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

Sideways adiabaticity: beyond ray optics for slowly varying metasurfaces

Open Access Open Access

Abstract

Optical metasurfaces (subwavelength-patterned surfaces typically described by variable effective surface impedances) are typically modeled by an approximation akin to ray optics: the reflection or transmission of an incident wave at each point of the surface is computed as if the surface were “locally uniform,” and then the total field is obtained by summing all of these local scattered fields via a Huygens principle. (Similar approximations are found in scalar diffraction theory and in ray optics for curved surfaces.) In this paper, we develop a precise theory of such approximations for variable-impedance surfaces. Not only do we obtain a type of adiabatic theorem showing that the “zeroth-order” locally uniform approximation converges in the limit as the surface varies more and more slowly, including a way to quantify the rate of convergence, but we also obtain an infinite series of higher-order corrections. These corrections, which can be computed to any desired order by performing integral operations on the surface fields, allow rapidly varying surfaces to be modeled with arbitrary accuracy, and also allow one to validate designs based on the zeroth-order approximation (which is often surprisingly accurate) without resorting to expensive brute-force Maxwell solvers. We show that our formulation works arbitrarily close to the surface, and can even compute coupling to guided modes, whereas in the far-field limit our zeroth-order result simplifies to an expression similar to what has been used by other authors.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Optical metasurfaces, subwavelength structures described by an effective sheet impedance [1–10], are now being designed for large-area optical devices using models in which the far-field reflection/transmission coefficients are computed at each point assuming a uniform (or periodic) surface—as explained below, we refer to these as “ray-optics” models [11–14]. This is a good approximation for surfaces (or unit cells) that are varying slowly, a fact that is closely connected to the “adiabatic theorem” [15,16] for waves propagating through slowly varying media. However, although there are countless papers and books on modeling propagation through slowly varying media [15,17,18], exploiting the rate of change as a small parameter ε, the “sideways” problem of scattering off a slowly varying surface (Fig. 1) is relatively unstudied. In this paper, we address the following key questions: how quickly does the ray-optics approximation converge as ε → 0, can we quickly compute the low-order corrections (both to improve accuracy and to validate ray optics), and how do we compute both far-field and near-field scattering (e.g. coupling to guided modes)? A typical metasurface has two scales: the subwavelength scale of the microstructure and the macroscale of the nonuniformity. In this paper, we address corrections due to the macroscale nonuniformity, which we allow to be completely arbitrary, while we follow other authors [1–10] in subsuming the microscale into effective surface impedances.

 figure: Fig. 1

Fig. 1 Scattering problem under consideration. A time-harmonic incident field uinc with angular frequency ω > 0 impinges on a metasurface with surface impedance Z and surface admittance Y, producing a reflected field propagating in the upper half-plane and a transmitted field propagating in the lower half-plane. The metasurface is surrounded by a homogeneous medium with electric permittivity 0 > 0 and magnetic permeability μ0 > 0.

Download Full Size | PDF

In particular, we use the technical machinery of surface-integral equations (SIEs) [19,20] and a “locally uniform approximation” of the metasurface to show that the ray-optics approximation is the far-field zero-th order term in a convergent series (Section 4 and Appendix C), that each successive correction can be computed simply by performing integrals (not by solving any PDE or other system of equations), and that the next-order correction scales as ε2. Moreover, our series allows us to compute the full Green’s function of the surface: the fields in response to arbitrary sources or incident fields, including the near-field terms (fields and/or sources close to the surface). We show that these near fields allow us to compute the coupling of an incident wave to guided modes on the surface [6,21] and that they also appear in the zeroth-order locally uniform approximation. For rapidly varying metasurfaces, such as those designed to reflect light at a very oblique angle [22], we show that even the far-field accuracy is substantially improved by including the next-order correction. Perhaps more importantly, the ability to compute the next-order correction provides a way to validate a ray-optics design for very large-area metasurfaces, where brute-force Maxwell simulations are impractical and there was previously no way to evaluate the ray-optics accuracy short of a laboratory experiment.

Since typical metasurface designs lead to large computational domains (often hundreds of wavelengths [23]) that are intractable by standard simulation techniques, e.g. finite-difference and finite-element methods, previous work on metasurfaces has made extensive use of numerical simulations based on ray-optics approximations. In particular, authors typically compute reflection/transmission coefficients for periodic surfaces with a variety of unit cells, they assume that these coefficients remain accurate even for an aperiodic surface where each unit cell is different, and then they select the unit cell at each point on the surface to achieve a desired optical functionality [10,24–30]. (For subwavelength unit cells where there is only a single “specular” reflected/transmitted wave, the reflection/transmission coefficients can also be fitted to an effective sheet impedance, giving a “homogenized” effective medium at each point [10,31–33].) Because these works described the surfaces by a single far-field (planewave) reflection and/or transmission coefficient at each point, they can be thought of as “ray-optics” approximations even if they were expressed in the language of wave optics. (A closely related approximation—curved surfaces treated as locally flat—is called a “tangent-plane” or “Kirchhoff” approximation [34]. Yet another closely related approximation is provided by scalar diffraction theory [35].) Here, we assume that the metasurface is subwavelength enough to be described as an effective sheet impedance at each point, but we do not only compute the scattering assuming that the impedance is locally uniform: our goal is to take the macro-scale spatial variation (the aperiodicity) explicitly into account by computing correction to the locally uniform approximation. (Potential extensions to slowly varying periodic structures where the micro-scale is treated explicitly, perhaps to include additional diffraction orders for large-period structures, are discussed in Section 5.)

Wave propagation through slowly varying media is usually treated by coupled-mode theory: one expands the wave in the basis of “instantaneous” [36] eigenfunctions of each cross-section [17] or period [16], and then obtains a set of coupled differential equations in the mode coefficients (typically truncated to only a few guided modes). As the medium varies more slowly, the coefficients tend to constants, corresponding to “adiabatic” transport of modes without intermodal scattering [17]. Unfortunately, this approach appears awkward to apply to the problem of scattering off of a slowly varying medium, both because there is a continuum of radiating modes and because one wants to describe the basis of incoming/outgoing planewaves independently of the varying surface in order to connect to a ray-optics (zero-th order) approximation. The fact that an incident planewave is not an eigenfunction of the cross-section at each point means that one cannot simply quote the standard adiabatic theorem to justify the metasurface ray-optics approximation, for example. Another type of technique for approximating the scattering from a weakly perturbed surfaces is a Born approximation [18, 37] (also known as a volume-current method [38], Kirchhoff approximation [34], etc.), which handles perturbations like surface roughness that are small in amplitude but not necessarily slowly varying, whereas our goal is to handle variations that are slow but not small.

A surface impedance Z(εx) (defined precisely in Section 2) varies more and more slowly as ε → 0, and our goal is an expansion with terms proportional (in a certain norm) to powers of ε [19,20]. This expansion is achieved through an SIE. In particular, since the media above and below the surface are homogeneous, we express the problem in terms of an SIE in which the unknowns reside only on the surface. Our derivations start with an approximate Green’s function Gp that is a building-block of the locally uniform approach (Section 4), and then we insert this into an exact SIE, obtained by enforcing to transition conditions on the metasurface, to derive a series of corrections (Appendix C)—like a Born–Dyson series [39], the corrections are expressed in terms of integrals involving Gp. These integrals must be computed numerically by a “quadrature” technique [40] (Appendix G), but such computations are simple summations on a computer that are far easier than solving the large systems of equations arising in brute-force computational methods, and also have the advantage of parallelizing perfectly (fields at different points can be computed completely independently). Truncating to the zeroth-order term in the series does not correspond to setting ε = 0 (a uniform surface), so even the lowest order locally uniform approximation captures to a certain extent the surface variation. In the far field, Gp simplifies to an expression Gpff that can be written in closed form (eliminating an integral), recovering the usual ray-optics approximation at zero-th order (Appendix F). Using this approach, we demonstrate through numerical experiments that the ray-optics approximation (i.e., the far-field of the locally uniform approximation) produces far-field errors that vanish as ε2, and more generally as ε2N+2 if we include Nth-order corrections (Fig. 5). In the presence of guided modes, which correspond to poles that appear in Gp at certain wavevectors [37], we show that this also simplifies the integrals in our perturbative expansion (via a steepest-descent approximation) if one is mainly interested in coupling to guided modes (Appendix F and Figs. 6 and 10).

2. Problem formulation

We consider a metasurface Γ in two spatial dimensions that divides the xy plane into two unbounded half-planes, Ω+ = {y > 0} and Ω = {y < 0}, occupying the regions above and below Γ, respectively. The media Ω+ and Ω surrounding the metasurface are assumed to be homogeneous with electric permittivity and magnetic permeability denoted by 0 > 0 and μ0 > 0, respectively (Fig. 1). The metasurface is characterized by the so-called generalized sheet transition conditions [1]:

Zn×(H+H)=12(E++E),Yn×(E+E)=12(H++H),
which for the sake of presentation simplicity are assumed to be given in terms of scalar quantities corresponding to surface impedance Z and the surface admittance Y [7]. The symbol n in Eq. (1) denotes the unit normal vector to Γ = {y = 0} pointing upwards, and E± and H± (resp. E± and H±) denote the tangential (resp. entire) fields at y = 0±.

It thus follows from Maxwell’s equations that in Ez polarization the total electromagnetic field (E, H) is given by E = (0, 0, E) and H = (H1, H2, 0), and can be obtained from the z-component of the electric field by means of the relations:

2E+k2E=0,H1=1ikηEy,H2=1ikηEx,
where k=ω0μ0 is the wavenumber of the surrounding media, ω > 0 is the angular frequency, and η=μ0/0 is the intrinsic free-space impedance. Similarly, in Hz polarization it holds that E = (E1, E2, 0) and H = (0, 0, H) where
2H+k2H=0,E1=ηikHy,E2=ηikHx.

Relations in Eqs. (2) and (3), on the other hand, yield that the transition conditions in Eq. (1) can be equivalently expressed as

[[Ey]]=ikη2Z{{E}}and{{Ey}}=2ikηY[[E]],
in Ez polarization, and
[[Hy]]=ik2Yη{{H}}and{{Hy}}=2ikZη[[H]],
in Hz polarization, where the notations
[[u]]=u+uand{{u}}=u++u,
have been introduced to refer to the jump and the sum of a scalar field u across Γ, where u+ (resp. u) denotes the limit value of u on Γ from Ω+ (resp. Ω).

In order to treat both Ez and Hz-polarization cases, we define the metasurface parameters α and β as

α=η2Zandβ=2ηY
in Ez polarization, and by
α=12Yηandβ=2Zη
in Hz polarization. Throughout this paper we assume that α and β are continuous complex-valued functions that satisfy Re α ≥ 0 and Re β ≥ 0, which correspond to assuming that both the surface impedance Z and the surface admittance Y are passive but not necessarily lossless. We will eventually consider these quantities to be slowly varying functions of the form α (x) = a(εx) and β(x) = b(εx), where ε > 0 is a small parameter.

Letting utot denote either the total electric field E in Ez polarization or the total magnetic field H in Hz polarization, it follows from Eqs. (4), (5) and (6) that the transition conditions can be equivalently expressed as

[[uytot]]=ikα{{utot}}and{{uytot}}=ikβ[[utot]]onΓ,
in terms of the metasurface parameters α and β introduced in Eq. (6), where we have used the notation uy = ∂u/∂y.

In this paper we consider the problem of scattering that arise when the metasurface is illuminated by a time-harmonic incident field uinc which is assumed to satisfy the Helmholtz equation ∇2uinc + k2uinc = 0 in Ω+ and Ω (Fig. 1). In order to properly formulate a scattering problem, we proceed to express the total field as utot = uscat + uinc, where uscat denotes the scattered field off of Γ. Replacing utot = uscat + uinc in the Helmholtz equation and the transition conditions in Eq. (7) we obtain that uscat satisfies

2uscat+k2uscat=0inΩ+Ω,
[[uyscat]]=ikα{{uscat}}2ikαuinconΓ,
{{uyscat}}=ikβ[[uscat]]2uyinconΓ.
In order for the problem in Eq. (8) to be a well-posed boundary value problem for uscat, the scattered field has to satisfy a certain radiation condition at infinity [19,41,42]. Such a radiation condition, which roughly speaking means that uscat corresponds to an up-going wave-field in Ω+ and a down-going wave-field in Ω, can be formally stated in terms of the angular spectral representation by requiring the existence of functions (or more generally, distributions) A+ and A such that
uscat(r)=CA±(kx)eikxx+ik2kx2|y|dkxforr=(x,y)Ω±,
where contour C corresponds to the real kx-axis that is suitably dented around the possible poles singularities of A± [43].

3. Exact and approximate surface integral representations

In this section we derive an exact and an approximate integral representation formulae for the scattered field uscat solution of the problem in Eq. (8). Such formulae involve the incident surface currents and the Green’s function of the boundary value problem in Eq. (8) and are given in terms of integrals on the metasurface only,

3.1. Exact integral representation

The Green’s function G of the boundary problem in Eq. (8) can be physically interpreted as the total field produced by a point source excitation placed off of the metasurface [37, 41]. In detail, letting r′ = (x′, y′) denote the location of a point source and r = (x, y) denote an observation point, the Green’s function G(r|r′) can be found by solving the following boundary value problem:

r2G(r|r)+k2G(r|r)=δ(rr),r=(x,y)Ω+Ω,
[[Gy(r|r)]]=ikα(x){{G(r|r)}},rΓ,
{{Gy(r|r)}}=ikβ(x)[[G(r|r)]],rΓ,
with δ denoting the Dirac’s delta distribution and where G is additionally required to satisfy the radiation condition.

As is shown in Appendix A, Green’s third identity together with Eqs. (8) and (9) can be combined to show that the scattered field uscat admits the integral representation

uscat(r)={G(r|s,0+)f+inc(s)G(r|s,0)finc(s)}ds,rΩ+Ω,
where the current densities f±inc in Eq. (10) are given by
f±inc(s)=uyinc(s,0)±ikα(s)uinc(s,0)
in terms of the incident field uinc, and where G(r|s, 0±) = limδ→0+ G(r|s, ±δ).

The integral representation formula in Eq. (10) provides an explicit expression for the scattered field which is valid everywhere (in the near and far fields). It has, however, little practical relevance unless an exact or approximate Green’s function is available. Unfortunately, a formula for the Green’s function in Eq. (9) cannot be easily obtained for “general” spatially varying metasurface parameters α and β, and thus, suitable approximations of G are needed in order to make proper use of Eq. (10) in scattering simulations. In the next section we derive an approximation for G based on ray-optics principles.

3.2. Ray-optics approximation

From the viewpoint of Huygens’ principle (formalized by the principal of equivalence), equation in Eq. (10) represents the scattered field by a source term corresponding to each point along the wavefront incident upon the surface (f±inc) [39]. The typical “ray-optics” approximation is to compute the reflection/transmission at each point x′ as if the surface were uniform in the vicinity of that point. That approach corresponds to approximating the integral expression Eq. (10) by a similar equation, but with the exact Green’s function G replaced by an approximate “proto”-Green’s function Gp defined by the scattering of the source at r′ = (x′, y′) from a uniform surface α(x′) and β(x′). This approximation yields

u0tot(r)=uinc(r)+{Gp(r|s,0+)f+inc(s)Gp(r|s,0)finc(s)}ds,rΩ+Ω,
which which turn out to be our zeroth order approximation in Sec. 4. We call Gp a proto-Green’s function because it is a building-block for our solution, but it is not the Green’s function one would get by putting a point source as the incident field in Eq. (12). We give an exact Green’s function Gp for reflection and transmission off a uniform surface in Appendix B. But in the far field (fields far from the surface), as is derived rigorously in Appendix F, this simplifies to a function Gpff that we present in a more elementary fashion here.

In order to construct Gpff, we consider the scattering configuration depicted in Fig. 2. With reference to that figure, the total wave field observed above the metasurface at a point r = (x, y), y > 0 is given by the superposition Gpff=Ginc+G˜r of the (primary) incident field Ginc(r|r)=i4H0(1)(k|rr|) produced by a point source placed above the metasurface, at r′ = (x′, y′), y′ > 0, and the (secondary) field r resulting from the reflection at r̃′ = (x̃′, 0) (on Γ) of the ray stemming from r′. (The function H0(1) is the Hankel function of first kind and order zero [44].) The magnitude and phase of the reflected field are characterized by the local reflection coefficient R(k cos θ, x̃′) that depends on the reflection angle ϑ = arctan(y/(xx̃′)) (which is measured with respect to tbe metasurface) of the reflected ray at r̃′ (Fig. 2). As is shown in Appendix B, the local reflection coefficient is given by

R(kx,s):=1Γ^α(kx,s)Γ^β(kx,s),
where
Γ^α(kx,s):=kα(s)ky+kα(s),Γ^β(kx,s):=kβ(s)ky+kβ(s)andky=k2kx2.
In order to account for both the magnitude and the direction of the reflected field, we consider the image point source r̄′ = (x′, −y′) ∈ Ω which allows the total field field above the metasurface to be approximated as
Gpff(r|r)=i4H0(1)(k|rr|)incidentfield+i4R(kcosϑ,x˜)H0(1)(k|rr¯|)reflectedfield(rΩ+).

 figure: Fig. 2

Fig. 2 Variables used in the derivation of the approximation in Eq. (16) of the Green’s function of the problem under consideration defined in Eq. (9). The field produced by a point source at r′ impinges on the metasurface at r̃′. The total field is observed at the point r far away from the metasurface that is denoted by Γ. The total field is decomposed in primary and secondary fields. The latter corresponds to the field reflected off of the metasurface, which is approximated by means of the local reflection coefficient R in Eq. (13) assuming specular reflection.

Download Full Size | PDF

Similarly, below the metasurface the total field at a point r ∈ Ω corresponds to the transmitted field which can be approximated as

Gpff(r|r)=i4T(kcosϑ,x˜)H0(1)(k|rr|)(rΩ),
in terms of the local transmission coefficient which is shown in Appendix B to be given by
T(kx,s):=Γ^α(kx,s)+Γ^β(kx,s).
in terms of Γ̂α and Γ̂β defined in Eq. (14).

For a point source placed below the metasurface at a point r′ in Ω, the approximate Green’s function Gpff can be derived from symmetry arguments. A rigorous derivation of Gpff based on asymptotic analysis is presented in Appendix F.

Note that Gpff provides a valid approximation of the exact Green’s function G in the far-field zone, and can be thought of as the approximate field scattered from a single point on the surface. In view of the Huygens’ principle, to get the total field we must add together all of the surface points resulting in what we refer to as the ray-optics approximation:

u0tot,ff(r)=uinc(r)+{Gpff(r|s,0+)f+inc(s)Gpff(r|s,0)finc(s)}ds,rΩ+Ω,
of the total field utot is achieved by replacing G by Gpff in the integral representation formula in Eq. (10), where the functions f±inc are defined in Eq. (11) and where the limits Gpff(r|s,0±)=limδ0+Gpff(r|s,±δ) are obtained by setting
R(k(xx)|rr|,x)andT(k(xx)|rr|,x)
in formulae in Eqs. (15a) and (15b), respectively.

We will show in Section 4 that the expression in Eq. (17) is a simplified version of the one in Eq. (12) in the limit of sources and fields far from the surface and, furthermore, that the expression in Eq. (17) is the zero-th order terms in a convergent series of corrections for slowly varying surfaces.

In order to examine the accuracy of the ray-optics approximation of the total-field (17), we consider a series of numerical examples. Figure 3 presents a comparison of the approximate and “exact” total field solution of the problem of scattering of a planewave that impinges at normal incidence upon three different mesaturfaces. The metasurface parameters α and β were selected so that the transmitted fields satisfy the so-called generalized laws of reflection and transmission [8,28,29]. As is shown in Appendix D, for a given incident planewave uinc in the direction inc = (cos θinc, sin θinc), −π/2 ≤ θinc ≤ 0, metasurface parameters of the form

α(x)=c0(1±eikxd)andβ(x)=c0(1eikxd),
with
c0=sinθincandd=cosθtcosθinc,
produce (to leading-order asymptotics) a transmitted field corresponding to a single planewave in the direction t = (cos θt, sin θt).

 figure: Fig. 3

Fig. 3 First column: Ray-optics approximation (17) of the total field utot = uinc + uscat solution of the problem of scattering of a planewave that impinges at normal incidence on metasurface’s that produce transmitted fields that are predominantly planewaves propagating in the directions θt (a) 45°, (d) 30° and (g) 22.5° with respect to the metasurface. Second column: Full “exact” total field solution of the corresponding for planewave transmitted fields in the directions θt (b) 45°, (e) 30° and (h) 22.2°. Third column: Absolute errors for planewave transmitted fields in the directions θt (c) 45°, (f) 30° and (i) 22.5°.

Download Full Size | PDF

Although the ray-optics approximation (17) seems to capture qualitatively the main features of the scattered field, quantitatively it exhibits large near-field errors—and also large far-field errors in some cases (Fig. 3(i), for example)—that are just one order of magnitude smaller than the scattered field itself. In what follows of this paper we present a methodology to produce both near- and far-field corrections to the ray-optics approximation (17), which turns out to be just the zeroth-order terms of a series approximation of the exact scattered field in the far-field.

4. Locally uniform approximation and corrections to ray-optics

This section presents an SIE formulation of the problem of scattering (8) from which corrections to the ray-optics approximation (17) can be easily obtained in the form of a Born (or Neumann) series [39]. We here follow a standard indirect integral equation formulation procedure [19,20] in which the field is represented by means of a Green’s function Gp that satisfies the Helmholtz equation in both upper and lower homogeneous media, but does not satisfy the correct sheet transition conditions at the metasurface, and we solve for effective source terms that restore the desired transition conditions. Note that, unfortunately, Gpff cannot be used for this purpose because it does not satisfy the Helmholtz equation, since both reflection and transmission coefficients depend on the location of the source and observation points.

Just as in Section 3.2, we begin by constructing a proto-Green’s function Gp, which is given in terms of Fourier-like integrals that we deem as Sommerfeld integrals (due to the similarities they share with layered-media Sommerfeld integrals [37]). This Green’s function possesses two important features. On one hand Gp—as G itself—satisfies the inhomogeneous Helmholtz equation (9a) with a point source excitation and, on the other hand, its far field equals the approximation Gpff of the exact Green’s function G, used in the ray-optics approximation (17). The former allows us to properly derive a second-kind SIE, while the latter guarantees that the zeroth-order approximation obtained by truncation of the Born series solution of the second-kind integral equation does indeed correspond to the ray-optics approximation (17) in the far-field zone. As in Section 3.2, we use the term “proto” because Gp is just a building block and is not equal to the Green’s function of our final zeroth-order approximation.

The key feature of the proto-Green’s function is that, instead of satisfying the non-local transition conditions (9b)(9c), it satisfies the following local transition conditions

[[Gp,y(r|r)]]=ikα(x){{Gp(r|r)}}and{{Gp,y(r|r)}}=ikβ(x)[[Gp(r|r)]],rΓ,
with metasurface parameters α and β depending on the source point r′ = (x′, y′). This local transition conditions formalize the slowly varying assumption used in Section 3.2 where the ray-optics approximation was derived. Since the metasurface parameters α and β that appear in local transition conditions (20) do not depend on the observation point r = (x, y), they can be treated as constants and, thus, an analytical expression for Gp in terms of Sommerfeld integrals can be easily obtained. The idea behind this calculation is to decompose the point-source incident field Ginc as a superposition of both propagative and evanescent planewaves. Since specular reflection takes place for each plane wave impinging on the metasurface, the resulting scattered field can be written down as a superposition of reflected and transmitted plane waves weighted by the reflection and transmission coefficients provided in (13) and (16), respectively. The details of this derivation are presented in Appendix B. Furthermore, it is shown in Appendix F by means of a detailed asymptotic analysis that, to leading asymptotics, Gp equals Gpff as |r| → ∞.

With an analytical expression for Gp in hand (i.e., formulae (42) and (44)) we proceed to derive an SIE for the solution of the scattering problem (8) from which corrections to the ray-optics approximation (17) can be computed. Following the exact integral representation (10) of the scattered field uscat we introduce an indirect integral formulation for the scattering problem (8) by setting

uscat(r)={Gp(r|σ,0+)φ(σ)Gp(r|σ,0)ψ(σ)}dσ,rΩ+Ω,
where φ and ψ are (so far) unknown surface density functions. Note that if Gp were the exact Green’s function G, then uscat in (21) would be the exact solution of (8) provided φ=f+inc and ψ=finc, where f+inc and finc are defined in (11). Note further that in virtue of the relationship between Gp and Gpff established in Appendix F, the substitutions φ=f+inc and ψ=finc in (21) would produce an approximation of uscat that, in the far-field zone, exhibits the same accuracy of the ray-optics approximation (17).

Continuing with the derivation of the SIE, we observe that in order for uscat (21) to be an exact solution of (8) it has to satisfy both the Helmholtz equation (8a) and the transition conditions (8b)(8c). The problem here is that, although uscat (21) does satisfy the Helmholtz equation (8a) in Ω+ and Ω for any admissible densities φ and ψ (since be construction Gp(r|r′), r′ ∈ Γ, satisfies it), it does not necessarily fulfill the correct transition conditions (9b)(9c) unless (φ, ψ) is solution of a certain SIE. Indeed, it is shown in Appendix C that, imposing the transition conditions (9b)(9c) on uscat in (21), an SIE for the unknown density functions (φ, ψ) is obtained. The resulting equations correspond to two decoupled second-kind SIEs:

μjTj[μj]=gjinc,j=1,2,
for the unknown auxiliary densities μj, j = 1, 2, which are directly related to the integral densities in (21) by
φ=μ1+μ22andψ=μ2μ12.
The precise definition of the integral operators Tj, j = 1, 2, is given in Appendix C and the functions gjinc, j = 1, 2, on right-hand-side of (22), are
g1inc(s)=2ikα(s)uinc(s,0)andg2inc(s)=2uyinc(s,0).

Clearly, the densities μj, j = 1, 2, can be determined by solving the SIEs (22) and from them, the desired densities φ and ψ that make uscat in (21) the exact solution of (8) can be readily obtained.

We now recall that we are here interested in slowly varying interface parameters α and β of the form α(x) = a(εx) and β(x) = b(εx) where ε > 0 is a small parameter. In view of definitions (50) and (51), we observe that both integral operators, T1 and T2, vanish as ε → 0 and, therefore, in the limit when ε = 0 the exact SIE solutions are simply μj=gjinc, j = 1, 2. For small but nonzero values of ε, in turn, convergent Neumann-series solutions

μj=n=0Tjngjinc,j=1,2,
of the SIEs (22) can be obtained because the integral operators satisfy ‖Tj< 1 in a certain operator norm for sufficiently small ε.

The Nth-order approximations of the density functions φ and ψ can thus be defined as

φN=μ1(N)+μ2(N)2andψN=μ1(N)μ2(N)2,
where μj(N), j = 1, 2, are the truncated Neumann series
μj(N)(s):=n=0NTjngjinc(s),j=1,2,N0.
From (26) we then define the Nth-order locally uniform approximation of the total near and far fields:
μNtot(r):=uinc(r)+{Gp(r|σ,0+)φN(σ)Gp(r|σ,0)ψN(σ)}dσand
μNtot,ff(r):=uinc(r)+{Gpff(r|σ,0+)φN(σ)Gpff(r|σ,0)ψN(σ)}dσ,
for r ∈ ℝ2 \ Γ, respectively. Note that the zeroth-order term (or any higher order term) in (28) does not correspond to setting ε = 0 (a uniform surface).

Finally, it follows from the definitions above that the ray-optics approximation (17) is simply the zeroth-order approximate far-field u0tot,ff, i.e., the N = 0 instance of the formula (28b). In order to see this it suffices to note that φ0=f+inc and ψ0=finc, which followed directly the definition of f±inc in (11) and the fact that

μ1(0)(s):=g1inc(s)=2ikαuinc(s,0)andμ2(0)(s):=g2inc(s)=2uyinc(s,0).

In summary, the Nth-order approximation of the total near and far fields resulting from the scattering of an incident field uinc off of a metasurface Γ, can obtained as follows:

  1. Evaluate the input data gjinc, j = 1, 2, defined in (24), using the prescribed incident field uinc.
  2. Compute the Nth-order approximate densities μj(N), j = 1, 2, defined in (27), by repeated application of the integral operators Tj to gjinc.
  3. Evaluate the approximate densities φN and ψN, defined in (26), by taking suitable linear combinations of μj(N), j = 1, 2, obtained in step 2.
  4. To produce the approximate near (resp. far) field, substitute the densities φN and ψN obtained in step 3 in the integral formula (28a) (resp. (28b)).

In order to illustrate the accuracy yielded by the higher-order corrections to the ray-optics (zeroth-order) approximation, we present Fig. 4 which concerns the scattering configuration considered above in Figs. 3(g)–(i), which corresponds to the scattering of a planewave that impinges at normal incidence on a metasurface that renders a transmitted planewave with wavevector forming an angle of 22.5° with respect to the metasurface (Fig. 4(a)). Figs. 4(b) and 4(c) display the real part of the total fields (incident + reflected, and transmitted) produced by the zeroth and first order approximations of the far field, corresponding to formula (28b) with N = 0, 1. In order to better visualize the convergence of the locally uniform approximations (28b) as N increases, we present Figs. 4(d)–(h) that display the absolute value of the zeroth, first, second, third and fourth order far-field errors. The reference “exact” far-field was computed by direct solution of the SIE system (22). These results indicate that the far-field error is roughly reduced by a factor of 0.5 as the order increases. This is explained by the fact that the spectral radii of the discrete versions of the integral operators Tj, j = 1, 2, are approximately 0.5. More details on the convergence of the Neumann series approximation are given in Appendix E.

 figure: Fig. 4

Fig. 4 High-order corrections to the ray optics approximation (17). (a) Geometrical configuration of the problem under consideration which corresponds to an unit amplitude planewave impinging at normal incidence on a metasurface that renders a transmitted planewave with wavevector forming an angle of 22.5° with respect to the metasurface. (b) and (c): Real part of the total field 0th and 1st order approximations. (d), (e), (f), (g) and (h): Absolute errors |uuNtot,ff|, for N = 0, 1, 2, 3 and 4, respectively, in the zeroth (no correction), first, second, third and fourth order corrections to the ray optics approximation u0tot,ff. The color scales were adjusted according to the maximum error displayed in each one of the figures.

Download Full Size | PDF

Our next example concerns the dependence of the rate of convergence of the Nth-order approximations (28a) and (28b) on the smoothness of the metasurface parameters. As it turns out, for constant metasurface parameters the zeroth-order near and far field approximations are exact. In this example we thus attempt to quantify how errors depart from zero as the metasurface parameters become non-constant. In order to so we consider slowly-varying metasurface real part of the total field parameters α(x) = a(εx) and β = b(εx)—which depend on a small parameter ε > 0—that tend to constants α0 = a(0) and β0 = b(0) as ε → 0. Figure 5 displays the near-field errors for vanishing values of the smoothness parameter ε > 0. Clearly, the zeroth-, first and third-order approximations exhibit errors of order O(ε2), O(ε4) and O(ε6), respectively, as ε → 0, i.e., as the metasurface parameters tend to constants.

Interestingly, the detailed asymptotic calculations presented in Appendix F also reveal that surface-wave modes appear in the asymptotic expansion of Gp for certain constant values of the metasurface parameters (note that for constant α and β, it holds that G = Gp). Such surface-wave modes are also present in the field scattered by metasurfaces with non-constant metasurface parameters. To demonstrate this fact, we present Fig. 6 which displays the total field solution of the problem of scattering of a Gaussian beam by a metasurface for which a surface-wave mode propagates from left to right along. Three difference solution are displayed in that figure: the exact solution, the ray optics approximation (17), and the zeroth-order locally uniform approximation (28a) with N = 0. Since Gpff is a far-field approximation (which is valid at a certain distance from the metasurface), u0tot,ff does not capture at all the aforementioned surface-wave modes.

 figure: Fig. 5

Fig. 5 Rate of convergence of the zeroth-, first-, second-, third- and fourth-order near-field approximations (28a) in terms of the smoothness of the metasurface parameters. The plot is in log-log scale. The problem under consideration is the scattering of a unit amplitude plane wave at normal incidence that impinges on a metasurface with metasurface parameters α(x) = a(εx) = c0{1 − eiεx} and β(x) = b(εx) = c0{1 − eiεx}. The parameter ε > 0 controls the smoothness of metasurface parameters. The color curves display the maximum of the absolute error |utot(r)uNtot(r)| in the near-field approximations (28a) evaluated at the spatial points r ∈ {−1, 0, 1} × {−10, 10}.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Total field solution of the problem of scattering of a Gaussian beam by a metasurface with interface parameters α and β engineered to allow for surface wave modes (69) to propagate along the metasurface.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Spectral radius of the discretized integral operator T1 as a function of the smoothness of the interface parameter α in Eq. (58). The yellow painted strip indicates the range of parameters ε for which the Neumann series diverges.

Download Full Size | PDF

5. Concluding remarks

We developed an SIE approach, based on a locally uniform approximation of a metasurface, to establish the accuracy and compute higher-order corrections to a ray-optics approximation commonly used in inverse metasurace design, where metasurfaces are modeled by means of slowly varying surface parameters.

This work opens many research directions that could be pursued in the future. The most important (and straightforward, in principle) is perhaps the extension of the proposed approach to three-spatial dimensions. As a practical matter, however, there are many subtle implementation aspects of this extension, such as the derivation of suitable three-dimensional SIE formulations and the efficient evaluation of the resulting two-dimensional surface integrals, that need to be addressed.

Another future research direction is the extension of the proposed approach to more general classes of metasurfaces that cannot be modeled by means of sheet transition conditions and in particular, to approximate metasurfaces as locally periodic rather than locally uniform. This extension, however, poses new theoretical challenges. Such an extension requires the knowledge of a certain proto-Green’s function associated with a periodic transmission problem, which does not admit an expression in terms of Sommerfeld integrals and must be computed numerically. Despite these theoretical challenges, there is both numerical and experimental evidence that a locally periodic approximation is sufficiently accurate for practical metasurface design [10, 25–30].

Finally, we mention that there remains considerable room for further asymptotic analysis of the integral operators Tj, j = 1, 2, and their convergence as ε → 0, to rigorously establish the convergence rate of the Neumann-series solution (25) (i.e. the corrections to ray optics). Although it is clear that ‖ Tj ‖ → 0 as ε → 0, an intricate analysis is required to obtain convergence rates, and to clearly specify for which function spaces convergence is obtained, especially for unbounded surfaces and incident fields where limiting processes are tricky to apply to the surface integrals (51).

Appendices

A. Exact integral representation

This appendix is devoted to the derivation of the integral representation formula (10). In order to achieve that, we show first the symmetry of the exact Green’s function (9). Consider then the functions w(r) = G(r|r1) and v(r) = G(r|r2) for the r1, r2 ∈ Ω+. By Green’s formula and the radiation condition we have

v(r1)w(r2)=Ω+{w2vv2w}dr={w+vy+v+wy+}ds,
and, similarly
0=Ω{w2vv2w}dr={wvyvwy}ds.
From the transition conditions (7), on the other hand, it follows that
w+vy+v+wy+=wvyvwy.
where v±(s) = v(s, 0±) = limδ→0± v(s, δ) and vy±(s)=vy(s,0±)=limδ0± vy(s, δ) and similarly for w. Combining (30), (31) and (32) it is obtained that v(r1)− w(r2) = G(r1|r2)−G(r2|r1) = 0, and thus G(r|r′) = G(r′|r) for all r, r′ ∈ Ω+. The identity G(r|r′) = G(r′|r) for all r, r′ ∈ Ω can be derived in a completely analogous way. Consider now the functions w(r) = G(r|r1) and v(r) = G(r|r2) but with r1 ∈ Ω+ and r2 ∈ Ω. Integration by parts yields the identity
v(r1)={w+vy+v+wy+}ds=w(r2)
in this case, which clearly implies that G(r1|r2) = G(r2|r1). The same result can be easily obtained in the case r1 ∈ Ω and r2 ∈ Ω+.

Finally, the identities

[[Gy(r|r)]]=kα(x){{G(r|r)}}and{{Gy(r|r)}}=kβ(x)[[G(r|r)]]
involving the normal derivatives of the Green’s function on the metasurface Γ for r ∈ Ω+ ∪ Ω, follow straightforwardly from the reciprocity condition G(r|r′) = G(r′|r) established above.

Having established the symmetry of the Green’s function, i.e., G(r|r′) = G(r′|r) for all r, r′ ∈ Ω ∪ Ω+, we can now use it to prove the identity (10). Indeed, it follows from (8), the symmetry of G, and Green’s theorem that

{Gy(r|s,0+)uscat,+(s)G(r|s,0+)uyscat,+(s)}ds={uscat(r),rΩ+,0,rΩ,
and
{Gy(r|s,0)uscat,(s)G(r|s,0)uyscat,(s)}ds={0,rΩ+,uscat(r),rΩ.
Writing the the normal derivatives of G and uscat in terms of their limit values from above and below Γ, it follows that
Gy(r|s,0)=A(s)G(r|s,0)+B(s)G(r|s,0+),Gy(r|s,0+)=B(s)G(r|s,0)A(s)G(r|s,0+),uyscat,(s)=A(s)uscat,(s)+B(s)uscat,+(s)finc(s)anduyscat,+(s)=B(s)uscat,(s)A(s)uscat,+(s)f+inc(s),
where
A(s)=ik{α(s)+β(s)2}andB=ik{α(s)β(s)2}
and f±inc defined in (11). Appropriately combining these expressions we arrive at
Gy(r|s,0)uscat,(s)G(r|s,0)uyscat,(s)=G(r|s,0)finc(s)+B(s)(G(r|s,0+)uscat,(s)G(r|s,0)uscat,+(s))
Gy(r|s,0+)uscat,+(s)G(r|s,0+)uyscat,+(s)=G(r|s,0+)f+inc(s)+B(s)(G(r|s,0+)uscat,(s)G(r|s,0)uscat,+(s))
Finally, from the identities above, and adding (35) and (36), the integral representation formula (10) for the field uscat solution of the boundary value problem (8) is obtained.

B. Sommerfeld-integral Green’s function approximation

This appendix in devoted to the derivation of a Sommerfeld-integral [37] representation of the proto-Green’s function Gp used in the SIE derivations presented in Section 4 above.

As was mentioned above in Section 4, Gp satisfies both the Helmholtz equation (9a) and the radiation condition, but instead of the transition conditions (9b)(9c), it satisfies the locally uniform transition conditions (20). In order to find an expression for Gp we first note that since both metasurface parameters α and β are taken to be functions of the source point r′ = (x′, y′), they are constant as functions of r = (x, y) and thus it is possible to find an exact expression for Gp by standard Fourier transform techniques. In fact, for a source point r′ ∈ Ω+ the proto-Green’s function Gp can be interpreted as the total field produced by the incident field

Ginc(r|r):=i4H0(1)(k|(rr|)=i4πeikx(xx)+iky|yy|kydkx.
The square root ky=ky(kx):=k2kx2 is defined in the complex kx plane as the product kkxk+kx where the first square root has a branch cut along the positive imaginary axis, and the second one has a branch cut along the negative imaginary axis. Figure 8 depicts the domain of definition of ky along with the curves in the complex plane where the real and imaginary parts of ky change sign.

 figure: Fig. 8

Fig. 8 Signs of the real and imaginary parts of square root ky(kx)=kx2k2 in the complex kx plane.

Download Full Size | PDF

Using Ginc as incident field, the total field—which corresponds to Gp—can be expressed as

Gp(r|r)={Ginc(r|r)+Gr(r|r),rΩ+,rΩ+,Gt(r|r),rΩ,rΩ+,
where the reflected and transmitted fields admit the integral representations
Gr(r|r):=i4πR(kx,x)eikx(xx)+iky|y+y|kydkx,
Gt(r|r):=i4πT(kx,x)eikx(xx)+iky|y+y|kydkx,
in terms of the reflection and transmission coefficient R and T defined in (13) and (16), respectively. The special integral sign “∮” introduced in (42) refers to the fact that the path of integration passes below (resp. above) any pole that the integrands may have on the positive (resp. negative) real kx-axis. For the sake of definiteness in what follows of this paper the integral sign ∮ refers to a contour integral along the path SC depicted in Fig. 9.

 figure: Fig. 9

Fig. 9 Sommerfeld contour SC (continuous blue line) and steepest descent contour SD (dashed red line) utilized in the evaluation of the integral in Eq. (59). Poles of the Γ̂α and Γ̂β lying in the shaded regions contribute to the far-field asymptotic expansion of the approximate Green’s function.

Download Full Size | PDF

In order to establish the validity of (42), we note that for Gp to satisfy the locally uniform transition conditions (20), R and T have to related by the equations

iky(1RT)=kα(1+R+T),iky(1R+T)=kβ(1+RT),
where for notational simplicity we have let R = R(kx, x′), T = T(kx, x′), α = α(x′) and β = β(x′). Solving for R and T from (43) we obtain the expressions in (13) and (16) utilized in the previous section.

Similarly, it follows from the symmetric of Gp established in Appendix A that for a point source r′ ∈ Ω the total field Gp takes the form

Gp(r|r)={Gt(r|r),rΩ+,rΩ,Ginc(r|r)+Gr(r|r),rΩ,rΩ.

C. Integral-equation formulation for corrections

This appendix is devoted to the derivations of the SIEs (22). To simplify the notation, we first define the functions gie(s,σ)=Gp(s,0e|σ,0i) and ygie(s,σ)=Gp,y(s,0e|σ,0i) for s, σ ∈ ℝ, where the indices i and e correspond to the symbols “+” or “−“ that refer to the limit values (from above and below, respectively) on Γ.

From the Sommerfeld-integral representation of Gp in (42) and (44) it thus follows that

gie(s,σ)=i4πg^ie(kx,σ)eikx(sσ)dkxandygie(s,σ)=i4πyg^ie(kx,σ)eikx(sσ)dkx,
where
g^++(kx,σ)=g^(kx,σ)=1ky{1+R(kx,σ)}=1ky{2Γ^α(kx,σ)Γ^β(kx,σ)},g^+(kx,σ)=g^+(kx,σ)=1kyT(kx,σ)=1ky{Γ^α(kx,σ)+Γ^β(kx,σ)},yg^++(kx,σ)=yg^(kx,σ)=i{1+R(kx,σ)}=2ii{Γ^α(kx,σ)+Γ^β(kx,σ)}yg^+(kx,σ)=yg^(kx,σ)=iT(kx,σ)=i{Γ^α(kx,σ)Γ^β(kx,σ)},
with Γ̂α and Γ̂β being defined in (14).

Next we introduce the boundary integral operators

(Sieϕ)(s)=gie(s,σ)ϕ(σ)dσand(ySie(ϕ))(s)=ygie(s,σ)ϕ(σ)dσ,s,
(that must be interpreted in the sense of distributions) which arise when taking limits of (21) and its normal derivative on Γ. From the integral representation (21) it follows that
[[uscat]](s)=(S++S+)(φ+ψ)(s),{{uscat}}(s)=(S+++S+)(φψ)(s),[[uyscat]](s)=(yS++yS+)(φψ)(s),{{uyscat}}(s)=(yS+++yS+)(φ+ψ)(s),
in terms of the integral operators (46), where we have utilized the identities S++=S, S+=S+, yS++=yS and yS+=yS+ that result directly from (45). The uncoupled system of SIEs
{yS++yS++ikα(s)(S+++S+)}μ1(s)=2ikα(s)uinc(s,0)
{yS+++yS++ikβ(s)(S++S+)}μ2(s)=2uyinc(s,0)
for the new density functions μ1 = φψ and μ2 = φ + ψ is thus obtained by substituting (47) in the transition conditions (8b)(8c).

To show that SIE system (48) is in fact of the second-kind, we need to further study the properties of the integral operators on the left-hand-side of (48). Such properties can be determined from the regularity of the integral kernels gie and ygie which can in turn be derived from decay estimates for their Fourier transforms in (45). Consequently, utilizing the properties of the Fourier transform, it can be shown that the integral kernels in (48) can be expressed as

yg++(s,σ)yg+(s,σ)+ikα(s)(g++(s,σ)+g+(s,σ))=δs+K1(s,σ),yg++(s,σ)+yg+(s,σ)+ikβ(s)(g++(s,σ)g+(s,σ))=δs+K2(s,σ),
where δs denotes the Dirac delta distribution supported at s and
K1(s,σ):=k2π{α(σ)α(s)}eikx(sσ)ky+kα(σ)dkx,
K2(s,σ):=k2π{β(σ)β(s)}eikx(sσ)ky+kβ(σ)dkx.
Using the properties of the Dirac delta distribution and defining the integral operators
Tj[μ](s):=Kj(s,σ)μ(σ)dσ,s,j=1,2,
we finally conclude that the relations in Eq. (48) can be equivalently expressed, in abstract form, as the SIEs in Eq. (22) for the μj, j = 1, 2.

The key fact about Tj for its use in our series of corrections is that Tj → 0 as ε → 0. We demonstrate this numerically in Appendix E. Analytically it occurs because the coefficients α(σ) − α(s) and β(σ) − β(s) vanish as ε → 0 for continuous functions α(x) = a(εx) and β(x) = b(εx). But, as discussed in Section 5, a technically challenging asymptotic analysis is required to rigorously demarcate the function spaces and norms for which Tj → 0 and to determine the rate of convergence, which we relegate to future work.

D. Generalized laws of reflection and transmission

This appendix is devoted to the derivation of the so-called generalized laws of reflection and transmission [28, 29] for general interface parameters for which standard Bloch–Floquet analysis [8] does not work. As it turns out, these laws can be derived from our zeroth-order approximation. To see this, consider metasurface parameters of the form

α(s)=α0+α˜(s)andβ(s)=β0+β˜(s),s,
where α0 and β0 are constants and α̃ and β̃ are bounded functions. The total field resulting from the scattering of a planewave uinc(r) = eikinc·r, kinc = k(cos θinc, sin θinc), −π/2 ≤ θinc ≤ 0, off of the metasurface can be expressed as u = v + u0 where
u0(r)={eikincr+R0(θinc)eikrr,rΩ+,T0(θinc)eikincr,rΩ,
with kr = k(cos θinc, − sin θinc) and
R0(θ):=1α0|sinθ|+α0β0|sinθ|+β0andT0(θ):=α0|sinθ|+α0+β0|sinθ|+β0,
is the total field resulting from the scattering of the planewave uinc off of a metasurface with constant interface parameters α0 and β0. The field v = uu0, on the other hand, satisfies the Helmholtz equation in ℝ2\Γ, the radiation condition, and the transition conditions
[[vy]]=ikα{{v}}ikα˜{{u0}}and{{vy}}=ikβ[[v]]ikβ˜[[u0]]onΓ.

Letting then G0 = Gp denote the Green’s function in Eqs. (42) and (44) corresponding to constant interface parameters α0 and β0, we obtain from the discussion in Section 4 that the zeroth-order approximation (i.e., Eq. (28a) with N = 0) of v is given by

v0(r)={G0(r|σ,0+)φ0(σ)G0(r|σ,0)ψ0(σ)}dσ,r2\Γ,
where letting
A˜(s)=ik{α˜(s)+β˜(s)2}andB˜(s)=ik{α˜(s)β˜(s)2},
the approximate densities φ0 and ψ0 are given by
φ0(s)={A˜(s)T0(θinc)+B˜(s)(1+R0(θinc))}eikscosθinc,ψ0(s)={B˜(s)T0(θinc)+A˜(s)(1+R0(θinc))}eikscosθinc.

Replacing G0 in Eq. (53) by its far-field approximation, derived in Appendix F, we find that

v0(r)~eik|r||r|v(θ,θinc)as|r|,
where, after some algebraic manipulations, the far-field pattern v(θ|θinc) can be expressed as
v(θ,θinc)=2kπei3π4|sin(θinc)sinθ|{vα(θ,θinc)+vβ(θ,θinc)θ(0,π),vα(θ,θinc)vβ(θ,θinc)θ(π,0),
where
vα(θ,θinc):=α˜(σ)eikσ(cosθinccosθ)(|sinθ|+α0)(|sinθinc|+α0)dσandvβ(θ,θinc):=β˜(σ)eikσ(cosθinccosθ)(|sinθ|+β0)(|sinθinc|+β0)dσ.

From Eq. (56) it thus follows that the far-field pattern v in Eq. (55) would correspond to a linear combination of planewaves with wavevectors kα = k(cos θα, sin θα) and kβ = k(cos θβ, sin θβ) if vα and vβ were Dirac delta distributions supported at angles θα and θβ, respectively. Formally, this can be achieved by selecting

α˜(s)=cαeikdαsandβ˜(s)=cβeikdβs,
with cα, cβ ∈ ℂ and dα and dβ being such that
dα=cosθαcosθincanddβ=cosθβcosθinc.

Finally, the expressions in Eq. (19) are obtained by letting α0 = β0 = − sin θinc, cα = ±α0, and cβ = ∓β0, for which a minimal reflection off of the metasurface is achieved.

E. On the convergence of the Neumann series in Eq. (25)

In our next appendix we consider an example to study the convergence of the Neumann series approximation in Eq. (27) by examining the dependence of the spectral radii of the discretized integral operators Tj, j = 1, 2, on the smoothness of the metasurface parameters α and β. We consider here the discretized version of the SIEs in Eq. (22) which take the form (ITj)μj = gj, where I is the identity matrix, Tj is the discretized integral operator by the method described in Appendix G, μj is the unknown vector, and gj is the discretized interface data. It is easy to show that if ‖Tj< 1 in some matrix norm, then the relative error in the discretized Neumann series approximation can be bounded by

μjμj(N)μjTjN+1c[ρ(Tj)]N+1
for some constant c > 0 where ρ(Tj) = maxn |λn(Tj)| denotes the spectral radius of Tj. This bound shows that the spectral radius provides an approximate rate of convergence of the Neumann series approximation as N increases. Moreover, it can be shown that the Neumann series for the discrete linear system converges if and only if the ρ(Tj) < 1. In the following example then, we consider a metasurface parameter α(x) = a(εx) given by the truncated Fourier series
α(x)=a(εx):==010ceiεx,x,
where the coefficients c are randomly generated from a uniform distribution and are also adjusted so that the constrain Re α ≥ 0 is satisfied. Clearly, ε > 0 is a parameter that controls the smoothness of α. Figure 7 displays the spectral radius of the matrix T1 in log-log scale for a range of values of ε. There results demonstrate that convergence of the Neumann series is in fact expected for a large range of values of ε, including some of those that give rise to quite rough interface parameters α.

F. The far field of Gp and its relationship with Gpff

This Appendix presents a detailed asymptotic analysis that establishes rigorously the relationship between Gp and Gpff, as well as the existence of guided modes.

To establish the relationship between Gpff and Gp, we derive the far-field asymptotic approximation (as |r| → ∞) of the proto-Green’s function Gp given in Eqs. (42) and (44). In order to do so we resort to the method steepest descents for which we follow the analysis of Sommerfeld integrals presented [41, Chapter 8]. Similar saddle point calculations can also be found in classical references on layered media scattering, such as [37,45].

Assuming first that r′ ∈ Ω+ and letting r = |r|(cos θ, sin θ), θ ∈ (−π, 0) ∪ (0, π), we have that the resulting reflected and transmitted fields in Eq. (42) can be expressed as

Gr(r|r)=SCqr(kx,r)e|r|ϕ(kx)dkxandGt(r|r)=SCqt(kx,r)e|r|ϕ(kx)dkx
in terms of the phase and amplitude functions defined as
ϕ(kx):=ikxcosθ+iky|sinθ|,
qr(kx,r):=iR(x,kx)4πeikxx+ikyyky,
qt(kx,r):=iT(x,kx)4πeikxx+ikyyky,
respectively. Note that in this case (r′ ∈ Ω+) we are interested in Gr for θ ∈ (0, π) and in Gt for θ ∈ (−π, 0).

Three kinds of critical points have to be taken in account in the steepest descent method approximation of the integrals in Eq. (59), namely, saddle points of the phase function ϕ, (possible) poles singularities of the integrands qr and qt, and the branch points of the square root ky=ky(kx)=k2kx2.

We first consider the saddle points of ϕ, which correspond to solutions kx* of the algebraic equation ϕ(kx*)=0. In view of ϕ′(kx) = i cos θikx| sin θ|/ky follows that there is only one saddle point on SC given by kx*=kcosθ=kx/|r| at which ϕ(kx*)=ik. The steepest descent directions from kx*, on the other hand, are given by the angles 3π/4 and −π/4, which were obtained from ϕ(kx*)=i/(ksin2θ)0. We then conclude that the steepest descent path SD is given implicitly by the equation Im ϕ(kx) = k from which it can be shown that SD intersects SC again at kx = k sec θ and that

Imkx=Rekx|cotθ|k|cscθ|as|kx|ifx>0,
and
Imkx=Rekx|cotθ|+k|cscθ|as|kx|ifx<0.
Figure 9 depicts the steepest descent paths for x > 0 and x < 0. SD coincides with SC when x = 0.

With this information in hand we then proceed to deform the Sommerfeld contour SC to the steepest descent contour SD that passes through the saddle point kx*. Note that SD does not intersect the branch cuts stemming from ±k and, thus, there is no contribution to the asymptotic expansions from the branch points at ±k. There might be, however, contributions arising from pole singularities of the integrands. In fact, from the expressions for the reflection and transmission coefficients in Eqs. (13) and (16), respectively, we have that the poles of both qr in Eq. (61) and qt in Eq. (62)—which correspond to the poles of the functions Γ̂α(x, x′) and Γ̂β(x, x′) defined in Eq. (14)—are solutions of the (independent) algebraic equations

ky(kx)=kα(x)orky(kx)=kβ(x).
To find necessary and sufficient conditions on the metasurface parameters α and β for such poles to exist, we first note that the conditions Re α ≥ 0, Re β ≥ 0 and the Eqs. (64) imply that poles of qr or qt could only exist in the regions of the complex kx plane where Re ky(kx) ≤ 0. It is easy to see that such regions amount to D = {Re kxk, Im kx ≥ 0} ∪ {Re kx ≤ −k, Im kx ≤ 0} (Fig. 8). Furthermore, since Imky=k2kx20 in D, poles will exist if and only Im α ≤ 0 or Im β ≤ 0.

By the Cauchy residue theorem we thus have that the poles of qr (resp. qt), if any, will only contribute to the far-field expansion of Gr (resp. Gt) if they lie within the region in the complex plane enclosed by SC and SD. In order to determine whether a pole of qr (resp. qt) lies inside that region, and in view of the fact that we do not have access to an explicit parametrization of SD, we utilize the asymptotic identities in Eq. (63). Doing so we conclude that the relevant poles—that are henceforth denoted by kx(p), p = 1, 2—must meet the conditions

kx(1)=k1α2orkx2=k1β2,and0Imkx(p)Rekx(p)|cotθ|k|cscθ|ifx>0,
and
kx(1)=k1α2orkx2=k1β2,andRekx(p)|cotθ|+k|cscθ|Imkx(p)0ifx<0.

Therefore, accounting for both saddle point and poles contributions, we obtain the following asymptotic expansions

Gr(r|r)=eik|r|ikrr¯/|r|+iπ48πk|r|R(kx|r|,x)+p=1,2Apreikx(p)|xx|+iky(kx(p))|y+y|+O(|r|3/2)
and
Gt(r|r)=eik|r|ikrr¯/|r|+iπ48πk|r|T(kx|r|,x)+p=1,2Apteikx(p)|xx|+iky(kx(p))|yy|+O(|r|3/2)
as |r| → ∞, where r̄′ is the image point source r̄′ = (x′, −y′) (Fig. 2). The amplitudes A1r and A1t of the guided waves in Eq. (66) are directly obtained from the residues of qt and qr and read as
A1r=A1t=α21α2
if a pole kx(1) in Eq. (65) exists, and they equal zero otherwise. Similarly,
A2r=A2t=β21β2
if a pole kx(2) in Eq. (65) exists, and they equal zero otherwise.

The contribution to the asymptotic expansions of the pole singularities corresponds to surface waves that travel away from the point source r′ = (x′, y′) and are confined to a narrow strip containing the metasurface—as they decay exponentially fast toward the upper and lower half-planes. For example, the contribution of the pole kx(1)=k1α2 to the asymptotic expansion of the reflected and transmitted fields equals

α21α2eik1α2|xx|ikα|y±y|,
with + and − corresponding to the reflected and transmitted fields, respectively.

 figure: Fig. 10

Fig. 10 Real part of the total field solution of the problem of scattering of the field produced by a point source by a metasurface with constant interface parameters α = −i and β = −1.2i. This selection of the interface parameters allow a surface wave mode in Eq. (69) to propagate along the metasurface. The point source is located at distance 0.25λ above the metasurface. (a) Gpff which in this case corresponds to the ray-optics approximation of the exact Green’s function G. (b) Rigorous far-field asymptotic approximation of Gp in Eq. (66). (c) Direct evaluation of the proto-Green’s function Gp defined in Eqs. (42) and (44).

Download Full Size | PDF

In order to establish the relationship between Gpff and Gp we recall the asymptotic identity

i4H0(1)(k|rr|)=eik|r|ikrr¯/|r|+iπ48πk|r|+O|r|3/2as|r|,
that follows from |rr′| = rr · r′/|r| + O(|r|−1) as |r| → ∞ and the standard asymptotic expansion of the Hankel function. Using the approximation in Eq. (70), the fact that (xx′)/|rr′| = x/|r| + O(|r|−1) as |r| → ∞, and comparing Eq. (15) with Eq. (42), we conclude that
Gpff(r|s,0±)=Gp(r|s,0±)+O(|r|3/2)as|y|.
Therefore, Gpff in Eq. (15) can be simply interpreted as an approximation of the proto-Green’s function Gp for point sources at the metasurface and observations points far away from the metasurface. Note that for such configuration of source and observation points, the surface wave modes do not have any significant contribution as they decay exponentially as |y| → ∞.

Finally, in order to demonstrate the validity of the asymptotic expansions derived in this section we present Fig. 10 that displays the real part of Gp(r|r′) for constant interface parameters α = −i and β = −1.2i. Note that for these interface parameters the terms Γ̂α(kx, s) and Γ̂β(kx, s) defined in Eq. (14) have poles on the real axis which make surface wave modes in Eq. (69) appear in the asymptotic expansions (66).

G. Numerics

In this appendix we briefly describe a high-order method for the numerical evaluation of the Sommerfeld integrals Gr and Gt in Eqs. (42)(44) and the integral kernels K1 and K2 in Eq. (50). This approach, which was originally developed for layered-media scattering problems [46] (see also [47, Section 2.3.5]), is a combination of the contour-integration method described in [48] and the the smooth-windowing approach put forth in [49] for the evaluation of oscillatory integrals.

Proto-Green’s function Consider the Sommerfeld integrals Gr and Gt in Eqs. (42)(44) which are given by linear combinations of integrals of the form

Φ(d1,d2)=i4πkγky+kγei(kxd1+kyd2)/kkydkx,d1d20,
where d1 = k|xx′|, d2 = k|y + y′| in the case of Gr or d2 = k|yy′| in the case of Gt, and γ = α(x′) or γ = β(x′). Note that dj, j = 1, 2, are dimensionless variables. To tackle the most challenging integration scenario, in what follows we consider the case when the integrand has a pole at k1γ2 on the real axis.

Making use of the change of variable ξ = kx/k, using the fact that ky(kx) = ky(−kx), and letting

f(ξ)=iγπcos(ξd1)1ξ2+γeid21ξ21ξ2,
we have that Φ can be expressed as Φ = I1 + I2 where Ij=Cjf(ξ)dξ, j = 1, 2, with C1 and C2 being the contours depicted in Fig. 11. The curve C1 is a simple curve in the fourth quadrant that is parametrized by a smooth complex-valued function ζ : [0, π] ↦ ℂ satisfying ζ(0) = 0 and ζ(π) = L := 1 + ξp, where ξp is the pole of f at 1γ2. For the sake of definiteness, the curve C1 is here selected as the semi-ellipse
ζ(t):={L(1+cos(t+π))2+iHsin(t+π):t[0,π]}
that passes below all the singularities of the integrand f. The contour C2, on the other hand, is simply the interval [L, ∞] on the real axis.

 figure: Fig. 11

Fig. 11 Integration contours used in the evaluation of the proto-Green’s function Gp, and the integral kernels K1 and K2.

Download Full Size | PDF

Note that on C1 the function f (ζ(t)) grows exponentially as t increases from 0 to π/2. Indeed, it can be shown [47] that |eid21ξ2|ed2H and |cos(d1ξ)| ≤ ed1H for ζC1. Thus, in order to control the exponential growth of the integrand on C1 we select H = (max{10, d1 + d2})−1. This simple procedure ensures that the exponential terms of f remain bounded by one along C1. The resulting expression for the contour integral I1 is then approximated by means of the Clenshaw–Curtis quadrature rule [40]—which, for the smooth integrand under consideration, yields rapid convergence. In view of the oscillatory behavior of the integrand and in order to maintain the same accuracy for all d1 and d2, the number of quadrature points is chosen to grow linearly with d1.

In order to evaluate the oscillatory integral I2, on the other hand, we utilize the windowing method put forth in [49]. Using this procedure I2 is approximated as

I2LA+Lf(t)wA(tL)dt,
where the window function wA is defined as wA(t) := η(t; cA, A), A > 0, 0 < c < 1, in terms of the C(ℝ) function
η(t;t0,t1)={1,|t|t0,exp(2e1/uu1),t0<|t|<t1,u=|t|t0t1t0,0,|t|>t1,
which equals one on [−t0, t0] and is supported on the (bounded) interval [−t1, t1].

In virtue of the oscillatory behavior of the integrand when d1 ≠ 0, and the exponential decay of the integrand when d2 ≠ 0, the integral on the right-hand-side of Eq. (75) converges to I2 faster than any negative power d12+d22A as A goes to infinity—as proved in [47, Proposition 2.3.4]. In the special case d1 = d2 = 0, however, the f is slowly decaying on C2 and does not oscillate, thus, it leads to slow (algebraic) convergence of the windowed-integral approximation in Eq. (75) to the integral I2 as A → ∞. In fact, in that case the error in the windowed-integral approximation decays as O((cA)−1) [47, Proposition 2.3.4]. Therefore, as in the case of I1, the integral on the right-hand-side of Eq. (75) is here approximated by using Clenshaw–Curtis quadrature. The super-algebraic/exponential convergence of the windowed integral allows I2 to be approximated with a fix accuracy and a fixed computational cost by choosing A inversely proportional to d12+d22.

Integral kernels In order to numerically evaluate the integral kernels K1 and K2 defined in Eq. (50), we resort to the contour-integration procedure described above. The evaluation of these kernels requires the approximation of integrals of the form

Ψ(d)=eikxd/kky+kγdkx,d0,
where d = k|sσ|, γ = α(σ) in the case of K1 and γ = β(σ) in the case of K2. Unlike the integral in Eq. (72), the Fourier integral in Eq. (77) is only conditionally convergent for d > 0 and diverges at d = 0. In order to separate the singularity of Ψ at d = 0 we use identity
H0(1)(d)=1πeikxd/kkydkx=1πeiξd1ξ2dξ,d0,
which follows from Eq. (41), to obtain
Ψ(d)=πH0(1)(d)kγeikxd/kky(ky+kγ)dkx.
Thus, the integral in Eq. (79)—that turns out to be a continuous function of d—can now be evaluated directly by means of the contour integration procedure presented above for all d ≥ 0. We thus have Ψ(d)=πH0(1)(d)γ(I1+I2) where Ij=Cjf(ξ)dξ with f now being given by f(ξ)=eiξd/[1ξ2(1ξ2+γ)]. Although f is absolutely integrable (it decays as ξ−2 as |ξ| → ∞) a large value of A > 0 is needed in the windowed approximation of I2 in Eq. (75) to achieve a prescribed accuracy when d = 0. In order to improve the slow O((cA)−1) convergence rate as A → ∞ when d = 0, we note further that using the identity
eikxd/kk2kx2dkx=iπ2keikxd/k,d0,
which follows directly from Jordan’s lemma and Cauchy’s residue theorem, the integral in Eq. (79) can be expressed as
eikxd/kky(ky+kγ)dkx=iπ2keikd/kkγeikxd/k(k2kx2)(ky+kγ)dkx.
Therefore, since the integrand on the right-hand-side of Eq. (81) decays as ξ−3 as |ξ| → ∞, the contour integration procedure described in the previous section yields super-algebraic convergence as A → ∞ for d > 0 and now yields an O((cA)−2) convergence rate when d = 0.

Integral operators and potentials Finally, we briefly mention that the (improper) oscillatory integrals in Eqs. (17), (21), (28), and in the definition of the operators T1 and T2 in Eq. (51), can be accurately truncated by means of the windowing procedure described above, in a manner similar to the windowed Green function method [50–52]. In order to handle the logarithmic singularity of the integral kernels in Eq. (51), on the other hand, standard singular integrations techniques, such as the spectrally accurate Martensen–Kussmaul quadrature rule [53], can be used.

Funding

Army Research Office and under Cooperative Agreement Number W911NF-18-2-0048.

References and links

1. E. F. Kuester, M. A. Mohamed, M. Piket-May, and C. L. Holloway, “Averaged transition conditions for electromagnetic fields at a metafilm,” IEEE Transactions on Antennas Propag. 51, 2641–2651 (2003). [CrossRef]  

2. 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 Propag. Mag. 54, 10–35 (2012). [CrossRef]  

3. C. L. Holloway, A. Dienstfrey, E. F. Kuester, J. F. O’Hara, A. K. Azad, and A. J. Taylor, “A discussion on the interpretation and characterization of metafilms/metasurfaces: The two-dimensional equivalent of metamaterials,” Metamaterials. 3, 100–112 (2009). [CrossRef]  

4. C. L. Holloway, E. F. Kuester, and A. Dienstfrey, “Characterizing metasurfaces/metafilms: The connection between surface susceptibilities and effective material properties,” IEEE Antennas Wirel. Propag. Lett. 10, 1507–1511 (2011). [CrossRef]  

5. S. A. Tretyakov, “Metasurfaces for general transformations of electromagnetic fields,” Proc. Royal Soc. Lond. A: Math. Phys. Eng. Sci. 373, 20140362 (2015).

6. S. N. Tcvetkova, D. H. Kwon, A. Díaz-Rubio, and S. A. Tretyakov, “Near-perfect conversion of a propagating plane wave into a surface wave using metasurfaces,” Phys. Rev. B 97, 115447 (2018). [CrossRef]  

7. A. Epstein and G. V. Eleftheriades, “Passive lossless Huygens metasurfaces for conversion of arbitrary source field to directive radiation,” IEEE Transactions on Antennas Propag. 62, 5680–5695 (2014). [CrossRef]  

8. A. Epstein and G. V. Eleftheriades, “Floquet-Bloch analysis of refracting Huygens metasurfaces,” Phys. Rev. B 90, 235127 (2014). [CrossRef]  

9. C. Pfeiffer and A. Grbic, “Metamaterial Huygens’ Surfaces: Tailoring Wave Fronts with Reflectionless Sheets,” Phys. Rev. Lett. 110, 197401 (2013). [CrossRef]  

10. K. Achouri, M. A. Salem, and C. Caloz, “General metasurface synthesis based on susceptibility tensors,” IEEE Transactions on Antennas Propag. 63, 2977–2991 (2015). [CrossRef]  

11. A. Díaz-Rubio, V. S. Asadchy, A. Elsakka, and S. A. Tretyakov, “From the generalized reflection law to the realization of perfect anomalous reflectors,” Sci. advances 3, e1602714 (2017). [CrossRef]  

12. V. S. Asadchy, M. Albooyeh, S. N. Tcvetkova, A. Díaz-Rubio, Y. Ra’di, and S. A. Tretyakov, “Perfect control of reflection and refraction using spatially dispersive metasurfaces,” Phys. Rev. B 94, 075142 (2016). [CrossRef]  

13. N. M. Estakhri and A. Alu, “Wave-front Transformation with Gradient Metasurfaces,” Phys. Rev. X 6, 041008 (2016).

14. B. Groever, W. T. Chen, and F. Capasso, “Meta-Lens Doublet in the Visible Region,” Nano Lett. 17, 4902–4907 (2017). [CrossRef]   [PubMed]  

15. B. Z. Katsenelenbaum, Theory of Nonuniform Waveguides: The Cross-Section Method (IET, 1998). [CrossRef]  

16. S. G. Johnson, P. Bienstman, M. Skorobogatiy, M. Ibanescu, E. Lidorikis, and J. Joannopoulos, “Adiabatic theorem and continuous coupled-mode theory for efficient taper transitions in photonic crystals,” Phys. Rev. E 66, 066608 (2002). [CrossRef]  

17. D. Marcuse, Theory of Dielectric Optical Waveguides (Academic Press, 1974).

18. A. W. Snyder and J. Love, Optical Waveguide Theory (Springer, 2012).

19. D. Colton and R. Kress, Integral Equation Methods in Scattering Theory, vol. 72 (SIAM, 2013). [CrossRef]  

20. J.-C. Nédélec, Acoustic and Electromagnetic Equations: Integral Representations for Harmonic Problems, vol. 144 (Springer, 2001). [CrossRef]  

21. E. Martini and S. Maci, “Metasurface transformation theory,” in “Transformation Electromagnetics and Metamaterials,” (Springer, 2014), pp. 83–116. [CrossRef]  

22. J. P. Wong, A. Epstein, and G. V. Eleftheriades, “Reflectionless wide-angle refracting metasurfaces,” IEEE Antennas Wirel. Propag. Lett. 15, 1293–1296 (2016). [CrossRef]  

23. M. Khorasaninejad and F. Capasso, “Metalenses: Versatile multifunctional photonic components,” Science. 358, 1–8 (2017). [CrossRef]  

24. L. Verslegers, P. B. Catrysse, Z. Yu, W. Shin, Z. Ruan, and S. Fan, “Phase front design with metallic pillar arrays,” Opt. Lett. 35, 844–846 (2010). [CrossRef]   [PubMed]  

25. C. Pfeiffer, N. K. Emani, A. M. Shaltout, A. Boltasseva, V. M. Shalaev, and A. Grbic, “Efficient light bending with isotropic metamaterial Huygens’ surfaces,” Nano Lett. 14, 2491–2497 (2014). [CrossRef]   [PubMed]  

26. F. Aieta, P. Genevet, N. Yu, M. A. Kats, Z. Gaburro, and F. Capasso, “Out-of-plane reflection and refraction of light by anisotropic optical antenna metasurfaces with phase discontinuities,” Nano Lett. 12, 1702–1706 (2012). [CrossRef]   [PubMed]  

27. N. Yu, P. Genevet, F. Aieta, M. A. Kats, R. Blanchard, G. Aoust, J.-P. Tetienne, Z. Gaburro, and F. Capasso, “Flat optics: controlling wavefronts with optical antenna metasurfaces,” IEEE J. Sel. Top. Quantum Electron. 19, 4700423 (2013). [CrossRef]  

28. N. Yu, P. Genevet, M. A. Kats, F. Aieta, J. P. Tetienne, F. Capasso, and Z. Gaburro, “Light propagation with phase discontinuities: Generalized laws of reflection and refraction,” Science. 334, 333–337 (2011). [CrossRef]   [PubMed]  

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

30. R. Pestourie, C. Pérez-Arancibia, Z. Lin, W. Shin, F. Capasso, and S. G. Johnson, “Inverse design of large-area metasurfaces,” arXiv preprint arXiv:1808.04215 (2018).

31. T. Niemi, A. O. Karilainen, and S. A. Tretyakov, “Synthesis of polarization transformers,” IEEE Transactions on Antennas Propag. 61, 3102–3111 (2013). [CrossRef]  

32. M. Selvanayagam and G. V. Eleftheriades, “Polarization control using tensor Huygens surfaces,” IEEE Transactions on Antennas Propag. 62, 6155–6168 (2014). [CrossRef]  

33. C. Pfeiffer and A. Grbic, “Bianisotropic metasurfaces for optimal polarization control: Analysis and synthesis,” Phys. Rev. Appl. 2, 044011 (2014). [CrossRef]  

34. A. G. Voronovich, Wave Scattering From Rough Surfaces, vol. 17 (Springer, 2013).

35. D. C. O’Shea, T. J. Suleski, A. D. Kathman, and D. W. Prather, Diffractive Optics: Design, Fabrication, and Test, vol. 62 (SPIE Press, 2004).

36. C. Cohen-Trannoudji, B. Diu, and F. Laloë, Quantum Mechanics. Volume II (Hermann, Paris, 1973).

37. W. C. Chew, Waves and Fields in Inhomogeneous Media, vol. 522 (IEEE pressNew York, 1995).

38. S. G. Johnson, M. Povinelli, M. Soljačić, A. Karalis, S. Jacobs, and J. Joannopoulos, “Roughness losses and volume-current methods in photonic-crystal waveguides,” Appl. Phys. B 81, 283–293 (2005). [CrossRef]  

39. M. Born and E. Wolf, Principles of Optics (Cambridge Univeristy Press, 1999), 7th ed. [CrossRef]  

40. P. J. Davis and P. Rabinowitz, Methods of Numerical Integration (Courier Corporation, 2007).

41. N. Bleistein, Mathematical Methods for Wave Phenomena (Academic Press, 2012).

42. A. Nosich, “Radiation conditions, limiting absorption principle, and general relations in open waveguide scattering,” J. Electromagn. Waves Appl. 8, 329–353 (1994). [CrossRef]  

43. J. A. DeSanto and P. A. Martin, “On angular-spectrum representations for scattering by infinite rough surfaces,” Wave Motion 24, 421–433 (1996). [CrossRef]  

44. M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions: With Formulas, Graphs, and Mathematical Tables, vol. 55 (Courier Corporation, 1964).

45. L. Brekhovskikh, Waves in Layered Media, vol. 16 (Elsevier, 2012).

46. C. Pérez-Arancibia and O. P. Bruno, “High-order integral equation methods for problems of scattering by bumps and cavities on half-planes,” J. Opt. Soc. Am. A 31, 1738–1746 (2014). [CrossRef]  

47. C. Pérez-Arancibia, “Windowed integral equation methods for problems of scattering by defects and obstacles in layered media,” Ph.D. thesis, California Institute of Technology (2017).

48. M. Paulus, P. Gay-Balmaz, and O. Martin, “Accurate and efficient computation of the Green’s tensor for stratified media,” Phys. Rev. E 62, 5797 (2000). [CrossRef]  

49. J. A. Monro, “A Super-Algebraically Convergent, Windowing-Based Approach to the Evaluation of Scattering from Periodic Rough Surfaces,” Ph.D. thesis, California Institute of Technology (2008).

50. O. P. Bruno, M. Lyon, C. Pérez-Arancibia, and C. Turc, “Windowed Green function method for layered-media scattering,” SIAM J. on Appl. Math. 76, 1871–1898 (2016). [CrossRef]  

51. O. P. Bruno and C. Pérez-Arancibia, “Windowed Green function method for the Helmholtz equation in the presence of multiply layered media,” Proc. Royal Soc. Lond. A: Math. Phys. Eng. Sci. 473, 20170161 (2017). [CrossRef]  

52. O. P. Bruno, E. Garza, and C. Pérez-Arancibia, “Windowed Green function method for nonuniform open-waveguide problems,” IEEE Transactions on Antennas Propag. 65, 4684–4692 (2017). [CrossRef]  

53. D. Colton and R. Kress, Inverse Acoustic and Electromagnetic Scattering Theory, vol. 93 (Springer, 2012), 3rd ed.

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 (11)

Fig. 1
Fig. 1 Scattering problem under consideration. A time-harmonic incident field uinc with angular frequency ω > 0 impinges on a metasurface with surface impedance Z and surface admittance Y, producing a reflected field propagating in the upper half-plane and a transmitted field propagating in the lower half-plane. The metasurface is surrounded by a homogeneous medium with electric permittivity 0 > 0 and magnetic permeability μ0 > 0.
Fig. 2
Fig. 2 Variables used in the derivation of the approximation in Eq. (16) of the Green’s function of the problem under consideration defined in Eq. (9). The field produced by a point source at r′ impinges on the metasurface at r̃′. The total field is observed at the point r far away from the metasurface that is denoted by Γ. The total field is decomposed in primary and secondary fields. The latter corresponds to the field reflected off of the metasurface, which is approximated by means of the local reflection coefficient R in Eq. (13) assuming specular reflection.
Fig. 3
Fig. 3 First column: Ray-optics approximation (17) of the total field utot = uinc + uscat solution of the problem of scattering of a planewave that impinges at normal incidence on metasurface’s that produce transmitted fields that are predominantly planewaves propagating in the directions θt (a) 45°, (d) 30° and (g) 22.5° with respect to the metasurface. Second column: Full “exact” total field solution of the corresponding for planewave transmitted fields in the directions θt (b) 45°, (e) 30° and (h) 22.2°. Third column: Absolute errors for planewave transmitted fields in the directions θt (c) 45°, (f) 30° and (i) 22.5°.
Fig. 4
Fig. 4 High-order corrections to the ray optics approximation (17). (a) Geometrical configuration of the problem under consideration which corresponds to an unit amplitude planewave impinging at normal incidence on a metasurface that renders a transmitted planewave with wavevector forming an angle of 22.5° with respect to the metasurface. (b) and (c): Real part of the total field 0th and 1st order approximations. (d), (e), (f), (g) and (h): Absolute errors | u u N tot , ff |, for N = 0, 1, 2, 3 and 4, respectively, in the zeroth (no correction), first, second, third and fourth order corrections to the ray optics approximation u 0 tot , ff. The color scales were adjusted according to the maximum error displayed in each one of the figures.
Fig. 5
Fig. 5 Rate of convergence of the zeroth-, first-, second-, third- and fourth-order near-field approximations (28a) in terms of the smoothness of the metasurface parameters. The plot is in log-log scale. The problem under consideration is the scattering of a unit amplitude plane wave at normal incidence that impinges on a metasurface with metasurface parameters α(x) = a(εx) = c0{1 − eiεx} and β(x) = b(εx) = c0{1 − eiεx}. The parameter ε > 0 controls the smoothness of metasurface parameters. The color curves display the maximum of the absolute error | u tot ( r ) u N tot ( r ) | in the near-field approximations (28a) evaluated at the spatial points r ∈ {−1, 0, 1} × {−10, 10}.
Fig. 6
Fig. 6 Total field solution of the problem of scattering of a Gaussian beam by a metasurface with interface parameters α and β engineered to allow for surface wave modes (69) to propagate along the metasurface.
Fig. 7
Fig. 7 Spectral radius of the discretized integral operator T1 as a function of the smoothness of the interface parameter α in Eq. (58). The yellow painted strip indicates the range of parameters ε for which the Neumann series diverges.
Fig. 8
Fig. 8 Signs of the real and imaginary parts of square root k y ( k x ) = k x 2 k 2 in the complex kx plane.
Fig. 9
Fig. 9 Sommerfeld contour SC (continuous blue line) and steepest descent contour SD (dashed red line) utilized in the evaluation of the integral in Eq. (59). Poles of the Γ̂α and Γ̂β lying in the shaded regions contribute to the far-field asymptotic expansion of the approximate Green’s function.
Fig. 10
Fig. 10 Real part of the total field solution of the problem of scattering of the field produced by a point source by a metasurface with constant interface parameters α = −i and β = −1.2i. This selection of the interface parameters allow a surface wave mode in Eq. (69) to propagate along the metasurface. The point source is located at distance 0.25λ above the metasurface. (a) G p ff which in this case corresponds to the ray-optics approximation of the exact Green’s function G. (b) Rigorous far-field asymptotic approximation of Gp in Eq. (66). (c) Direct evaluation of the proto-Green’s function Gp defined in Eqs. (42) and (44).
Fig. 11
Fig. 11 Integration contours used in the evaluation of the proto-Green’s function Gp, and the integral kernels K1 and K2.

Equations (106)

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

Z n × ( H + H ) = 1 2 ( E + + E ) , Y n × ( E + E ) = 1 2 ( H + + H ) ,
2 E + k 2 E = 0 , H 1 = 1 i k η E y , H 2 = 1 i k η E x ,
2 H + k 2 H = 0 , E 1 = η i k H y , E 2 = η i k H x .
[ [ E y ] ] = i k η 2 Z { { E } } and { { E y } } = 2 i k η Y [ [ E ] ] ,
[ [ H y ] ] = i k 2 Y η { { H } } and { { H y } } = 2 i k Z η [ [ H ] ] ,
[ [ u ] ] = u + u and { { u } } = u + + u ,
α = η 2 Z and β = 2 η Y
α = 1 2 Y η and β = 2 Z η
[ [ u y tot ] ] = i k α { { u tot } } and { { u y tot } } = i k β [ [ u tot ] ] on Γ ,
2 u scat + k 2 u scat = 0 in Ω + Ω ,
[ [ u y scat ] ] = i k α { { u scat } } 2 i k α u inc on Γ ,
{ { u y scat } } = i k β [ [ u scat ] ] 2 u y inc on Γ .
u scat ( r ) = C A ± ( k x ) e i k x x + i k 2 k x 2 | y | d k x for r = ( x , y ) Ω ± ,
r 2 G ( r | r ) + k 2 G ( r | r ) = δ ( r r ) , r = ( x , y ) Ω + Ω ,
[ [ G y ( r | r ) ] ] = i k α ( x ) { { G ( r | r ) } } , r Γ ,
{ { G y ( r | r ) } } = i k β ( x ) [ [ G ( r | r ) ] ] , r Γ ,
u scat ( r ) = { G ( r | s , 0 + ) f + inc ( s ) G ( r | s , 0 ) f inc ( s ) } d s , r Ω + Ω ,
f ± inc ( s ) = u y inc ( s , 0 ) ± i k α ( s ) u inc ( s , 0 )
u 0 tot ( r ) = u inc ( r ) + { G p ( r | s , 0 + ) f + inc ( s ) G p ( r | s , 0 ) f inc ( s ) } d s , r Ω + Ω ,
R ( k x , s ) : = 1 Γ ^ α ( k x , s ) Γ ^ β ( k x , s ) ,
Γ ^ α ( k x , s ) : = k α ( s ) k y + k α ( s ) , Γ ^ β ( k x , s ) : = k β ( s ) k y + k β ( s ) and k y = k 2 k x 2 .
G p ff ( r | r ) = i 4 H 0 ( 1 ) ( k | r r | ) incident field + i 4 R ( k cos ϑ , x ˜ ) H 0 ( 1 ) ( k | r r ¯ | ) reflected field ( r Ω + ) .
G p ff ( r | r ) = i 4 T ( k cos ϑ , x ˜ ) H 0 ( 1 ) ( k | r r | ) ( r Ω ) ,
T ( k x , s ) : = Γ ^ α ( k x , s ) + Γ ^ β ( k x , s ) .
u 0 tot , ff ( r ) = u inc ( r ) + { G p ff ( r | s , 0 + ) f + inc ( s ) G p ff ( r | s , 0 ) f inc ( s ) } d s , r Ω + Ω ,
R ( k ( x x ) | r r | , x ) and T ( k ( x x ) | r r | , x )
α ( x ) = c 0 ( 1 ± e i k x d ) and β ( x ) = c 0 ( 1 e i k x d ) ,
c 0 = sin θ inc and d = cos θ t cos θ inc ,
[ [ G p , y ( r | r ) ] ] = i k α ( x ) { { G p ( r | r ) } } and { { G p , y ( r | r ) } } = i k β ( x ) [ [ G p ( r | r ) ] ] , r Γ ,
u scat ( r ) = { G p ( r | σ , 0 + ) φ ( σ ) G p ( r | σ , 0 ) ψ ( σ ) } d σ , r Ω + Ω ,
μ j T j [ μ j ] = g j inc , j = 1 , 2 ,
φ = μ 1 + μ 2 2 and ψ = μ 2 μ 1 2 .
g 1 inc ( s ) = 2 i k α ( s ) u inc ( s , 0 ) and g 2 inc ( s ) = 2 u y inc ( s , 0 ) .
μ j = n = 0 T j n g j inc , j = 1 , 2 ,
φ N = μ 1 ( N ) + μ 2 ( N ) 2 and ψ N = μ 1 ( N ) μ 2 ( N ) 2 ,
μ j ( N ) ( s ) : = n = 0 N T j n g j inc ( s ) , j = 1 , 2 , N 0 .
μ N tot ( r ) : = u inc ( r ) + { G p ( r | σ , 0 + ) φ N ( σ ) G p ( r | σ , 0 ) ψ N ( σ ) } d σ and
μ N tot , ff ( r ) : = u inc ( r ) + { G p ff ( r | σ , 0 + ) φ N ( σ ) G p ff ( r | σ , 0 ) ψ N ( σ ) } d σ ,
μ 1 ( 0 ) ( s ) : = g 1 inc ( s ) = 2 i k α u inc ( s , 0 ) and μ 2 ( 0 ) ( s ) : = g 2 inc ( s ) = 2 u y inc ( s , 0 ) .
v ( r 1 ) w ( r 2 ) = Ω + { w 2 v v 2 w } d r = { w + v y + v + w y + } d s ,
0 = Ω { w 2 v v 2 w } d r = { w v y v w y } d s .
w + v y + v + w y + = w v y v w y .
v ( r 1 ) = { w + v y + v + w y + } d s = w ( r 2 )
[ [ G y ( r | r ) ] ] = k α ( x ) { { G ( r | r ) } } and { { G y ( r | r ) } } = k β ( x ) [ [ G ( r | r ) ] ]
{ G y ( r | s , 0 + ) u scat , + ( s ) G ( r | s , 0 + ) u y scat , + ( s ) } d s = { u scat ( r ) , r Ω + , 0 , r Ω ,
{ G y ( r | s , 0 ) u scat , ( s ) G ( r | s , 0 ) u y scat , ( s ) } d s = { 0 , r Ω + , u scat ( r ) , r Ω .
G y ( r | s , 0 ) = A ( s ) G ( r | s , 0 ) + B ( s ) G ( r | s , 0 + ) , G y ( r | s , 0 + ) = B ( s ) G ( r | s , 0 ) A ( s ) G ( r | s , 0 + ) , u y scat , ( s ) = A ( s ) u scat , ( s ) + B ( s ) u scat , + ( s ) f inc ( s ) and u y scat , + ( s ) = B ( s ) u scat , ( s ) A ( s ) u scat , + ( s ) f + inc ( s ) ,
A ( s ) = i k { α ( s ) + β ( s ) 2 } and B = i k { α ( s ) β ( s ) 2 }
G y ( r | s , 0 ) u scat , ( s ) G ( r | s , 0 ) u y scat , ( s ) = G ( r | s , 0 ) f inc ( s ) + B ( s ) ( G ( r | s , 0 + ) u scat , ( s ) G ( r | s , 0 ) u scat , + ( s ) )
G y ( r | s , 0 + ) u scat , + ( s ) G ( r | s , 0 + ) u y scat , + ( s ) = G ( r | s , 0 + ) f + inc ( s ) + B ( s ) ( G ( r | s , 0 + ) u scat , ( s ) G ( r | s , 0 ) u scat , + ( s ) )
G inc ( r | r ) : = i 4 H 0 ( 1 ) ( k | ( r r | ) = i 4 π e i k x ( x x ) + i k y | y y | k y d k x .
G p ( r | r ) = { G inc ( r | r ) + G r ( r | r ) , r Ω + , r Ω + , G t ( r | r ) , r Ω , r Ω + ,
G r ( r | r ) : = i 4 π R ( k x , x ) e i k x ( x x ) + i k y | y + y | k y d k x ,
G t ( r | r ) : = i 4 π T ( k x , x ) e i k x ( x x ) + i k y | y + y | k y d k x ,
i k y ( 1 R T ) = k α ( 1 + R + T ) , i k y ( 1 R + T ) = k β ( 1 + R T ) ,
G p ( r | r ) = { G t ( r | r ) , r Ω + , r Ω , G inc ( r | r ) + G r ( r | r ) , r Ω , r Ω .
g i e ( s , σ ) = i 4 π g ^ i e ( k x , σ ) e i k x ( s σ ) d k x and y g i e ( s , σ ) = i 4 π y g ^ i e ( k x , σ ) e i k x ( s σ ) d k x ,
g ^ + + ( k x , σ ) = g ^ ( k x , σ ) = 1 k y { 1 + R ( k x , σ ) } = 1 k y { 2 Γ ^ α ( k x , σ ) Γ ^ β ( k x , σ ) } , g ^ + ( k x , σ ) = g ^ + ( k x , σ ) = 1 k y T ( k x , σ ) = 1 k y { Γ ^ α ( k x , σ ) + Γ ^ β ( k x , σ ) } , y g ^ + + ( k x , σ ) = y g ^ ( k x , σ ) = i { 1 + R ( k x , σ ) } = 2 i i { Γ ^ α ( k x , σ ) + Γ ^ β ( k x , σ ) } y g ^ + ( k x , σ ) = y g ^ ( k x , σ ) = i T ( k x , σ ) = i { Γ ^ α ( k x , σ ) Γ ^ β ( k x , σ ) } ,
( S i e ϕ ) ( s ) = g i e ( s , σ ) ϕ ( σ ) d σ and ( y S i e ( ϕ ) ) ( s ) = y g i e ( s , σ ) ϕ ( σ ) d σ , s ,
[ [ u scat ] ] ( s ) = ( S + + S + ) ( φ + ψ ) ( s ) , { { u scat } } ( s ) = ( S + + + S + ) ( φ ψ ) ( s ) , [ [ u y scat ] ] ( s ) = ( y S + + y S + ) ( φ ψ ) ( s ) , { { u y scat } } ( s ) = ( y S + + + y S + ) ( φ + ψ ) ( s ) ,
{ y S + + y S + + i k α ( s ) ( S + + + S + ) } μ 1 ( s ) = 2 i k α ( s ) u inc ( s , 0 )
{ y S + + + y S + + i k β ( s ) ( S + + S + ) } μ 2 ( s ) = 2 u y inc ( s , 0 )
y g + + ( s , σ ) y g + ( s , σ ) + i k α ( s ) ( g + + ( s , σ ) + g + ( s , σ ) ) = δ s + K 1 ( s , σ ) , y g + + ( s , σ ) + y g + ( s , σ ) + i k β ( s ) ( g + + ( s , σ ) g + ( s , σ ) ) = δ s + K 2 ( s , σ ) ,
K 1 ( s , σ ) : = k 2 π { α ( σ ) α ( s ) } e i k x ( s σ ) k y + k α ( σ ) d k x ,
K 2 ( s , σ ) : = k 2 π { β ( σ ) β ( s ) } e i k x ( s σ ) k y + k β ( σ ) d k x .
T j [ μ ] ( s ) : = K j ( s , σ ) μ ( σ ) d σ , s , j = 1 , 2 ,
α ( s ) = α 0 + α ˜ ( s ) and β ( s ) = β 0 + β ˜ ( s ) , s ,
u 0 ( r ) = { e i k inc r + R 0 ( θ inc ) e i k r r , r Ω + , T 0 ( θ inc ) e i k inc r , r Ω ,
R 0 ( θ ) : = 1 α 0 | sin θ | + α 0 β 0 | sin θ | + β 0 and T 0 ( θ ) : = α 0 | sin θ | + α 0 + β 0 | sin θ | + β 0 ,
[ [ v y ] ] = i k α { { v } } i k α ˜ { { u 0 } } and { { v y } } = i k β [ [ v ] ] i k β ˜ [ [ u 0 ] ] on Γ .
v 0 ( r ) = { G 0 ( r | σ , 0 + ) φ 0 ( σ ) G 0 ( r | σ , 0 ) ψ 0 ( σ ) } d σ , r 2 \ Γ ,
A ˜ ( s ) = i k { α ˜ ( s ) + β ˜ ( s ) 2 } and B ˜ ( s ) = i k { α ˜ ( s ) β ˜ ( s ) 2 } ,
φ 0 ( s ) = { A ˜ ( s ) T 0 ( θ inc ) + B ˜ ( s ) ( 1 + R 0 ( θ inc ) ) } e i k s cos θ inc , ψ 0 ( s ) = { B ˜ ( s ) T 0 ( θ inc ) + A ˜ ( s ) ( 1 + R 0 ( θ inc ) ) } e i k s cos θ inc .
v 0 ( r ) ~ e i k | r | | r | v ( θ , θ inc ) as | r | ,
v ( θ , θ inc ) = 2 k π e i 3 π 4 | sin ( θ inc ) sin θ | { v α ( θ , θ inc ) + v β ( θ , θ inc ) θ ( 0 , π ) , v α ( θ , θ inc ) v β ( θ , θ inc ) θ ( π , 0 ) ,
v α ( θ , θ inc ) : = α ˜ ( σ ) e i k σ ( cos θ inc cos θ ) ( | sin θ | + α 0 ) ( | sin θ inc | + α 0 ) d σ and v β ( θ , θ inc ) : = β ˜ ( σ ) e i k σ ( cos θ inc cos θ ) ( | sin θ | + β 0 ) ( | sin θ inc | + β 0 ) d σ .
α ˜ ( s ) = c α e i k d α s and β ˜ ( s ) = c β e i k d β s ,
d α = cos θ α cos θ inc and d β = cos θ β cos θ inc .
μ j μ j ( N ) μ j T j N + 1 c [ ρ ( T j ) ] N + 1
α ( x ) = a ( ε x ) : = = 0 10 c e i ε x , x ,
G r ( r | r ) = SC q r ( k x , r ) e | r | ϕ ( k x ) d k x and G t ( r | r ) = SC q t ( k x , r ) e | r | ϕ ( k x ) d k x
ϕ ( k x ) : = i k x cos θ + i k y | sin θ | ,
q r ( k x , r ) : = i R ( x , k x ) 4 π e i k x x + i k y y k y ,
q t ( k x , r ) : = i T ( x , k x ) 4 π e i k x x + i k y y k y ,
Im k x = Re k x | cot θ | k | csc θ | as | k x | if x > 0 ,
Im k x = Re k x | cot θ | + k | csc θ | as | k x | if x < 0 .
k y ( k x ) = k α ( x ) or k y ( k x ) = k β ( x ) .
k x ( 1 ) = k 1 α 2 or k x 2 = k 1 β 2 , and 0 Im k x ( p ) Re k x ( p ) | cot θ | k | csc θ | if x > 0 ,
k x ( 1 ) = k 1 α 2 or k x 2 = k 1 β 2 , and Re k x ( p ) | cot θ | + k | csc θ | Im k x ( p ) 0 if x < 0 .
G r ( r | r ) = e i k | r | i k r r ¯ / | r | + i π 4 8 π k | r | R ( k x | r | , x ) + p = 1 , 2 A p r e i k x ( p ) | x x | + i k y ( k x ( p ) ) | y + y | + O ( | r | 3 / 2 )
G t ( r | r ) = e i k | r | i k r r ¯ / | r | + i π 4 8 π k | r | T ( k x | r | , x ) + p = 1 , 2 A p t e i k x ( p ) | x x | + i k y ( k x ( p ) ) | y y | + O ( | r | 3 / 2 )
A 1 r = A 1 t = α 2 1 α 2
A 2 r = A 2 t = β 2 1 β 2
α 2 1 α 2 e i k 1 α 2 | x x | i k α | y ± y | ,
i 4 H 0 ( 1 ) ( k | r r | ) = e i k | r | i k r r ¯ / | r | + i π 4 8 π k | r | + O | r | 3 / 2 as | r | ,
G p ff ( r | s , 0 ± ) = G p ( r | s , 0 ± ) + O ( | r | 3 / 2 ) as | y | .
Φ ( d 1 , d 2 ) = i 4 π k γ k y + k γ e i ( k x d 1 + k y d 2 ) / k k y d k x , d 1 d 2 0 ,
f ( ξ ) = i γ π cos ( ξ d 1 ) 1 ξ 2 + γ e i d 2 1 ξ 2 1 ξ 2 ,
ζ ( t ) : = { L ( 1 + cos ( t + π ) ) 2 + i H sin ( t + π ) : t [ 0 , π ] }
I 2 L A + L f ( t ) w A ( t L ) d t ,
η ( t ; t 0 , t 1 ) = { 1 , | t | t 0 , exp ( 2 e 1 / u u 1 ) , t 0 < | t | < t 1 , u = | t | t 0 t 1 t 0 , 0 , | t | > t 1 ,
Ψ ( d ) = e i k x d / k k y + k γ d k x , d 0 ,
H 0 ( 1 ) ( d ) = 1 π e i k x d / k k y d k x = 1 π e i ξ d 1 ξ 2 d ξ , d 0 ,
Ψ ( d ) = π H 0 ( 1 ) ( d ) k γ e i k x d / k k y ( k y + k γ ) d k x .
e i k x d / k k 2 k x 2 d k x = i π 2 k e i k x d / k , d 0 ,
e i k x d / k k y ( k y + k γ ) d k x = i π 2 k e i k d / k k γ e i k x d / k ( k 2 k x 2 ) ( k y + k γ ) d k x .
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.