting in significant reductions (more than 10 dB difference on average) to RCS. Reprinted with permission from Ref. [199]. Copyright 2014 Nature Publishing Group.

8 Outlook

A key reason for nanotechnology’s prominence in recent research efforts lies in its twin promise of unprecedented device miniaturization, which in turn leads to performance enhancements by virtue of improved spatial resolution, as well as the development of completely new functionalities. The field of optics is a prime example: driven by the ability to create artificial materials with nearly arbitrary optical responses, it has undergone paradigm shifts from traditional ray optics to metadevices, and more recently to planar metasurfaces.With each step, novel functionalities for a wide range of applications have emerged, accounting for phenomena as diverse as cloaking, holographic imaging and even computing. In this article, we have highlighted multiple areas of interest for further research from a materials perspective. Nevertheless it can be argued that in this era of rapid technological advancement and rise of consumer-driven demand, merely being “faster” or “smaller” is not enough. One needs to open up to the possibility of “multi-physics” devices, which integrate multiple diverse functionalities (such as electrical, thermal, and acoustic) on a single platform synergistically. Such devices could redefine the limits of what is possible, and result in the creation of disruptive technologies. On the one hand, the efficient integration of (planar) photonics and electronics could deliver on the long-held vision of vastly improved telecommunications, and even handheld consumer products, beyond the bulk nano-scale optical components and Si interconnects available today. On the other hand, it could also offer a number of new solutions that represent the first step towards conformal and/orwearable photonic gadgets for diverse applications ranging from beam shaping, smart materials, all the way to the implementation of virtual and augmented reality applications.

Acknowledgement

A.Y.Z., A.I.K. and B.L. have been supported by DSI core funds and A*STAR Science and Engineering Research Council (SERC) Pharos grant #1527300025. A.Y.Z. also acknowledges support under the A*STAR National Science Scholarship (BS-PhD) scheme. N. E. acknowledges partial supports from the US office of Naval Research (ONR) Multidisciplinary University Research Initiative (MURI) grant number N00014-10-1-0942 and the US Air Force Office of Scientific Research (AFOSR) MURI grant number FA9550-14-1-0389. P. G. gratefully acknowledges financial supports from the European Research Council under grant no. 639109.

References

[1] Munk BA. Frequency selective surfaces: theory and design. John Wiley and Sons, 2000.10.1002/0471723770Search in Google Scholar

[2] Huang J. Reflectarray antenna. John Wiley and Sons, 2005.10.1002/0471654507.eme515Search in Google Scholar

[3] Kuester EF, Mohamed M, Piket-May M, Holloway CL. Averaged transition conditions for electromagnetic fields at a metafilm. IEEE Trans Antennas Propag 2003; 51: 2641-51.10.1109/TAP.2003.817560Search in Google Scholar

[4] Holloway CL, Dienstfrey A, Kuester EF, O’Hara JF, Azad AK, Taylor AJ. A discussion on the interpretation and characterization of metafilms/metasurfaces: The two-dimensional equivalent of metamaterials. Metamaterials 2009; 3: 100-12.10.1016/j.metmat.2009.08.001Search in Google Scholar

[5] Holloway CL, Kuester EF, Dienstfrey A. Characterizing metasurfaces/metafilms: The connection between surface susceptibilities and effective material properties. IEEE Antenn Wireless Propag Lett 2011;10:1507-11.10.1109/LAWP.2011.2182591Search in Google Scholar

[6] Holloway CL, Love DC, Kuester EF, Gordon J, Hill D. Use of generalized sheet transition conditions to model guided waves on metasurfaces/metafilms. IEEE Trans Antennas Propag 2012; 60: 5173-86.10.1109/TAP.2012.2207668Search in Google Scholar

[7] Smith DR, Padilla WJ, Vier DC, Nemat-Nasser SC, Schultz S. Composite medium with simultaneously negative permeability and permittivity. Phys Rev Lett 2000; 84: 4184.10.1103/PhysRevLett.84.4184Search in Google Scholar PubMed

[8] Fang N, Lee H, Sun C, Zhang X. Sub–diffraction-limited optical imaging with a silver superlens. Science 2005; 308: 534-7.10.1126/science.1108759Search in Google Scholar PubMed

[9] Engheta N, Salandrino A. Far-field subdiffraction optical microscopy using metamaterial crystals: Theory and simulations. Phys Rev B 2006; 74: 075103.10.1103/PhysRevB.74.075103Search in Google Scholar

[10] Jacob Z, Alekseyev LV, Narimanov E. Optical hyperlens: far-field imaging beyond the diffraction limit. Opt Express 2006; 14: 8247-56.10.1364/OE.14.008247Search in Google Scholar

[11] Leonhardt U. Optical conformal mapping. Science 2006; 312: 1777-80.10.1126/science.1126493Search in Google Scholar PubMed

[12] Pendry JB, Schurig D, Smith DR. Controlling electromagnetic fields. Science 2006; 312: 1780-2.10.1126/science.1125907Search in Google Scholar PubMed

[13] Schurig D, Mock J, Justice B, Cummer SA, Pendry JB, Starr A, Smith D. Metamaterial electromagnetic cloak at microwave frequencies. Science 2006; 314: 977-80.10.1126/science.1133628Search in Google Scholar PubMed

[14] Chen H, Chan C, Sheng P. Transformation optics and metamaterials. Nat Mater 2010; 9: 387-96.10.1038/nmat2743Search in Google Scholar PubMed

[15] Datta S, Chan CT, Ho KM, Soukoulis CM. Effective dielectric constant of periodic composite structures. Phys Rev B 1993; 48: 14936.10.1103/PhysRevB.48.14936Search in Google Scholar

[16] Smith D, Schultz S, Markoš P, Soukoulis C. Determination of effective permittivity and permeability of metamaterials from reflection and transmission coefficients. Phys Rev B 2002; 65: 195104.10.1103/PhysRevB.65.195104Search in Google Scholar

[17] Sjöberg D. Dispersive effective material parameters. Microw Opt Technol Lett 2006; 48: 2629-32.10.1002/mop.21983Search in Google Scholar

[18] Simovski C. Material parameters of metamaterials (a review). Opt Spectrosc 2009; 107: 726-53.10.1134/S0030400X09110101Search in Google Scholar

[19] Alů A. First-principles homogenization theory for periodic metamaterials. Phys Rev B 2011; 84: 075153.10.1103/PhysRevB.84.075153Search in Google Scholar

[20] Bomzon Ze, Kleiner V, Hasman E. Computer-generated space-variant polarization elements with subwavelength metal stripes. Opt Lett 2001; 26: 33-5.10.1364/OL.26.000033Search in Google Scholar

[21] Bomzon Ze, Kleiner V, Hasman E. Pancharatnam-Berry phase in space-variant polarization state manipulations with subwavelength gratings. Opt Lett 2001; 26: 1424-26.10.1364/OL.26.001424Search in Google Scholar PubMed

[22] Biener G, Niv A, Kleiner V, Hasman E. Formation of helical beams by use of Pancharatnam-Berry phase optical elements. Opt Lett 2002; 27: 1875-77.10.1364/OL.27.001875Search in Google Scholar PubMed

[23] Hasman E, Kleiner V, Biener G, Niv A. Polarization dependent focusing lens by use of quantized Pancharatnam-Berry phase diffractive optics. Appl Phys Lett 2003; 82: 328-30.10.1063/1.1539300Search in Google Scholar

[24] Yu N, Genevet P, Kats MA, Aieta F, Tetienne J-P, Capasso F, Gaburro Z. Light propagation with phase discontinuities: generalized laws of reflection and refraction. Science 2011; 334: 333-7.10.1126/science.1210713Search in Google Scholar PubMed

[25] Zhao Y, Alu A. Manipulating light polarization with ultrathin plasmonic metasurfaces. Phys Rev B 2011; 84: 205428.10.1103/PhysRevB.84.205428Search in Google Scholar

[26] Ni X, Emani NK, Kildishev AV, Boltasseva A, Shalaev VM. Broadband light bending with plasmonic nanoantennas. Science 2012; 335: 427.10.1126/science.1214686Search in Google Scholar PubMed

[27] Kildishev AV, Boltasseva A, Shalaev VM. Planar photonics with metasurfaces. Science 2013; 339: 1232009.10.1126/science.1232009Search in Google Scholar PubMed

[28] Lin D, Fan P, Hasman E, Brongersma ML. Dielectric gradient metasurface optical elements. Science 2014; 345: 298-302.10.1126/science.1253213Search in Google Scholar PubMed

[29] Yu N, Capasso F. Flat optics with designer metasurfaces. Nat Mater 2014; 13: 139-50.10.1038/nmat3839Search in Google Scholar PubMed

[30] Achouri K, Salem M, Caloz C. General metasurface synthesis based on susceptibility tensors. IEEE Trans Antennas Propag 2014; 63: 2977-91.10.1109/TAP.2015.2423700Search in Google Scholar

[31] Genevet P, Yu N, Aieta F, Lin J, Kats MA, Blanchard R, Scully MO, Gaburro Z, Capasso F. Ultra-thin plasmonic optical vortex plate based on phase discontinuities. Appl Phys Lett 2012; 100: 013101.10.1063/1.3673334Search in Google Scholar

[32] Tetienne J, Blanchard R, Yu N, Genevet P, Kats M, Fan J, Edamura T, Furuta S, Yamanishi M, Capasso F. Dipolar modeling and experimental demonstration of multi-beam plasmonic collimators. New J Phys 2011; 13: 053057.10.1088/1367-2630/13/5/053057Search in Google Scholar

[33] Yu N, Aieta F, Genevet P, Kats MA, Gaburro Z, Capasso F. A broadband, background-free quarter-wave plate based on plasmonic metasurfaces. Nano Lett 2012; 12: 6328-33.10.1021/nl303445uSearch in Google Scholar PubMed

[34] Aieta F, Genevet P, Kats MA, Yu N, Blanchard R, Gaburro Z, Capasso F. Aberration-free ultrathin flat lenses and axicons at telecom wavelengths based on plasmonic metasurfaces. Nano Lett 2012; 12: 4932-6.10.1021/nl302516vSearch in Google Scholar PubMed

[35] Ni X, Ishii S, Kildishev AV, Shalaev VM. Ultra-thin, planar, Babinet-inverted plasmonic metalenses. Light Sci Appl 2013;2:e72.10.1038/lsa.2013.28Search in Google Scholar

[36] Genevet P, Capasso F. Holographic optical metasurfaces: a review of current progress. Reports on Rep Prog Phys 2015; 78: 024401.10.1088/0034-4885/78/2/024401Search in Google Scholar PubMed

[37] Huang K, Liu H, Garcia-Vidal FJ, Hong M, Luk’yanchuk B, Teng J, Qiu CW. Ultrahigh-capacity non-periodic photon sieves operating in visible light. Nat Commun 2015; 6: 7059.10.1038/ncomms8059Search in Google Scholar PubMed

[38] Scheuer J, Yifat Y. Holography: Metasurfaces make it practical. Nat Nanotechnol 2015; 10: 296-8.10.1038/nnano.2015.26Search in Google Scholar PubMed

[39] Sun S, He Q, Xiao S, Xu Q, Li X, Zhou L. Gradient-index metasurfaces as a bridge linking propagating waves and surface waves. Nat Mater 2012; 11: 426-31.10.1038/nmat3292Search in Google Scholar PubMed

[40] Huang L, Chen X, Mühlenbernd H, Zhang H, Chen S, Bai B, Tan Q, Jin G, Cheah K-W, Qiu CW. Three-dimensional optical holography using a plasmonic metasurface. Nat Commun 2013; 4: 2808.10.1038/ncomms3808Search in Google Scholar

[41] Genevet P, Wintz D, Ambrosio A, She A, Blanchard R, Capasso F. Controlled steering of Cherenkov surface plasmon wakes with a one-dimensional metamaterial. Nat Nanotechnol 2015; 10: 804-9.10.1364/CLEO_QELS.2015.FW4E.4Search in Google Scholar

[42] Engheta N. Circuits with light at nanoscales: optical nanocircuits inspired by metamaterials. Science 2007; 317: 1698-1702.10.1126/science.1133268Search in Google Scholar PubMed

[43] Alů A, Engheta N. All optical metamaterial circuit board at the nanoscale. Phys Rev Lett 2009; 103: 143902.10.1103/PhysRevLett.103.143902Search in Google Scholar PubMed

[44] Sun Y, Edwards B, Alů A, Engheta N. Experimental realization of optical lumped nanocircuits at infrared wavelengths. Nat Mater 2012; 11: 208-12.10.1038/nmat3230Search in Google Scholar PubMed

[45] Cai W, Chettiar UK, Kildishev AV, Shalaev VM. Optical cloaking with metamaterials. Nat Photon 2007; 1: 224-7.10.1038/nphoton.2007.28Search in Google Scholar

[46] Alu A, Engheta N. Plasmonic and metamaterial cloaking: physical mechanisms and potentials. J Opt 2008; 10: 093002.Search in Google Scholar

[47] Chen PY, Alů A. Mantle cloaking using thin patterned metasurfaces. Phys Rev B 2011; 84: 205110.10.1103/PhysRevB.84.205110Search in Google Scholar

[48] Zhao Y, Liu X-X, Alů A. Recent advances on optical metasurfaces. J Opt 2014; 16: 123001.10.1088/2040-8978/16/12/123001Search in Google Scholar

[49] Vahala KJ. Optical microcavities. Nature 2003; 424: 839-46.10.1038/nature01939Search in Google Scholar PubMed

[50] Lipson A, Lipson SG, Lipson H. Optical physics (4th ed.). Cambridge University Press 2010.10.1017/CBO9780511763120Search in Google Scholar

[51] Yamamoto Y, Slusher RE. Optical processes in microcavities, in Confined Electrons and Photons, New York: Springer 1995:871-8.10.1007/978-1-4615-1963-8_46Search in Google Scholar

[52] Armani AM, Kulkarni RP, Fraser SE, Flagan RC, Vahala KJ. Label-free, single-molecule detection with optical microcavities. Science 2007; 317: 783-7.10.1126/science.1145002Search in Google Scholar PubMed

[53] Del’Haye P, Schliesser A, Arcizet O, Wilken T, Holzwarth R, Kippenberg T. Optical frequency comb generation from a monolithic microresonator. Nature 2007; 450: 1214-7.10.1038/nature06401Search in Google Scholar PubMed

[54] Skivesen N, Trtu A, Kristensen M, Kjems J, Frandsen LH, Borel PI. Photonic-crystal waveguide biosensor. Opt Express 2007; 15: 3169-76.10.1109/OFC.2008.4528627Search in Google Scholar

[55] Vollmer F, Arnold S. Whispering-gallery-mode biosensing: label-free detection down to single molecules. Nat Methods 2008; 5: 591-6.10.1038/nmeth.1221Search in Google Scholar PubMed

[56] Kippenberg TJ, Holzwarth R, Diddams S. Microresonator-based optical frequency combs. Science 2011; 332: 555-9.10.1126/science.1193968Search in Google Scholar PubMed

[57] Maier SA, Atwater HA. Plasmonics: Localization and guiding of electromagnetic energy in metal/dielectric structures. J Appl Phys 2005; 98: 011101.10.1063/1.1951057Search in Google Scholar

[58] Maier SA. Plasmonics: fundamentals and applications. Springer 2007.10.1007/0-387-37825-1Search in Google Scholar

[59] Schuller JA, Barnard ES, Cai W, Jun YC, White JS, Brongersma ML. Plasmonics for extreme light concentration and manipulation. Nat Mater 2010; 9: 193-204.10.1038/nmat2630Search in Google Scholar PubMed

[60] Novotny L, Hecht B. Principles of nano-optics. Cambridge University Press 2012.10.1017/CBO9780511794193Search in Google Scholar

[61] Tassin P, Koschny T, Kafesaki M, Soukoulis CM. A comparison of graphene, superconductors and metals as conductors for metamaterials and plasmonics. Nat Photon 2012; 6: 259-64.10.1038/nphoton.2012.27Search in Google Scholar

[62] Blaber MG, Arnold MD, Ford MJ. Designing materials for plasmonic systems: the alkali-noble intermetallics. J Phys Condens Matter 2010; 22: 095501.10.1088/0953-8984/22/9/095501Search in Google Scholar PubMed

[63] Blaber MG, Arnold MD, Ford MJ. A review of the optical properties of alloys and intermetallics for plasmonics. J Phys Condens Matter 2010; 22: 143201.10.1088/0953-8984/22/14/143201Search in Google Scholar PubMed

[64] Palik ED. Handbook of optical constants of solids. Academic press 1998.Search in Google Scholar

[65] High AA, Devlin RC, Dibos A, Polking M, Wild DS, Perczel J, de Leon NP, Lukin MD, Park H. Visible-frequency hyperbolic metasurface. Nature 2015; 522: 192-6.10.1038/nature14477Search in Google Scholar PubMed

[66] Feigenbaum E, Diest K, Atwater HA. Unity-order index change in transparent conducting oxides at visible frequencies. Nano Lett 2010; 10: 2111-6.10.1021/nl1006307Search in Google Scholar PubMed

[67] Adachi S. The Handbook on Optical Constants of Metals: In Tables and Figures. Singapore: World Scientific, 2012.10.1142/8479Search in Google Scholar

[68] Falkovsky L. Optical properties of graphene. J Phys Conf Ser 2008; 129: 012004.10.1088/1742-6596/129/1/012004Search in Google Scholar

[69] Jablan M, Buljan H, Soljačić M. Plasmonics in graphene at infrared frequencies. Phys Rev B 2009; 80: 245435.10.1103/PhysRevB.80.245435Search in Google Scholar

[70] Naik GV, Schroeder JL, Ni X, Kildishev AV, Sands TD, Boltasseva A. Titanium nitride as a plasmonic material for visible and nearinfrared wavelengths. Opt Mater Express 2012; 2: 478-89.10.1364/OME.2.000478Search in Google Scholar

[71] Guler U, Boltasseva A, Shalaev VM. Refractory plasmonics. Science 2014; 344: 263-4.10.1126/science.1252722Search in Google Scholar PubMed

[72] Armani D, Kippenberg T, Spillane S, Vahala K. Ultra-high-Q toroid microcavity on a chip. Nature 2003; 421: 925-8.10.1038/nature01371Search in Google Scholar PubMed

[73] Grudinin IS, Matsko AB, Savchenkov AA, Strekalov D, Ilchenko VS, Maleki L. Ultra high Q crystalline microcavities. Opt Commun 2006; 265: 33-8.10.1109/IQEC.2005.1560894Search in Google Scholar

[74] Zhao Y, Belkin M, Alů A. Twisted optical metamaterials for planarized ultrathin broadband circular polarizers. Nat Commun 2012; 3: 870.10.1038/ncomms1877Search in Google Scholar PubMed

[75] West PR, Ishii S, Naik GV, Emani NK, Shalaev VM, Boltasseva A. Searching for better plasmonic materials. Laser Photon Rev 2014; 4: 795:808.10.1002/lpor.200900055Search in Google Scholar

[76] Boltasseva A, Atwater HA. Low-loss plasmonic metamaterials. Science 2011; 331: 290-1.10.1126/science.1198258Search in Google Scholar PubMed

[77] Naik GV, Shalaev VM, Boltasseva A. Alternative plasmonic materials: beyond gold and silver. Adv Mater 2013; 25: 3264-94.10.1002/adma.201205076Search in Google Scholar PubMed

[78] Zheng G, Muhlenbernd H, Kenney M, Li G, Zentgraf T, Zhang S. Metasurface holograms reaching 80% efficiency. Nat Nanotechnol 2015; 10: 308-12.10.1038/nnano.2015.2Search in Google Scholar PubMed

[79] Pancharatnam S. Generalized theory of interference, and its applications. Proc Ind Acad Sci A 1956; 44: 247-62.10.1007/BF03046050Search in Google Scholar

[80] Bomzon Ze, Biener G, Kleiner V, Hasman E. Space-variant Pancharatnam-Berry phase optical elements with computer-generated subwavelength gratings. Opt Lett 2002; 27: 1141-3.10.1364/OL.27.001141Search in Google Scholar PubMed

[81] Li G, Kang M, Chen S, Zhang S, Pun EY-B, Cheah KW, Li J. Spinenabled plasmonic metasurfaces for manipulating orbital angular momentum of light. Nano Lett 2013; 13: 4148-51.10.1021/nl401734rSearch in Google Scholar PubMed

[82] Yin X, Ye Z, Rho J, Wang Y, Zhang X. Photonic spin Hall effect at metasurfaces. Science 2013; 339: 1405-7.10.1126/science.1231758Search in Google Scholar PubMed

[83] Yang Y, Wang W, Moitra P, Kravchenko II, Briggs DP, Valentine J. Dielectric meta-reflectarray for broadband linear polarization conversion and optical vortex generation. Nano Lett 2014; 14: 1394-9.10.1021/nl4044482Search in Google Scholar PubMed

[84] Lee S-Y, Kim K, Kim S-J, Park H, Kim K-Y, Lee B. Plasmonic metaslit: shaping and controlling near-field focus. Optica 2015; 2: 6-13.10.1364/OPTICA.2.000006Search in Google Scholar

[85] Zhang S, Wei H, Hakanson U, Halas NJ, Nordlander P, Xu H. Chiral surface plasmon polaritons on metallic nanowires. Phys Rev Lett 2011; 107: 096801.10.1103/PhysRevLett.107.096801Search in Google Scholar PubMed

[86] Wu C, Burton Neuner I, Shvets G, John J, Milder A, Zollars B, Savoy S. Large-area wide-angle spectrally selective plasmonic absorber. Phys Rev B 2011; 84: 075102.10.1103/PhysRevB.84.075102Search in Google Scholar

[87] Pors A, Nielsen MG, Eriksen RL, Bozhevolnyi SI. Broadband focusing flat mirrors based on plasmonic gradient metasurfaces. Nano Lett 2013; 13: 829-34.10.1021/nl304761mSearch in Google Scholar PubMed

[88] Pors A, Nielsen MG, Bozhevolnyi SI. Broadband plasmonic half-wave plates in reflection. Opt Lett 2013; 38: 513-5.10.1364/OL.38.000513Search in Google Scholar PubMed

[89] Pors A, Albrektsen O, Radko IP, Bozhevolnyi SI. Gap plasmon-based metasurfaces for total control of reflected light. Sci Rep 2013; 3: 02155.10.1038/srep02155Search in Google Scholar PubMed PubMed Central

[90] Engheta N, Saladrino A, Alů A. Circuit elements at optical frequencies: nano-inductor, nano-capacitor, and nano-resistor. Phys Rev Lett 2005; 95: 095504.10.1103/PhysRevLett.95.095504Search in Google Scholar PubMed

[91] Monticone F, Estakhri NM, Alů A. Full control of nanoscale optical transmission with a composite metascreen. Phys Rev Lett 2013; 110: 203903.10.1103/PhysRevLett.110.203903Search in Google Scholar PubMed

[92] Pfeiffer C, Grbic A. Metamaterial Huygens’ surfaces: tailoring wave fronts with reflectionless sheets. Phys Rev Lett 2013; 110: 197401.10.1103/PhysRevLett.110.197401Search in Google Scholar PubMed

[93] Pfeiffer C, Zhang C, Ray V, Guo LJ, Grbic A. High performance bianisotropic metasurfaces: asymmetric transmission of light. Phys Rev Lett 2014; 113: 023902.10.1103/PhysRevLett.113.023902Search in Google Scholar PubMed

[94] Pfeiffer C, Emani NK, Shaltout AM, Boltasseva A, Shalaev VM, Grbic A. Efficient light bending with isotropic metamaterial huygens’ surfaces. Nano Lett 2014; 14: 2491-7.10.1021/nl5001746Search in Google Scholar PubMed

[95] Evlyukhin AB, Reinhardt C, Seidel A, Luk’yanchuk BS, Chichkov BN. Optical response features of Si-nanoparticle arrays. Phys Rev B 2010; 82: 045404.10.1103/PhysRevB.82.045404Search in Google Scholar

[96] García-Etxarri A, Gόmez-Medina R, Froufe-Pérez LS, Lόpez C, Chantada L, Scheffold F, Aizpurua J, Nieto-Vesperinas M, Sáenz JJ. Strong magnetic response of submicron silicon particles in the infrared. Opt Express 2011; 19: 4815-26.10.1364/OE.19.004815Search in Google Scholar PubMed

[97] Evlyukhin AB, Novikov SM, Zywietz U, Eriksen RL, Reinhardt C, Bozhevolnyi SI, Chichkov BN. Demonstration of magnetic dipole resonances of dielectric nanospheres in the visible region. Nano Lett 2012; 12: 3749-55.10.1021/nl301594sSearch in Google Scholar PubMed

[98] Kuznetsov AI, Miroshnichenko AE, Fu YH, Zhang J, Luk’yanchuk B. Magnetic light. Sci Rep 2012; 2: 492.10.1038/srep00492Search in Google Scholar PubMed PubMed Central

[99] Swanson GJ. Binary optics technology: the theory and design of multi-level diffractive optical elements. MIT Lincoln Lab Technical Report 854, NTIS Publ. AD-A213-404, 1989.10.21236/ADA213404Search in Google Scholar

[100] Lalanne P, Astilean S, Chavel P, Cambril E, Launois H. Design and fabrication of blazed binary diffractive elements with sampling periods smaller than the structural cutoff. J Opt Soc Am A 1999; 16: 1143-56.10.1364/JOSAA.16.001143Search in Google Scholar

[101] Fattal D, Li J, Peng Z, Fiorentino M, Beausoleil RG. Flat dielectric grating reflectors with focusing abilities. Nat Photon 2010; 4: 466-70.10.1038/nphoton.2010.116Search in Google Scholar

[102] West PR, Stewart JL, Kildishev AV, Shalaev VM, Shkunov VV, Strohkendl F, Zakharenkov YA, Dodds RK, Byren R. All-dielectric subwavelength metasurface focusing lens. Opt Express 2014; 22: 26212-21.10.1364/OE.22.026212Search in Google Scholar PubMed

[103] Levy U, Kim HC, Tsai C-H, Fainman Y. Near-infrared demonstration of computer-generated holograms implemented by using subwavelength gratings with space-variant orientation. Opt Lett 2005; 30: 2089-91.10.1364/OL.30.002089Search in Google Scholar PubMed

[104] Mie G. Beiträge zur Optik trüber Medien, speziell kolloidaler Metallösungen. Ann. Phys. (Berlin) 1908; 330: 377-445.10.1002/andp.19083300302Search in Google Scholar

[105] Lewin L. The electrical constants of a material loaded with spherical particles. J. Inst. Electr. Eng. 3 1947; 94: 65-8.10.1049/ji-3-2.1947.0013Search in Google Scholar

[106] Kerker M, Wang DS, Giles C. Electromagnetic scattering by magnetic spheres. J Opt Soc Am 1983; 73: 765-7.10.1364/JOSA.73.000765Search in Google Scholar

[107] Videen G, Bickel WS. Light-scattering resonances in small spheres. Phys Rev A 1992; 45: 6008.10.1103/PhysRevA.45.6008Search in Google Scholar PubMed

[108] Cao L, White JS, Park JS, Schuller JA, Clemens BM, Brongersma ML. Engineering light absorption in semiconductor nanowire devices. Nat Mater 2009; 8: 643-7.10.1038/nmat2477Search in Google Scholar PubMed

[109] Cao L, Fan P, Barnard ES, Brown AM, Brongersma ML. Tuning the color of silicon nanostructures. Nano Lett 2010; 10: 2649-54.10.1021/nl1013794Search in Google Scholar PubMed

[110] Person S, Jain M, Lapin Z, Sáenz JJ, Wicks G, Novotny L. Demonstration of zero optical backscattering from single nanoparticles. Nano Lett 2013; 13: 1806-9.10.1021/nl4005018Search in Google Scholar PubMed

[111] Ginn JC, Brener I, Peters DW, Wendt JR, Stevens JO, Hines PF, Basilio LI, Warne LK, Ihlefeld JF, Clem PG, Sinclair MB. Realizing optical magnetism from dielectric metamaterials. Phys Rev Lett 2012; 108: 097402.10.1103/PhysRevLett.108.097402Search in Google Scholar PubMed

[112] Geffrin JM, García-Cámara B, Gόmez-Medina R, Albella P, Froufe-Pérez L, Eyraud C, Litman A, Vaillon R, González F, Nieto-Vesperinas M. Magnetic and electric coherence in forward-and back-scattered electromagnetic waves by a single dielectric subwavelength sphere. Nat Commun 2012; 3: 1171.10.1038/ncomms2167Search in Google Scholar PubMed

[113] Fu YH, Kuznetsov AI, Miroshnichenko AE, Yu YF, Luk’yanchuk B. Directional visible light scattering by silicon nanoparticles. Nature Commun 2013; 4: 1527.10.1038/ncomms2538Search in Google Scholar PubMed

[114] Evlyukhin AB, Reinhardt C, Chichkov BN. Multipole light scattering by nonspherical nanoparticles in the discrete dipole approximation. Phys Rev B 2011; 84: 235429.10.1103/PhysRevB.84.235429Search in Google Scholar

[115] Staude I, Miroshnichenko AE, Decker M, Fofang NT, Liu S, Gonzales E, Dominguez J, Luk TS, Neshev DN, Brener I. Tailoring directional scattering through magnetic and electric resonances in subwavelength silicon nanodisks. ACS Nano 2013; 7: 7824-32.10.1021/nn402736fSearch in Google Scholar PubMed

[116] Luk’yanchuk BS, Voshchinnikov NV, Paniagua-Domínguez R, Kuznetsov AI. Optimum forward light scattering by spherical and spheroidal dielectric nanoparticles with high refractive index. ACS Photon 2015; 2: 993-999.10.1021/acsphotonics.5b00261Search in Google Scholar

[117] Decker M, Staude I, Falkner M, Dominguez J, Neshev DN, Brener I, Pertsch T, Kivshar YS. High-Efficiency Dielectric Huygens’ Surfaces. Adv Opt Mater 2015; 3: 813-20.10.1002/adom.201400584Search in Google Scholar

[118] Yu YF, Zhu AY, Paniagua-Domínguez R, Fu YH, Luk’yanchuk B, Kuznetsov AI. High-transmission dielectric metasurface with 2π phase control at visible wavelengths. Laser Photon Rev 2015; 9:412-418.10.1002/lpor.201500041Search in Google Scholar

[119] Chong KE, Stuade I, James A, Dominguez J, Liu S, Campione S, Subramania GS, Luk TS, Decker M, Neshev DN, Brener I, Kivshar YS. Polarization-independent silicon metadevices for efficient optical wavefront control. Nano Lett 2015; 15: 5369-74.10.1021/acs.nanolett.5b01752Search in Google Scholar PubMed

[120] Shalaev MI, Sun J, Tsukernik A, Pandey A, Nikolskiy K, Litchinitser NM. High-efficiency all-dielectric metasurfaces for ultracompact beam manipulation in transmission mode. Nano Lett 2015; 15: 6261-6.10.1021/acs.nanolett.5b02926Search in Google Scholar PubMed

[121] Arbabi A, Horie Y, Bagheri M, Faraon A. Dielectric metasurfaces for complete control of phase and polarization with sub-wavelength spatial resolution and high transmission. Nat Nanotechnol 2015; 10: 937-43.10.1038/nnano.2015.186Search in Google Scholar PubMed

[122] Arbabi A, Horie Y, Ball AJ, Bagheri M, Faraon A. Subwavelength-thick lenses with high numerical apertures and large efficiency based on high-contrast transmitarrays. Nat Commun 2015; 6: 7069.10.1038/ncomms8069Search in Google Scholar PubMed

[123] Moitra P, Slovick BA, Yu ZG, Krishnamurthy S, Valentine J. Experimental demonstration of a broadband all-dielectric meta-material perfect reflector. Appl Phys Lett 2014; 104: 171102.10.1063/1.4873521Search in Google Scholar

[124] Yang Y, Kravchenko II, Briggs DP, Valentine J. All-dielectric metasurface analogue of electromagnetically induced transparency. Nat Commun 2014; 5: 5753.10.1038/ncomms6753Search in Google Scholar PubMed

[125] Liu S, Sinclair MB, Mahony TS, Jun YC, Campione S, Ginn J, Bender DA, Wendt JR, Ihlefeld JF, Clem PG. Optical magnetic mirrors without metals. Optica 2014; 1: 250-6.10.1364/OPTICA.1.000250Search in Google Scholar

[126] Paniagua-Domínguez R, Yu YF, Miroshnichenko AE, Krivitsky LA, Fu YH, Valuckas V, Gonzaga L, Toh YT, Kay YSA, Luk’yanchuk B, Kuznetsov AI. Generalized Brewster effect in dielectric metasurfaces. Nat Commun 2015; 6: 10362.10.1038/ncomms10362Search in Google Scholar PubMed PubMed Central

[127] Schuller JA, Zia R, Taubner T, Brongersma M. Dielectric meta-materials based on electric and magnetic resonances of silicon carbide particles. Phys Rev Lett 2008; 99: 107401.10.1103/PhysRevLett.99.107401Search in Google Scholar PubMed

[128] Zheludev NI, Kivshar YS. From metamaterials to metadevices. Nat Mater 2012; 11: 917-24.10.1038/nmat3431Search in Google Scholar PubMed

[129] Tao H, Strikwerda A, Fan K, Padilla W, Zhang X, Averitt R. Reconfigurable terahertz metamaterials. Phys Rev Lett 2009; 103: 147401.10.1103/PhysRevLett.103.147401Search in Google Scholar PubMed

[130] Zhu WM, Liu AQ, Zhang XM, Tsai DP, Bourouina T, Teng JH, Zhang XH, Guo HC, Tanoto H, Mei T. Switchable magnetic metamaterials using micromachining processes. Adv Mater 2011; 23: 1792-6.10.1002/adma.201004341Search in Google Scholar PubMed

[131] Ou JY, Plum E, Zhang J, Zheludev NI. An electromechanically reconfigurable plasmonic metamaterial operating in the nearinfrared. Nat Nanotechnol 2013; 8: 252-5.10.1038/nnano.2013.25Search in Google Scholar PubMed

[132] Chen HT, Padilla WJ, Zide JM, Gossard AC, Taylor AJ, Averitt RD. Active terahertz metamaterial devices. Nature 2006; 444: 597-600.10.1038/nature05343Search in Google Scholar PubMed

[133] Padilla WJ, Taylor AJ, Highstrete C, Lee M, Averitt RD. Dynamical electric and magnetic metamaterial response at terahertz frequencies. Phys Rev Lett 2006; 96: 107401.10.1109/CLEO.2006.4627768Search in Google Scholar

[134] Yi F, Shim E, Zhu AY, Zhu H, Reed JC, Cubukcu E. Voltage tuning of plasmonic absorbers by indium tin oxide. Appl Phys Lett 2013; 102: 221102.10.1063/1.4809516Search in Google Scholar

[135] Watts CM, Shrekenhamer D, Montoya J, Lipworth G, Hunt J, Sleasman T, Krishna S, Smith DR, Padilla WJ. Terahertz compressive imaging with metamaterial spatial light modulators. Nat Photon 2014.10.1038/nphoton.2014.139Search in Google Scholar

[136] Lee HW, Papadakis G, Burgos SP, Chander K, Kriesch A, Pala R, Peschel U, Atwater HA. Nanoscale Conducting Oxide Plas-MOStor. Nano Lett 2014; 14: 6463-8.10.1021/nl502998zSearch in Google Scholar PubMed

[137] Krasavin AV, Zheludev N. Active plasmonics: Controlling signals in Au/Ga waveguide using nanoscale structural transformations. Appl Phys Lett 2004; 84: 1416-8.10.1063/1.1650904Search in Google Scholar

[138] Driscoll T, Kim HT, Chae B-G, Kim B-J, Lee Y-W, Jokerst NM, Palit S, Smith DR, Di Ventra M, Basov DN. Memory metamaterials. Science 2009; 325: 1518-21.10.1126/science.1176580Search in Google Scholar PubMed

[139] Dicken MJ, Aydin K, Pryce IM, Sweatlock LA, Boyd EM, Walavalkar S, Ma J, Atwater HA. Frequency tunable nearinfrared metamaterials based on VO 2 phase transition. Opt Express 2009; 17: 18330-9.10.1364/OE.17.018330Search in Google Scholar PubMed

[140] Samson Z, MacDonald K, De Angelis F, Gholipour B, Knight K, Huang C, Di Fabrizio E, Hewak D, Zheludev N. Metamaterial electro-optic switch of nanoscale thickness. Appl Phys Lett 2010; 96: 143105.10.1063/1.3355544Search in Google Scholar

[141] Liu N, Tang ML, Hentschel M, Giessen H, Alivisatos AP. Nanoantenna-enhanced gas sensing in a single tailored nanofocus. Nat Mater 2011; 10: 631-6.10.1038/nmat3029Search in Google Scholar PubMed

[142] Kats MA, Sharma D, Lin J, Genevet P, Blanchard R, Yang Z, Qazilbash MM, Basov D, Ramanathan S, Capasso F. Ultra-thin perfect absorber employing a tunable phase change material. Appl Phys Lett 2012; 101: 221101.10.1063/1.4767646Search in Google Scholar

[143] Kats MA, Blanchard R, Zhang S, Genevet P, Ko C, Ramanathan S, Capasso F. Vanadium dioxide as a natural disordered meta-material: perfect thermal emission and large broadband negative differential thermal emittance. Phys Rev X 2013; 3: 041004.Search in Google Scholar

[144] Strohfeldt N, Tittl A, Scha “ferling M, Neubrech F, Kreibig U, Griessen R, Giessen H. Yttrium hydride nanoantennas for active plasmonics. Nano Lett 2014; 14: 1140-7.10.1021/nl403643vSearch in Google Scholar PubMed

[145] Wang Q, Rogers ETF, Gholipour B, Wang CM, Yuan G, Teng JH, Zheludev NI. Optically reconfigurable metasurfaces and photonic devices based on phase change materials. Nat Photon 2016; 10: 60-6.10.1038/nphoton.2015.247Search in Google Scholar

[146] Zhao Q, Kang L, Du B, Li B, Zhou J, Tang H, Liang X, Zhang B. Electrically tunable negative permeability metamaterials based on nematic liquid crystals. Appl Phys Lett 2007; 90: 011112.10.1063/1.2430485Search in Google Scholar

[147] Shrekenhamer D, Chen W-C, Padilla WJ. Liquid crystal tunable metamaterial absorber. Phys Rev Lett 2013; 110: 177403.10.1103/PhysRevLett.110.177403Search in Google Scholar PubMed

[148] Buchnev O, Podoliak N, Kaczmarek M, Zheludev NI, Fedotov VA. Electrically controlled nanostructured metasurface loaded with liquid crystal: toward multifunctional photonic switch. Adv Opt Mater 2015; 3: 674-9.10.1002/adom.201400494Search in Google Scholar

[149] Ziegler K. Robust transport properties in graphene. Phys Rev Lett 2006; 97: 266802.10.1103/PhysRevLett.97.266802Search in Google Scholar PubMed

[150] Geim AK, Novoselov KS. The rise of graphene. Nat Mater 2007; 6: 183-91.10.1142/9789814287005_0002Search in Google Scholar

[151] Neto AC, Guinea F, Peres N, Novoselov KS, Geim AK. The electronic properties of graphene. Rev Mod Phys 2009; 81: 109.10.1103/RevModPhys.81.109Search in Google Scholar

[152] Novoselov KS, Geim AK, Morozov S, Jiang D, Zhang Y, Dubonos Sa, Grigorieva I, Firsov A. Electric field effect in atomically thin carbon films. Science 2004; 306: 666-9.10.1126/science.1102896Search in Google Scholar PubMed

[153] Yu YJ, Zhao Y, Ryu S, Brus LE, Kim KS, Kim P. Tuning the graphene work function by electric field effect. Nano Lett 2009; 9: 3430-4.10.1021/nl901572aSearch in Google Scholar PubMed

[154] Lin YM, Dimitrakopoulos C, Jenkins KA, Farmer DB, Chiu H-Y, Grill A, Avouris P. 100-GHz transistors from wafer-scale epitaxial graphene. Science 2010; 327: 662-.10.1126/science.1184289Search in Google Scholar PubMed

[155] Luk’yanchuk IA, Kopelevich Y. Phase analysis of quantum oscillations in graphite. Phys Rev Lett 2004; 93: 166402.10.1103/PhysRevLett.93.166402Search in Google Scholar PubMed

[156] Wang F, Zhang Y, Tian C, Girit C, Zettl A, Crommie M, Shen YR. Gate-variable optical transitions in graphene. Science 2008; 320: 206-9.10.1126/science.1152793Search in Google Scholar PubMed

[157] Ju L, Geng B, Horng J, Girit C, Martin M, Hao Z, Bechtel HA, Liang X, Zettl A, Shen YR. Graphene plasmonics for tunable terahertz metamaterials. Nat Nanotechnol 2011; 6: 630-4.10.1038/nnano.2011.146Search in Google Scholar PubMed

[158] Vakil A, Engheta N.Transformation optics using graphene. Science 2011; 332: 1291-4.10.1126/science.1202691Search in Google Scholar PubMed

[159] Yan H, Li X, Chandra B, Tulevski G, Wu Y, Freitag M, Zhu W, Avouris P, Xia F. Tunable infrared plasmonic devices using graphene/insulator stacks. Nat Nanotechnol 2012; 7: 330-4.10.1038/nnano.2012.59Search in Google Scholar PubMed

[160] Fei Z, Rodin AS, Andreev GO, Bao W, McLeod AS, Wagner M, Zhang LM, Zhao Z, Thiemens M, Dominguez G, Fogler MM, Castro Neto AH, Lau CN, Keilmann F, Basov DN. Gate-tuning of graphene plasmons revealed by infrared nano-imaging. Nature 2012; 487: 82-5.10.1038/nature11253Search in Google Scholar PubMed

[161] Chen J, Badioli M, Alonso-Gonzalez P, Thongrattanasiri S, Huth F, Osmond J, Spasenovic M, Centeno A, Pesquera A, Godignon P, Elorza AZ, Camara N, Garcia de Abajo FJ, Hillenbrand R, Kop-pens FH. Optical nano-imaging of gate-tunable graphene plasmons. Nature 2012; 487: 77-81.10.1038/nature11254Search in Google Scholar PubMed

[162] Yan H, Low T, Zhu W, Wu Y, Freitag M, Li X, Guinea F, Avouris P, Xia F. Damping pathways of mid-infrared plasmons in graphene nanostructures. Nat Photon 2013; 7: 394-9.10.1038/nphoton.2013.57Search in Google Scholar

[163] Emani NK, Chung T-F, Ni X, Kildishev AV, Chen YP, Boltasseva A. Electrically tunable damping of plasmonic resonances with graphene. Nano Lett 2012; 12: 5202-6.10.1021/nl302322tSearch in Google Scholar PubMed

[164] Yao Y, Kats MA, Shankar R, Song Y, Kong J, Loncar M, Capasso F. Wide wavelength tuning of optical antennas on graphene with nanosecond response time. Nano Lett 2014; 14: 214-9.10.1021/nl403751pSearch in Google Scholar PubMed

[165] Mousavi SH, Kholmanov I, Alici KB, Purtseladze D, Arju N, Tatar K, Fozdar DY, Suk JW, Hao Y, Khanikaev AB. Inductive tuning of Fano-resonant metasurfaces using plasmonic response of graphene in the mid-infrared. Nano Lett 2013; 13: 1111-7.10.1021/nl304476bSearch in Google Scholar PubMed

[166] Zhu AY, Yi F, Reed JC, Zhu H, Cubukcu E. Optoelectromechanical multimodal biosensor with graphene active region. Nano Lett 2014; 14: 5641-9.10.1021/nl502279cSearch in Google Scholar PubMed

[167] Fei Z, Rodin A, Andreev G, Bao W, McLeod A, Wagner M, Zhang L, Zhao Z, Thiemens M, Dominguez G. Gate-tuning of graphene plasmons revealed by infrared nano-imaging. Nature 2012; 487: 82-5.10.1038/nature11253Search in Google Scholar PubMed

[168] Grigorenko A, Polini M, Novoselov K. Graphene plasmonics. Nat Photon 2012; 6: 749-58.10.1038/nphoton.2012.262Search in Google Scholar

[169] Koppens FH, Chang DE, Garcia de Abajo FJ. Graphene plasmonics: a platform for strong light–matter interactions. Nano Lett 2011; 11: 3370-7.10.1021/nl201771hSearch in Google Scholar PubMed

[170] Brar VW, Jang MS, Sherrott M, Lopez JJ, Atwater HA. Highly confined tunable mid-infrared plasmonics in graphene nanoresonators. Nano Lett 2013; 13: 2541-7.10.1021/nl400601cSearch in Google Scholar PubMed

[171] Xia F, Wang H, Xiao D, Dubey M, Ramasubramaniam A. Two-dimensional material nanophotonics. Nat Photon 2014; 8: 899-907.10.1038/nphoton.2014.271Search in Google Scholar

[172] Jang MS, Brar VW, Sherrott MC, Lopez JJ, Kim L, Kim S, Choi M, Atwater HA. Tunable large resonant absorption in a midinfrared graphene Salisbury screen. Phys Rev B 2014; 90: 165409.10.1103/PhysRevB.90.165409Search in Google Scholar

[173] Stauber T. Plasmonics in Dirac systems: from graphene to topological insulators. J Phys Condens Matter 2014; 26: 123201.10.1088/0953-8984/26/12/123201Search in Google Scholar PubMed

[174] Thongrattanasiri S, Koppens FH, Garcia de Abajo FJ. Complete optical absorption in periodically patterned graphene. Phys Rev Lett 2012; 108: 047401.10.1103/PhysRevLett.108.047401Search in Google Scholar PubMed

[175] Yoon H, Yeung KY, Umansky V, Ham D. A Newtonian approach to extraordinarily strong negative refraction. Nature 2012; 488: 65-9.10.1038/nature11297Search in Google Scholar PubMed

[176] Andress WF, Yoon H, Yeung KY, Qin L, West K, Pfeiffer L, Ham D. Ultra-subwavelength two-dimensional plasmonic circuits. Nano Lett 2012; 12: 2272-7.10.1021/nl300046gSearch in Google Scholar PubMed

[177] Zhang L, Hao J, Qiu M, Zouhdi S, Yang JKW, Qiu CW. Anomalous behavior of nearly-entire visible band manipulated with degenerated image dipole array. Nanoscale 2014; 6: 12303.10.1039/C4NR03163FSearch in Google Scholar PubMed

[178] Fang Z, Wang Y, Schlather AE, Liu Z, Ajayan PM, García de Abajo FJ, Nordlander P, Zhu X, Halas NJ. Active tunable absorption enhancement with graphene nanodisk arrays. Nano Lett 2013; 14: 299-304.10.1021/nl404042hSearch in Google Scholar PubMed

[179] Kumar K, Duan H, Hegde RS, Koh SCW, Wei JN, Yang JKW. Printing colour at the optical diffraction limit. Nat Nanotechnol 2012; 7: 557-61.10.1038/nnano.2012.128Search in Google Scholar PubMed

[180] Gu Y, Zhang L, Yang JKW, Yeo SP, Qiu CW. Color generation via subwavelength plasmonic nanostructures. Nanoscale 2015; 7: 6409.10.1039/C5NR00578GSearch in Google Scholar

[181] Tan SJ, Zhang L, Zhu D, Goh XM, Wang YM, Kumar K, Qiu CW, Yang JKW. Plasmonic color palettes for photorealistic printing with aluminum nanostructures. Nano Lett 2014; 14: 4023-9.10.1021/nl501460xSearch in Google Scholar PubMed

[182] Aieta F, Kats MA, Genevet P, Capasso F. Multiwavelength achromatic metasurfaces by dispersive phase compensation. Science 2015; 347: 1342-5.10.1126/science.aaa2494Search in Google Scholar PubMed

[183] Khorasaninejad M, Aieta F, Kanhaiya P, Kats MA, Genevet P, Rousso D, Capasso F. Achromatic metasurface lens at telecommunication wavelengths. Nano Lett 2015; 15: 5358-62.10.1021/acs.nanolett.5b01727Search in Google Scholar PubMed

[184] Bakker RM, Permyakov D, Yu YF, Markovich D, Paniagua-Domínguez R, Gonzaga L, Samusev A, Kivshar YS, Luk’yanchuk B, Kuznetsov AI. Magnetic and electric hotspots with silicon nanodimers. Nano Lett 2015; 15: 2137-42.10.1021/acs.nanolett.5b00128Search in Google Scholar PubMed

[185] Hasan MZ, Kane CL. Colloquium: topological insulators. Rev Mod Phys 2010; 82: 3045.10.1103/RevModPhys.82.3045Search in Google Scholar

[186] Kane CL, Mele EJ. Z 2 topological order and the quantum spin Hall effect. Phys Rev Lett 2005; 95: 146802.10.1103/PhysRevLett.95.146802Search in Google Scholar PubMed

[187] Hsieh D, Xia Y, Qian D, Wray L, Dil J, Meier F, Osterwalder J, Patthey L, Checkelsky J, Ong N. A tunable topological insulator in the spin helical Dirac transport regime. Nature 2009; 460: 1101-5.10.1038/nature08234Search in Google Scholar PubMed

[188] Zhang T, Cheng P, Chen X, Jia J-F, Ma X, He K, Wang L, Zhang H, Dai X, Fang Z. Experimental demonstration of topological surface states protected by time-reversal symmetry. Phys Rev Lett 2009; 103: 266803.10.1103/PhysRevLett.103.266803Search in Google Scholar PubMed

[189] Efimkin DK, Lozovik YE, Sokolik AA. Collective excitations on a surface of topological insulator. Nanoscale Res Lett 2012; 7: 1-10.10.1186/1556-276X-7-163Search in Google Scholar PubMed PubMed Central

[190] Raghu S, Chung SB, Qi X-L, Zhang S-C. Collective modes of a helical liquid. Phys Rev Lett 2010; 104: 116401.10.1103/PhysRevLett.104.116401Search in Google Scholar PubMed

[191] Shitrit N, Yulevich I, Maguid E, Ozeri D, Veksler D, Kleiner V, Hasman E. Spin-optical metamaterial route to spin-controlled photonics. Science 2013; 340: 724-6.10.1126/science.1234892Search in Google Scholar PubMed

[192] Bliokh KY, Smirnova D, Nori F. Quantum spin Hall effect of light. Science 2015; 348: 1448-51.10.1126/science.aaa9519Search in Google Scholar PubMed

[193] Wolf S, Awschalom D, Buhrman R, Daughton J, Von Molnar S, Roukes M, Chtchelkanova AY, Treger D. Spintronics: a spin-based electronics vision for the future. Science 2001; 294: 1488-95.10.1126/science.1065389Search in Google Scholar PubMed

[194] žutić I, Fabian J, Sarma SD. Spintronics: Fundamentals and applications. Rev Mod Phys 2004; 76: 323.10.1103/RevModPhys.76.323Search in Google Scholar

[195] Liao ZM, Han B-H, Wu HC, Yashina L, Yan Y, Zhou YB, Bie YQ, Bozhko S, Fleischer K, Shvets I. Surface plasmon on topological insulator/dielectric interface enhanced ZnO ultraviolet photoluminescence. AIP Adv 2012; 2: 022105.10.1063/1.3703320Search in Google Scholar

[196] Ou JY, So JK, Adamo G, Sulaev A, Wang L, Zheludev NI. Ultraviolet and visible range plasmonics in the topological insulator Bi1.5Sb0.5Te1.8Se1.2. Nat Commun 2014; 5: 5139.10.1038/ncomms6139Search in Google Scholar PubMed

[197] Di Pietro P, Ortolani M, Limaj O, Di Gaspare A, Giliberti V, Giorgianni F, Brahlek M, Bansal N, Koirala N, Oh S. Observation of Dirac plasmons in a topological insulator. Nat Nanotechnol 2013; 8: 556-60.10.1038/nnano.2013.134Search in Google Scholar PubMed

[198] Della Giovampaola C, Engheta N. Digital metamaterials. Nat Mater 2014; 13: 1115-21.10.1038/nmat4082Search in Google Scholar PubMed

[199] Cui TJ, Qi MQ, Wan X, Zhao J, Cheng Q. Coding metamaterials, digital metamaterials and programmable metamaterials. Light Sci Appl 2014;3:e218.10.1038/lsa.2014.99Search in Google Scholar

[200] Gao LH et al. Broadband diffusion of terahertz waves by multibit coding metasurfaces. Light Sci Appl 2015;4:e324.10.1038/lsa.2015.97Search in Google Scholar

Received: 2015-9-30
Accepted: 2016-1-27
Published Online: 2017-2-6
Published in Print: 2017-3-1

© 2017 Alexander Y. Zhu et al.

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 3.0 License.

Downloaded on 19.3.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2016-0032/html
Scroll to top button