Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Analyzing the polarization response of a chiral metasurface stack by semi-analytic modeling

Open Access Open Access

Abstract

We investigate a class of stacked metasurfaces where the interaction between layers is dominated by their respective far-field response. Using a semi-analytic scattering matrix approach, we exploit the Fabry-Perot-type response for different layer distances to show the spectral tunability of the resonant effect. This method presents a faster and more intuitive route to modeling Fabry-Perot-type effects than rigorous numerical simulations. The results are illustrated for a chiral metasurface stack that exhibits asymmetric transmission. Here, the effect of asymmetric transmission is highly sensitive to the layer distance, which is used as a free parameter in our model. To prove our theoretical findings we fabricate two variants of the stack with different layer distances and show that far-field interaction between layers is sufficient to generate the effect while being accessible by semi-analytic modeling. The analyticity of the approach is promising for designing sophisticated layered media containing stacks of arbitrary metasurfaces.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The interaction of electromagnetic radiation with anisotropic media is an essential part of modern electrodynamics that has sparked many ideas on how polarization states can be manipulated with different materials. Whether a material is isotropic, anisotropic or even chiral is ultimately determined by its symmetry properties. These can have their origin either in molecular properties or lattice properties of periodic media such as crystals. Of particular interest are materials with artificial subwavelength structures in various arrangements, known as metamaterials [1]. They allow anisotropy to be designed freely from the geometry, material and distribution of their constituent subwavelength structures [2, 3]. In recent years the two-dimensional versions of metamaterials, so called metasurfaces [4, 5], manifested themselves as the most widely used architecture [6, 7].

Anisotropic metasurfaces can be engineered to manipulate polarization generating effects such as asymmetric transmission [8–10], dichroism [11, 12] or optical activity [13, 14]. If the geometry of the structures used is such that they cannot be superposed with their mirror, image they become chiral and can produce a chiral polarization response in turn [15]. To be geometrically chiral, structures have to be three-dimensional [16–19]. Nevertheless, two-dimensional structure designs have been demonstrated, which can also achieve a chiral polarization response [20, 21]. A prominent example are resonant particles such as asymmetric split ring resonators [22, 21].Yet another approach uses bianisotropic bilayers of metasurfaces where the structures are neither chiral nor bianisotropic themselves. Here, bianisotropy is induced through near-field coupling between the layers and can create or enhance chiral polarization effects [10, 23–25]. Following this approach achiral particles like wires [26, 27] or crosses [28, 29] can be rotated with respect to each other to achieve chiral polarization effects [30].

Taking a similar path, it was shown by Zhao et.al. [31] that multiple layers of successively rotated nano-wires can form a so called twisted metamaterial with a chiral polarization response in the absence of near-field coupling. In addition, their work demonstrated that lateral layer alignment during fabrication could be ignored in this type of stacked metasurface [26]. This made apparent that stacking different geometries is a promising route towards designing polarization properties using only basic structure types (e.g. wires, Ls or crosses) and, thus, less challenging fabrication schemes. However, when combining metasurfaces vertically, it is frequently overlooked that spacer layers do not only separate the functional layers of the stack but determine their physical interaction as well. Here, we aim to describe spacer layers as an extra degree of freedom in the design of stacked metasurfaces.

Indeed, the understanding of inter-layer interactions controlled by spacer layers is of great importance from the point of homogenization. It was shown that a periodic metasurface can be considered to be an effective medium if and only if its far-field response is governed by a fundamental mode [32, 33]. For multiple layers of metasurfaces this means that the inter-layer interaction in the stack is solely determined via the fundamental modes of the layers [34]. This differs from the before mentioned case of closely stacked bilayers. In that case, due to the near-field coupling their inter-layer interaction is also influenced by higher order evanescent modes [35, 36]. On the other hand, if distances between metasurface layers are large enough such that all higher order evanescent modes have decayed, interaction on the far-field level via fundamental modes can be assumed [36, 37]. The resulting stack can then be described by the properties of its constituent layers [34]. This, in turn, gives an intuitive perspective on their physical role in the stack, including the influence of spacer layers on anisotropic polarization effects. Ideologically and from a point analytic modeling, this far-field stacking of multiple layers is similar to the simplified analytical description of nano-structures using laterally coupled dipoles [38] or the coupling of resonant modes between periodic metasurfaces [39].

In this paper we consider a twisted metamaterial consisting of a stack of two nano-wire metasurfaces separated by a dielectric spacer, where the wires of one layer are rotated with respect to the ones of the other. We investigate the influence of the spacer layer on the polarization response of the stack and demonstrate how semi-analytics can give a precise understanding of Fabry-Perot-type interactions between stacked metasurfaces. We start with an analysis of the stack’s polarization properties based on the symmetry of its constituent layers. We demonstrate how to derive these properties by analyzing the structure of scattering matrices (S-matrices) describing the system (Sec. 2). Based on these considerations we introduce and investigate a fabricated stack that was designed to have pronounced asymmetric transmission, a strong difference in cross-polarization for orthogonal input polarization. Two versions of the stack are shown, with identical order and type of layers but different spacer thickness (Sec. 3). One refers to the case of far-field coupling and is analyzed using a semi-analytic stacking algorithm (SASA) considering only the fundamental mode of each layer. For comparison, the other incorporates near-field coupling. The focus of our analysis lies on the far-field coupled version to showcase the benefits of a semi-analytic approach. Using SASA we reveal and discuss the role of Fabry-Perot-type interactions for tailoring the polarization properties of the device.

 figure: Fig. 1

Fig. 1 Rendered sketch of a stack of gold nano-wire metasurfaces embedded in glass.

Download Full Size | PDF

2. Deducing polarization from S-matrix structure

We consider the case of two metasurfaces stacked along the z-axis, symmetrically embedded in glass, separated by a spacer of thickness dsp, and with a cladding layer of thickness dc

as a cover. The lower metasurface is comprised of parallel nano-wires, measuring w×l×h, aligned to the y-axis, and arranged in a square lattice. The upper metasurface is of the same type as the lower one but its wires are rotated as shown in Fig. 1. To describe the response of the stack and its constituent layers for excitation with normal incident linearly polarized light we make use of optical scattering matrices (S-matrices). An S-matrix is a 2-by-2 block matrix that contains the Jones-matrices T^ for forward (f) and backward (b) transmission on its diagonal and the forward and backward reflection matrices R^ on its off-diagonal. They have the meaning of Jones matrices in reflection. Thus, an arbitrary S-matrix S is defined as [34]

S=(T^fR^bR^fT^b).

Each Jones-matrix contains the four complex transmission or reflection coefficients Txx, Txy, Tyx, and Tyy or Rxx, Rxy, Ryx, and Ryy, respectively. Here, the right index of each index pair denotes the ingoing polarization state and the left index the outgoing one.

Our formulation relies on the condition that coupling between layers solely happens in the far-field via a propagating fundamental mode [34]. For periodic metasurfaces with period Λ this is represented by the zeroth diffraction order, where all higher orders decay evanescently. Generally, this is known as the fundamental mode approximation (FMA) [32, 35, 36]. If the FMA is valid for a metasurface, its S-matrix takes the form of a 4-by-4 matrix with two-dimensional block matrices. Furthermore, if all layers of a stack can be considered to be in the FMA, the S-matrix of the entire system can be calculated analytically as a combination of its constituent S-matrices using Redheffer’s starproduct * [40, 41]. As long as the FMA is valid any type of metasurface can be described and stacked in this way using either analytic, numerical or measured S-matrices. Utilizing this as a semi-analytic stacking algorithm (SASA) we can model stacks with an arbitrary number of layers of almost any type. In the following, we will use SASA to analyze the polarization behavior of the stack qualitatively from the structure of its S-matrices using only symbolic representations of their coefficients.

As the simplest part of the stack, the S-matrix of a dielectric spacer with refractive index n can be written as

Ssp=diag(P,P,P,P),
where P=exp(ink0dsp) and k0=2π/λ0 [34]. We begin our structural analysis of the stack’s S-matrices by adding such a spacer to an arbitrary layer S-matrix S using the starproduct, which results in
SSsp=(PTxxfPTxyfP2RxxbP2RxybPTyxfPTyyfP2RyxbP2RyybRxxfRxyfPTxxbPTxybRyxfRyyfPTyxbPTyyb).

There are serval things to note. First, as intuition would tell us, all transmission coefficients are multiplied by an identical phase term such that reciprocity is not affected [42]. Second, reflection coefficients add twice the phase depending on the side of incidence. Additionally, the overall structure of the S-matrix will not be changed by the spacer S-matrix because of the block-wise application of the phase terms.

The S-matrices of two wire-based metasurfaces from Fig. 1 can be constructed analytically. The only assumptions we need are the following: the particle geometry is identical in both metasurfaces, where one has wires aligned to the y-axis, the other wires rotated by an angle α about the z-axis, and both are symmetrically embedded. Similar to the discussion of unit cell symmetries and the resulting Jones matrix symmetries in references [2] and [34] we can derive each of our S-matrices by considering the polarization properties of nano-wires. We know that a nano-wire will have a different polarization response depending along which axis it is excited. It is also clear from the metasurfaces’ symmetric embedding that its response is reciprocal and, thus, its S-matrix is symmetric [43]. Accordingly, it is straight forward to write the S-matrix of the wires Sw as

Sw=(Tx0Rx00Ty0RyRx0Tx00Ry0Ty),
where the single-letter subscript of the coefficients gives credit to the cross-polarization free response. For the second metasurface, we can analytically approximate the rotation by an angle α of the wires in the unit cell by applying a rotation on Sw [34, 37]. This implies a rotation of the medium itself. If no near-field coupling occurs between the two metasurfaces, the same polarization characteristic is achieved as in the unit cell rotated case. However, the complex coefficients of the S-matrix will differ slightly due to different alignment of the wires towards each other in the lattice [37]. Nevertheless, our main interest lies in the structure of the S-matrix. For that purpose this approximation suffices. Thus, the S-matrix of rotated wires is given by
S˜w=(T˜xxT˜xyR˜xxR˜xyT˜xyT˜yyR˜xyR˜yyR˜xxR˜xyT˜xxT˜xyR˜xyR˜yyT˜xyT˜yy).

For readability we chose placeholder variables defined by T˜xx=Txcos2α+Tysin2α, T˜yy=Tycos2α+Txsin2α, T˜xy=(TyTx)cos αsin α, R˜xx=Rxcos2α+Rysin2α, R˜yy=Rycos2α+Rxsin2α, and R˜xy=(RyRx)cos αsin α. The structure of S˜w already shows that rotated wires may produce cross-polarization. However, this alone does not produce asymmetric transmission which is defined for reciprocal media and incident x-polarization (analogously for y-polarization) as the difference of the cross-polarization components in forward (or backward) direction [2],

Δx=|Txyf|2|Tyxf|2.

In case of eq. (5) the cross-polarization terms are identical (T˜xy=T˜yx and R˜xy=R˜yx) and Δx vanishes.

Eqs. (2), (4), and (5) give us the layers of the desired stack of rotated nano-wires on top of axis-aligned nano-wires. Performing the starproduct on the three S-matrices consecutively results in

S¯=Sw*SSp*S˜w=(T¯xxT¯xyR¯xxbR¯xybT¯yxT¯yyR¯xybR¯yybR¯xxfR¯xyfT¯xxT¯yxR¯xyfR¯yyfT¯xyT¯yy)

The explicit expressions of the resulting coefficients are given in the appendix and reveal that T¯xyT¯yx, i.e. asymmetric transmission Δx emerges from the stacking of the individual S-matrices. It is noteworthy that we can determine the general polarization behavior of the complete stack only from the symmetry properties of the individual layers.

 figure: Fig. 2

Fig. 2 (a) SEM picture of a FIB cut revealing the metasurface layers of the stack. The nano-particles in this image were colored golden during postprocessing to enhance visibility. (b) SEM image of the lower layer showing axis-aligned nano-wires. (c) SEM image of the upper layer showing nano-wires rotated counterclockwise by 60.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Experimental and simulation results of the stack in Jones matrix form: a) Txx, b) Txy, c) Tyx, d) Tyy. Blue curves show transmittance and orange ones phase. Measured data is represented by solid and dashed lines showing results for the 345 nm stack and 40 nm stack, respectively. Results obtained via SASA are marked with crosses and those from rigorous FMM calculations with circles.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Sketch of the stack construction with S-matrices using eq. (9). Dashed lines mark the interfaces of the stack to the surrounding medium. Arrows point in the direction of (forward) propagation along the z-axis. While spacer (spac.), and cladding (clad.) have defined thicknesses, the two metasurface layers (MS l, MS u) are assumed to be infinitely thin.

Download Full Size | PDF

3. Analysis of a fabricated stack by semi-analytic modeling

We fabricated a stack resembling the one shown in Fig. 1 and analyzed it both experimentally and employing SASA. For simplicity we chose equal periods with Λ=400nm and aligned unit cells for both metasurfaces. The rotation angle of the upper layer metasurface’s wires was set to α = 60°. During fabrication of each metasurface we applied a two-step electron beam lithography process. They were structured by exposure of a two layer electron beam resist (150 nm Allresist ARP617.08 and on top of that 100 nm Allresist ARP6200.4) with a variably shaped electron beam (Vistec, SB 350). This was followed by chemical development and coating by gold evaporation. Finally, we applied a lift-off process to remove the resist. During this process we produced two fields of the same metasurface. The spacer layer was spin-coated (Futurrex IC1-200) on top of both fields, tempered at 200°C for half an hour and etched to the desired spacer height. We produced two versions of the stack: spacer thickness dsp=40 nm on one field and dsp= 345 nm on the other. The former showcases the effect of evanescent coupling as opposed to the latter, coupling only in the far-field. The second metasurface was fabricated with the same electron beam lithography process on top of the respective spacer. In this way, we made sure that both versions were comparable with nearly identical structures. Finally, a cladding layer was spin-coated to dc= 900 nm. In total the field size of each fabricated sample was 2 × 2 mm2.

A scanning electron microscope (SEM) image of the resulting stack with dsp= 345 nm is shown in Fig. 2(a), where parts of the layers were removed with a focused ion beam (FIB) to reveal the structures underneath. Average dimensions of the nano-wires of both layers, Figs. 2(b) and (c), are almost identical, with: ll= 235 nm, wl= 75 nm, hl= 50 nm, lu= 225 nm, wu= 75 nm, and hu= 75 nm, with l, w, and h defined in Fig. 1. The subscripts l and u denote the lower and upper layer, respectively.

We characterized the stack in a wavelength range from 600 nm to 1620 nm using an interferometric setup which measures both the transmitted intensity (transmittance) and phase in a linear polarization basis [17, 44]. The diameter of the characterization beam on the sample was 2 mm, making sure that it was fully illuminated and no finite-size effects occurred. The results for both variants of the stack are plotted in Fig. 3, arranged in the form of a Jones matrix. The co-polarization components for x- and y-polarization, Figs. 3(a) and (d), vary only slightly in transmittance and are almost identical in phase. The difference between the respective cross-polarization components, Figs. 3(c) and (b), is more noticeable since they have much lower overall transmittance. In general, the functionality of both versions of the stack is similar. However, the peak in Tyx, Fig. 3(c), exhibits a redshift of 150 nm for the thin spacer stack. This indicates near-field coupling between the metasurfaces of the dc= 40 nm stack.

To investigate the dependence of the stack’s polarization behavior on the spacer we developed a model of the dsp= 345 nm version as a sequence of layers similar to the semi-analytic model eq. (7) using SASA. Now, interfaces to substrate and cladding of the stack have to be included, as shown in Fig. 4. Since the sample was fabricated and characterized on a glass waver, the substrate was assumed to be infinite. Both the spacer and the cladding layer on top of the stack were modeled analytically using equation (2). Likewise, the interfaces were calculated using the interface S-matrix Si [34],

Si=(2n1n1+n20n1n2n1+n2002n1n1+n20n1n2n1+n2n1n2n1+n202n2n1+n200n1n2n1+n202n2n1+n2),
where n1 and n2 are the refractive indices of two adjacent media when propagating from medium 1 to 2. The stack’s S-matrix Sst is therefore composed by the cascaded starproduct
Sst=SisSlSspSuScSic,
where Sc and Ssp denote the S-matrices of cladding and spacer and Sic and Sis their corresponding interfaces. The metasurfaces’ S-matrices are represented by Sl and Su, referring to the upper and lower metasurface, respectively.

 figure: Fig. 5

Fig. 5 (a)-(b) Linear asymmetric transmission (a) and ellipticity (b) derived from measurement (solid lines) and SASA results (dashed lines). Red indicates x-polarized and blue y-polarized incident light. (c)-(e) Asymmetric transmission and ellipticity scanned over the spacer thickness from 40 nm to 1000 nm. (c) Asymmetric transmission for x-polarization (y-polarization as its negative counterpart was omitted); (d) and (e) display ellipticity for x- and y-polarization. The red line in each surface plot marks a spacer thickness of dsp= 345 nm, corresponding to solid lines in (a) and (b).

Download Full Size | PDF

For quantitative modeling, the S-matrix coefficients of both metasurfaces Sl and Su were determined numerically as single layers, using a self-implemented Fourier-Modal-Method (FMM) [45, 46] and the geometric parameters from SEM measurement. Based on a beforehand done convergence test the FMM was truncated at 15

Fourier orders (in x- and y-direction). In case of the upper metasurface the nano-wires were rotated individually in their unit cell in adaption to the fabricated sample. To account for the corner rounding of the fabricated structures a curvature radius of 25 nm was assumed for the corners of our model particles. As material parameters we used measured ellipsometric data of evaporated gold and spin-on glass as they result from our fabrication process [47].

In addition to SASA a rigorous FMM of the full stack was performed to compare and determined the correctness of the semi-analytical model. Since a rigorous FMM takes all evanescent and propagating orders into account up to a set truncation limit (here: 15 Fourier orders) [41, 46], a direct comparison to SASA will help to indicate the validity of the FMA. As a side note, the computation times of the rigorous FMM and SASA are compared in the appendix.

The resulting transmittance and phase plots of both methods are shown together with the experimental results in Fig. 3. SASA and FMM results coincide almost perfectly both in transmittance and phase, indicating the validity of the FMA. Notably, all the main features of the measured curves were reproduced and the overall agreement between experimental and theoretical results is very good.

Next, in order to analyze the polarization behavior of the stack we derived its asymmetric transmission Δx, eq. (6), and ellipticity ϵ in forward transmission. For incident x-polarization (analogously for y-polarization) ϵx is defined as the ratio of the major and minor half axis of the polarization ellipse,

ϵx=2|Txxf||Tyxf||Txxf|2+|Tyxf|2sin δx.
The term δx=arg(Txxf)arg(Tyxf) represents the phase difference between the elements of the outgoing field. If ϵx=±1, outgoing light is circularly polarized and if ϵx=0, it is linearly polarized. Values in between give elliptic polarization to varying degree.

We note that asymmetric transmission, Fig. 5(a), is strongest in the spectral range of the peak around λ0= 950 nm, reaching up to ± 12 %. In the same spectral region, we see a strong conversion from x-polarized to elliptically polarized light close to being circular, Fig. 5(b). The ellipticity for incident y-polarized light is more shallow and does not go beyond ±0.5. This is explained by the presence of asymmetric transmission giving light a different twist depending on the incident polarization.

To further clarify the role of the spacer for asymmetric transmission we varied its thickness from 40 nm to 1000 nm in 1 nm steps using SASA. Applying the associativity of the starproduct [41], such that

S(dsp)=(SisSl)Ssp(dsp)((SuSc)Sic),
we have to calculate the complex parts of the stack only once and can vary dsp separately[48], giving it an advantage over rigorous methods where the entire stack would have to be recalculated for each variation.

This enabled us to determine Δx and ϵx/y as a function of dsp, as shown in Fig. 5(c)-(e). The resulting Fabry-Perot type patterns show that the polarization characteristic of this system is highly sensitive with respect to the spacer thickness. Both the emerging polarization states and the dispersion of asymmetric transmission change as the Fabry-Perot condition changes with the spacer thickness. Moreover, the SASA result reveals that the ellipticity in y-polarization is not deviating from its flat characteristic from 1000 nm to longer wavelengths for different spacer thicknesses. We can deduce that in this spectral region y-polarized light is converted to a slightly elliptical state mostly determined by the geometry of the stack. However, it should also be noted that in this particular region both cross-polarization components of the stack almost vanish.

Our analysis of the stack shows that the resulting asymmetric transmission Δx is sensitive to the chosen spacer thickness dsp. Indeed, a different choice of dsp could almost result in Δx vanishing. Thus, when designing chiral stacks of this type it is not sufficient to define the geometry of the structure but necessary to also optimize the separation between the metasurfaces. In conclusion, the spacer thickness of complex metasurface stacks represents another degree of freedom in the design of sophisticated layered media. Besides simply controlling the occurrence of Fabry-Perot resonances it can be utilized to enhance otherwise negligible polarization effects. It is also imaginable to overlap Fabry-Perot resonances with other resonant effects to modify a stack’s response for different functionalities.

4. Summary

Using a scattering matrix based approach we derived the polarization behavior of a twisted metamaterial from basic symmetry considerations. The twisted metamaterial consisted of two stacked gold nano-wire metasurfaces with a dielectric spacer in between, where the wires of the upper metasurface were rotated with respect to the lower ones. At first, we derived the stack’s polarization properties without specifying its constituent materials or their exact geometric measures, demonstrating how to construct polarization effects from general symmetry considerations. Viewing the stack layer by layer we could show analytically that it exhibits asymmetric transmission even though the symmetry of each layer alone could not produce this effect. The predicted polarization behavior was realized in two fabricated variants of the modeled stack with different spacer thickness. One targeted the regime of near-field coupling between the metasurfaces and the other coupling in the far-field. The latter had the advantage of being accessible by fast semi-analytic modeling, whereas the former was used to showcase the difference and similarities between the two cases. Our results showed almost perfect agreement between full-wave simulations and our semi-analytic stacking algorithm, and a very good match to measurements. It was thus possible to numerically investigate the physical role of the spacer, revealing a Fabry-Perot-type modal resonance which can be spectrally tailored. This showed a distinct influence of the spacer height on the polarization characteristics of the stack. With this we demonstrated the important role of spacers in the design of stacked metamaterials which are often considered to be nothing but placeholder between functional layers. With the ability of using the spacer thickness as a free design parameter far-field coupled stacks of arbitrary metasurfaces represent ideal candidates for easily designable sophisticated layered media.

Appendix

Computation time

Since SASA computes purely algebraic equations, it took only about 8×105 s to simulate the stack for a single wavelength point on a standard laptop (2.7 GHz Sandy Bridge Intel Core i7 (2620M) and 16 GB 1333 MHz DDR3 memory). Moreover, the FMM for the single layer metasurfaces took about 790 s per wavelength point on a cluster node (using 2 threads on a Harpertown Intel Xeon-L5420 processor and 16 GB DDR memory). On the other hand, the rigorous FMM of the entire stack took about 1.7×103 s per wavelength using the same cluster nodes. Adding the time for the algebraic calculations and the single layer FMMs together and taking into account that simulations on a cluster can be run in parallel the semi-analytic approach is twice as fast for a single calculation as the rigorous FMM. When varying stack parameters such as spacer thickness, the order of layers or their orientation, the saved computation time by using SASA increases significantly. Nevertheless, a proper benchmark test would be needed for a thorough comparison of computation times.

Stack S-matrix coefficients

The explicit expressions for the coefficients of equation (7) are as follows,

T¯xx=PTx(P2Ry(R˜xyT˜xy+R˜yyT˜xx)-T˜xx)P4RxRy(R˜xy2-R˜xxR˜yy)+P2(RxR˜xx+RyR˜yy)-1T¯xy=PTy(P2Rx(R˜xyT˜xx+R˜xxT˜xy)-T˜xy)P4RxRy(R˜xy2-R˜xxR˜yy)+P2(RxR˜xx+RyR˜yy)-1T¯yx=PTx(P2Ry(R˜xyT˜yy+R˜yyT˜xy)-T˜xy)P4RxRy(R˜xy2-R˜xxR˜yy)+P2(RxR˜xx+RyR˜yy)-1T¯yy=PTy(P2Rx(R˜xyT˜xy+R˜xxT˜yy)-T˜yy)P4RxRy(R˜xy2-R˜xxR˜yy)+P2(RxR˜xx+RyR˜yy)-1R¯xxb=(P4RxRy(2R˜xyT˜xxT˜xy-R˜xxT˜xy2-R˜yyT˜xx2-R˜xx(R˜xy2-R˜xxR˜yy))+P2(RxT˜xx2+RyT˜xy2-R˜xx(RxR˜xx+RyR˜yy))+R˜xx)/(P4RxRy(R˜xxR˜yy-R˜xy2)-P2(RxR˜xx+RyR˜yy)+1)R¯xyb=(P4RxRy(R˜xy(R˜xxR˜yy+T˜xy2-R˜xy2+T˜xxT˜yy)-T˜xy(R˜yyT˜xx+R˜xxT˜yy))+P2(RxT˜xxT˜xy+RyT˜xyT˜yy-R˜xy(RxR˜xx+RyR˜yy))+R˜xy)/(P4RxRy(R˜xxR˜yy-R˜xy2)-P2(RxR˜xx+RyR˜yy)+1)R¯yyb=(P4(-Rx)Ry(R˜yy(-R˜xxR˜yy+R˜xy2+T˜xy2)+R˜xxT˜yy2-2R˜xyT˜xyT˜yy)+P2(RxT˜xy2+RyT˜yy2-R˜yy(RxR˜xx+RyR˜yy))+R˜yy)/(P4RxRy(R˜xxR˜yy-R˜xy2)-P2(RxR˜xx+RyR˜yy)+1)R¯xxf=Rx2-Tx2Rx+Tx2(1-P2RyR˜yy)Rx(P4RxRy(R˜xxR˜yy-R˜xy2)-P2(RxR˜xx+RyR˜yy)+1)R¯xyf=-P2TxTyR˜xyP4RxRy(R˜xy2-R˜xxR˜yy)+P2(RxR˜xx+RyR˜yy)-1R¯yyf=Ry2-Ty2Ry+Ty2(1-P2RxR˜xx)Ry(P4RxRy(R˜xxR˜yy-R˜xy2)-P2(RxR˜xx+RyR˜yy)+1).

Funding

Bundesministerium für Bildung und Forschung (03ZZ0466); Deutsche Forschungsgemeinschaft (STA 1426/1-1,PE 1524/10-1).

Acknowledgments

We would like to thank Michael Steinert for his support on SEM-imaging and FIB milling as well as Waltraud Gräf, Holger Schmidt, and Daniel Voigt for support with fabrication. Furthermore, we gratefully acknowledge financial support by the German Federal Ministry of Education and Research in the program Zwanzig20 ― Partnership for Innovation as part of the research alliance 3Dsensation (grant number 03ZZ0466) as well as the German Research Foundation through the Priority Program SPP 1839 “Tailored Disorder” (STA 1426/1-1 and PE 1524/10-1).

References

1. A. Sihvola, “Metamaterials in electromagnetics,” Metamaterials 1, 2–11 (2007). [CrossRef]  

2. C. Menzel, C. Rockstuhl, and F. Lederer, “Advanced Jones calculus for the classification of periodic metamaterials,” Physical Review A - Atomic, Molecular, and Optical Physics 82, 1–9 (2010). [CrossRef]  

3. C. L. Holloway, E. F. Kuester, J. A. Gordon, J. O’Hara, J. Booth, and D. R. Smith, “An overview of the theory and applications of metasurfaces: the two-dimensional equivalents of metamaterials,” IEEE Antennas and Propagation Magazine 54, 10–35 (2012). [CrossRef]  

4. I. Staude, A. E. Miroshnichenko, M. Decker, N. T. Fofang, S. Liu, E. Gonzales, J. Dominguez, T. S. Luk, D. N. Neshev, I. Brener, and Y. Kivshar, “Tailoring directional scattering through magnetic and electric resonances in subwavelength silicon nanodisks,” ACS Nano 7, 7824–7832 (2013). [CrossRef]   [PubMed]  

5. M. Decker, I. Staude, M. Falkner, J. Dominguez, D. N. Neshev, I. Brener, T. Pertsch, and Y. S. Kivshar, “High-efficiency dielectric Huygens’ surfaces,” Advanced Optical Materials 3, 813–820 (2015). [CrossRef]  

6. N. Yu and F. Capasso, “Flat optics with designer metasurfaces,” Nature Materials 13, 139–150 (2014). [CrossRef]   [PubMed]  

7. P. Genevet, F. Capasso, F. Aieta, M. Khorasaninejad, and R. Devlin, “Recent advances in planar optics: from plasmonic to dielectric metasurfaces,” Optica 4, 139 (2017). [CrossRef]  

8. C. Menzel, C. Helgert, C. Rockstuhl, E. B. Kley, A. Tünnermann, T. Pertsch, and F. Lederer, “Asymmetric transmission of linearly polarized light at optical metamaterials,” Physical Review Letters 104, 1–4 (2010). [CrossRef]  

9. E. Plum, V. A. Fedotov, and N. I. Zheludev, “Asymmetric transmission: a generic property of two-dimensional periodic patterns,” J. Opt. 13, 024006 (2011). [CrossRef]  

10. S. Fang, K. Luan, H. F. Ma, W. Lv, Y. Li, Z. Zhu, C. Guan, J. Shi, and T. J. Cui, “Asymmetric transmission of linearly polarized waves in terahertz chiral metamaterials,” Journal of Applied Physics 121, 033103 (2017). [CrossRef]  

11. K. Dietrich, D. Lehr, C. Helgert, A. Tünnermann, and E.-B. Kley, “Circular dichroism from chiral nanomaterial fabricated by on-edge lithography,” Advanced Materials 24, OP321–OP325 (2012). [CrossRef]   [PubMed]  

12. Z. Wang, B. H. Teh, Y. Wang, G. Adamo, J. Teng, and H. Sun, “Enhancing circular dichroism by super chiral hot spots from a chiral metasurface with apexes,” Applied Physics Letters 110, 221108 (2017). [CrossRef]  

13. E. Plum, V. A. Fedotov, and N. I. Zheludev, “Optical activity in extrinsically chiral metamaterial,” Applied Physics Letters 93, 1–4 (2008). [CrossRef]  

14. M. V. Gorkunov, A. A. Ezhov, V. V. Artemov, O. Y. Rogov, and S. G. Yudin, “Extreme optical activity and circular dichroism of chiral metal hole arrays,” Applied Physics Letters 104, 8–12 (2014). [CrossRef]  

15. M. Hentschel, M. Schäferling, X. Duan, H. Giessen, and N. Liu, “Chiral plasmonics,” Science Advances 3, e1602735 (2017). [CrossRef]   [PubMed]  

16. M. Thiel, M. Decker, M. Deubel, M. Wegener, S. Linden, and G. von Freymann, “Polarization stop bands in chiral polymeric three-dimensional photonic crystals,” Advanced Materials 19, 207–210 (2007). [CrossRef]  

17. C. Helgert, E. Pshenay-Severin, M. Falkner, C. Menzel, C. Rockstuhl, E.-B. Kley, A. Tünnermann, F. Lederer, and T. Pertsch, “Chiral metamaterial composed of three-dimensional plasmonic nanostructures,” Nano Letters 11, 4400–4404 (2011). [CrossRef]   [PubMed]  

18. J. Kaschke, J. K. Gansel, and M. Wegener, “On metamaterial circular polarizers based on metal N-helices,” Optics Express 20, 26012 (2012). [CrossRef]   [PubMed]  

19. M. Schäferling, D. Dregely, M. Hentschel, and H. Giessen, “Tailoring enhanced optical chirality: design principles for chiral plasmonic nanostructures,” Physical Review X 2, 031010 (2012). [CrossRef]  

20. K. Konishi, B. Bai, Y. Toya, J. Turunen, Y. P. Svirko, and M. Kuwata-Gonokami, “Surface-plasmon enhanced optical activity in two-dimensional metal chiral networks,” Optics Letters 37, 4446 (2012). [CrossRef]   [PubMed]  

21. E. Plum and N. I. Zheludev, “Chiral mirrors,” Applied Physics Letters 106, 221901 (2015). [CrossRef]  

22. L. Cong, N. Xu, W. Zhang, and R. Singh, “Polarization control in terahertz metasurfaces with the lowest order rotational symmetry,” Advanced Optical Materials 3, 1176–1183 (2015). [CrossRef]  

23. G. Kenanakis, R. Zhao, A. Stavrinidis, G. Konstantinidis, N. Katsarakis, M. Kafesaki, C. M. Soukoulis, and E. N. Economou, “Flexible chiral metamaterials in the terahertz regime: a comparative study of various designs,” Optical Materials Express 2, 1702 (2012). [CrossRef]  

24. C. Pfeiffer, C. Zhang, V. Ray, L. J. Guo, and A. Grbic, “High performance bianisotropic metasurfaces: Asymmetric transmission of light,” Physical Review Letters 113, 1–5 (2014). [CrossRef]  

25. J. H. Shi, H. F. Ma, C. Y. Guan, Z. P. Wang, and T. J. Cui, “Broadband chirality and asymmetric transmission in ultrathin 90°-twisted Babinet-inverted metasurfaces,” Physical Review B - Condensed Matter and Materials Physics 89, 1–7 (2014). [CrossRef]  

26. Y. Zhao, J. Shi, L. Sun, X. Li, and A. Alù, “Alignment-free three-dimensional optical metamaterials,” Advanced Materials 26, 1439–1445 (2014). [CrossRef]  

27. J.-G. Yun, S.-J. Kim, H. Yun, K. Lee, J. Sung, J. Kim, Y. Lee, and B. Lee, “Broadband ultrathin circular polarizer at visible and near-infrared wavelengths using a non-resonant characteristic in helically stacked nano-gratings,” Optics Express 25, 14260 (2017). [CrossRef]   [PubMed]  

28. M. Decker, M. Ruther, C. E. Kriegler, J. Zhou, C. M. Soukoulis, S. Linden, and M. Wegener, “Strong optical activity from twisted-cross photonic metamaterials,” Opt. Lett. 34, 2501–2503 (2009). [CrossRef]   [PubMed]  

29. K. Hannam, D. A. Powell, I. V. Shadrivov, and Y. S. Kivshar, “Broadband chiral metamaterials with large optical activity,” Physical Review B - Condensed Matter and Materials Physics 89, 1–6 (2014). [CrossRef]  

30. Z. Wu and Y. Zheng, “Moiré Chiral Metamaterials,” Advanced Optical Materials 1700034, 1700034 (2017). [CrossRef]  

31. Y. Zhao, M. Belkin, and A. Alù, “Twisted optical metamaterials for planarized ultrathin broadband circular polarizers,” Nature Communications 3, 870 (2012). [CrossRef]   [PubMed]  

32. C. R. Simovski, “Bloch material parameters of magneto-dielectric metamaterials and the concept of Bloch lattices,” Metamaterials 1, 62–80 (2007). [CrossRef]  

33. C. R. Simovski and S. A. Tretyakov, “Local constitutive parameters of metamaterials from an effective-medium perspective,” Physical Review B 75, 195111 (2007). [CrossRef]  

34. C. Menzel, J. Sperrhake, and T. Pertsch, “Efficient treatment of stacked metasurfaces for optimizing and enhancing the range of accessible optical functionalities,” Physical Review A 93, 063832 (2016). [CrossRef]  

35. A. Andryieuski, C. Menzel, C. Rockstuhl, R. Malureanu, F. Lederer, and A. Lavrinenko, “Homogenization of resonant chiral metamaterials,” Physical Review B 82, 235107 (2010). [CrossRef]  

36. T. Paul, C. Menzel, W. Śmigaj, C. Rockstuhl, P. Lalanne, and F. Lederer, “Reflection and transmission of light at periodic layered metamaterial films,” Physical Review B 84, 115142 (2011). [CrossRef]  

37. Y. Zhao, N. Engheta, and A. Alù, “Homogenization of plasmonic metasurfaces modeled as transmission-line loads,” Metamaterials 5, 90–96 (2011). [CrossRef]  

38. J. Petschulat, A. Chipouline, A. Tünnermann, T. Pertsch, C. Menzel, C. Rockstuhl, T. Paul, and F. Lederer, “Simple and versatile analytical approach for planar metamaterials,” Physical Review B 82, 075102 (2010). [CrossRef]  

39. N. A. Gippius, T. Weiss, S. G. Tikhodeev, and H. Giessen, “Resonant mode coupling of optical resonances in stacked nanostructures,” Optics Express 18, 7569 (2010). [CrossRef]   [PubMed]  

40. R. M. Redheffer, “On a certain linear fractional transformation,” Journal of Mathematics and Physics 39, 269–286 (1960). [CrossRef]  

41. L. Li, “Formulation and comparison of two recursive matrix algorithms for modeling layered diffraction gratings,” Journal of the Optical Society of America A 13, 1024–1035 (1996). [CrossRef]  

42. R. J. Potton, “Reciprocity in optics,” Reports on Progress in Physics 67, 717–754 (2004). [CrossRef]  

43. N. A. Gippius, S. G. Tikhodeev, and T. Ishihara, “Optical properties of photonic crystal slabs with an asymmetrical unit cell,” Physical Review B 72, 045138 (2005). [CrossRef]  

44. E. Pshenay-Severin, M. Falkner, C. Helgert, and T. Pertsch, “Ultra broadband phase measurements on nanostructured metasurfaces,” Applied Physics Letters 104, 8–13 (2014). [CrossRef]  

45. L. Li, “New formulation of the Fourier modal method for crossed surface-relief gratings,” Journal of the Optical Society of America A 14, 2758 (1997). [CrossRef]  

46. E. Noponen and J. Turunen, “Eigenmode method for electromagnetic synthesis of diffractive elements with three-dimensional profiles,” Journal of the Optical Society of America A 11, 2494 (1994). [CrossRef]  

47. D. Lehr, R. Alaee, R. Filter, K. Dietrich, T. Siefke, C. Rockstuhl, F. Lederer, E. B. Kley, and A. Tünnermann, “Plasmonic nanoring fabrication tuned to pitch: efficient, deterministic, and large scale realization of ultra-small gaps for next generation plasmonic devices,” Applied Physics Letters 105, 143110 (2014). [CrossRef]  

48. T. Weiss, N. A. Gippius, S. G. Tikhodeev, G. Granet, and H. Giessen, “Efficient calculation of the optical properties of stacked metamaterials with a Fourier modal method,” Journal of Optics A: Pure and Applied Optics 11, 114019 (2009). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 Rendered sketch of a stack of gold nano-wire metasurfaces embedded in glass.
Fig. 2
Fig. 2 (a) SEM picture of a FIB cut revealing the metasurface layers of the stack. The nano-particles in this image were colored golden during postprocessing to enhance visibility. (b) SEM image of the lower layer showing axis-aligned nano-wires. (c) SEM image of the upper layer showing nano-wires rotated counterclockwise by 60 .
Fig. 3
Fig. 3 Experimental and simulation results of the stack in Jones matrix form: a) T xx, b) T xy, c) T yx, d) T yy. Blue curves show transmittance and orange ones phase. Measured data is represented by solid and dashed lines showing results for the 345 nm stack and 40 nm stack, respectively. Results obtained via SASA are marked with crosses and those from rigorous FMM calculations with circles.
Fig. 4
Fig. 4 Sketch of the stack construction with S-matrices using eq. (9). Dashed lines mark the interfaces of the stack to the surrounding medium. Arrows point in the direction of (forward) propagation along the z-axis. While spacer (spac.), and cladding (clad.) have defined thicknesses, the two metasurface layers (MS   l, MS   u) are assumed to be infinitely thin.
Fig. 5
Fig. 5 (a)-(b) Linear asymmetric transmission (a) and ellipticity (b) derived from measurement (solid lines) and SASA results (dashed lines). Red indicates x-polarized and blue y-polarized incident light. (c)-(e) Asymmetric transmission and ellipticity scanned over the spacer thickness from 40 nm to 1000 nm. (c) Asymmetric transmission for x-polarization (y-polarization as its negative counterpart was omitted); (d) and (e) display ellipticity for x- and y-polarization. The red line in each surface plot marks a spacer thickness of d sp = 345 nm, corresponding to solid lines in (a) and (b).

Equations (12)

Equations on this page are rendered with MathJax. Learn more.

S = ( T ^ f R ^ b R ^ f T ^ b ) .
S sp = diag ( P , P , P , P ) ,
S S sp = ( P T xx f P T xy f P 2 R xx b P 2 R xy b P T yx f P T yy f P 2 R yx b P 2 R yy b R xx f R xy f P T xx b P T xy b R yx f R yy f P T yx b P T yy b ) .
S w = ( T x 0 R x 0 0 T y 0 R y R x 0 T x 0 0 R y 0 T y ) ,
S ˜ w = ( T ˜ xx T ˜ xy R ˜ xx R ˜ xy T ˜ xy T ˜ yy R ˜ xy R ˜ yy R ˜ xx R ˜ xy T ˜ xx T ˜ xy R ˜ xy R ˜ yy T ˜ xy T ˜ yy ) .
Δ x = | T xy f | 2 | T yx f | 2 .
S ¯ = S w * S Sp * S ˜ w = ( T ¯ xx T ¯ xy R ¯ xx b R ¯ xy b T ¯ yx T ¯ yy R ¯ xy b R ¯ yy b R ¯ xx f R ¯ xy f T ¯ xx T ¯ yx R ¯ xy f R ¯ yy f T ¯ xy T ¯ yy )
S i = ( 2 n 1 n 1 + n 2 0 n 1 n 2 n 1 + n 2 0 0 2 n 1 n 1 + n 2 0 n 1 n 2 n 1 + n 2 n 1 n 2 n 1 + n 2 0 2 n 2 n 1 + n 2 0 0 n 1 n 2 n 1 + n 2 0 2 n 2 n 1 + n 2 ) ,
S st = S is S l S sp S u S c S ic ,
ϵ x = 2 | T xx f | | T yx f | | T xx f | 2 + | T yx f | 2 sin  δ x .
S ( d sp ) = ( S is S l ) S sp ( d sp ) ( ( S u S c ) S ic ) ,
T ¯ x x = P T x ( P 2 R y ( R ˜ x y T ˜ x y + R ˜ y y T ˜ x x ) - T ˜ x x ) P 4 R x R y ( R ˜ x y 2 - R ˜ x x R ˜ y y ) + P 2 ( R x R ˜ x x + R y R ˜ y y ) - 1 T ¯ x y = P T y ( P 2 R x ( R ˜ x y T ˜ x x + R ˜ x x T ˜ x y ) - T ˜ x y ) P 4 R x R y ( R ˜ x y 2 - R ˜ x x R ˜ y y ) + P 2 ( R x R ˜ x x + R y R ˜ y y ) - 1 T ¯ y x = P T x ( P 2 R y ( R ˜ x y T ˜ y y + R ˜ y y T ˜ x y ) - T ˜ x y ) P 4 R x R y ( R ˜ x y 2 - R ˜ x x R ˜ y y ) + P 2 ( R x R ˜ x x + R y R ˜ y y ) - 1 T ¯ y y = P T y ( P 2 R x ( R ˜ x y T ˜ x y + R ˜ x x T ˜ y y ) - T ˜ y y ) P 4 R x R y ( R ˜ x y 2 - R ˜ x x R ˜ y y ) + P 2 ( R x R ˜ x x + R y R ˜ y y ) - 1 R ¯ x x b = ( P 4 R x R y ( 2 R ˜ x y T ˜ x x T ˜ x y - R ˜ x x T ˜ x y 2 - R ˜ y y T ˜ x x 2 - R ˜ x x ( R ˜ x y 2 - R ˜ x x R ˜ y y ) ) + P 2 ( R x T ˜ x x 2 + R y T ˜ x y 2 - R ˜ x x ( R x R ˜ x x + R y R ˜ y y ) ) + R ˜ x x ) / ( P 4 R x R y ( R ˜ x x R ˜ y y - R ˜ x y 2 ) - P 2 ( R x R ˜ x x + R y R ˜ y y ) + 1 ) R ¯ x y b = ( P 4 R x R y ( R ˜ x y ( R ˜ x x R ˜ y y + T ˜ x y 2 - R ˜ x y 2 + T ˜ x x T ˜ y y ) - T ˜ x y ( R ˜ y y T ˜ x x + R ˜ x x T ˜ y y ) ) + P 2 ( R x T ˜ x x T ˜ x y + R y T ˜ x y T ˜ y y - R ˜ x y ( R x R ˜ x x + R y R ˜ y y ) ) + R ˜ x y ) / ( P 4 R x R y ( R ˜ x x R ˜ y y - R ˜ x y 2 ) - P 2 ( R x R ˜ x x + R y R ˜ y y ) + 1 ) R ¯ y y b = ( P 4 ( - R x ) R y ( R ˜ y y ( - R ˜ x x R ˜ y y + R ˜ x y 2 + T ˜ x y 2 ) + R ˜ x x T ˜ y y 2 - 2 R ˜ x y T ˜ x y T ˜ y y ) + P 2 ( R x T ˜ x y 2 + R y T ˜ y y 2 - R ˜ y y ( R x R ˜ x x + R y R ˜ y y ) ) + R ˜ y y ) / ( P 4 R x R y ( R ˜ x x R ˜ y y - R ˜ x y 2 ) - P 2 ( R x R ˜ x x + R y R ˜ y y ) + 1 ) R ¯ x x f = R x 2 - T x 2 R x + T x 2 ( 1 - P 2 R y R ˜ y y ) R x ( P 4 R x R y ( R ˜ x x R ˜ y y - R ˜ x y 2 ) - P 2 ( R x R ˜ x x + R y R ˜ y y ) + 1 ) R ¯ x y f = - P 2 T x T y R ˜ x y P 4 R x R y ( R ˜ x y 2 - R ˜ x x R ˜ y y ) + P 2 ( R x R ˜ x x + R y R ˜ y y ) - 1 R ¯ y y f = R y 2 - T y 2 R y + T y 2 ( 1 - P 2 R x R ˜ x x ) R y ( P 4 R x R y ( R ˜ x x R ˜ y y - R ˜ x y 2 ) - P 2 ( R x R ˜ x x + R y R ˜ y y ) + 1 ) .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.