Volume 2016, Issue 1 7436136
Research Article
Open Access

Upscaling of Helmholtz Equation Originating in Transmission through Metallic Gratings in Metamaterials

Hari Shankar Mahato

Corresponding Author

Hari Shankar Mahato

College of Engineering, University of Georgia, Athens, GA 30602, USA uga.edu

Search for more papers by this author
First published: 21 September 2016
Academic Editor: Tongxing Li

Abstract

We investigate the transmission properties of a metallic layer with narrow slits. We consider (time-harmonic) Maxwell’s equations in the H-parallel case with a fixed incident wavelength. We denote η > 0 as the typical size of the complex structure and obtain the effective equations by letting η → 0. For metallic permittivities with negative real part, plasmonic waves can be excited on the surfaces of the slits. For the waves to be in resonance with the height of the metallic layer, the corresponding results can be perfect transmission through the layer.

1. Introduction

Negative refraction of electromagnetic waves in metamaterials has become of major interest in recent years, compare [1, 2], especially to construct small scale optical devices for technical applications in the fields of micro- and nanooptics. Metamaterials are the materials that are not found in nature; instead they are created by the composition of several metals or plastics or both. Due to their precise shape, size, geometry, and arrangement of metals, these metamaterials are capable of influencing the electromagnetic waves by absorbing, bending, or refracting. To create the metamaterials, the composite materials are arranged in repeated (periodic) fashion with periodicity scales smaller than the wavelength of waves. Negative index metamaterial or negative index material (NIM) is a metamaterial where the refractive index (in optics theory, the refractive index of a material is a dimensionless number which describes how light propagates through that medium and is defined as the ratio c/υ, where c is the speed of light in vacuum and υ is the phase velocity of light in the medium) has a negative value over some frequency range when an electromagnetic wave passes through it. Negative index materials are extensively studied in the fields on optics, electromagnetics, microwave engineering, material sciences, semiconductor engineering, and several others.

In this work, we study the phenomena of light wave passing through the subwavelength metallic structure; that is, we investigate the high transmission of light wave through a metamaterial with thin holes inside it. We consider a thin metallic structure (inside a medium) with holes smaller than the wavelength of incident photon which shows the high transmission of light waves through this metallic structure. This high transmission contradicts the classical aperture theory and shows an important feature of metamaterials. To demonstrate the geometry assumed in this work, let us consider Figure 1 where the light wave emerging from a source (l.h.s of the figure) is passing through a metamaterial with negative refractive index and its image is given on the r.h.s. (cf. this figure to that of [3]).

Details are in the caption following the image
The light wave is originating from the left side and is getting refracted through the metallic plate and its image is on the right (this image is taken from [2]).
Details are in the caption following the image
The light wave is originating from the left side and is getting refracted through the metallic plate and its image is on the right (this image is taken from [2]).

The holes inside the metallic layer are periodically distributed with period η > 0 smaller than the wavelength λ of incident light wave. This layer can be considered as a heterogeneous or perforated media and our goal is to give a physically consistent approach to transmission properties of heterogeneous media using the techniques from homogenization theory and applied analysis. We obtain an effective (upscaled) scaterring problem where the metallic layer with holes is replaced by a homogenized structure with effective permittivity εeff and permeability μeff. We also obtain the tranmission coefficient T in terms of incident wave number k and incident angle θ. We will see that, for lossless materials with (real) negative permittivity εη, perfect transmission T = 1 can be obtained for every θ and suitable value for k. In the recent times, several significant investigations for metamaterials have been done. In [4] the connection between the high transmission and the excitation of surface plasmon polaritons has been established. The photonic band structure of the surface plasmons is evaluated numerically, In [5] the authors have calculated the transmission coefficients for the lamellar gratings, while the effect of surface plasmons on the upper and lower boundary of the layer is investigated in [6]. In [7], the effect of finite conductivity is studied. In [8], the relation between the high transmission effect and the negative index material is obtained with a fishnet like structure. A homogenization method is proposed in [9] where the author accentuates the connection between the skin depth of evanescent modes in the metallaic structure and the period of the gratings. Some results in this direction can be found in [1013].

Two-scale convergence has proven to be a very efficient tool in homogenization theory while dealing with the problems where the underlying medium is heterogenous. The concept of two-scale convergence is first introduced by Nguetseng in [14]. This convergence criterion and the results related to it have been used extensively in the homogenization of partial differential equations; see Allaire [15, 16], Cioranescu and Donato [17], and Mahato and Böhm [18]. In this work too, we have used the two-scale convergence of an oscillating sequence and its gradient; see Section 1.4. At this point we would like to point out that the geometry of metallic structure in this work is generalized compared to that considered in [3]. In [3] due to the rectangular shapes of the metallic gratings, the coefficients of the effective system were determined by the help of a scalar, one-dimensional shape function given by the hyperbolic functions; however in our work where the considered geometry is more realistic, such nice representation is not possible. To deal with this problem in this work an eigenvalue approach has been proposed and this eigenvalue problem in the unit cell Y helps us to determine the effective parameters of the problem.

Although this paper can be compared with [3] in some way, the major difference in this work is that our limit function of (see Section 1.5 for details) and the limit function in [3] are totally different. In [3], the authors worked with a rectangular metallic subpart Σ of type (−γ, γ)×(−1/2, 1/2) and therefore by defining a suitable test function, they have shown that the first component of j0 vanishes and they obtain . This is clearly not the case in this paper as the metallic subpart Σ is chosen to be sinusoidal along y2-axis due to the geometry of the metallic structure given by Figure 1. This will lead to nonvanishing componenets of j0 and we end up having a different j0 compared to that in [3] and hence, we will obtain a different upscaled equation. Also no explicit representation of can be obtained due to the geometry of Σ.

In Sections 1.1, 1.2, and 1.3, we will outline the model in detail. In Section 2, we gather some mathematical tools required to do the analysis and we state our main results. In Sections 3 and 4 we will prove the main results.

1.1. Model

We investigate the time-harmonic solutions of a Maxwell equations with a fixed wave number k and the corresponding wave length λ = 2π/k. Let the metallic structure remain unchanged towards x3-direction and the metallic field, denoted by , is parellel to x3; that is, , where .

The heterogenoeus domain Ω has a metallic structure of finite length and finite height in , and the slits (vacuum) are repeated periodically with a small period η > 0, compare Figure 1. The period η is assumed to be infinitesimally small with respect to the wavelength λ. The relative permittivity of the metal is denoted by ε. Since the permittivity of conductors has large absolute values, we assume that it depends on η and consider ε = εη. We obtain nontrivial effects due to plasmonic resonance for |εη| ∝ η−2, compare [3, 13]. If Ση denotes the matallic part in Ω, we set
()
where . Due to ohmic losses inside the metal, Im⁡(εη) is always assumed to be positive in a physical system which means we always take Im⁡(εr) ≥ 0 and εr ≠ 0. A material is called a lossless material if Im⁡(εr) = 0. Our particular interest is to study a lossless material with negative relative permittivity; that is, Im⁡(εr) = 0 and Re⁡(εr) < 0. For such εr transverse evanescent modes will be generated in the metal. Since Re⁡(εr) < 0 and Im(εr) = 0, then from (14) we have ; that is, we can obtain wave like solutions and waves cannot penetrate the metallic grating. These evanescent modes can penetrate only in a region which is given by the skin depth of order η. The evanascent mode is related to a surface plasmon solution (in this case a solution which is nonvanishing in the grating but which has exponential decay in the metal). The main aspect of the current work is to generalize the geometric structure of the metallic slab inside Ω given in [3].

1.2. Geometry

Let η > 0 be a small scale parameter and Ω be the domain under investigation which is bounded in . Let Y≔(−1/2, 1/2)×(−1/2, 1/2) be the representative unit cell in and Σ be an open set in Y such that and YΣ ∪ (YΣ). Let us choose Σ in such a way that it follows a sinusoidal profile along y2-axis; that is,
()
Keeping physics of the problem in mind, Σ denotes the mettalic part which lies between the two columns of holes in the metallic structure of type introduced in Figure 1.
The relative aperture volume α = 1 − 3γ/2 ∈ (1/4, 1) and relative metal volume is (3/2)γ, where γ ∈ (0, 1/2). We define 2Nη + 1≔2l/2ηγ. We assume that the compact rectangle contains (2Nη + 1) number of small rectangles of type (nηηγ, nη + ηγ) × (−h, 0), that is, of width 2γη and height h, which include the η-scaled versions of the metallic part Σ (cf. Figures 2 and 3), where each Σ is of width (3/2 )γη and height h. The collection of these small η-scaled versions of the metallic part is the metallic domain Ση and assume that the two-dimensional heterogeneous metallic structure introduced in Figure 1, denoted by Ση and parellel to x1-axis, is contained in the closure of the set R≔(−l, l)×(−h, 0) ⊂ Ω with ; that is;
()
see Figure 3.
Details are in the caption following the image
The representative cell (cf. this figure with figure  1 in [3]).
Details are in the caption following the image
The metallic structure Ση inside the domain Ω.

As η → 0, Nη. Due to nondimensionalization, we are, however, only interested in h = 1.

1.3. Function Spaces

Let θ ∈ [0,1] and 1 ≤ r, s be such that 1/r + 1/s = 1. Assume that Ξ ∈ {Ω, Ση} and ; then as usual Lr(Ξ) and Hl,r(Ξ) denote the Lebesgue and Sobolev spaces with their usual norms and they are denoted by ‖·‖r and ‖·‖l,r. For the sake of clarity if ϕLr(Ξ), then
()
and if ϕHl,r(Ξ), then
()
where is a multi-index, |α | = α1 + α2 + ⋯+αn, and . Similarly, , (·, ·) θ,r, and [·, ·] θ are the Hölder, real, and complex interpolation spaces, respectively, endowed with their standard norms; for definition confer [19, 20]. denotes the set of all Y-periodic α-times continuously differentiable functions in y for . In particular, C#(Y) is the space of all the Y-periodic continuous function in y. The C-spaces are as usual equipped with their maximum norm whereas the space of all continuous functions C(Ξ) is furnished with supremum norm, compare in [19].

1.4. Two-Scale Convergence

Definition 1. A sequence of functions (uη) η>0 in Lr((0, T) × Ω) is said to be two-scale convergent to a limit uLr((0, T) × Ω × Y) if

()
for all ϕLs((0, T) × Ω; C#(Y)).

By ,, and → we denote the two-scale, weak, and strong convergence of a sequence, respectively. Finally, S = (0, T) denotes the time interval.

Lemma 2 (cf. [21]). For every bounded sequence (uη) η>0 in Lr(S × Ω) there exists a subsequence (uη) η>0 (still denoted by same symbol) and uLr((0, T) × Ω × Y) such that .

Lemma 3 (cf. [21]). Let (uη) η>0 be strongly convergent to uLr((0, T) × Ω), and then , where u1(t, x, y) = u(t, x).

Lemma 4 (cf. [21]). Let (uη) η>0 be a sequence in Lr((0, T); H1,r(Ω)) such that in Lr((0, T); H1,r(Ω)). Then and there exists a subsequence (uη) η>0, still denoted by same symbol, and such that .

Lemma 5. Let (uη) η>0 be a bounded sequence of functions in Lr(S × Ω) such that ηuη and η1/2uη are bounded in Lr(S × Ω)n. Then there exists some functions such that , , and .

Proof. (i) Since uη and ηuη are bounded sequence of functions in Lr(S × Ω) and Lr(S; Lr(Ω)) n, respectively, then there exists uLr(S × Ω × Y) and ULr(S × Ω × Y) n such that and as η → 0. This means that, for the sequence ηuη, we have

()
for all . We integrate by parts the l.h.s, and which gives
()
And it follows from (7) and (8) that U(t, x, y) = ∇yu(t, x, y).

(ii) To prove the second part of the lemma, let us choose , where and . Note that the boundedness of η1/2uη implies the boundedness of ηuη in Lr(S × Ω) and hence, by part (i) there exists such that and . Now let us assume that , and then by definition

()
We integrate by parts the l.h.s.; then
()
We compare (9) and (10) which leads us to
()
Since ϕ1 is independent of y and ∇y · ϕ1(t, x) = 0, from (11) it follows that U must be the gradient of some function such that U(t, x, y) = ∇yu1(t, x, y); that is, . This completes the proof.

1.5. Mathematical Formulation and Statement of the Main Results

We study the Maxwell equations in a complex geometry with highly oscillating permittivities. By η we denote (i) the dimensionless positive scale parameter which represents the small length scale in the geometry and (ii) the oscillations of large absolute values of the permittivity. We follow the standard nondimenionsalization techniques; for instance, see [3, 18, 22], and so forth and from here on all the quantities considered in this work are dimensionless unless stated otherwise. For the electric field and magnetic field , the time-harmonic Maxwell equations are
()
with fixed positive real constants ω,    μ0, and ε0 denoting the frequency of the incident waves and the permeability and the permittivity of vacuum, respectively. We postulate that all the quantities are x3-independent and the polarized magnetic field is given by , where . By orthogonal property of and , we have . Then (12) reduce to
()
By (13), a straightforward calculation yields
()
where we have set k2 = ω2ε0μ0. We define the coefficient which can have a negative real part and that it vanishes in the metal as η → 0. Thus we have the desired Helmholtz equation which we will study in this paper and is given below. We study solutions of
()
where the coefficient aη is given by
()
The set Ση ⊂ ⊂RΩ describes the complex geometry of the metallic inclusion in Ω; see Figure 3.

Remark 6 (scattering problem). We will investigate the effective behavior of solutions of (15) in two different cases. In the first case we will study an arbitrary bounded sequence of solutions on a bounded domain Ω while the second one concerns the scattering problem. In other words we consider (15) in whole of . For a given incident wave ui, which solves ∇2ui = −k2ui in , we take the Sommerfeld condition as the boundary condition which says that the scattered field satisfies

()
for r = |x | → , uniformly in the angle variable.

Remark 7. Note that for (15) we have not given any boundary conditions; instead we have considered an arbitrary sequence of solutions; however, the uniqueness of solution of the scattering problem will be proven for every η. To state the main results, we rewrite (15) as a system:

()
Comparing with (12), we see that jη represents (up to a factor and perhaps a rotation) the horizontal electric field and since the magnetic field , system (18) is nothing but (12) itself.

Theorem 8 (upscaled equations). Let the matallic geometry be given by Ση (Figure 3) on a domain and let the coefficient be as in (16). On we assume that either Im⁡(εr) > 0 or εr < 0 = Im⁡(εr). Let (uη) η>0 be the sequence of solutions of (15) such that in L2(Ω) for η → 0. We define UL2(Ω) as the function

()
where Nw is defined by (36). Then the function ∇UL2(Ω). The field jη = aηuη converges weakly to some j in which is given by
()
Moreover, the limit functions satisfy the system
()
where
()

By applying Theorem 8 for with a large radius r0 > 0, we can treat the scattering problem with an incoming wave generated at infinity. We obtain the strong convergence of the scattered field outside the metallic obstacle and we identify the limit U(x) as the solution of the effective diffraction problem. We define the exterior domain outside of R as .

Theorem 9 (effective scattering problem). Let the metallic gratings be given by Ση (Figure 1) and the coefficient be as in (16). Assume further that ui is an incident wave solving the free space equation ∇2ui = −k2ui on and uη is the unique sequence of solutions to (15) such that satisfies (17) and that the solution sequence satisfies the uniform bound

()
Then uηU  strongly in with uniform convergence for all derivatives on any compact subset of Rext. The effective field is determined as the unique solution of the upscaled equation
()
with (17) for the scattered field (Uui).

1.5.1. Interface Conditions

The homogenized equation (24) should be understood in the sense of distributions on the whole of . The exterior field for every large radius r; hence its trace on R from outside, denoted by U+, is a well-defined element of H1/2(R). Note that as ∇U belongs to L2(Br), the function U(·, ·) is an element of . This helps us to define traces of U on the horizontal boundary parts from the inside. Moreover, we have the information that the distributional divergence of the vector field j = aeffU is of class .

We define the transmission condition on the boundary R of R with using traces from inside and outside of R. We denote by superscript + (resp., by −) traces from outside (resp., by inside); then problem (24) can be rewritten as
()
with the transmission (interface) conditions
()
where Nw is defined in (36).

2. Derivation of the Effective Model

2.1. A Priori Estimates

Lemma 10. For an with Im(εr) > 0, let aη be defined as in (16). Then there exists a such that

()
where C0 > 0 is independent of η.

Proof. Let xΣη (if εr = a + ib with b > 0, then aη = η2((aib)/(a2 + b2))). For an arbitrary small δ > 0, let us define λ≔−1 + δi such that δ|Re(aη)| ≤ −(1/2)Im⁡(λaη). There exists a constant C0 > 0 such that

()

For xΩΣη, we have

()

Lemma 11 (gradient estimate). Suppose that the solution (uη) η>0 of (15) is a bounded sequence in L2(Ω); that is, . Then for every compactly contained subdomain Ω ⊂ ⊂Ω, the following estimate holds:

()
where C is independent of the scale parameter η.

Proof. Since , there exists a subdomain Ω ⊂ ⊂Ω such that . Without loss of generality, let us assume that ΣηΩ and take a cut-off function , where Θ(x) = 1 on Ω. We test (15) with , where is the complex conjugate of uη(·). This gives

()
We employ Lemma 10. For a , we multiply (31) by λ and equate its imaginary part and rearrange the factors of the second integrand which will yield
()
where in the second step we used Young’s inequality. We see that the first integral on the r.h.s. of (32) is bounded by the L2-boundedness assumption on uη whereas the second integral on the r.h.s. is bounded by the boundedness of |aη| and . Using the fact that Θ(x) = 1 on Ω, we have , where C is independent of η and uη.

2.2. An Eigenvalue Problem in the Unit Cell Y

Let us consider the eigenvalues 0 < λ1λ2 ≤ ⋯≤λn ≤ ⋯ of the problem
()
and we denote {ζn} the associated normalized eigenfunctions in , so that {ζn} is an orthonormal basis of L2(Σ). Since with Im(εr) > 0, k2εr satisfies the condition
()
We set
()
()
Let us consider the following boundary value problem:
()
By [23, theorem  8.22], it follows that (i) Mw(y) is a solution of (37) and this solution is unique if k2εrλn for all n and (ii) if condition (34) is not fulfilled then (37) has no solution.

In the next theorem we will analyze the behavior of uη as η → 0 in the sense of two-scale convergence, compare [15]. We notice that the geometry is not only periodic in the x1-direction but it is also periodic with respect to to the cell Y≔(−1/2, 1/2)×(−1/2, 1/2). The metal part in the cell Y is given by ΣY; see Figure 2.

We recall that the sequnce (uη) η>0 is weakly convergent to uL2(Ω). We define a function u0(x, y)≔u0(x1, x2, y1, y2) as
()
where w(y) is a Y-periodic function defined in (35). We have defined u0 in such a way that, for every xΩ, there holds u(x) = ∫Yu0(x, y)dy. We will show in next theorem that as η → 0.

Lemma 12 (two-scale limit). Let (uη) η>0, weakly converging to u in L2(Ω), be a sequence of solutions of (15). Then for the function u0 defined in (38) it holds that .

Outside of R, the strong convergence uηu holds in L2(ΩR). More precisely, uη together with all its derivatives converges uniformly on every compact subset .

Proof. We divide the proof into three steps.

(i) From the assumption on uη and the estimate (30), the sequences (uη) η>0 and (ηuη) η>0 are bounded in L2(Ω). Then there exists such that, up to a subsequence, and as η → 0. As a Y-periodic function, u0 and ∇yu0 can be extended by periodicity to all . This shows that which implies that u0(x, ·) belongs to , in particular, in H1(Σ) and has a trace on Σ. In other words, does not jump accross Σ by trace theorem (cf. [19, theorem  5.5.1]).

Next, we investigate the coefficient aη = 1 on the set ΩΣη. From (30), it follows that which implies strongly in L2(Ω). Since strong convergence implies the two-scale convergence, by localisation Lemma (cf. [3]) the two-scale limit χ0 vanishes a.e. in R × (YΣ) and in (ΩR) × Y. Due to ∇yu0 = χ0, it implies that the function u0(x, ·) is constant in YΣ and for xR; and it is constant everywhere for xR. We use this y-independence to define a function UL2(Ω) as

()
We note that, at this stage of the proof, u0 and U are defined as the two-scale limit of uη and by (39), respectively.

(ii) Characterisation of Two-Scale Limit for xR. We claim that, for a.e. xR, the function u0(x, ·), which belongs to H1(Σ), solves the linear boundary value problem

()
()
()
where (40a), (40b), and (40c) hold in the distributional sense in Σ. To verify this, we choose φ(x, y) = Θ(x)ψ(y), where and ψC(Y; [0,1]) a periodic function on Σ with supp⁡(ψ)∩(YΣ) = ϕ. Using φη(x) = φ(x, x/η) as the test function in (15), we obtain
()
Passing the two-scale limit as η → 0
()
Since Θ was chosen arbitrarily, (40a), (40b), and (40c) hold. For every xR, we write u0(x, ·) = U(x)(1 + w(·)), where and w(·) is Y-periodic. Clearly, satisfies the equation
()
()
()

Then, for k2εrλn, as shown in Section 2.2, we express w uniquely in terms of the orthonormal basis {ζn}. Note that if the condition (34) is violated, the equation in w has no solution and we are led to u0(x, y) = U(x) = 0.

Therefore, to sum up, we obtain the two-scale limit as

ζn(y)∫Σζn(y)dy), provided (34) holds. Consequently, for xR, the weak limit u satisfies

()
Therefore the two-scale limit is given by
()
where .

(iii) Strong Convergence Outside of R. We know that u0(x, y) = u(x) = U(x) holds for a.e. xΩR and for all yY. Moreover, by the assumption on uη and estimate (30), we have . This then implies that uη, up to a subsequence, is strongly convergent to U in L2(ΩR) by Aubin-Lion’s Lemma, compare [24]. The uniform convergence on compact subsets of uη and of all its derivatives is a consequence of the fact that uη Helmholtz equation Δuη + k2uη = 0.

With the help of Lemma 12, we can completely determine the two-scale limit of the sequence (uη) η>0 if we know the function U(x) which is defined in (39). Now we collect the properties of , its weak limit , and its two-scale limit .

Proposition 13. Let be as in Lemma 12 and U be given by (39). For , we suppose that in . Then is characterized as follows:

  • (i)

    The sequence converges in the sense of two scales to which is given by

    ()

  • (ii)

    The limit ∇UL2(Ω) and it holds:

    ()

Remark 14. We would like to point out a major difference in our j0 and the limit function j0 in [3]. In [3], the authors worked with a rectangular metallic subpart Σ of type (−γ, γ)×(−1/2, 1/2) and therefore by defining a suitable test function, they have shown that the first component of j0 vanishes and they obtain . This is clearly not the case in this paper as the metallic subpart Σ is chosen to be sinusoidal along y2-axis due to geometry of the metallic structure given by Figure 1. This will lead to nonvanishing componenets of j0 and we end up having a different j0 compared to that in [3] and hence, we will obtain a different upscaled equation.

Proof. By (30), it follows that is bounded in which implies that up to a subsequence two-scale converges to some . The weak limit would then be given as .

The Field outside of R. For , . Then by Lemma 12, ∇uη → ∇U uniformly on compact subsets of . This leads to

()

The Field in the Metal Part of R. We note that |aη| ≤ Cη2 in Ση and |aη| ≤ C in RΣη; therefore (30) gives and . This implies and by [21, theorem  17], we have a.e. in R × Σ. Moreover, j0(x, y) = j0(x) = ∇U(x) for a.e. (x, y) ∈ R × (YΣ).

Divergence of j0. Due to boundedness assumption on uη, by (18) we have . For and , we test (18) by Θ(x)ψ(x/η) which gives

()
Since is arbitrary, ∫Y ∇yψ(y) · j0(x, y)  dy = 0 for a.e. xΩ which implies ∇y · j0(x, ·) = 0 for yY in distributional sense. This shows that is independent of y; that is, , some function in x only.

Next we determine the relation between and U as shown in [3]. We define , ψ is Y-periodic, and ψ = 0  in  Σ}. We choose a test function φ(x, x/η)≔Θ(x)ψ(x/η), where and ψ ∈ Ξ. We use and ∇·φ(x, x/η)≔∇·(Θ(x)ψ(x/η)) = ψ(x/η)·∇Θ(x) as ∇y · ψ(y) = 0. Then

()
Since ψ(y)|Σ = 0 and is nonvanishing in YΣ and by (39) it implies that u0(x, y) = U(x), all these lead to
()
Therefore, j0(x, y) = ∇U(x) for and for reminder j0(x, y) = 0 for (x, y) ∈ R × Σ.

Proof of (i). To conclude this part, the arguments rely on that of [3]. We consider . We intend to show that for almost every xR. To show this, we define a function . We notice that (i) ∫Yj0(x, y)dy = ∫Yj(x)dy = j(x) and , (ii) ∇y · j0(·, y) = 0 and ∇y · j1(·, y) = 0, and (iii) . This implies that (51) holds good for as well as for the conjugate of . Therefore using the fact that from (51), we have

()
()
Substraction of (52a) and (52b) gives . Since Θ is arbitrary, therefore ψ0(y) = 0, which shows that . This completes the proof of part (i).

Proof of (ii). To verify the claim, let us choose and . Note that . Then from (51), we have

()
It follows that ∇UL2(Ω) and we find also that for xR and for xΩR; that is, , where a(x) = α for xR and a(x) = 1 for xΩR.

3. Proofs of Theorems 8 and 9

Proof of Theorem 8. The proof of Theorem 8 is a straightforward consequence of Lemmas 11 and 12 and Proposition 13. It is being shown that if, for any subdomain Ω with RΩ ⊂ ⊂Ω, is bounded in L2(Ω), then, up to a subsequence, is weakly convergent to some in L2(Ω).

By Proposition 13, we have the relation between U and ; that is, the weak and the two-scale limits of are given in terms of U; see (46) and (47). Since Ω is arbitrary, the results of Proposition 13 hold good in all Ω. Now we obatin the limit problem by dividing the proof into two following cases.

Case 1. Let ; then for from (15) we have

()

Case 2. Let xR; then again for from (15) we have

()

The combination of (54) and (55) gives the limit problem as

()
where
()

Here we can compare our upscaled equation with the limit problem obtained in [3], especially for xR. Due to their rectangular metallic gratings inside R, the component along x1 direction vanishes; that is, the first component of and thus the authors obtained their upscaled equation as .

Proof of Theorem 9. The proof is devided into three steps which are demonstrated below.

(i) Uniqueness of the Limit Problem. With a fixed incident field ui we will show that the limit problem (24) has a unique solution. On the contrary, let us assume that U1 and U2 are the two solutions of (24) and set u = U1U2. We consider the equations satisfied by difference of two solutions as

()
()
We claim that . The main ingredient for this uniqueness result is Rellich’s first lemma and the fact that aeff is real and μeff has positive imaginary part. In fact aeff is identity and μeff is 1 outside of R. Let us denote the surface of a sphere Br(0) of radius r by Sr  (≔Br(0)), where r is chosen so large such that RBr(0). Let r0 be such r; then by (59), we have
()
This gives
()
Now we multiply (58) by and integrate over . Since (58) holds only in the sense of distributions and due to possible jumps on Γhor, we approximate by smooth functions. By divergence theorem we have
()
The surface integral on r.h.s. of (62) is well defined. This can be argued as follows: outside of R, u is a solution of the Helmholtz equation ∇2u = −k2u and so it is analytic in the exterior of R. Therefore the traces of u and ru are well defined in , compare [25]. Comparing the imaginary parts of (62) and investing the knowledge of Im(μeff) > 0, then
()
Therefore from (63) we have u = 0 in R. Since r0 is chosen arbitrarly, for every r from (60) it follows that
()

Thus by Rellich’s first lemma (which states that the solutions u of the Helmholtz equation on an exterior domain satisfying property (64) vanish) we obtain u = 0 in all of which concludes the proof of the uniqueness property, compare [25].

(ii) Convergence to the Limit Problem Assuming an -Bound. Let the radius r0 > 0 be such that and set . We begin with the assumption that

()
The proof basically follows as the one for Theorem 8. Using (65), up to a subsequence, passing the limit as η → 0, we obtain that solves (24).

We only need to verify the radiation condition (17). By Lemma 12 it follows that uη and ∇uη are uniformly convergent on every compact subset of . Let us choose r < r0 such that R ⊂ ⊂Br(0)⊂⊂Ω. By [25, theorem  2.4] and end remark of that theorem, we have from the Sommerfeld radiation condition that the scattered field coincides on with its Helmholtz representation through values and derivatives of uηui on Br(0). By the similar representation formula, using the values and derivatives of Uui on Br(0), we can extend U into all of to a solution of the Helmholtz equation ∇2U = −k2U outside of R. Thus this construction of U shows that Uui satisfies the Sommerfeld radiation condition. The uniform convergence of uηU and ∇uη → ∇U on Br(0) implies the uniform convergence of uη and its derivatives on all compact subsets of exterior of R. Finally by uniqueness of the limit from part (i), as η → 0 for the whole sequence. This shows that the Sommerfeld radiation condition holds for r = |x | → . which establishes (17).

(iii) Boundedness of tη. In the previous step the limit problem is obtained assuming (65) is true. We will prove that (65) holds true by the method of contradiction. We suppose that tη, up to a subsequence, as η → 0. Now we consider the normalized sequence

()
Due to linearity, vη solves the original scaterring field problem with incident field as η. Following the proofs of Lemma 12 and parts (i) and (ii), the function is the unique solution of (24) and satisfies the Sommerfeld wave condition. By the construction of vη, we obtain V = 0 and therefore vη → 0 weakly in L2(Ω).

For outside of R, the gradient estimate (30) for vη remains valid and hence, vη|ΩR remains in a bounded subset of H1(ΩR). Then by Rellich compactness lemma as η → 0. For inside of R, we use the estimate (23) on uη and since tη, as η → 0. Therefore as η → 0 but this contradicts the fact that . Thus tη has to be bounded.

4. Transmission Properties of the Effective Layer

By Theorems 8 and 9 we have obtained the upscaled Helmholtz equation with effective coefficients. In this section we calculate the corresponding effective reflection and transmission properties of the metallic structure.

Let the rectangle R be for h > 0. We assume planar front of waves that reaches the metallic slab (−h < x2 < 0) from above (x2 > 0). The incoming waves would be partially reflected and partially transmitted through the metallic structure. Before we proceed any further we define the following parameters:
  • M = amplitude of the incident wave, where M = 1

  • θ≔ incident angle, where θ ∈ (−π/2, π/2)

  • T≔ complex amplitude and phase shift, where

  •   

    Ai, Bi≔ complex amplitudes in the structure, where

  • R≔ complex amplitude of the reflected wave, where

We write the solution U of (24) as
()
where τ≔√(Nw/α). We are yet to determine the coefficients Ai,   Bi,   R, and T by using the interface (x2 = 0 and x2 = −h) and transmission conditions (see (26)).
The Transfer Matrix M. We will calculate a transfer matrix M which basically gives a transformation relation between the solutions on the upper boundary x2 = 0 and the lower boundary x2 = −h. To be precise, we define a map (i.e., ) as
()
where is outer normal at x2 = 0 and x2 = −h which we choose as (0,1) and (0, −1), respectively. In short the matrix M maps the vector on the upper boundary to the vector on the lower boundary. As the map is seen to be linear, M can be expressed as a matrix. Now we determine the transfer matrix M where the two columns are obtained by M · (1,0) t and M · (0,1) t.

Columns of M. To obtain the first column of M, we study a solution U of the effective system such that and . The solution U in the interval (−h, 0) is given by (67). By transmission conditions we have , , and . With the help of these conditions, we obtain A2 = B2 = 0 and (1/2)A1B1cos(τkx1) = 1 which gives U(x1, x2) = cos(τkx2). With the help of similar transmission condition we obtain

, , and . This gives first column of M as (cos⁡(τkh), ατksin⁡(τkh)) t. A similar computation by taking M · (0,1) t in account will yield the second column of M as (−(ατk) −1sin⁡(τkh), cos⁡(τkh)) t. Thus the required transfer matrix is given by
()
where τ≔√(Nw/α) and α = 1 − (3/2)γ.
The Transmission Coefficient. After having the matrix M in hand, our next step is to calculate the transfer coefficient T. With the help of matrix M, we map the values at x2 = 0+ to the values at x2 = −h−; that is, will get mapped to . In other words,
()
Here since we are only interested in the transmission coefficient T, we eliminate the unknown R. Now we follow a simple elimination technique shown in [3] and introduce two vectors and by
()
Since the left hand side of (70) is , multiplying it with v will result in the elimination of R from (70). This leads to
()
and a straightforward computation yields
()

By (73), we have determined the transmission coefficient T which depends on wave number k, height of the metallic structure h, the aperture volume α, effective material parameter τ, and the angle θ. We note that τ = √(Nw/α), where Nw is defined by the help of an eigenvalue problem in the metallic part Σ and we also notice that T depends on the wave number k by the relation Nw = Nw(k). For a rather simple Nw, the graph of |T|2 against the wave number k is shown in figure  4 in [3].

Let us focus again on the case of a material that permits perfect plasmon waves, that is, of a lossless material with negative permittivity, εr < 0; then Nw ∈ (0,1). Also α = 1 − (3/2)γ ∈ (0,1), where 0 < γ < (1/2). This implies that the term ((ατ/cos⁡(θ)) + (cos⁡(θ)/ατ)) in (73) is greater than or equal to 2. Consequently, |T| ≤ 1 and we get |T| = 1⇔cos⁡(τkh) = 1. This corresponds to a resonance of the plasmon waves in the metallic structure (by solving ∇2U = −k2τ2U for x2 ∈ (−h, 0)) with height h.

We see that this effect can also be deduced from the transfer matrix M of (68), since for cos⁡(τkh) = 1, sin⁡(τkh) = 0 and we get the transfer matrix M = I, the Identity matrix, corresponding to perfect transmission.

Competing Interests

The author declares that they have no competing interests.

Acknowledgments

The author appreciates the financial support provided by Chair of Analysis (at TU Dortmund) during his stay there which made this work possible.

      The full text of this article hosted at iucr.org is unavailable due to technical difficulties.