This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Time-harmonic scattering by locally perturbed periodic structures with Dirichlet and Neumann boundary conditions

Guanghui Hu Guanghui Hu: School of Mathematical Sciences and LPMC
Nankai University
Tianjin 300071, China
ghhu@nankai.edu.cn
 and  Andreas Kirsch Andreas Kirsch: Department of Mathematics
Karlsruhe Institute of Technology (KIT)
76131 Karlsruhe, Germany
andreas.kirsch@kit.edu
Abstract.

The paper is concerned with well-posedness of TE and TM polarizations of time-harmonic electromagnetic scattering by perfectly conducting periodic surfaces and periodically arrayed obstacles with local perturbations. The classical Rayleigh Expansion radiation condition does not always lead to well-posedness of the Helmholtz equation even in unperturbed periodic structures. We propose two equivalent radiation conditions to characterize the radiating behavior of time-harmonic wave fields incited by a source term in an open waveguide under impenetrable boundary conditions. With these open waveguide radiation conditions, uniqueness and existence of time-harmonic scattering by incoming point source waves, plane waves and surface waves from locally perturbed periodic structures are established under either the Dirichlet or Neumann boundary condition. A Dirichlet-to-Neumann operator without using the Green’s function is constructed for proving well-posedness of perturbed scattering problems.

Keywords: Helmholtz equation, periodic structures, radiation condition, uniqueness, existence, Dirichlet boundary condition, Neumann boundary condition.

1. Introduction

The electromagnetic scattering theory in periodic structures has many applications in micro-optics, radar imaging and non-destructive testing. We refer to [23] for historical remarks and details of these applications. As a standard model, we consider a time-harmonic electromagnetic plane wave incident onto a perfectly reflecting periodic surface or periodically arrayed conducting obstacles which remain invariant in the x3x_{3}-direction. Without loss of generality the direction of periodicity is supposed to be x1x_{1} and the arrayed obstacles lie in a layer of finite height in the x2x_{2}-direction. We consider both the TE polarization case where the electric field is transversal to the ox1x2ox_{1}x_{2}-plane by assuming E(x)=(0,0,u(x1,x2))E(x)=(0,0,u(x_{1},x_{2})) and the TM polarization case where the magnetic field is transversal to the ox1x2ox_{1}x_{2}-plane by assuming H(x)=(0,0,u(x1,x2))H(x)=(0,0,u(x_{1},x_{2})). The background medium above the periodic surface or in the exterior of the periodically arrayed obstacles is supposed to be homogeneous and isotropic. The time-harmonic Maxwell’s equations for (E(x),H(x))(E(x),H(x)) will be reduced to the scalar Helmholtz equation for u(x1,x2)u(x_{1},x_{2}) over the ox1x2ox_{1}x_{2}-plane together with the Dirichlet/Neumann boundary condition in TE/TM case and with proper radiation conditions as |x2||x_{2}|\rightarrow\infty; see Figure 1 (a) and (b) for illustration of the scattering problems.

Refer to caption
(a) A Lipschitz periodic curve
Refer to caption
(b) Periodically arrayed obstacles
Figure 1. Illustration of wave scattering from (a) a perfectly reflecting periodic curve and (b) perfectly conducting obstacles. Guided waves might exist in (a)-(b), leading to difficulties in establishing well-posedness of the scattering problem with the classical Rayleigh Expansion radiation condition (3b).

In periodic structures, a frequently used radiation condition is the so-called quasi-periodic Rayleigh expansion (see (3b)), which was firstly used by Lord Rayleigh in 1907 [21] for plane wave incidence. The Rayleigh expansion consists of a finite number of plane waves and infinitely many evanescent waves. However, such a radiation condition does not always lead to uniqueness of solutions for all frequencies due to the presence of evanescent/surface waves propagating along the unbounded periodic curve, or due to the existence of guided waves propagating between the arrayed obstacles, both of them decaying exponentially in x2x_{2}. Examples of surface waves for unbounded periodic curves of Dirichlet kind were constructed in [24] where the reflecting curve is not a graph and in [14] under the Neumann boundary condition. We also refer to [1] for non-uniqueness examples of solutions incited by periodically arrayed obstacles immersed in a dielectric layer. On the other hand, it is well known that surface waves do not exist if a Dirichlet periodic curve is given by the graph of some function or satisfies the geometrical condition (21); see [4, 6, 15] for different regularity and geometry assumptions made on the reflecting curve. We also mention that the Rayleigh expansion condition does not apply to incoming source waves given by the fundamental solution of the Helmholtz equation and does not hold for scattering by compactly supported source terms. In these cases the incident waves lose the quasi-periodicity in x1x_{1}. It was firstly discussed in [2] that the radiated field should satisfy a Sommerfeld-type radiation condition and was recently proved in [11] for Dirichlet rough surfaces given by graphs and in [19] for periodic inhomogeneous layers. Hence, the radiating behavior of wave fields in periodic structures also depends on the type of incident waves. To sum up, precise and sharp radiation conditions are still needed in order to mathematically interpret the radiating behavior of time-harmonic wave fields in periodic structures, in particular for non-quasiperiodic incoming waves or when guided waves exist.

In recent years, a new radiation condition has been derived from the limiting absorption principle for scattering by layered periodic media in 2\mathbb{R}^{2} and by periodic tubes in 3\mathbb{R}^{3}; see [9, 16, 17, 18, 19, 20]. Such a radiation condition turns out to be equivalent to the radiation condition based on dispersion curves for closed periodic wave guides (see e.g., [8] and [19, Remark 2.4]). By this new radiation condition, the diffracted fields caused by a compactly supported source term or a local defect can be decomposed into the sum of a radiating part and a propagating (guided) part. The former decays as |x1|3/2|x_{1}|^{-3/2} in the horizontal direction x1x_{1} and decays as |x|1/2|x|^{-1/2} in the radical direction, whereas the latter is a finite number of quasi-periodic left-going and right-going evanescent modes which decay exponentially in the vertical direction x2x_{2} ([19]). Moreover, this new radiation condition is stronger than the angular spectrum representation [4] and the upward propagating radiation condition [3] for rough surface scattering problems. It can also be used for proving well-posedness of scattering by locally perturbed inhomogeneous layers in the presence of guided waves; see [9, 18, 19].

The aim of this paper is to investigate well-posedness of time-harmonic scattering by locally perturbed periodic curves and periodically arrayed obstacles of Dirichlet and Neumann kinds. The main results of this paper are summarized as follows.

  • (i)

    Propose two equivalent radiation conditions to prove uniqueness of weak solutions for periodic Lipschitz interfaces with local perturbations. The first radiation condition was adapted from [18, 19] for characterizing left-going and right-going evanescent waves of the propagating part of wave fields. It is referred to as the open waveguide radiation condition, in comparision with the closed waveguide radiation condition of [8]. The second radiation condition, which modifies the asymptotic behavior of radiating part of the first one, was motivated by the Sommerfeld radiation condition justified in [11] and [19, Section 6] for point source waves. The second radiation condition extends the well-posedness result of [11] to general periodic Lipschitz curves of Dirichlet or Neumann kind, in particular when guides waves are present. Since the decaying condition of Sommerfeld type contains more information on the radiating part, the second radiation condition yields a simplified proof of the uniqueness; see Theorem 2.15.

  • (ii)

    Existence of solutions for incoming plane waves, surfaces waves and point source waves in a locally perturbed periodic structure under a priori assumptions (Sections 4). Unlike the scattering by inhomogeneous periodic layers with local perturbations [18, 19, 9], there is no analogue of the Lippmann-Schwinger integral equation under the Dirichlet and Neumann boundary conditions. This leads to difficulties in the analysis of wave scattering from perfectly reflecting periodic curves with local perturbations. Our idea is to reduce the scattering problem to a bounded domain enclosing the perturbed part by constructing the DtN operator. For this purpose, we construct a Dirichlet-to-Neumann operator without using the Green’s function for proving well-posedness of the perturbed scattering problem.


The remaining part of the paper is organized as follows. We first consider the perturbed/unperturbed scattering problem due to a compact source term. In Section 2, we describe an open waveguide radiation condition and its equivalent version, and use them to prove the uniqueness results. In comparison with the results for layered media [18, 19], a more general transmission problem and the scattering by exponentially decaying source terms without a compact support will be investigated in the unperturbed periodic domain (see Theorems 3.4 and 3.5). In Section 4, we prove well-posedness results for incoming point source waves, plane waves as well as surface waves in the perturbed setting. Finally, concluding remarks will made in Section 5 on how to carry out the analysis for unbounded periodic Dirichlet curves to Neumann curves and to periodically arrayed obstacles with boundary conditions.

2. Scattering by Dirichlet periodic curves with local perturbations: radiation condition and uniqueness

2.1. Notations

Let D2D\subset{\mathbb{R}}^{2} be a 2π2\pi-periodic domain with respect to the x1x_{1}-direction. The boundary Γ:=D\Gamma:=\partial D is supposed to be given by a non-self-intersecting Lipschitz curve which is bounded in x2x_{2}-direction and 2π2\pi-periodic with respect to x1x_{1}. Therefore, in this paper we exclude the case of Figure 1 (b) but refer to Section 5. Let D~\tilde{D} be a local perturbation of DD in the way that ΓΓ~\Gamma\setminus\tilde{\Gamma} and Γ~Γ\tilde{\Gamma}\setminus\Gamma are bounded where Γ~=D~\tilde{\Gamma}=\partial\tilde{D} is the perturbed boundary which is also assumed to be a non-self-intersecting curve. Suppose that D~\tilde{D} is filled by a homogeneous and isotropic medium and that Γ~\tilde{\Gamma} is a perfectly reflecting curve of Dirichlet kind. Denote by fL2(D~)f\in L^{2}(\tilde{D}) a source term of compact support which radiates wave fields at the wavenumber k>0k>0.

We consider the problem of determining the radiated wave uHloc1(D~):={w|D~:wHloc1(2)}u\in H^{1}_{loc}(\tilde{D}):=\bigl{\{}w|_{\tilde{D}}:w\in H^{1}_{loc}(\mathbb{R}^{2})\bigr{\}} such that

(1) Δu+k2u=f in D~,u=0 on Γ~,\Delta u+k^{2}u\ =\ -f\mbox{ in }\tilde{D}\,,\quad u=0\mbox{ on }\tilde{\Gamma}\,,

and complemented by the open waveguide radiation condition explained in the next section. Without loss of generality (changing the period of the periodic structure if otherwise) we can assume that the perturbations ΓΓ~\Gamma\setminus\tilde{\Gamma} and Γ~Γ\tilde{\Gamma}\setminus\Gamma and also the support of ff are contained in the disc {x2:(x1π)2+x22<π2}\{x\in\mathbb{R}^{2}:(x_{1}-\pi)^{2}+x_{2}^{2}<\pi^{2}\}. We fix R>πR>\pi and h0>πh_{0}>\pi throughout this paper and use the following notations for h>πh>\pi (see Figure 1 (a) and Figure 2).

Qh\displaystyle Q_{h}\ :=\displaystyle:= {xD:0<x1<2π,x2<h},\displaystyle\{x\in D:0<x_{1}<2\pi,\ x_{2}<h\}\,,
Q\displaystyle\quad Q_{\infty}\ :=\displaystyle:= {xD:0<x1<2π},\displaystyle\ \{x\in D:0<x_{1}<2\pi\}\,,
Γh\displaystyle\Gamma_{h} :=\displaystyle:= (0,2π)×{h},\displaystyle(0,2\pi)\times\{h\}\,,\quad
Wh\displaystyle W_{h}\ :=\displaystyle:= {xD:x2<h},\displaystyle\ \{x\in D:x_{2}<h\}\,,\quad
Uh\displaystyle U_{h}\ :=\displaystyle:= {xD:x2>h},\displaystyle\ \{x\in D:x_{2}>h\}\,,
CR\displaystyle C_{R} :=\displaystyle:= {xD:(x1π)2+x22=R2},\displaystyle\{x\in D:(x_{1}-\pi)^{2}+x_{2}^{2}=R^{2}\},\quad
ΣR\displaystyle\Sigma_{R}\ :=\displaystyle:= {xD:(x1π)2+x22>R2},\displaystyle\ \{x\in D:(x_{1}-\pi)^{2}+x_{2}^{2}>R^{2}\}\,,
DR\displaystyle D_{R} :=\displaystyle:= {xD:(x1π)2+x22<R2},\displaystyle\{x\in D:(x_{1}-\pi)^{2}+x_{2}^{2}<R^{2}\}\,,\quad
D~R\displaystyle\tilde{D}_{R}\ :=\displaystyle:= {xD~:(x1π)2+x22<R2}.\displaystyle\ \{x\in\tilde{D}:(x_{1}-\pi)^{2}+x_{2}^{2}<R^{2}\}.

In the unperturbed setting we introduce the following function spaces 111The definitions hold also for DD instead of D~\tilde{D}.

Hloc,01(D~)\displaystyle H^{1}_{loc,0}(\tilde{D}) :=\displaystyle:= {uHloc1(D~):u=0 on D~},\displaystyle\bigl{\{}u\in H^{1}_{loc}(\tilde{D}):u=0\mbox{ on }\partial\tilde{D}\bigr{\}}\,,
H1(D~)\displaystyle H^{1}_{\ast}(\tilde{D}) :=\displaystyle:= {uHloc1(D~):u|WhD~H1(WhD~) for all h>h0,},\displaystyle\bigl{\{}u\in H^{1}_{loc}(\tilde{D}):u|_{W_{h}\cap\tilde{D}}\in H^{1}(W_{h}\cap\tilde{D})\mbox{ for all }h>h_{0},\bigr{\}}\,,
H1(ΣR)\displaystyle H^{1}_{\ast}(\Sigma_{R}) :=\displaystyle:= {uHloc1(ΣR):u|WhΣRH1(WhΣR)for allh>h0,u=0onΣRD},\displaystyle\biggl{\{}u\in H^{1}_{loc}(\Sigma_{R}):\begin{array}[]{l}u|_{W_{h}\cap\Sigma_{R}}\in H^{1}(W_{h}\cap\Sigma_{R})\;\mbox{for all}\;h>h_{0},\\ u=0\;\mbox{on}\;\partial\Sigma_{R}\cap\partial D\end{array}\biggr{\}}\,,
Hα,loc1(D)\displaystyle H^{1}_{\alpha,loc}(D) :=\displaystyle:= {uHloc1(D):u(,x2) is α-quasi-periodic},\displaystyle\bigl{\{}u\in H^{1}_{loc}(D):u(\cdot,x_{2})\mbox{ is $\alpha$-quasi-periodic}\bigr{\}}\,,
Hα,loc,01(D)\displaystyle H^{1}_{\alpha,loc,0}(D) :=\displaystyle:= {uHα,loc1(D):u=0 on D}.\displaystyle\bigl{\{}u\in H^{1}_{\alpha,loc}(D):u=0\mbox{ on }\partial D\bigr{\}}\,.

2.2. The Open Waveguide Radiation Condition And An Energy Formula

As mentioned in the introduction part the diffracted field will have a decomposition into a (guided) propagating part and a radiating part. The loss of exponential decay of the radiating part is a consequence of the existence of cut-off values while the propagative wave numbers determine the behavior of the guided part along the waveguide. We first recall that a function ϕLloc2()\phi\in L^{2}_{loc}(\mathbb{R}) is called α\alpha-quasi-periodic if ϕ(x1+2π)=e2παiϕ(x1)\phi(x_{1}+2\pi)=e^{2\pi\alpha i}\phi(x_{1}) for all x1x_{1}\in\mathbb{R}.

Definition 2.1.

(i) α[1/2,1/2]\alpha\in[-1/2,1/2] is called a cut-off value if there exists \ell\in\mathbb{Z} such that |α+|=k|\alpha+\ell|=k.
(ii) α[1/2,1/2]\alpha\in[-1/2,1/2] is called a propagative wave number if there exists a non-trivial ϕHα,loc,01(D)\phi\in H^{1}_{\alpha,loc,0}(D) such that

(3a) Δϕ+k2ϕ= 0 in D,\Delta\phi+k^{2}\phi\ =\ 0\text{ in }D\,,

and ϕ\phi satisfies the upward Rayleigh expansion

(3b) ϕ(x)=ϕei(+α)x1eik2(+α)2(x2h0)for x2>h0\phi(x)\ =\ \sum_{\ell\in\mathbb{Z}}\phi_{\ell}\,e^{i(\ell+\alpha)x_{1}}\,e^{i\sqrt{k^{2}-(\ell+\alpha)^{2}}(x_{2}-h_{0})}\quad\mbox{for }x_{2}>h_{0}

for some ϕ\phi_{\ell}\in\mathbb{C} where the convergence is uniform for x2h0+εx_{2}\geq h_{0}+\varepsilon for every ε>0\varepsilon>0. The functions ϕ\phi are called guided (or propagating or Floquet) modes.

In all of the paper, we choose the square root function to be holomorphic in the cutted plane (i0)\mathbb{C}\setminus(i\mathbb{R}_{\leq 0}). In particular, t=i|t|\sqrt{t}=i\sqrt{|t|} for t<0t\in\mathbb{R}_{<0}. In Definition 2.1 we restrict the quasi-periodic parameter α\alpha to the interval [1/2,1/2][-1/2,1/2], because an α\alpha-quasi-periodic function must be also (α+j)(\alpha+j)-quasi-periodic for any jj\in{\mathbb{N}}. Throughout this paper we make the following assumptions.

Assumption 2.2.

Let |+α|k|\ell+\alpha|\not=k for every propagative wave number α[1/2,1/2]\alpha\in[-1/2,1/2] and every \ell\in\mathbb{Z}; that is, no cut-off value is a propagative wave number.

Under Assumption 2.2 it can be shown (see, e.g. [18] for the case of a flat curve Γ=Γ0\Gamma=\Gamma_{0} and an additional index of refraction) that at most a finite number of propagative wave numbers exists in the interval [1/2,1/2][-1/2,1/2]. Furthermore, if α\alpha is a propagative wave number with mode ϕ\phi then α-\alpha is a propagative wave number with mode ϕ¯\overline{\phi}. Therefore, we can number the propagative wave numbers in [1/2,1/2][-1/2,1/2] such that they are given by {α^j:jJ}\{\hat{\alpha}_{j}:j\in J\} where JJ\subset\mathbb{Z} is finite and symmetric with respect to 0 and α^j=α^j\hat{\alpha}_{-j}=-\hat{\alpha}_{j} for jJj\in J. Furthermore, it is known that (under Assumption 2.2) every mode ϕ\phi is evanescent; that is, exponentially decaying as x2x_{2} tends to infinity in DD; that is, satisfies |ϕ(x)|ceδ|x2||\phi(x)|\leq c\,e^{-\delta|x_{2}|} for x2h0x_{2}\geq h_{0} and some c,δ>0c,\delta>0 which are independent of xx. The corresponding space

(4) Xj:={ϕHα^j,loc,01(D):u satisfies (3a) and (3b) for α=α^j}X_{j}\ :=\ \bigl{\{}\phi\in H^{1}_{\hat{\alpha}_{j},loc,0}(D):u\mbox{ satisfies (\ref{exc:a}) and (\ref{exc:b}) for }\alpha=\hat{\alpha}_{j}\bigr{\}}

of modes is finite dimensional with some dimension mj>0m_{j}>0. On XjX_{j} we define the sesqui-linear form B:Xj×XjB:X_{j}\times X_{j}\to\mathbb{C} by

(5) B(ϕ,ψ):=2iQϕx1ψ¯𝑑x,ϕ,ψXj.B(\phi,\psi)\ :=\ -2i\int\limits_{Q_{\infty}}\frac{\partial\phi}{\partial x_{1}}\,\overline{\psi}\,dx\,,\quad\phi,\psi\in X_{j}\,.

Note that BB is hermitian. We make the assumption that BB is non-degenerated on every XjX_{j}; that is,

Assumption 2.3.

For every jJj\in J and ψXj\psi\in X_{j}, ψ0\psi\not=0, the linear form B(,ψ):XjB(\cdot,\psi):X_{j}\to\mathbb{C} is non-trivial on XjX_{j}; that is, there exists ϕXj\phi\in X_{j} with B(ϕ,ψ)0B(\phi,\psi)\not=0.

The hermitian sesqui-linear form BB defines the cones {ψXj:B(ψ,ψ)0}\{\psi\in X_{j}:B(\psi,\psi)\gtrless 0\} of propagating waves traveling to the right and left, respectively. We construct a basis of XjX_{j} with elements in these cones by taking any inner product (,)Xj(\cdot,\cdot)_{X_{j}} and consider the following eigenvalue problem in XjX_{j} for every fixed jJj\in J. Determine λ,j\lambda_{\ell,j}\in\mathbb{R} and non-trivial ϕ^,jXj\hat{\phi}_{\ell,j}\in X_{j} with

(6) B(ϕ^,j,ψ)=2iQϕ^,jx1ψ¯𝑑x=λ,j(ϕ^,j,ψ)Xjfor all ψXjB(\hat{\phi}_{\ell,j},\psi)\ =\ -2i\int\limits_{Q_{\infty}}\frac{\partial\hat{\phi}_{\ell,j}}{\partial x_{1}}\,\overline{\psi}\,dx\ =\ \lambda_{\ell,j}\,\bigl{(}\hat{\phi}_{\ell,j},\psi\bigr{)}_{X_{j}}\quad\mbox{for all }\psi\in X_{j}

and =1,,mj\ell=1,\ldots,m_{j}. We normalize the basis such that (ϕ^,j,ϕ^,j)Xj=δ,\bigl{(}\hat{\phi}_{\ell,j},\hat{\phi}_{\ell^{\prime},j}\bigr{)}_{X_{j}}=\delta_{\ell,\ell^{\prime}} for ,=1,,mj\ell,\ell^{\prime}=1,\ldots,m_{j}. Then λ,j=B(ϕ^,j,ϕ^,j)\lambda_{\ell,j}=B(\hat{\phi}_{\ell,j},\hat{\phi}_{\ell,j}) and the function ψXj\psi\in X_{j} in Assumption 2.3 must take the form ψ==1mjcϕ^,j\psi=\sum_{\ell=1}^{m_{j}}c_{\ell}\hat{\phi}_{\ell,j} with c0c_{\ell}\neq 0 for some =1,2,,mj\ell=1,2,\cdots,m_{j}. Choosing ϕ=ϕ^,j\phi=\hat{\phi}_{\ell,j}, one deduces B(ϕ,ψ)=cλ,jB(\phi,\psi)=c_{\ell}\lambda_{\ell,j}. Hence, the Assumption 2.3 is equivalent to λ,j0\lambda_{\ell,j}\not=0 for all =1,,mj\ell=1,\ldots,m_{j} and jJj\in J.

Remark 2.4.
  • (i)

    The set of propagative wave numbers obviously depends on k+k\in{\mathbb{R}}_{+}. Analogously, one may define k=k(α)k_{\ell}=k_{\ell}(\alpha) for α[1/2,1/2]\alpha\in[-1/2,1/2] as the wave number if the problem (3a) and (3b) admits a non-trivial solution. Since the solutions are in Hα,loc,01(D)H^{1}_{\alpha,loc,0}(D) the values k(α)k_{\ell}(\alpha) are just eigenvalues of Δ-\Delta with respect to α\alpha-quasi-periodic boundary conditions on the vertical boundary of QQ_{\infty} and homogeneous Dirichlet boundary condition on Γ\Gamma. The functions αk(α)\alpha\rightarrow k_{\ell}(\alpha) are well known as the dispersion relations/curves. Throughout our paper the wavenumber kk is fixed. Under the Assumption 2.2, the set {(α^j(k0),k0)}jJ\{(\hat{\alpha}_{j}(k_{0}),k_{0})\}_{j\in J} constitutes the intersection points of the dispersion curves with the line k=k0k=k_{0} in the (α,k)(\alpha,k)-plane. Assumption 2.2 implies the absence of flat dispersion curves.

  • (ii)

    The eigenvalue problem (6) originates from the limiting absorption principle (LAP) by applying an abstract functional theorem that goes back to [18]. We refer to [8, 20] for detailed discussions in justifying the radiation conditions for closed full and half-waveguide problems. Note that the choice of the inner product in XjX_{j} relies on the way how to perturb the original scattering problem by applying the LAP. For example, if the LAP is applied to the wavenumber kk then (ϕ,ψ)Xj=Qϕψ¯𝑑x(\phi,\psi)_{X_{j}}=\int_{Q_{\infty}}\phi\overline{\psi}dx.

In all of the paper we make Assumptions 2.2 and 2.3 without mentioning this always. The one-dimensional Fourier transform is defined as

(ϕ)(ω):=12πϕ(s)eisω𝑑s,ω.(\mathcal{F}\phi)(\omega)\ :=\ \frac{1}{\sqrt{2\pi}}\int\limits_{-\infty}^{\infty}\phi(s)\,e^{-is\omega}\,ds\,,\quad\omega\in\mathbb{R}\,.

It can be considered as an unitary operator from L2()L^{2}(\mathbb{R}) onto itself. Now we are able to formulate the radiation condition caused by compactly supported source terms, which will also serve as the radiation condition of the Green’s function to perturbed and unperturbed scattering problems (see Theorem 4.1 and Remark 4.2).

Definition 2.5.

Let ψ+,ψC()\psi_{+},\psi_{-}\in C^{\infty}(\mathbb{R}) be any functions with ψ±(x1)=1\psi_{\pm}(x_{1})=1 for ±x1σ0\pm x_{1}\geq\sigma_{0} (for some σ0>max{R,2π}+1\sigma_{0}>\max\{R,2\pi\}+1) and ψ±(x1)=0\psi_{\pm}(x_{1})=0 for ±x1σ01\pm x_{1}\leq\sigma_{0}-1.

A solution uHloc1(ΣR)u\in H^{1}_{loc}(\Sigma_{R}) of (1) satisfies the open waveguide radiation condition with respect to an inner product (,)Xj(\cdot,\cdot)_{X_{j}} in XjX_{j} if uu has in ΣR\Sigma_{R} a decomposition into u=urad+upropu=u_{rad}+u_{prop} which satisfy the following conditions.

  • (a)

    The propagating part upropu_{prop} has the form

    (7) uprop(x)=jJ[ψ+(x1):λ,j>0a,jϕ^,j(x)+ψ(x1):λ,j<0a,jϕ^,j(x)]u_{prop}(x)\ =\ \sum_{j\in J}\biggl{[}\psi_{+}(x_{1})\sum_{\ell:\lambda_{\ell,j}>0}a_{\ell,j}\,\hat{\phi}_{\ell,j}(x)\ +\ \psi_{-}(x_{1})\sum_{\ell:\lambda_{\ell,j}<0}a_{\ell,j}\,\hat{\phi}_{\ell,j}(x)\biggr{]}

    for xΣRx\in\Sigma_{R} and some a,ja_{\ell,j}\in\mathbb{C}. Here, for every jJj\in J the scalars λ,j\lambda_{\ell,j}\in\mathbb{R} and ϕ^,jX^j\hat{\phi}_{\ell,j}\in\hat{X}_{j} for =1,,mj\ell=1,\ldots,m_{j} are given by the eigenvalues and corresponding eigenfunctions, respectively, of the self adjoint eigenvalue problem (6). Note that by the choice of ψ±\psi_{\pm} the propagating part vanishes for |x1|<σ01|x_{1}|<\sigma_{0}-1 and is therefore well defined in ΣR\Sigma_{R}.

  • (b)

    The radiating part uradH1(ΣR)u_{rad}\in H^{1}_{\ast}(\Sigma_{R}) satisfies the generalized angular spectrum radiation condition

    (8) |(urad)(ω,x2)x2ik2ω2(urad)(ω,x2)|2𝑑ω 0,x2.\int\limits_{-\infty}^{\infty}\left|\frac{\partial(\mathcal{F}u_{rad})(\omega,x_{2})}{\partial x_{2}}-i\sqrt{k^{2}-\omega^{2}}\,(\mathcal{F}u_{rad})(\omega,x_{2})\right|^{2}d\omega\ \longrightarrow\ 0\,,\quad x_{2}\to\infty\,.

The radiation condition (8) can be used to prove well-posedness of the Helmholtz equation with a source term which is supported in x1x_{1}-direction and exponentially decays in x2x_{2} (see (10a)). It has been shown in [18] for the case of a half plane problem with an inhomogeneous period layer that the radiation condition of Definition 2.5 for the inner product (ϕ,ψ)Xj=2kQnϕψ¯𝑑x(\phi,\psi)_{X_{j}}=2k\int_{Q_{\infty}}n\,\phi\,\overline{\psi}\,dx is a consequence of the limiting absorption principle by replacing kk with k+iϵk+i\epsilon, ϵ>0\epsilon>0. In this paper we will not justify this radiation condition, although we are sure that this can be done in the same way as [18, 20]. A second motivation of our radiation condition is the following result on the direction of the energy flow which will play a central role in the proof of uniqueness.

Lemma 2.6.

Let u=jJ=1mja,jϕ^,ju=\sum_{j\in J}\sum_{\ell=1}^{m_{j}}a_{\ell,j}\hat{\phi}_{\ell,j} for some a,ja_{\ell,j}\in\mathbb{C} and write q+Q:={xD:q<x1<q+2π}q+Q_{\infty}:=\{x\in D:q<x_{1}<q+2\pi\} for qq\in{\mathbb{R}}. Then we have

2Imq+Qu¯ux1𝑑x=jJ=1mjλ,j|a,j|2.2\,\operatorname{Im}\int\limits_{q+Q_{\infty}}\overline{u}\,\frac{\partial u}{\partial x_{1}}\,dx\ =\ \sum_{j\in J}\sum_{\ell=1}^{m_{j}}\lambda_{\ell,j}\,|a_{\ell,j}|^{2}\,.

By Lemma 2.6, the propagating part upropu_{prop} satisfies the energy formula

2Imq+Quprop¯upropx1𝑑x={jJλ,j>0λ,j|a,j|2,q>σ0,jJλ,j<0λ,j|a,j|2,q<σ0,2\,\operatorname{Im}\int\limits_{q+Q_{\infty}}\overline{u_{prop}}\,\frac{\partial u_{prop}}{\partial x_{1}}\,dx\ =\ \left\{\begin{array}[]{cl}\displaystyle\sum_{j\in J}\sum_{\lambda_{\ell,j}>0}\lambda_{\ell,j}\,|a_{\ell,j}|^{2}\,,&q>\sigma_{0}\,,\\ \displaystyle\sum_{j\in J}\sum_{\lambda_{\ell,j}<0}\lambda_{\ell,j}\,|a_{\ell,j}|^{2}\,,&q<-\sigma_{0}\,,\end{array}\right.

where σ0>2π+1\sigma_{0}>2\pi+1 is the number specified in Definition 2.5. To prove Lemma 2.6, we have to modify the arguments of [19] for inhomogeneous layered media, because solutions of the Dirichlet and Neumann boundary value problems are in Hloc1(D~)H^{1}_{loc}(\tilde{D}) but fail to be in Hloc2(D~)H^{2}_{loc}(\tilde{D}) if Γ~\tilde{\Gamma} is Lipschitz. For C2C^{2}-smooth boundaries, the quantity in Lemma 2.6 also equals to 4πImD{x1=q}u¯ux1𝑑s4\pi\operatorname{Im}\int_{D\cap\{x_{1}={\color[rgb]{0,0,0}q}\}}\overline{u}\,\frac{\partial u}{\partial x_{1}}\,ds; see [18, Lemma 6.3] and [19, Lemma 2.6].

Proof of Lemma 2.6. We recall the following form of Green’s formula valid in any Lipschitz domain Ω\Omega: For uH1(Ω)u\in H^{1}(\Omega) with ΔuL2(Ω)\Delta u\in L^{2}(\Omega) we have

Ω[uψ+ψΔu]𝑑x= 0for all ψH01(Ω).\int\limits_{\Omega}[\nabla u\cdot\nabla\psi+\psi\,\Delta u]\,dx\ =\ 0\quad\mbox{for all }\psi\in H^{1}_{0}(\Omega)\,.

Let j{1,2}j\in\{1,2\}. First we show for αj\alpha_{j}-quasi-periodic solutions ujHloc1(D)u_{j}\in H^{1}_{loc}(D) of Δuj+k2uj=0\Delta u_{j}+k^{2}u_{j}=0 in DD with uj=0u_{j}=0 on Γ\Gamma and αj(1/2,1/2]\alpha_{j}\in(-1/2,1/2] with α1α2\alpha_{1}\not=\alpha_{2} that

(9) q+Q[u2¯u1x1u1u2¯x1]𝑑x= 0.\int\limits_{q+Q_{\infty}}\biggl{[}\overline{u_{2}}\,\frac{\partial u_{1}}{\partial x_{1}}-u_{1}\,\frac{\partial\overline{u_{2}}}{\partial x_{1}}\biggr{]}\,dx\ =\ 0\,.

Indeed, defining ψ(x1):=1|x1q|/(2π)\psi(x_{1}):=1-|x_{1}-q|/(2\pi) and applying Green’s theorem in Ω:={xD:q2π<x1<q+2π}\Omega:=\{x\in D:q-2\pi<x_{1}<q+2\pi\} yields (note that uju_{j} decay exponentially as x2x_{2} tends to infinity)

0\displaystyle 0 =\displaystyle= Ω[u1(ψu2¯)k2(ψu2¯)u1]𝑑x\displaystyle\int\limits_{\Omega}[\nabla u_{1}\cdot\nabla(\psi\overline{u_{2}})-k^{2}(\psi\overline{u_{2}})\,u_{1}]\,dx
=\displaystyle= Ωψ[u1u2¯k2u2¯u1]𝑑x+Ωψu2¯u1x1𝑑x.\displaystyle\int\limits_{\Omega}\psi\,[\nabla u_{1}\cdot\nabla\overline{u_{2}}-k^{2}\overline{u_{2}}\,u_{1}]\,dx\ +\ \int\limits_{\Omega}\psi^{\prime}\,\overline{u_{2}}\,\frac{\partial u_{1}}{\partial x_{1}}\,dx\,.

Interchanging the roles of u1u_{1} and u2¯\overline{u_{2}} and subtraction yields

0\displaystyle 0 =\displaystyle= Ωψ[u2¯u1x1u1u2¯x1]𝑑x\displaystyle\int\limits_{\Omega}\psi^{\prime}\,\biggl{[}\overline{u_{2}}\,\frac{\partial u_{1}}{\partial x_{1}}-u_{1}\,\frac{\partial\overline{u_{2}}}{\partial x_{1}}\biggr{]}\,dx
=\displaystyle= q+Qψ[u2¯u1x1u1u2¯x1]𝑑x+q2π+Qψ[u2¯u1x1u1u2¯x1]𝑑x\displaystyle\int\limits_{q+Q_{\infty}}\psi^{\prime}\,\biggl{[}\overline{u_{2}}\,\frac{\partial u_{1}}{\partial x_{1}}-u_{1}\,\frac{\partial\overline{u_{2}}}{\partial x_{1}}\biggr{]}\,dx+\int\limits_{q-2\pi+Q_{\infty}}\psi^{\prime}\,\biggl{[}\overline{u_{2}}\,\frac{\partial u_{1}}{\partial x_{1}}-u_{1}\,\frac{\partial\overline{u_{2}}}{\partial x_{1}}\biggr{]}\,dx
=\displaystyle= 12πq+Q[u2¯u1x1u1u2¯x1]𝑑x12πq2π+Q[u2¯u1x1u1u2¯x1]𝑑x\displaystyle\frac{1}{2\pi}\int\limits_{q+Q_{\infty}}\biggl{[}\overline{u_{2}}\,\frac{\partial u_{1}}{\partial x_{1}}-u_{1}\,\frac{\partial\overline{u_{2}}}{\partial x_{1}}\biggr{]}\,dx-\frac{1}{2\pi}\int\limits_{q-2\pi+Q_{\infty}}\biggl{[}\overline{u_{2}}\,\frac{\partial u_{1}}{\partial x_{1}}-u_{1}\,\frac{\partial\overline{u_{2}}}{\partial x_{1}}\biggr{]}\,dx
=\displaystyle= (1e2πi(α2α1))12πq+Q[u2¯u1x1u1u2¯x1]𝑑x\displaystyle\bigl{(}1-e^{2\pi i(\alpha_{2}-\alpha_{1})}\bigr{)}\,\frac{1}{2\pi}\int\limits_{q+Q_{\infty}}\biggl{[}\overline{u_{2}}\,\frac{\partial u_{1}}{\partial x_{1}}-u_{1}\,\frac{\partial\overline{u_{2}}}{\partial x_{1}}\biggr{]}\,dx

where we used the quasi-periodicity of uju_{j}. This yields (9).

Now we rewrite uu as

u=jJ=1mja,jϕ^,j=jJujwithuj:==1mja,jϕ^,j.u\ =\ \sum_{j\in J}\sum_{\ell=1}^{m_{j}}a_{\ell,j}\,\hat{\phi}_{\ell,j}\ =\ \sum_{j\in J}u_{j}\quad\mbox{with}\quad u_{j}\ :=\ \sum_{\ell=1}^{m_{j}}a_{\ell,j}\,\hat{\phi}_{\ell,j}.

Then uju_{j} is α^j\hat{\alpha}_{j}-quasi-periodic. Using (9) and the orthonormalization of ϕ^,j\hat{\phi}_{\ell,j}, we arrive at

2iImq+Qu¯ux1𝑑x=q+Q[u¯ux1uu¯x1]𝑑x\displaystyle 2i\operatorname{Im}\int\limits_{q+Q_{\infty}}\overline{u}\,\frac{\partial u}{\partial x_{1}}\,dx\ =\ \int\limits_{q+Q_{\infty}}\biggl{[}\overline{u}\,\frac{\partial u}{\partial x_{1}}-u\,\frac{\partial\overline{u}}{\partial x_{1}}\biggr{]}\,dx
=\displaystyle= jJjJq+Q[uj¯ujx1ujuj¯x1]𝑑x=jJq+Q[uj¯ujx1ujuj¯x1]𝑑x\displaystyle\sum_{j\in J}\sum_{j^{\prime}\in J}\int\limits_{q+Q_{\infty}}\biggl{[}\overline{u_{j}}\,\frac{\partial u_{j^{\prime}}}{\partial x_{1}}-u_{j^{\prime}}\,\frac{\partial\overline{u_{j}}}{\partial x_{1}}\biggr{]}\,dx\ =\ \sum_{j\in J}\int\limits_{q+Q_{\infty}}\biggl{[}\overline{u_{j}}\,\frac{\partial u_{j}}{\partial x_{1}}-u_{j}\,\frac{\partial\overline{u_{j}}}{\partial x_{1}}\biggr{]}\,dx
=\displaystyle= 2iImjJq+Quj¯ujx1𝑑x=iRe[2ijJQuj¯ujx1𝑑x]=ijJ=1mjλ,j|a,j|2,\displaystyle 2i\operatorname{Im}\sum_{j\in J}\int\limits_{q+Q_{\infty}}\overline{u_{j}}\,\frac{\partial u_{j}}{\partial x_{1}}\,dx\ =\ i\operatorname{Re}\biggl{[}-2i\sum_{j\in J}\int\limits_{Q_{\infty}}\overline{u_{j}}\,\frac{\partial u_{j}}{\partial x_{1}}\,dx\biggr{]}\ =\ i\sum_{j\in J}\sum_{\ell=1}^{m_{j}}\lambda_{\ell,j}\,|a_{\ell,j}|^{2}\,,

which proves the lemma. ∎


Below we review a result on the asymptotic behavior of uradu_{rad} which will be needed in the proof of uniqueness. By (1) and (7), the radiating part uradu_{rad} to the scattering problem satisfies

(10a) Δurad+k2urad=fjJ=1mja,jφ,jin D~,urad=0 on Γ~,\Delta u_{rad}\ +\ k^{2}\,u_{rad}\ =\ -f\ -\ \sum_{j\in J}\sum_{\ell=1}^{m_{j}}a_{\ell,j}\varphi_{\ell,j}\quad\mbox{in }\tilde{D}\,,\quad u_{rad}=0\mbox{ on }\tilde{\Gamma}\,,

where

(10b) φ,j(x)={2ψ+(x1)ϕ^,j(x)x1+ψ+′′(x1)ϕ^,j(x)if λ,j>0,2ψ(x1)ϕ^,j(x)x1+ψ′′(x1)ϕ^,j(x)if λ,j<0.\varphi_{\ell,j}(x)\ =\ \left\{\begin{array}[]{cl}2\,\psi_{+}^{\prime}(x_{1})\,\frac{\partial\hat{\phi}_{\ell,j}(x)}{\partial x_{1}}+\psi_{+}^{\prime\prime}(x_{1})\,\hat{\phi}_{\ell,j}(x)&\mbox{if }\lambda_{\ell,j}>0\,,\\ 2\,\psi_{-}^{\prime}(x_{1})\,\frac{\partial\hat{\phi}_{\ell,j}(x)}{\partial x_{1}}+\psi_{-}^{\prime\prime}(x_{1})\,\hat{\phi}_{\ell,j}(x)&\mbox{if }\lambda_{\ell,j}<0\,.\end{array}\right.

We note that ff has compact support in Qh0Q_{h_{0}} and φ,j\varphi_{\ell,j} vanish for |x1|σ01|x_{1}|\leq\sigma_{0}-1 and |x1|σ0|x_{1}|\geq\sigma_{0}, and are evanescent; that is, there exist c^,δ>0\hat{c},\delta>0 with |φ,j(x)|c^exp(δx2)|\varphi_{\ell,j}(x)|\leq\hat{c}\exp(-\delta x_{2}) for all x2h0x_{2}\geq h_{0}. Furthermore, uradu_{rad} satisfies the generalized angular spectrum radiation condition (8). In [19] the following result has been shown.222These properties are consequences of the differential equation and radiation condition above the line x2=h0x_{2}=h_{0} solely and are therefore independent of the differential equation or boundary condition below this line.

Lemma 2.7.

Let Assumptions 2.2 and 2.3 hold, and let uHloc1(D)u\in H_{loc}^{1}(D) be a solution of (1) satisfying the radiation condition of Definition 2.5. Then the radiating part uradu_{rad} satisfies a stronger form of the radiation condition (8), namely,

(11) |(urad)(ω,x2)x2ik2ω2(urad)(ω,x2)|cδ+|ω2k2|eδx2\left|\frac{\partial(\mathcal{F}u_{rad})(\omega,x_{2})}{\partial x_{2}}-i\sqrt{k^{2}-\omega^{2}}\,(\mathcal{F}u_{rad})(\omega,x_{2})\right|\leq\ \frac{c}{\delta+\sqrt{|\omega^{2}-k^{2}|}}\,e^{-\delta x_{2}}

for almost all ω\omega\in\mathbb{R} and x2>h0x_{2}>h_{0} where c>0c>0 is independent of ω\omega and xx.

Furthermore, there exists c>0c>0 with

(12) |urad(x)|+|urad(x)|c(1+|x2|)ρ(x1)\bigl{|}u_{rad}(x)\bigr{|}\ +\ \bigl{|}\nabla u_{rad}(x)\bigr{|}\ \leq\ c\,(1+|x_{2}|)\,\rho(x_{1})

for all xD~x\in\tilde{D} with x2h0+1x_{2}\geq h_{0}+1, where ρL2()L()\rho\in L^{2}(\mathbb{R})\cap L^{\infty}(\mathbb{R}) is given by

(13) ρ(x1):=|urad(y1,h0)|(1+|x1y1|)3/2𝑑y1+11+|x1|3/2,x1.\rho(x_{1})\ :=\ \int\limits_{\mathbb{R}}\frac{|u_{rad}(y_{1},h_{0})|}{(1+|x_{1}-y_{1}|)^{3/2}}\,dy_{1}\ +\ \frac{1}{1+|x_{1}|^{3/2}}\,,\quad x_{1}\in\mathbb{R}\,.

2.3. A Modified Open Waveguide Radiation Condition

In this subsection we propose another open waveguide radiation condition that is equivalent to the Def. 2.5. We first define the half-plane Sommerfeld radiation condition used in [11, 19]. Introduce the weighted Sobolev space Hρ1(Ω)H_{\rho}^{1}(\Omega) by

Hρ1(Ω):={u:(|1+|x1|2)ρ/2uH1(Ω)},ρ.\displaystyle H_{\rho}^{1}(\Omega)\ :=\ \bigl{\{}u:(|1+|x_{1}|^{2})^{\rho/2}u\in H^{1}(\Omega)\bigr{\}}\,,\quad\rho\in{\mathbb{R}}\,.
Definition 2.8.

A function vC(Uh0ΣR)v\in C^{\infty}(U_{h_{0}}\cap\Sigma_{R}) satisfies the Sommerfeld radiation condition in Uh0ΣRU_{h_{0}}\cap\Sigma_{R} if vHρ1(WhΣR)v\in H^{1}_{\rho}(W_{h}\cap\Sigma_{R}) for all h>h0h>h_{0} and all ρ<0\rho<0 and

(14) supxCaUh|x|1/2|v(x)rikv(x)| 0,a,supxUh|x|1/2|v(x)|<\displaystyle\sup\limits_{x\in C_{a}\cap U_{h}}|x|^{1/2}\bigl{|}\frac{\partial v(x)}{\partial r}-ikv(x)\bigr{|}\ \rightarrow\ 0\,,\quad a\rightarrow\infty\,,\qquad\sup\limits_{x\in U_{h}}|x|^{1/2}|v(x)|\ <\ \infty

for all h>h0h>h_{0} where r=|x|r=|x|.

Remark 2.9.

Since Φ(x,y)=𝒪(|x|1/2)\Phi(x,y)=\mathcal{O}(|x|^{-1/2}) and Φ(x,y)rikΦ(x,y)=𝒪(|x|3/2)\frac{\partial\Phi(x,y)}{\partial r}-ik\Phi(x,y)=\mathcal{O}(|x|^{-3/2}) as r=|x|r=|x|\rightarrow\infty, it holds that Φ(,y)Hρ1(WhΣR)\Phi(\cdot,y)\in H^{1}_{\rho}(W_{h}\cap\Sigma_{R}) for all ρ<0\rho<0 if R>|y1π|R>|y_{1}-\pi|. Hence, the above Sommerfeld radiation condition covers two-dimensional point source waves, but excludes plane waves and surface (evanescent) waves, which do not decay along the horizontal direction.

If Γ\Gamma is a Lipschitz function, it was shown in [11] that the scattered field caused by a point source source must satisfy the above Sommerfeld radiation condition. However, the total field (i.e., the Green’s function to the rough surface scattering problem) satisfies an analogous condition but with the weighted index ρ<1\rho<1 in place of ρ<0\rho<0. Motivated by this fact, we define a modified open waveguide radiation condition by changing the generalized angular spectrum radiation condition of the radiating part of Def. 2.5.

Definition 2.10.

A solution uHloc1(ΣR)u\in H^{1}_{loc}(\Sigma_{R}) of (1) satisfies the modified open waveguide radiation condition with respect to an inner product (,)Xj(\cdot,\cdot)_{X_{j}} in XjX_{j} if uu has a decomposition into u=urad+upropu=u_{rad}+u_{prop} in ΣR\Sigma_{R} where upropu_{prop} satisfied the same condition specified as in Def. 2.5 (a) and uradu_{rad} fulfills the Sommerfeld radiation condition of Def. 2.8 but with the index ρ<1\rho<1.

Below we prove the equivalence of the two open waveguide radiation conditions.

Theorem 2.11.

The open waveguide radiation condition of Def. 2.5 and the modified one given by Def. 2.10 are equivalent.

Proof. Write u=urad+upropu=u_{rad}+u_{prop} where uradH1(D~)u_{rad}\in H^{1}_{\ast}(\tilde{D}) denotes the radiating part and upropu_{prop} the propagating part. First we suppose that uradu_{rad} fulfills the generalized angular spectrum radiation condition (8). By arguing analogously to [19, Theorem 6.2] for compact source terms, one can show the asymptotics urad(x)=O(|x1|3/2)u_{rad}(x)=O(|x_{1}|^{-3/2}) as |x1||x_{1}|\rightarrow\infty in WhW_{h}. This gives uradHρ1(WhΣR)u_{rad}\in H^{1}_{\rho}(W_{h}\cap\Sigma_{R}) for all h>h0h>h_{0}, ρ<1\rho<1 and proves the modified open waveguide radiation condition of Definition 2.10; see [19, Section 6] for details.

Now it remains to justify the generalized angular spectrum radiation condition of uradu_{rad}, under the assumption that uradu_{rad} satisfies the Sommerfeld radiation condition of Def. 2.8 but with the index ρ<1\rho<1. Since urad|Γh0Hρ1/2()u_{rad}|_{\Gamma_{h_{0}}}\in H^{1/2}_{\rho}({\mathbb{R}}) for all 1/2<ρ<11/2<\rho<1, we recall from [11, Lemma A.2, Appendix] (see also [19]) that the function

v(x)=2Γh0G(x,y)y2urad(y)𝑑s(y),x2>h0,v(x)=2\int_{\Gamma_{h_{0}}}\frac{\partial G(x,y)}{\partial y_{2}}u_{rad}(y)\,ds(y),\quad x_{2}>h_{0},

satisfies the homogeneous Helmholtz equation together the Sommerfeld radiation conditions 14 and the boundary value v=uradv=u_{rad} on x2=h0x_{2}=h_{0}. Hence, the function w:=uradvw:=u_{rad}-v satisfies 14 in x2>h0x_{2}>h_{0} and the boundary value problem

Δw+k2w=φinx2>h0,w=0onx2=h0,\Delta w+k^{2}w=\varphi\quad\mbox{in}\quad x_{2}>h_{0},\qquad w=0\quad\mbox{on}\quad x_{2}=h_{0},

where φ\varphi is given by the right hand side of (10a).This implies that ww can be represented as

w(x)=σ0σ0h0[G(x,y)G(x,y)]φ(y)𝑑y2𝑑y1,x2>h0,w(x)=\int_{-\sigma_{0}}^{\sigma_{0}}\int_{h_{0}}^{\infty}[G(x,y)-G(x,y^{*})]\,\varphi(y)\,dy_{2}dy_{1},\quad x_{2}>h_{0},

with y:=(y12h0y2)y^{*}:=(y_{1}-2h_{0}-y_{2})^{\top}. Now, following the proof of [19, Lemma 7.1] one can show that ww satisfies the stronger form (11) of the radiation condition (8).This proves the generalized angular spectrum radiation condition of uradu_{rad}. ∎

We would like to extend the Sommerfeld radiation condition up to the boundary Γ\Gamma. However, since Γ\Gamma is only Lipschitz, in general the derivatives v/r\partial v/\partial r do not exist up to the boundary. We can, however, define a weaker form which models the integral form of the Sommerfeld radiation condition as follows. The connection between these two radiation conditions will be described in Lemma 2.13.

Definition 2.12.

Let aja_{j} be a sequence in \mathbb{R} such that aja_{j}\to\infty and D~aj\tilde{D}_{a_{j}} are Lipschitz domains. A solution vHloc1(ΣR)v\in H^{1}_{loc}(\Sigma_{R}) satisfies the Sommerfeld radiation condition in integral form if

vrikvH1/2(Caj) 0,j,\biggl{\|}\frac{\partial v}{\partial r}-ikv\biggr{\|}_{H^{-1/2}(C_{a_{j}})}\ \longrightarrow\ 0\,,\quad j\to\infty\,,

where r=|x|r=|x|.

Lemma 2.13.

If vv satisfies the Sommerfeld radiation condition of Definition 2.8 with the index ρ0\rho\geq 0, then v also fulfills the integral form of the radiation condition defined by Definition 2.12.

Proof. Without loss of generality we suppose that aj+1aj1a_{j+1}-a_{j}\geq 1 for all jj\in{\mathbb{N}}. Let h0>0h_{0}>0 be the number specified in Definition 2.8. We set h:=h0+1h:=h_{0}+1 and choose ψC(2)\psi\in C^{\infty}(\mathbb{R}^{2}) such that ψ(x)=0\psi(x)=0 for xUhx\in U_{h} and ψ(x)=1\psi(x)=1 for xUhεx\notin U_{h-\varepsilon}. We decompose CajC_{a_{j}} into Caj=(CajUh)(CajUh)C_{a_{j}}=\bigl{(}C_{a_{j}}\cap U_{h}\bigr{)}\cup\bigl{(}C_{a_{j}}\setminus U_{h}\bigr{)}. Then

(15) vrikvH1/2(Caj)\displaystyle\biggl{\|}\frac{\partial v}{\partial r}-ikv\biggr{\|}_{H^{-1/2}(C_{a_{j}})} =\displaystyle= ψ(vrikv)H1/2(Caj)+(1ψ)(vrikv)H1/2(Caj)\displaystyle\biggl{\|}\psi\biggl{(}\frac{\partial v}{\partial r}-ikv\biggr{)}\biggr{\|}_{H^{-1/2}(C_{a_{j}})}\ +\ \biggl{\|}(1-\psi)\biggl{(}\frac{\partial v}{\partial r}-ikv\biggr{)}\biggr{\|}_{H^{-1/2}(C_{a_{j}})}
\displaystyle\leq ψvrH1/2(Caj)+kψvH1/2(Caj)\displaystyle\biggl{\|}\psi\frac{\partial v}{\partial r}\biggr{\|}_{H^{-1/2}(C_{a_{j}})}\ +\ k\bigl{\|}\psi v\bigr{\|}_{H^{-1/2}(C_{a_{j}})}
+(1ψ)(vrikv)H1/2(Caj).\displaystyle+\ \biggl{\|}(1-\psi)\biggl{(}\frac{\partial v}{\partial r}-ikv\biggr{)}\biggr{\|}_{H^{-1/2}(C_{a_{j}})}.

The last integral converges to zero because, by the Sommerfeld radiation condition of (14),

(1ψ)(vrikv)H1/2(Caj)CvrikvL2(CajUhϵ) 0\displaystyle\biggl{\|}(1-\psi)\biggl{(}\frac{\partial v}{\partial r}-ikv\biggr{)}\biggr{\|}_{H^{-1/2}(C_{a_{j}})}\ \leq\ C\,\biggl{\|}\frac{\partial v}{\partial r}-ikv\biggr{\|}_{L^{2}(C_{a_{j}}\cap U_{h-\epsilon})}\ \longrightarrow\ 0

as jj\rightarrow\infty. It remains to discuss the first two integrals on the right hand side of (15). Let Ej:H01/2(Caj)H1(Daj+1Daj¯)E_{j}:H^{1/2}_{0}(C_{a_{j}})\to H^{1}(D_{a_{j}+1}\setminus\overline{D_{a_{j}}}) be extension operators which are uniformly bounded with respect to jj. In fact, given φH01/2(Caj)\varphi\in H^{1/2}_{0}(C_{a_{j}}) we define Ejφ=wjE_{j}\varphi=w_{j} in Daj+1Daj¯D_{a_{j+1}}\setminus\overline{D_{a_{j}}} where wjw_{j} is the unique solution to the boundary value problem

Δwj=0inDaj+1Daj¯,\displaystyle\Delta w_{j}=0\quad\mbox{in}\quad D_{a_{j+1}}\setminus\overline{D_{a_{j}}},
wj=φonCaj,wj=0on(Daj+1Daj¯)Caj.\displaystyle w_{j}=\varphi\quad\mbox{on}\quad C_{a_{j}},\qquad w_{j}=0\quad\mbox{on}\quad\partial(D_{a_{j+1}}\setminus\overline{D_{a_{j}}})\setminus C_{a_{j}}.

The norm of such an extension operator depends only on the Lipschitz constants of Daj+1Daj¯D_{a_{j+1}}\setminus\overline{D_{a_{j}}}, which are uniformly bounded in jj. Then we have

ψvrH1/2(Caj)+kψvH1/2(Caj)\displaystyle\biggl{\|}\psi\,\frac{\partial v}{\partial r}\biggr{\|}_{H^{-1/2}(C_{a_{j}})}\ +\ k\bigl{\|}\psi v\bigr{\|}_{H^{-1/2}(C_{a_{j}})}
\displaystyle\leq (ψv)rH1/2(Caj)+vψrH1/2(Caj)+kψvH1/2(Caj)\displaystyle\biggl{\|}\frac{\partial(\psi v)}{\partial r}\biggr{\|}_{H^{-1/2}(C_{a_{j}})}\ +\ \biggl{\|}v\,\frac{\partial\psi}{\partial r}\biggr{\|}_{H^{-1/2}(C_{a_{j}})}\ +\ k\bigl{\|}\psi v\bigr{\|}_{H^{-1/2}(C_{a_{j}})}
\displaystyle\leq supφH01/2(Caj)=1r(ψv),φ+cvH1/2(CajUh)\displaystyle\sup_{\|\varphi\|_{H^{1/2}_{0}(C_{a_{j}})}=1}\langle\partial_{r}(\psi v),\varphi\rangle\ +\ c\,\bigl{\|}v\bigr{\|}_{H^{1/2}(C_{a_{j}}\setminus U_{h})}
=\displaystyle= supφH01/2(Caj)=1Daj+1Daj¯[(ψv)Ejφ¯k2ψvEjφ¯]𝑑x+cvH1/2(CajUh)\displaystyle\sup_{\|\varphi\|_{H^{1/2}_{0}(C_{a_{j}})}=1}\int\limits_{D_{a_{j+1}}\setminus\overline{D_{a_{j}}}}\bigl{[}\nabla(\psi v)\cdot\nabla\overline{E_{j}\varphi}-k^{2}\psi v\,\overline{E_{j}\varphi}\bigr{]}\,dx\ +\ c\,\bigl{\|}v\bigr{\|}_{H^{1/2}(C_{a_{j}}\setminus U_{h})}
\displaystyle\leq cvH1(Zj)\displaystyle c\,\|v\|_{H^{1}(Z_{j})}

where Zj={xDaj+1:|x|>aj,x2<h}Z_{j}=\{x\in D_{a_{j+1}}:|x|>a_{j}\,,\ x_{2}<h\} and c>0c>0 is independent of jj. Simple estimates show that ZjZ_{j} is contained in the set {xD:ajε<x1<aj+1,x2<h}\{x\in D:a_{j}-\varepsilon<x_{1}<a_{j+1},\ x_{2}<h\}. From vH1(Wh)v\in H^{1}(W_{h}) we conclude that vH1(Zj)\|v\|_{H^{1}(Z_{j})} tends to zero. ∎

2.4. Uniqueness Of Solutions Of The Perturbed And Unperturbed Problems

First we show that the propagating part upropu_{prop} of the open waveguide radiation condition 2.5 has to vanish, if f=0f=0.

Theorem 2.14.

Let uHloc,01(D~)u\in H^{1}_{loc,0}(\tilde{D}) be a solution of Δu+k2u=0\Delta u+k^{2}u=0 in D~\tilde{D} satisfying the open waveguide radiation condition of Definition 2.5. Then upropu_{prop} vanishes; that is, all the coefficients a,ja_{\ell,j} vanish.

Proof. Choose ψNC()\psi_{N}\in C^{\infty}(\mathbb{R}) and φHC()\varphi_{H}\in C^{\infty}(\mathbb{R}) with ψN(x1)=1\psi_{N}(x_{1})=1 for |x1|N|x_{1}|\leq N and ψN(x1)=0\psi_{N}(x_{1})=0 for |x1|N+1|x_{1}|\geq N+1 and φH(x2)=0\varphi_{H}(x_{2})=0 for x2H+1x_{2}\geq H+1 and φH(x2)=1\varphi_{H}(x_{2})=1 for x2Hx_{2}\leq H.
For N>σ0+1N>\sigma_{0}+1 and H>h0+1H>h_{0}+1 we define the regions DN,H:={xD~:|x1|<N,x2<H}D_{N,H}:=\{x\in\tilde{D}:|x_{1}|<N,\ x_{2}<H\} and WN,H:={xD~:N1<x1<N,x2<H}W^{-}_{N,H}:=\{x\in\tilde{D}:-N-1<x_{1}<-N,\ x_{2}<H\} and WN,H+:={xD~:N<x1<N+1,x2<H}W^{+}_{N,H}:=\{x\in\tilde{D}:N<x_{1}<N+1,\ x_{2}<H\} and the horizontal line segments ΓN,H:=(N,N)×{H}\Gamma_{N,H}:=(-N,N)\times\{H\}. We apply Green’s theorem in DN+1,H+1D_{N+1,H+1} to v(x):=ψN(x1)u(x)v(x):=\psi_{N}(x_{1})\,u(x) and v(x)¯φH(x2)\overline{v(x)}\,{\color[rgb]{0,0,0}\varphi_{H}}(x_{2}). First we note that ΔvL2(DN+1,H+1)\Delta v\in L^{2}(D_{N+1,H+1}) because Δu=k2u\Delta u=-k^{2}u. Furthermore vφH01(DN+1,H+1)v\varphi\in H^{1}_{0}(D_{N+1,H+1}), therefore,

0\displaystyle 0 =\displaystyle= DN+1,H+1[v(v¯φH)+(v¯φH)Δv]𝑑x\displaystyle\int\limits_{D_{N+1,H+1}}\bigl{[}\nabla v\cdot\nabla(\overline{v}\,{\color[rgb]{0,0,0}\varphi_{H}})+(\overline{v}\,{\color[rgb]{0,0,0}\varphi_{H}})\,\Delta v\bigr{]}\,dx
=\displaystyle= DN+1,H[|v|2+v¯Δv]𝑑x+DN+1,H+1DN+1,H[v(v¯φH)+(v¯φH)Δv]𝑑x\displaystyle\int\limits_{D_{N+1,H}}\bigl{[}|\nabla v|^{2}+\overline{v}\,\Delta v\bigr{]}\,dx\ +\int\limits_{D_{N+1,H+1}\setminus D_{N+1,H}}\bigl{[}\nabla v\cdot\nabla(\overline{v}\,{\color[rgb]{0,0,0}\varphi_{H}})+(\overline{v}\,{\color[rgb]{0,0,0}\varphi_{H}})\,\Delta v\bigr{]}\,dx
=\displaystyle= DN+1,H[|v|2+v¯Δv]𝑑xΓN+1,Hv¯vx2𝑑s\displaystyle\int\limits_{D_{N+1,H}}\bigl{[}|\nabla v|^{2}+\overline{v}\,\Delta v\bigr{]}\,dx\ -\int\limits_{\Gamma_{N+1,H}}\overline{v}\,\frac{\partial v}{\partial x_{2}}\,ds

where we applied the classical Green’s theorem in the rectangle (N1,N+1)×(H,H+1)(-N-1,N+1)\times(H,H+1) to the second integral for the smooth function vv. Therefore,

ΓN+1,HψN2u¯ux2𝑑s=ΓN+1,Hv¯vx2𝑑s=DN+1,H[|v|2+v¯Δv]𝑑x\displaystyle\int\limits_{\Gamma_{N+1,H}}\psi_{N}^{2}\,\overline{u}\,\frac{\partial u}{\partial x_{2}}\,ds=\int\limits_{\Gamma_{N+1,H}}\overline{v}\,\frac{\partial v}{\partial x_{2}}\,ds\ =\ \int\limits_{D_{N+1,H}}\bigl{[}\bigl{|}\nabla v\bigr{|}^{2}+\overline{v}\,\Delta v\bigr{]}\,dx
=\displaystyle= DN,H[|u|2+u¯Δu]𝑑x+WN,H+[|v|2+v¯Δv]𝑑x+WN,H[|v|2+v¯Δv]𝑑x;\displaystyle\int\limits_{D_{N,H}}\bigl{[}\bigl{|}\nabla u\bigr{|}^{2}+\overline{u}\,\Delta u\bigr{]}\,dx+\ \int\limits_{W^{+}_{N,H}}\bigl{[}\bigl{|}\nabla v\bigr{|}^{2}+\overline{v}\,\Delta v\bigr{]}\,dx\ +\ \int\limits_{W^{-}_{N,H}}\bigl{[}\bigl{|}\nabla v\bigr{|}^{2}+\overline{v}\,\Delta v\bigr{]}\,dx\,;

that is, with Δu=k2u\Delta u=-k^{2}u,

(16) ImΓN+1,HψN2u¯ux2𝑑s=ImWN,H+[|v|2+v¯Δv]𝑑x+ImWN,H[|v|2+v¯Δv]𝑑x.\operatorname{Im}\int\limits_{\Gamma_{N+1,H}}\psi_{N}^{2}\,\overline{u}\,\frac{\partial u}{\partial x_{2}}\,ds\ =\ \operatorname{Im}\int\limits_{W^{+}_{N,H}}\bigl{[}\bigl{|}\nabla v\bigr{|}^{2}+\overline{v}\,\Delta v\bigr{]}\,dx\ +\ \operatorname{Im}\int\limits_{W^{-}_{N,H}}\bigl{[}\bigl{|}\nabla v\bigr{|}^{2}+\overline{v}\,\Delta v\bigr{]}\,dx\,.

The decomposition u=urad+upropu=u_{rad}+u_{prop} yields four terms in each of the integrals of (16).
(a) First, we look at the two integrals on the right hand side of (16). We define v(1):=ψNuradv^{(1)}:=\psi_{N}u_{rad} and v(2)=ψNupropv^{(2)}=\psi_{N}u_{prop} and estimate the terms

aN,H±(j,):=WN,H±[v(j)¯v()+v(j)¯Δv()]𝑑xa^{\pm}_{N,H}(j,\ell)\ :=\ \int\limits_{W^{\pm}_{N,H}}\bigl{[}\nabla\overline{v^{(j)}}\cdot\nabla v^{(\ell)}+\overline{v^{(j)}}\,\Delta v^{(\ell)}\bigr{]}\,dx

for j,{1,2}j,\ell\in\{1,2\}. Then |aN,H±(1,1)||a^{\pm}_{N,H}(1,1)|, |aN,H±(1,2)||a^{\pm}_{N,H}(1,2)|, and |aN,H±(2,1)||a^{\pm}_{N,H}(2,1)| are estimated as in the proof of [19, Theorem 2.2]:

|aN,H±(1,1)|cγN,H,|aN,H±(1,2)|+|aN,H±(2,1)|cγN,H|a^{\pm}_{N,H}(1,1)|\ \leq\ c\,\gamma_{N,H}\,,\qquad|a^{\pm}_{N,H}(1,2)|\ +\ |a^{\pm}_{N,H}(2,1)|\ \leq\ c\,\sqrt{\gamma_{N,H}}

with

(17) γN,H:=uradH1(QN)2+H3N<|x1|<N+1ρ(x1)2𝑑x1\gamma_{N,H}\ :=\ \|u_{rad}\|_{H^{1}(Q_{N})}^{2}\ +\ H^{3}\int\limits_{N<|x_{1}|<N+1}\rho(x_{1})^{2}\,dx_{1}

and QN:={xD~:N<|x1|<N+1,x2<h0+1}Q_{N}:=\{x\in\tilde{D}:N<|x_{1}|<N+1,\ x_{2}<h_{0}+1\}.
For aN,H±(2,2)a^{\pm}_{N,H}(2,2) we need to argue differently as in the proof of [19, Theorem 3.2] to avoid the integral over the vertical boundaries of WN,H±W_{N,H}^{\pm}. We recall that

aN,H+(2,2)=WN,H+[|(ψNuprop)|2+(ψNuprop¯)Δ(ψNuprop)]𝑑xa^{+}_{N,H}(2,2)\ =\ \int\limits_{W^{+}_{N,H}}\bigl{[}|\nabla(\psi_{N}\,u_{prop})|^{2}+(\psi_{N}\,\overline{u_{prop}})\,\Delta(\psi_{N}u_{prop})\bigr{]}\,dx

and note that Δ(ψNuprop)=k2ψNuprop+2ψNupropx1+ψN′′uprop\Delta(\psi_{N}\,u_{prop})=-k^{2}\psi_{N}\,u_{prop}+2\psi_{N}^{\prime}\,\frac{\partial u_{prop}}{\partial x_{1}}+\psi_{N}^{\prime\prime}\,u_{prop}. Therefore,

ImaN,H+(2,2)\displaystyle\operatorname{Im}a^{+}_{N,H}(2,2) =\displaystyle= 2ImWN,H+ψNψNuprop¯upropx1𝑑x=ImWN,H+ddx1ψN2uprop¯upropx1𝑑x\displaystyle 2\operatorname{Im}\int\limits_{W^{+}_{N,H}}\psi_{N}\,\psi_{N}^{\prime}\,\overline{u_{prop}}\,\frac{\partial u_{prop}}{\partial x_{1}}\,dx\ =\ \operatorname{Im}\int\limits_{W^{+}_{N,H}}\frac{d}{dx_{1}}\psi_{N}^{2}\,\overline{u_{prop}}\,\frac{\partial u_{prop}}{\partial x_{1}}\,dx
=\displaystyle= ImWN,H+[uprop(ψN2uprop¯)k2(ψN2uprop¯)uprop]𝑑x\displaystyle\operatorname{Im}\int\limits_{W^{+}_{N,H}}\bigl{[}\nabla u_{prop}\cdot\nabla(\psi_{N}^{2}\,\overline{u_{prop}})-k^{2}(\psi_{N}^{2}\,\overline{u_{prop}})\,u_{prop}\bigr{]}\,dx
=\displaystyle= ImN+Q[uprop(ψN2uprop¯)k2(ψN2uprop¯)uprop]𝑑xβN,H+\displaystyle\operatorname{Im}\int\limits_{N+Q_{\infty}}\bigl{[}\nabla u_{prop}\cdot\nabla(\psi_{N}^{2}\,\overline{u_{prop}})-k^{2}(\psi_{N}^{2}\,\overline{u_{prop}})\,u_{prop}\bigr{]}\,dx\ -\ \beta^{+}_{N,H}

where again N+Q={xD~:N<x1<N+2π}N+Q_{\infty}=\{x\in\tilde{D}:N<x_{1}<N+2\pi\} and

|βN,H+|=|(N+Q)WN,H+[uprop(ψN2uprop¯)k2(ψN2uprop¯)uprop]𝑑x|ce2δH.|\beta^{+}_{N,H}|\ =\ \biggl{|}\int\limits_{(N+Q_{\infty})\setminus W^{+}_{N,H}}\bigl{[}\nabla u_{prop}\cdot\nabla(\psi_{N}^{2}\,\overline{u_{prop}})-k^{2}(\psi_{N}^{2}\,\overline{u_{prop}})\,u_{prop}\bigr{]}\,dx\biggr{|}\ \leq\ c\,e^{-2\delta H}\,.

Now we set φ(x1)=1(x1N)/(2π)\varphi(x_{1})=1-(x_{1}-N)/(2\pi) and observe that ψN2φ\psi_{N}^{2}-\varphi vanishes for x1=Nx_{1}=N and x1=N+2πx_{1}=N+2\pi. Green’s theorem implies that

N+Q[uprop((ψN2φ)uprop¯)k2((ψN2φ)uprop¯)uprop]𝑑x= 0\int\limits_{N+Q_{\infty}}\bigl{[}\nabla u_{prop}\cdot\nabla\bigl{(}(\psi_{N}^{2}-\varphi)\,\overline{u_{prop}}\bigr{)}-k^{2}\bigl{(}(\psi_{N}^{2}-\varphi)\,\overline{u_{prop}}\bigr{)}\,u_{prop}\bigr{]}\,dx\ =\ 0

and thus

ImaN,H+(2,2)\displaystyle\operatorname{Im}a^{+}_{N,H}(2,2) =\displaystyle= ImN+Q[uprop(φuprop¯)k2(φuprop¯)uprop]𝑑xβN,H+\displaystyle\operatorname{Im}\int\limits_{N+Q_{\infty}}\bigl{[}\nabla u_{prop}\cdot\nabla(\varphi\,\overline{u_{prop}})-k^{2}(\varphi\,\overline{u_{prop}})\,u_{prop}\bigr{]}\,dx\ -\ \beta^{+}_{N,H}
=\displaystyle= 12πImN+Quprop¯upropx1𝑑xβN,H+\displaystyle-\frac{1}{2\pi}\operatorname{Im}\int\limits_{N+Q_{\infty}}\overline{u_{prop}}\,\frac{\partial u_{prop}}{\partial x_{1}}\,dx\ -\ \beta^{+}_{N,H}
=\displaystyle= 14πjJλ,j>0λ,j|a,j|2βN,H+\displaystyle-\frac{1}{4\pi}\sum_{j\in J}\sum_{\lambda_{\ell,j}>0}\lambda_{\ell,j}\,|a_{\ell,j}|^{2}\ -\ \beta^{+}_{N,H}

where we used the results of Lemma 2.7 above. The same estimates hold for aN,H(j,)a_{N,H}^{-}(j,\ell); that is, the integrals over WN,HW^{-}_{N,H}. Therefore, we have shown that

ImWN,H+[|v|2+v¯Δv]𝑑x+ImWN,H[|v|2+v¯Δv]𝑑x\displaystyle\operatorname{Im}\int\limits_{W^{+}_{N,H}}\bigl{[}\bigl{|}\nabla v\bigr{|}^{2}+\overline{v}\,\Delta v\bigr{]}\,dx\ +\ \operatorname{Im}\int\limits_{W^{-}_{N,H}}\bigl{[}\bigl{|}\nabla v\bigr{|}^{2}+\overline{v}\,\Delta v\bigr{]}\,dx
\displaystyle\leq 14πjJλ,j>0λ,j|a,j|2+14πjJλ,j<0λ,j|a,j|2+ce2δH+c[γN,H+γN,H].\displaystyle-\frac{1}{4\pi}\sum_{j\in J}\sum_{\lambda_{\ell,j}>0}\lambda_{\ell,j}\,|a_{\ell,j}|^{2}\ +\ \frac{1}{4\pi}\sum_{j\in J}\sum_{\lambda_{\ell,j}<0}\lambda_{\ell,j}\,|a_{\ell,j}|^{2}\ +\ c\,e^{-2\delta H}\ +\ c\,[\gamma_{N,H}+\sqrt{\gamma_{N,H}}]\,.

(b) Now we look at the left hand side of (16). The line integrals are outside of the layer Wh0W_{h_{0}}. Their estimates in [19] (proof of Theorem 3.2) are independent of the equation or boundary condition below the line x2=h0x_{2}=h_{0}. In [19] we have shown the existence of sequences (Nm)(N_{m}) and (Hm)(H_{m}) converging to infinity such that γNm,Hm0\gamma_{N_{m},H_{m}}\to 0 and

(19) lim supm[ImΓNm+1,HmψNm2u¯ux2𝑑s] 0.\limsup_{m\to\infty}\biggl{[}\operatorname{Im}\int\limits_{\Gamma_{N_{m}+1,H_{m}}}\psi_{N_{m}}^{2}\,\overline{u}\,\frac{\partial u}{\partial x_{2}}\,ds\biggr{]}\ \geq\ 0\,.

From (2.4) we conclude that

lim supm[ImWNm,Hm+[|v|2+v¯Δv]𝑑x+ImWNm,Hm[|v|2+v¯Δv]𝑑x]\displaystyle\limsup\limits_{m\to\infty}\biggl{[}\operatorname{Im}\int\limits_{W^{+}_{N_{m},H_{m}}}\bigl{[}\bigl{|}\nabla v\bigr{|}^{2}+\overline{v}\,\Delta v\bigr{]}\,dx\ +\ \operatorname{Im}\int\limits_{W^{-}_{N_{m},H_{m}}}\bigl{[}\bigl{|}\nabla v\bigr{|}^{2}+\overline{v}\,\Delta v\bigr{]}\,dx\biggr{]}
\displaystyle\leq 14πjJλ,j>0λ,j|a,j|2+14πjJλ,j<0λ,j|a,j|2.\displaystyle-\frac{1}{4\pi}\sum_{j\in J}\sum_{\lambda_{\ell,j}>0}\lambda_{\ell,j}\,|a_{\ell,j}|^{2}\ +\ \frac{1}{4\pi}\sum_{j\in J}\sum_{\lambda_{\ell,j}<0}\lambda_{\ell,j}\,|a_{\ell,j}|^{2}\,.

Combining this estimate with (19) and (16) yields that a,j=0a_{\ell,j}=0 for all \ell and jj. ∎

Below we sketch another proof based on the modified open waveguide radiation condition.

Theorem 2.15.

Let uHloc,01(D~)u\in H^{1}_{loc,0}(\tilde{D}) be a solution of Δu+k2u=0\Delta u+k^{2}u=0 in D~\tilde{D} satisfying the modified open waveguide radiation condition of Definition 2.10. Then upropu_{prop} vanishes.

Proof. Choose a>0a>0 and suppose without loss of generality that D~a\tilde{D}_{a} is a Lipschitz domain. Applying Green’s formula for uu to D~a\tilde{D}_{a} gives

D~a|u|2k2|u|2dx+Caνuu¯ds=0.\int_{\tilde{D}_{a}}|\nabla u|^{2}-k^{2}|u|^{2}\,dx+\int_{C_{a}}\partial_{\nu}u\overline{u}\,ds=0.

Here we have used the Dirichlet boundary condition on D~\partial\tilde{D}. Taking the imaginary part and using u=urad+upropu=u_{rad}+u_{prop} yields

0=ImCaνuu¯ds=ImCa[νuradu¯rad+νuradu¯prop+νupropu¯rad+νupropu¯prop]𝑑s0=\operatorname{Im}\int_{C_{a}}\partial_{\nu}u\overline{u}\,ds=\operatorname{Im}\int_{C_{a}}[\partial_{\nu}u_{rad}\overline{u}_{rad}+\partial_{\nu}u_{rad}\overline{u}_{prop}+\partial_{\nu}u_{prop}\overline{u}_{rad}+\partial_{\nu}u_{prop}\overline{u}_{prop}]\,ds

Recalling the Sommerfeld radiation condition of uradu_{rad}, one can show that the integrals involving the term uradu_{rad} on the right hand side all vanish as aa\rightarrow\infty; see the proof of [12, Theorem 3.1] for details. Therefore, one arrives at

0=limaImCaνupropu¯propds=limaImD~a(Δ+k2)upropu¯prop𝑑x,0=\lim_{a\rightarrow\infty}\operatorname{Im}\int_{C_{a}}\partial_{\nu}u_{prop}\,\overline{u}_{prop}\,ds=\lim_{a\rightarrow\infty}\operatorname{Im}\int_{\tilde{D}_{a}}(\Delta+k^{2})u_{prop}\,\overline{u}_{prop}\,dx,

because upropu_{prop} vanishes on D~\partial\tilde{D}. Rcalling the representation of upropu_{prop} (see (7)) and using the definition of ψ±\psi^{\pm}, one deduces that (see e.g., [12, Lemma 2.3])

(20) D~a(Δ+k2)upropu¯prop𝑑x=(γa+γa)uprop¯upropx1ds+Sa+Sauprop¯upropν𝑑s\displaystyle\int_{\tilde{D}_{a}}(\Delta+k^{2})u_{prop}\,\overline{u}_{prop}\,dx=\left(\int\limits_{\gamma_{a}^{+}}-\int\limits_{\gamma_{a}^{-}}\right)\overline{u_{prop}}\,\frac{\partial u_{prop}}{\partial x_{1}}\,ds\ +\int\limits_{S_{a}^{+}\cup S_{a}^{-}}\overline{u_{prop}}\,\frac{\partial u_{prop}}{\partial\nu}\,ds

where γa±=D~a{x1=±σ0}\gamma_{a}^{\pm}=\tilde{D}_{a}\cap\{x_{1}=\pm\sigma_{0}\} and Sa±={xD~a:|x|=a,σ01<±x1<σ0}S_{a}^{\pm}=\{x\in\tilde{D}_{a}:|x|=a,\sigma_{0}-1<\pm x_{1}<\sigma_{0}\}. Note that here have supposed that {xD~a:σ01<±x1<σ0}\{x\in\tilde{D}_{a}:\sigma_{0}-1<\pm x_{1}<\sigma_{0}\} are both Lipschitz domains by the choice of σ0>max{R,2π}+1\sigma_{0}>\max\{R,2\pi\}+1 and that the integrals over γa±\gamma_{a}^{\pm} are understood in the dual form between H1/2(γa±)H^{-1/2}(\gamma_{a}^{\pm}) and H01/2(γa±)H^{1/2}_{0}(\gamma_{a}^{\pm}). The second term on the right hand side of (20) tends to zero as aa\rightarrow\infty, due to the exponential decay of upropu_{prop} as x2x_{2}\rightarrow\infty. For the imaginary part of the first term, we have the limit (see [19, Lemma 2.6])

ImD~{x1=±σ0}uprop¯upropx1𝑑s\displaystyle\operatorname{Im}\int\limits_{\tilde{D}\cap\{x_{1}=\pm\sigma_{0}\}}\overline{u_{prop}}\,\frac{\partial u_{prop}}{\partial x_{1}}\,ds =\displaystyle= 12πImD~{±σ0+Q}uprop¯upropx1𝑑x\displaystyle\frac{1}{2\pi}\operatorname{Im}\int\limits_{\tilde{D}\cap\{\pm\sigma_{0}+Q_{\infty}\}}\overline{u_{prop}}\,\frac{\partial u_{prop}}{\partial x_{1}}dx
=\displaystyle= 14πjJ±λ,j>0λ,j|a,j|2.\displaystyle\frac{1}{4\pi}\displaystyle\sum_{j\in J}\sum_{\pm\lambda_{\ell,j}>0}\lambda_{\ell,j}\,|a_{\ell,j}|^{2}.

In this step we have used the fact ±σ0+Q\pm\sigma_{0}+Q_{\infty} are Lipschitz domains and Lemma 2.6. Finally, taking the imaginary part in (20) and letting aa\rightarrow\infty, we obtain al,j=0a_{l,j}=0 for all jJj\in J and l=1,2,,mjl=1,2,\cdots,m_{j}. This proves uprop0u_{prop}\equiv 0. ∎


Having proved the unique determination of the propagating part, we can show uniqueness of solutions of the unperturbed and perturbed boundary value problems following almost the same lines in the proof of [19, Theorem 3.3]. We omit the proof of Theorem 2.16 below.

Theorem 2.16.

Let the Assumptions 2.2 and 2.3 hold.

  • (i)

    Let uHloc,01(D)u\in H^{1}_{loc,0}(D) be a solution of Δu+k2u=0\Delta u+k^{2}u=0 in DD satisfying the open waveguide radiation condition of Definition 2.5. Then u0u\equiv 0.

  • (ii)

    In the perturbed case we have u0u\equiv 0, if there are no bound states to the problem (1), that is, any solution uH01(D~)u\in H^{1}_{0}(\tilde{D}) of Δu+k2u=0\Delta u+k^{2}u=0 in D~\tilde{D} must vanish identically.

In the remaining part we suppose that there are no bound states for the perturbed scattering problem, so that uniqueness always holds true by Theorem 2.16 (ii). Note that this assumption can be removed, if the domain D~\tilde{D} fulfills the following condition (see [4]):

(21) (x1,x2)D~(x1,x2+s)D~for all s>0.(x_{1},x_{2})\in\tilde{D}\quad\Longrightarrow\quad(x_{1},x_{2}+s)\in\tilde{D}\quad\mbox{for all }s>0\,.

Obviously, the geometrical condition (21) can be fulfilled if the boundary Γ~\tilde{\Gamma} is given by the graph of some continuous function. But then also the existence of guided modes is excluded.

3. Construction of the Dirichlet-to-Neumann (DtN) operator

For simplicity we suppose that there is an open arc of the form CR:={xD:|x1π|2+|x2|2=R2}C_{R}:=\{x\in D:|x_{1}-\pi|^{2}+|x_{2}|^{2}=R^{2}\} for some R>πR>\pi such that the domain DRD_{R} is Lipschitz (otherwise we can replace CRC_{R} by an open curve with a slightly different shape). This implies that the perturbed defect ΓΓ~\Gamma\setminus\tilde{{\Gamma}} always lies below CRC_{R}. We refer to Figure 2 for a typical situation.

To reduce the scattering problem to a bounded domain, we need Sobolev spaces defined on an open arc. Define the Sobolev spaces (see [22])

H01/2(CR)\displaystyle H^{1/2}_{0}(C_{R}) :=\displaystyle:= {fH1/2(DR):f=0 on DCR},\displaystyle\bigl{\{}f\in H^{1/2}(\partial D_{R}):f=0\mbox{ on }\partial D\setminus C_{R}\bigr{\}}\,,
H1/2(CR)\displaystyle H^{1/2}(C_{R}) :=\displaystyle:= {f|CR:fH1/2(DR)}.\displaystyle\bigl{\{}f|_{C_{R}}:f\in H^{1/2}(\partial D_{R})\bigr{\}}\,.

An important property of H01/2(CR)H^{1/2}_{0}(C_{R}) is that the zero extension of uu to DR\partial D_{R} belongs to H1/2(DR)H^{1/2}(\partial D_{R}). We remark that in the previous definitions the closed boundary DR\partial D_{R} can be replaced by other closed boundaries. If uH1(DR)u\in H^{1}(D_{R}) with u=0u=0 on ΓDR¯\Gamma\cap\overline{D_{R}}, then we have the traces u|CRH01/2(CR)u|_{C_{R}}\in H_{0}^{1/2}(C_{R}). The spaces H01/2(CR){\color[rgb]{0,0,0}H^{1/2}_{0}(C_{R})} and H1/2(CR)H^{-1/2}(C_{R}) are (anti-linear) dual spaces in the sense that

ϕ,ψH01/2(CR),H1/2(CR)=ϕ~,ψH1/2(DR),H1/2(DR),\displaystyle\langle\phi,\psi\rangle_{H^{1/2}_{0}(C_{R}),H^{-1/2}(C_{R})}\ =\ \langle\tilde{\phi},\psi\rangle_{H^{1/2}(\partial D_{R}),H^{-1/2}(D_{R})}\,,

where ϕ~\tilde{\phi} denotes the zero extension of ϕ\phi to DR\partial D_{R}. We further remark that for any a>Ra>R there exists a bounded extension operator EE from H01/2(CR)H^{1/2}_{0}(C_{R}) into H01(Da)H^{1}_{0}(D_{a}). Indeed, extending ψH01/2(CR)\psi\in H^{1/2}_{0}(C_{R}) by zero in DRΓ\partial D_{R}\cap\Gamma we observe that this extension is in H1/2(DR)H^{1/2}(\partial{D}_{R}). By well known results for Lipschitz domains there exists a bounded extension operator E1E_{1} from H1/2(DR)H^{1/2}(\partial D_{R}) into H1(DR)H^{1}(D_{R}). In the same way one extends ψH01/2(CR)\psi\in H^{1/2}_{0}(C_{R}) by zero in DaDR¯\partial D_{a}\setminus\overline{D_{R}} and constructs an extension operator E2E_{2} from H1/2((DaDR))H^{1/2}(\partial(D_{a}\setminus D_{R})) into H1(DaDR)H^{1}(D_{a}\setminus D_{R}) with zero boundary values for |x|=a|x|=a.

Below we recall the definition of the Floquet-Bloch transform to be used later.

Definition 3.1.

For gC0()g\in C_{0}^{\infty}({\mathbb{R}}), the Floquet-Bloch transform FF is defined by

(Fg)(x1,α):=ng(x1+2πn)ei2πnα,x1,α[1/2,1/2].(Fg)(x_{1},\alpha)\ :=\ \sum_{n\in{\mathbb{Z}}}g(x_{1}+2\pi n)\,e^{-i2\pi n\alpha}\,,\qquad x_{1}\in{\mathbb{R}},\ \alpha\in[-1/2,1/2]\,.

The Floquet-Bloch transform FF extends to an unitary operator from L2()L^{2}({\mathbb{R}}) to L2((0,2π)×(1/2,1/2))L^{2}((0,2\pi)\times(-1/2,1/2)). If gg depends on two variables x1x_{1} and x2x_{2} then the symbol FF means the Floquet-Bloch transform with respect to x1x_{1}.

In the next subsection we prepare several auxiliary results before constructing the DtN operator.

3.1. Existence Results For Some Unperturbed Problems

The first result is well known and a simple application of the Theorem of Riesz. Define the weighted Sobolev spaces H(ρ)s={uH0s(D):wρuHs(D)}H^{s}_{(\rho)}=\bigl{\{}u\in H^{s}_{0}(D):w_{\rho}u\in H^{s}(D)\bigr{\}} where wρ(x)=eρ|x|w_{\rho}(x)=e^{\rho|x|} for ρ0\rho\geq 0.

Theorem 3.2.

Let φH1/2(CR)\varphi\in H^{-1/2}(C_{R}) and ρ(0,1)\rho\in(0,1). Then there exists a unique solution vH01(D)v\in H^{1}_{0}(D) of

(22) D[vψ¯+vψ¯]𝑑x=CRφψ¯𝑑sfor all ψH01(D).\int\limits_{D}[\nabla v\cdot\nabla\overline{\psi}+v\,\overline{\psi}]\,dx\ =\ \int\limits_{C_{R}}\varphi\,\overline{\psi}\,ds\quad\mbox{for all }\psi\in H^{1}_{0}(D)\,.

Note that we have written the dual form φ,ψ\langle\varphi,\psi\rangle on the right hand side as integral. Here we need that the trace ψ|CRH01/2(CR)\psi|_{C_{R}}\in H^{1/2}_{0}(C_{R}). Furthermore, vH(ρ)1(D)v\in H^{1}_{(\rho)}(D) and φv\varphi\mapsto v is bounded from H1/2(CR)H^{-1/2}(C_{R}) into H(ρ)1(D)H^{1}_{(\rho)}(D) and even compact from H1/2(CR)H^{-1/2}(C_{R}) into L(ρ)2(D)L^{2}_{(\rho^{\prime})}(D) for all ρ<ρ\rho^{\prime}<\rho.

Proof: The left hand side is just the inner product in H1(D)H^{1}(D), and the right hand side is estimated by

|CRφψ¯𝑑s|φH1/2(CR)ψH01/2(CR)cφH1/2(CR)ψH1(D).\displaystyle\bigl{|}\int_{C_{R}}\varphi\,\overline{\psi}\,ds\bigr{|}\ \leq\ \|\varphi\|_{H^{-1/2}(C_{R})}\,\|\psi\|_{H^{1/2}_{0}(C_{R})}\ \leq\ c\|\varphi\|_{H^{-1/2}(C_{R})}\,\|\psi\|_{H^{1}(D)}.

Therefore, Riesz’s theorem implies uniqueness and existence of a solution in H01(D)H^{1}_{0}(D). Set v~=wρv\tilde{v}=w_{\rho}v and ψ~=1wρψ\tilde{\psi}=\frac{1}{w_{\rho}}\psi. Then ψ=wρψ~+wρψ~\nabla\psi=\nabla w_{\rho}\tilde{\psi}+w_{\rho}\nabla\tilde{\psi} and v=v~wρ2wρ+1wρv~\nabla v=-\frac{\tilde{v}}{w_{\rho}^{2}}\nabla w_{\rho}+\frac{1}{w_{\rho}}\nabla\tilde{v}. Substituting this into the variational equation yields

D[v~ψ~¯+ψ~¯wρwρv~v~wρwρψ~¯|wρ|2wρ2v~ψ~¯+v~ψ~¯]𝑑x=CRφψ~¯wρ𝑑s.\int\limits_{D}\biggl{[}\nabla\tilde{v}\cdot\nabla\overline{\tilde{\psi}}+\overline{\tilde{\psi}}\,\frac{\nabla w_{\rho}}{w_{\rho}}\cdot\nabla\tilde{v}-\tilde{v}\,\frac{\nabla w_{\rho}}{w_{\rho}}\cdot\nabla\overline{\tilde{\psi}}-\frac{|\nabla w_{\rho}|^{2}}{w_{\rho}^{2}}\,\tilde{v}\,\overline{\tilde{\psi}}+\tilde{v}\,\overline{\tilde{\psi}}\biggr{]}\,dx\ =\ \int\limits_{C_{R}}\varphi\,\overline{\tilde{\psi}}\,w_{\rho}\,ds\,.

We observe that the left hand side defines a sesqui-linear form on H1(D)H^{1}(D) which is coercive for ρ<1\rho<1 because |wρ|wρ=ρ\frac{|\nabla w_{\rho}|}{w_{\rho}}=\rho. The right hand side defines again a bounded linear form on H1(D)H^{1}(D). Therefore, Lax-Milgram yields existence and uniqueness. This proves that vH(ρ)1(D)v\in H^{1}_{(\rho)}(D) and that φv\varphi\mapsto v is bounded from H1/2(CR)H^{-1/2}(C_{R}) into H(ρ)1(D)H^{1}_{(\rho)}(D).

Finally we show that H(ρ)1(D)H^{1}_{(\rho)}(D) is compactly imbedded in L(ρ)2(D)L^{2}_{(\rho^{\prime})}(D) for all ρ<ρ\rho^{\prime}<\rho. Let (vj)(v_{j}) be a sequence in H(ρ)1(D)H^{1}_{(\rho)}(D) which converges weakly to zero. Set again v~j=wρvj\tilde{v}_{j}=w_{\rho}v_{j}. Then (v~j)(\tilde{v}_{j}) converges weakly to zero in H1(D)H^{1}(D) and is thus bounded. Therefore, there exists c>0c>0 with v~jL2(D)c\|\tilde{v}_{j}\|_{L^{2}(D)}\leq c for all jj. We estimate for any a>0a>0

DDawρ2(x)|vj(x)|2𝑑x\displaystyle\int\limits_{D\setminus D_{a}}w^{2}_{\rho^{\prime}}(x)\,|v_{j}(x)|^{2}\,dx =\displaystyle= DDawρ2(x)|vj(x)|2e2(ρρ)|x|𝑑x\displaystyle\int\limits_{D\setminus D_{a}}w^{2}_{\rho}(x)\,|v_{j}(x)|^{2}\,e^{-2(\rho-\rho^{\prime})|x|}\,dx
\displaystyle\leq e2(ρρ)aDwρ2(x)|vj(x)|2𝑑xc2e2(ρρ)a.\displaystyle e^{-2(\rho-\rho^{\prime})a}\int\limits_{D}w^{2}_{\rho}(x)\,|v_{j}(x)|^{2}\,dx\ \leq\ c^{2}\,e^{-2(\rho-\rho^{\prime})a}\,.

Given ε>0\varepsilon>0 we choose a>0a>0 with c2e2(ρρ)a<ε22c^{2}\,e^{-2(\rho-\rho^{\prime})a}<\frac{\varepsilon^{2}}{2} and keep rr fixed. Since (v~j)(\tilde{v}_{j}) tends to zero weakly in H1(D)H^{1}(D) it tends to zero weakly in H1(Da)H^{1}(D_{a}). Therefore, v~jL2(Da)\|\tilde{v}_{j}\|_{L^{2}(D_{a})} tends to zero and thus also Dawρ2|vj|2𝑑x\int_{D_{a}}w_{\rho^{\prime}}^{2}|v_{j}|^{2}dx because on DaD_{a} the norms wηvL2(Da)\|w_{\eta}v\|_{L^{2}(D_{a})} are all equivalent. Thus, for sufficiently large jj the term Dawρ2|vj|2𝑑x\int_{D_{a}}w_{\rho^{\prime}}^{2}|v_{j}|^{2}dx is less than ε22\frac{\varepsilon^{2}}{2}. ∎


The proofs of most existence results for the Helmholtz equation in periodic structures are based on the following result for quasi-periodic problems. For a proof we refer to [19, Theorems 4.2, 4.3, and Remark 4.4] adopted to the present situation.

Theorem 3.3.

Let Assumptions 2.2 and 2.3 hold and let gαL2(Q)g_{\alpha}\in L^{2}(Q_{\infty}) for α[1/2,1/2]\alpha\in[-1/2,1/2] depend continuously differentiable on α\alpha in [1/2,1/2][-1/2,1/2]. Let there exist c^>0\hat{c}>0 and δ>0\delta>0 with |gα(x)|+|gα(x)/α|c^eδx2|g_{\alpha}(x)|+|\partial g_{\alpha}(x)/\partial\alpha|\leq\hat{c}e^{-\delta x_{2}} for almost all xQx\in Q_{\infty} with x2>h0x_{2}>h_{0} and all α[1/2,1/2]\alpha\in[-1/2,1/2]. Furthermore, let GH1(Qh0)=H01(Qh0)G\in H^{-1}(Q_{h_{0}})=H^{1}_{0}(Q_{h_{0}})^{\ast} and assume that for any propagative wave number α^j[1/2,1/2]\hat{\alpha}_{j}\in[-1/2,1/2] the orthogonality condition

(23) G,ϕ^¯+Qgα^j(x)ϕ^(x)¯𝑑x= 0\langle G,\overline{\hat{\phi}}\rangle\ +\ \int\limits_{Q_{\infty}}g_{\hat{\alpha}_{j}}(x)\,\overline{\hat{\phi}(x)}\,dx\ =\ 0

hold for all modes ϕ^Xj\hat{\phi}\in X_{j} corresponding to the propagative wave number α^j\hat{\alpha}_{j}. Here, ,\langle\cdot,\cdot\rangle denotes the dual (bi-linear) form.

Then for every α[1/2,1/2]\alpha\in[-1/2,1/2] there exists an α\alpha-quasi-periodic solution vαHα,loc,01(D)v_{\alpha}\in H_{\alpha,loc,0}^{1}(D) of the equation

(24) Δvα+k2vα=gαGin Q\Delta v_{\alpha}+k^{2}v_{\alpha}\ =\ -g_{\alpha}\ -\ G\quad\mbox{in }Q_{\infty}

satisfying the generalized Rayleigh radiation condition

(25) n|vα,n(x2)x2ik2(α+n)2vα,n(x2)|2 0,x2+.\sum_{n\in\mathbb{Z}}\biggl{|}\frac{\partial v_{\alpha,n}(x_{2})}{\partial x_{2}}-i\sqrt{k^{2}-(\alpha+n)^{2}}\,v_{\alpha,n}(x_{2})\biggr{|}^{2}\ \rightarrow\ 0\,,\qquad x_{2}\rightarrow+\infty\,.

Here, vα,n(x2)=12π02πvα(x1,x2)ei(n+α)x1𝑑x1v_{\alpha,n}(x_{2})=\frac{1}{\sqrt{2\pi}}\int_{0}^{2\pi}v_{\alpha}(x_{1},x_{2})e^{-i(n+\alpha)x_{1}}dx_{1} are the Fourier coefficients of vα(,x2)v_{\alpha}(\cdot,x_{2}), and (24) is understood in the variational sense

Q[vαψ¯k2vαψ¯]𝑑x=G,ψ¯+Qgαψ¯𝑑x\int\limits_{Q_{\infty}}\bigl{[}\nabla v_{\alpha}\cdot\nabla\overline{\psi}-k^{2}v_{\alpha}\,\overline{\psi}\bigr{]}\,dx\ =\ \langle G,\overline{\psi}\rangle\ +\ \int\limits_{Q_{\infty}}g_{\alpha}\,\overline{\psi}\,dx

for all ψHα,loc,01(D)\psi\in H^{1}_{\alpha,loc,0}(D) which vanish for x2>hx_{2}>h for some h>h0h>h_{0}.

Furthermore, vαv_{\alpha} can be chosen to depend continuously on α\alpha, and for every h>h0h>h_{0} there exists ch>0c_{h}>0 with

vαH1(Qh)ch[supβ[1/2,1/2]gβL(1,2)(Q)+supβ[1/2,1/2]gβ/βL(1,2)(Q)+GH1(Qh0)]\|v_{\alpha}\|_{H^{1}(Q_{h})}\ \leq\ c_{h}\bigl{[}\sup\limits_{\beta\in[-1/2,1/2]}\|g_{\beta}\|_{L^{(1,2)}(Q_{\infty})}+\sup\limits_{\beta\in[-1/2,1/2]}\|\partial g_{\beta}/\partial\beta\|_{L^{(1,2)}(Q_{\infty})}\ +\ \|G\|_{H^{-1}(Q_{h_{0}})}\bigr{]}

for all α[1/2,1/2]\alpha\in[-1/2,1/2] where we used the notation ϕL(1,2)(Q):=ϕL1(Q)+ϕL2(Q)\|\phi\|_{L^{(1,2)}(Q_{\infty})}:=\|\phi\|_{L^{1}(Q_{\infty})}+\|\phi\|_{L^{2}(Q_{\infty})}.


We will apply this result to the following two problems.

Given φH1/2(CR)\varphi\in H^{-1/2}(C_{R}), consider the problem of determining uHloc1(D)u\in H^{1}_{loc}(D) such that

(26) Δu+k2u=0 in DCR,u=0 on Γ,uνu+ν=φ on CR,\Delta u+k^{2}u=0\mbox{ in }D\setminus C_{R}\,,\quad u=0\mbox{ on }\Gamma\,,\quad\frac{\partial u_{-}}{\partial\nu}-\frac{\partial u_{+}}{\partial\nu}=\varphi\mbox{ on }C_{R}\,,

and that uu satisfies the open waveguide radiation condition of Definition 2.5. Here the normal direction ν\nu is supposed to direct into the exterior ΣR\Sigma_{R}. Well-posedness of the variational formulation corresponding to the problem (26) is stated as follows.

Theorem 3.4.

Let φH1/2(CR)\varphi\in H^{-1/2}(C_{R}). Then there exists a unique solution uHloc,01(D)u\in H^{1}_{loc,0}(D) of

(27) D[uψ¯k2uψ¯]𝑑x=CRφψ¯𝑑sfor all ψH0,c1(D)\int\limits_{D}[\nabla u\cdot\nabla\overline{\psi}-k^{2}u\,\overline{\psi}]\,dx\ =\ \int\limits_{C_{R}}\varphi\,\overline{\psi}\,ds\quad\mbox{for all }\psi\in H^{1}_{0,c}(D)

satisfying the open waveguide radiation condition. Here,

H0,c1(D):={ψH01(D):there exists a>0 with ψ(x)=0 for |x|>a}.H^{1}_{0,c}(D)\ :=\ \bigl{\{}\psi\in H^{1}_{0}(D):\mbox{there exists $a>0$ with $\psi(x)=0$ for }|x|>a\bigr{\}}\,.

Furthermore, the mapping φu|CR\varphi\mapsto u|_{C_{R}} is bounded from H1/2(CR)H^{-1/2}(C_{R}) into H01/2(CR)H^{1/2}_{0}(C_{R}).

Proof: Uniqueness follows directly from Theorem 2.14, part (i). To prove existence, we suppose without loss of generality that CRC_{R} is chosen to lie in Qh0Q_{h_{0}} and define the coefficients a,ja_{\ell,j} explicitly as

(28) a,j:=2πi|λ,j|CRφ(x)ϕ^,j(x)¯𝑑s(x),=1,,mj,jJ.a_{\ell,j}\ :=\ \frac{2\pi i}{|\lambda_{\ell,j}|}\int\limits_{C_{R}}\varphi(x)\,\overline{\hat{\phi}_{\ell,j}(x)}\,ds(x)\,,\ \ell=1,\ldots,m_{j}\,,\ j\in J\,.

Then the propagating part upropu_{prop} is defined, and the radiating part uradu_{rad} has to satisfy

D[uradψ¯k2uradψ¯]𝑑x\displaystyle\int\limits_{D}[\nabla u_{rad}\cdot\nabla\overline{\psi}-k^{2}u_{rad}\,\overline{\psi}]\,dx
=\displaystyle= D[upropψ¯k2upropψ¯]𝑑x+CRφψ¯𝑑s\displaystyle-\int\limits_{D}[\nabla u_{prop}\cdot\nabla\overline{\psi}-k^{2}u_{prop}\,\overline{\psi}]\,dx\ +\ \int\limits_{C_{R}}\varphi\,\overline{\psi}\,ds
=\displaystyle= jJ=1mja,jDφ,jψ¯𝑑x+CRφψ¯𝑑sfor all ψH0,c1(D)\displaystyle\sum_{j\in J}\sum_{\ell=1}^{m_{j}}a_{\ell,j}\int\limits_{D}\varphi_{\ell,j}\,\overline{\psi}\,dx\ +\ \int\limits_{C_{R}}\varphi\,\overline{\psi}\,ds\quad\mbox{for all }\psi\in H^{1}_{0,c}(D)

and the generalized angular spectrum radiation condition (8). Here, φ,j\varphi_{\ell,j} are given by (10b). Defining the distribution fφH1(D)f_{\varphi}\in H^{-1}(D) as

fφ,ψ:=CRφψ¯𝑑sfor allψH01(D),\langle f_{\varphi},\psi\rangle\ :=\ \int\limits_{C_{R}}\varphi\,\overline{\psi}\,ds\quad\mbox{for all}\quad\psi\in H^{1}_{0}(D)\,,

where the right hand side is understood as the duality between H1/2(CR)H^{-1/2}(C_{R}) and H01/2(CR)H^{1/2}_{0}(C_{R}), we observe that the variational equation represents the differential equation

Δurad+k2urad=jJ=1mja,jφ,jfφ.\Delta u_{rad}+k^{2}u_{rad}\ =\ -\sum_{j\in J}\sum_{\ell=1}^{m_{j}}a_{\ell,j}\,\varphi_{\ell,j}\ -\ f_{\varphi}\,.

One now applies Theorem 3.3 to gα=jJ=1mja,jFφ,jg_{\alpha}=\sum_{j\in J}\sum_{\ell=1}^{m_{j}}a_{\ell,j}\,F\varphi_{\ell,j} and G=fφG=f_{\varphi}. The orthogonality condition (23) is satisfied by the choice of the coefficients (28) (see [19, Lemma 5.1]). Therefore, for all α[1/2,1/2]\alpha\in[-1/2,1/2] there exists a solution v^(,α)Hα,loc,01(D)\hat{v}(\cdot,\alpha)\in H^{1}_{\alpha,loc,0}(D) of the α\alpha-quasi-periodic problems

Q[v^(,α)ψ^¯k2v^(,α)ψ^¯]𝑑x\displaystyle\int\limits_{Q_{\infty}}\bigl{[}\nabla\hat{v}(\cdot,\alpha)\cdot\nabla\overline{\hat{\psi}}-k^{2}\hat{v}(\cdot,\alpha)\,\overline{\hat{\psi}}\bigr{]}\,dx
=\displaystyle= jJ=1mja,jQ(Fφ,j)(,α)ψ^¯𝑑x+CRφψ^¯𝑑s\displaystyle\sum_{j\in J}\sum_{\ell=1}^{m_{j}}a_{\ell,j}\int\limits_{Q_{\infty}}(F\varphi_{\ell,j})(\cdot,\alpha)\,\overline{\hat{\psi}}\,dx\ +\ \int\limits_{C_{R}}\varphi\,\overline{\hat{\psi}}\,ds

for all ψ^Hα1(D)\hat{\psi}\in H^{1}_{\alpha}(D) which vanish for x2>hx_{2}>h for some h>h0h>h_{0} satisfying the generalized Rayleigh radiation condition (25). Furthermore, v^(,α)\hat{v}(\cdot,\alpha) depends continuously on α\alpha and for every h>h0h>h_{0} there exists ch>0c_{h}>0 with

(31) v^(,α)H1(Qh)ch[jJ=1mj|a,j|+supψH1(Q)=1|CRφψ¯𝑑s|]chφH1/2(CR).\|\hat{v}(\cdot,\alpha)\|_{H^{1}(Q_{h})}\ \leq\ c_{h}\,\biggl{[}\sum_{j\in J}\sum_{\ell=1}^{m_{j}}|a_{\ell,j}|\ +\sup\limits_{\|\psi\|_{H^{1}(Q_{\infty})=1}}\biggl{|}\int\limits_{C_{R}}\varphi\,\overline{\psi}\,ds\biggr{|}\biggr{]}\ \leq\ c_{h}\,\|\varphi\|_{H^{-1/2}(C_{R})}\,.

By the properties of the Floquet-Bloch transform the inverse transform

urad(x):=(F1v^)(x)=1/21/2v^(x,α)𝑑αu_{rad}(x)\ :=\ (F^{-1}\hat{v})(x)\ =\ \int\limits_{-1/2}^{1/2}\hat{v}(x,\alpha)\,d\alpha

is in H1(D)H^{1}_{\ast}(D). Furthermore, taking ψC0(D)\psi\in C^{\infty}_{0}(D) we substitute ψ^:=(Fψ)(,α)\hat{\psi}:=(F\psi)(\cdot,\alpha) into the variational equation (3.1) and integrate with respect to α\alpha; that is,

1/21/2Q[v^(x,α)(Fψ)(x,α)¯k2v^(x,α)(Fψ)(x,α)¯]𝑑x𝑑α\displaystyle\int\limits_{-1/2}^{1/2}\int\limits_{Q_{\infty}}\bigl{[}\nabla\hat{v}(x,\alpha)\cdot\nabla\overline{(F\psi)(x,\alpha)}-k^{2}\hat{v}(x,\alpha)\,\overline{(F\psi)(x,\alpha)}\bigr{]}\,dx\,d\alpha
=\displaystyle= jJ=1mja,j1/21/2Q(Fφ,j)(x,α)(Fψ)(x,α)¯𝑑x𝑑α+1/21/2CRφ(x)(Fψ)(x,α)¯𝑑s(x)𝑑α\displaystyle\sum_{j\in J}\sum_{\ell=1}^{m_{j}}a_{\ell,j}\int\limits_{-1/2}^{1/2}\int\limits_{Q_{\infty}}(F\varphi_{\ell,j})(x,\alpha)\,\overline{(F\psi)(x,\alpha)}\,dx\,d\alpha+\int\limits_{-1/2}^{1/2}\int\limits_{C_{R}}\varphi(x)\,\overline{(F\psi)(x,\alpha)}\,ds(x)\,d\alpha

Noting that 1/21/2(Fψ)(,α)𝑑s=ψ\int_{-1/2}^{1/2}(F\psi)(\cdot,\alpha)\,ds=\psi and using the unitarity of the Floquet-Bloch transform we observe that this is exactly the equation (3.1). Boundedness of φu|CR\varphi\rightarrow u|_{C_{R}} is now easily seen from (31) and the unitarity of FF and the fact that upropu_{prop} depends explicitly on φ\varphi through a,ja_{\ell,j}. ∎


A second application is the following result where the source fails to be compactly supported.

Theorem 3.5.

Let fL(ρ)2(D)f\in L^{2}_{(\rho)}({D}) for some ρ(0,1)\rho\in(0,1). Then there exists a unique solution wHloc,01(D)w\in H^{1}_{loc,0}(D) of Δw+k2w=f\Delta w+k^{2}w=-f in DD satisfying the open waveguide radiation condition. Furthermore, for every a>Ra>R the mappings fw|Daf\mapsto w|_{D_{a}} and fw|CRf\mapsto w|_{C_{R}} are bounded from L(ρ)2(D)L^{2}_{(\rho)}(D) into H1(Da)H^{1}(D_{a}) and H01/2(CR)H^{1/2}_{0}(C_{R}), respectively.

Proof: Since ff decays exponentially, its Floquet-Bloch transform FfFf is well defined and continuously differentiable with respect to α\alpha. Instead of (3.1) we now solve

Q[w^(,α)ψ^¯k2w^(,αψ^¯]dx\displaystyle\int\limits_{Q_{\infty}}\bigl{[}\nabla\hat{w}(\cdot,\alpha)\cdot\nabla\overline{\hat{\psi}}-k^{2}\hat{w}(\cdot,\alpha\,\overline{\hat{\psi}}\bigr{]}\,dx
=\displaystyle= jJ=1mja,jQ(Fφ,j)(,α)ψ^¯𝑑x+Q(Ff)(,α)ψ^¯𝑑x\displaystyle\sum_{j\in J}\sum_{\ell=1}^{m_{j}}a_{\ell,j}\int\limits_{Q_{\infty}}(F\varphi_{\ell,j})(\cdot,\alpha)\,\overline{\hat{\psi}}\,dx\ +\ \int\limits_{Q_{\infty}}(Ff)(\cdot,\alpha)\,\overline{\hat{\psi}}\,dx

for all ψ^Hα1(D)\hat{\psi}\in H^{1}_{\alpha}(D) which vanish for x2>hx_{2}>h for some h>h0h>h_{0} and the generalized Rayleigh radiation condition (25).

The coefficients a,ja_{\ell,j} have to be chosen as

a,j=2πi|λ,j|Q(Ff)(x,α^j)ϕ^,j(x)¯𝑑x,=1,,mj,jJ,a_{\ell,j}\ =\ \frac{2\pi i}{|\lambda_{\ell,j}|}\int\limits_{Q_{\infty}}(Ff)(x,\hat{\alpha}_{j})\,\overline{\hat{\phi}_{\ell,j}(x)}\,dx\,,\quad\ell=1,\cdots,m_{j}\,,\ j\in J\,,

so that the right hand side is always orthogonal to the nullspace of the homogeneous equation. Using the estimate

(Ff)(,β)L2(Q)2\displaystyle\|(Ff)(\cdot,\beta)\|^{2}_{L^{2}(Q_{\infty})} =\displaystyle= Q|(Ff)(x,β|2dxQ[|f(x1+2π,x2)|]2dx\displaystyle\int\limits_{Q_{\infty}}\bigl{|}(Ff)(x,\beta\bigr{|}^{2}dx\ \leq\ \int\limits_{Q_{\infty}}\biggl{[}\sum_{\ell\in\mathbb{Z}}|f(x_{1}+2\pi\ell,x_{2})|\biggr{]}^{2}dx
=\displaystyle= Q[(1+2)1/2|(1+2)1/2f(x1+2π,x2)|]2𝑑x\displaystyle\int\limits_{Q_{\infty}}\biggl{[}\sum_{\ell\in\mathbb{Z}}(1+\ell^{2})^{-1/2}\bigl{|}(1+\ell^{2})^{1/2}f(x_{1}+2\pi\ell,x_{2})\bigr{|}\biggr{]}^{2}dx
\displaystyle\leq Q11+2(1+2)|f(x1+2π,x2)|2dx\displaystyle\int\limits_{Q_{\infty}}\sum_{\ell\in\mathbb{Z}}\frac{1}{1+\ell^{2}}\sum_{\ell\in\mathbb{Z}}(1+\ell^{2})|f(x_{1}+2\pi\ell,x_{2})|^{2}\,dx
\displaystyle\leq cD(1+x12)|f(x)|2𝑑xcfL(ρ)2(D)2\displaystyle c\int\limits_{D}(1+x_{1}^{2})\,|f(x)|^{2}\,dx\ \leq\ c\,\|f\|^{2}_{L^{2}_{(\rho)}(D)}

and analogously for (Ff)(,β)L1(Q)2\|(Ff)(\cdot,\beta)\|^{2}_{L^{1}(Q_{\infty})} and the derivatives with respect to β\beta we can repeat the proof of Theorem 3.4. ∎

3.2. The DtN Operator

Now we turn to the construction of the Dirichlet-to-Neumann operator on the artificial boundary CRC_{R}. In the remaining part of this paper we make the following assumption.

Assumption 3.6.

Assume that k2k^{2} is not the Dirichlet eigenvalue of Δ-\Delta in the Lipschitz domain DRD_{R} and there are no bound states of the Helmholtz equation over the domain ΣR\Sigma_{R}; that is, if uH01(ΣR)u\in H^{1}_{0}(\Sigma_{R}) solves Δu+k2u=0\Delta u+k^{2}u=0 in ΣR\Sigma_{R}, then uu must vanish identically.

As usual, the DtN operator Λ\Lambda should be defined as follows.

Definition 3.7.

The Dirichlet-to-Neumann operator Λ:H01/2(CR)H1/2(CR)\Lambda:H_{0}^{1/2}(C_{R})\rightarrow H^{-1/2}(C_{R}) is defined by Λg=νu|CR\Lambda\,g=\partial_{\nu}u|_{C_{R}} where uHloc1(ΣR)u\in H_{loc}^{1}(\Sigma_{R}) is the unique solution to

(32) Δu+k2u=0 in ΣRu=g on CR,u=0 on ΓΣR,\Delta u+k^{2}u=0\mbox{ in }\Sigma_{R}\,\quad u=g\mbox{ on }C_{R}\,,\quad u=0\mbox{ on }\Gamma\cap\partial\Sigma_{R}\,,

which fulfills the open waveguide radiation condition of Definition 2.5. Here the unit normal vector ν\nu at CRC_{R} is supposed to direct into ΣR\Sigma_{R}.

The above definition assumes already the solvability of a boundary value problem in the perturbed region ΣR\Sigma_{R} – which to show is the purpose of the forthcoming Section 4. However, the perturbed region ΣR\Sigma_{R} is a subset of DD (in contrast to the more general perturbation D~\tilde{D}) which allows the application of an integral equation approach with the Dirichlet-Green’s function of DD. Before we explain the construction we note that an explicit representation in form of a series can be obtained if Γ\Gamma is a straight line parallel to the x1x_{1}-axis. In this exceptional case the propagating part (guided waves) vanishes identically and the radiating part fulfills the classical Sommerfeld radiation condition. Consequently, the function gH01/2(CR)g\in H_{0}^{1/2}(C_{R}) can be expanded into g(θ)=n0gnsinnθg(\theta)=\sum_{n\in{\mathbb{N}}_{0}}g_{n}\sin n\theta with θ(0,π)\theta\in(0,\pi) and the DtN operator takes the explicit form

(Λg)(θ)=n0gnkHn(1)(kR)Hn(1)(kR)sinnθfor θ(0,π).(\Lambda g)(\theta)\ =\ \sum_{n\in{\mathbb{N}}_{0}}g_{n}\frac{k\,H_{n}^{(1)\prime}(kR)}{H_{n}^{(1)}(kR)}\,\sin n\theta\quad\mbox{for }\theta\in(0,\pi)\,.

Here, H0(1)H_{0}^{(1)} denotes the Hankel function of the first kind of order zero. In the general case that Γ\Gamma is a periodic curve, we will express the field in ΣR\Sigma_{R} as a single layer potential with density φ\varphi and the Green’s function as kernel. As usual, φ\varphi is determined from gg by solving an integral equation for the single layer boundary operator. We divide our arguments into two steps.

(A) Construction of the single layer boundary operator. As a motivation we recall that for smooth data the single layer boundary operator with the Green’s function as kernel is given by Sφ=u|CRS\varphi=u|_{C_{R}} where uHloc1(D)u\in H^{1}_{loc}(D) satisfies the transmission problem (26) and the open waveguide radiation condition. In this way we avoid the explicit use of the Green’s function. For given φH1/2(CR)\varphi\in H^{-1/2}(C_{R}), the variational form of (26) is given by (27) and has been studied in Theorem 3.4.

We take the solution of this transmission problem as the definition of the single layer operator, namely Sφ:=u|CRS\varphi:=u|_{C_{R}} where uHloc,01(D)u\in H^{1}_{loc,0}(D) is the unique solution of (27) satisfying the open waveguide radiation condition. Then SS is bounded from H1/2(CR)H^{-1/2}(C_{R}) into H01/2(CR)H^{1/2}_{0}(C_{R}) by Theorem 3.4. To show the injectivity of SS, we suppose that Sφ=0S\varphi=0. Then u=0u=0 in ΣR\Sigma_{R} and u=0u=0 in DRD_{R} by the Assumption 3.6 and the uniqueness result of Theorem 2.16. From the variational equation (27) we conclude that CRφψ¯𝑑s=0\int_{C_{R}}\varphi\,\overline{\psi}\,ds=0 for all ψ\psi which implies that φ\varphi vanishes. This proves the injectivity of SS. Next we show that SS is boundedly invertible.

Let SiS_{i} be the operator corresponding to wave number k=ik=i. Then, setting Siφ=v|CRS_{i}\varphi=v|_{C_{R}}, we get φ,Siφ=CRφv¯𝑑s\langle\varphi,S_{i}\varphi\rangle=\int_{C_{R}}\varphi\,\overline{v}\,ds by the definition of the dual form ,\langle\cdot,\cdot\rangle (see Theorem 3.2) and vH01(D)v\in H^{1}_{0}(D) solves (22). Setting ψ=vϕa\psi=v\,\phi_{a} in (22) where ϕaC0(D)\phi_{a}\in C^{\infty}_{0}(D) satisfies ϕa=1\phi_{a}=1 for |x|a|x|\leq a and letting aa tend to infinity shows that

φ,Siφ=CRφv¯𝑑s=D[|v|2+|v|2]𝑑x=vH1(D)2.\langle\varphi,S_{i}\varphi\rangle\ =\ \int\limits_{C_{R}}\varphi\,\overline{v}\,ds\ =\ \int\limits_{D}[|\nabla v|^{2}+|v|^{2}]\,dx\ =\ \|v\|^{2}_{H^{1}(D)}\,.

Next we note that φH1/2(CR)=sup{|φ,ψ|:ψH01/2(CR)1}\|\varphi\|_{H^{-1/2}(C_{R})}=\sup\bigl{\{}|\langle\varphi,\psi\rangle|:\|\psi\|_{H^{1/2}_{0}(C_{R})}\leq 1\bigr{\}}. For ψH01/2(CR)\psi\in H^{1/2}_{0}(C_{R}) with ψH01/2(CR)1\|\psi\|_{H^{1/2}_{0}(C_{R})}\leq 1 we set ψ~=Eψ\tilde{\psi}=E\psi with the extension operator EE from H01/2(CR)H^{1/2}_{0}(C_{R}) into H01(Da)H^{1}_{0}(D_{a}) for some a>Ra>R and estimate

|φ,ψ|=|CRφψ¯𝑑s|=|D[vψ~¯+vψ~¯]𝑑x|cψ~H1(D)vH1(D)cEvH1(D)|\langle\varphi,\psi\rangle|\ =\biggl{|}\int\limits_{C_{R}}\varphi\,\overline{\psi}\,ds\biggr{|}=\biggl{|}\int\limits_{D}[\nabla v\cdot\nabla\overline{\tilde{\psi}}+v\,\overline{\tilde{\psi}}]\,dx\biggr{|}\leq\ c\,\|\tilde{\psi}\|_{H^{1}(D)}\,\|v\|_{H^{1}(D)}\ \leq\ c\,\|E\|\,\|v\|_{H^{1}(D)}

for ψH01/2(CR)1\|\psi\|_{H^{1/2}_{0}(C_{R})}\leq 1 and thus φH1/2(CR)cEvH1(D)\|\varphi\|_{H^{-1/2}(C_{R})}\leq c\,\|E\|\,\|v\|_{H^{1}(D)}. Combining this with the previous estimate yields coercivity of SiS_{i}; that is,

φ,Siφ1c2E2φH1/2(CR)2.\langle\varphi,S_{i}\varphi\rangle\ \geq\ \frac{1}{c^{2}\|E\|^{2}}\,\|\varphi\|_{H^{-1/2}(C_{R})}^{2}\,.

Now we show that SSiS-S_{i} is compact. We observe that (SSi)φ=w|CR(S-S_{i})\varphi=w|_{C_{R}} where w=uvHloc1(D)w=u-v\in H^{1}_{loc}(D) satisfies

Δw+k2w=(k2+1)v in D,w=0 on Γ,\Delta w+k^{2}w\ =\ -(k^{2}+1)\,v\mbox{ in }D\,,\quad w=0\mbox{ on }\Gamma\,,

and the open waveguide radiation conditions. Here, vv corresponds to the solution of (22) with k=ik=i as before. By Theorem 3.2 we know that φv\varphi\mapsto v is compact from H1/2(CR)H^{-1/2}(C_{R}) into L(ρ)2(D)L^{2}_{(\rho^{\prime})}(D) for all ρ<ρ\rho^{\prime}<\rho. Furthermore, by Theorem 3.5 (for ρ\rho^{\prime} replacing ρ\rho) the mapping (1+k2)vw|CR(1+k^{2})v\mapsto w|_{C_{R}} is bounded from L(ρ)2(D)L^{2}_{(\rho^{\prime})}(D) into H01/2(CR)H^{1/2}_{0}(C_{R}). Combining this yields compactness of φw|CR\varphi\mapsto w|_{C_{R}}; that is, compactness of SSiS-S_{i} from H1/2(CR)H^{-1/2}(C_{R}) into H01/2(CR)H^{1/2}_{0}(C_{R}).

Therefore, the operator equation Sφ=gS\varphi=g can be written as Siφ+(SSi)φ=gS_{i}\varphi+(S-S_{i})\varphi=g. This shows that SS is a Fredholm operator with index zero. By the Fredholm alternative, the injectivity implies the invertibility of SS.

(B) Construction of the Dirichlet-to-Neumann operator. Given gH01/2(CR)g\in H^{1/2}_{0}(C_{R}) we define φ:=S1gH1/2(CR)\varphi:=S^{-1}g\in H^{-1/2}(C_{R}). Then, by definition, g=Sφ=u|CRg=S\varphi=u|_{C_{R}} where uu satisfies (27); in particular Δu+k2u=0\Delta u+k^{2}u=0 in ΣR\Sigma_{R} and u=0u=0 on ΓΣR\Gamma\cap\partial\Sigma_{R}, complemented by the open waveguide radiation condition. Consequently, the Neumann boundary data can be defined by Green’s first formula; that is, the DtN operator Λ\Lambda from H01/2(CR)H^{1/2}_{0}(C_{R}) into H1/2(CR)=(H01/2(CR))H^{-1/2}(C_{R})=\bigl{(}H^{1/2}_{0}(C_{R})\bigr{)}^{\ast} can be defined as follows.

Definition 3.8.

Let a>Ra>R be fixed. Then Λ:H01/2(CR)H1/2(CR)=(H01/2(CR))\Lambda:H^{1/2}_{0}(C_{R})\to H^{-1/2}(C_{R})=\bigl{(}H^{1/2}_{0}(C_{R})\bigr{)}^{\ast} is defined as

(33) Λg,ψ=DaDR[u(Eψ)¯k2u(Eψ)¯]𝑑x,ψH01/2(CR),\langle\Lambda g,\psi\rangle\ =\ -\int\limits_{D_{a}\setminus D_{R}}\bigl{[}\nabla u\cdot\nabla\overline{(E\psi)}-k^{2}u\,\overline{(E\psi)}\bigr{]}\,dx\,,\quad\psi\in H^{1/2}_{0}(C_{R})\,,

where E:H01/2(CR)H01(Da)E:H^{1/2}_{0}(C_{R})\to H^{1}_{0}(D_{a}) is again a fixed extension operator and uHloc1(D)u\in H^{1}_{loc}(D) is the single layer potential with density φ:=S1gH1/2(CR)\varphi:=S^{-1}g\in H^{-1/2}(C_{R}); that is, the unique solution of (27) studied in Theorem 3.4.

We note that the definition is independent of a>Ra>R or the choice of the extension operator EE. This follows from Green’s identity DaDR[u(ψ1¯ψ2¯)k2u(ψ1¯ψ2¯)]𝑑x=0\int_{D_{a}\setminus D_{R}}\bigl{[}\nabla u\cdot\nabla(\overline{\psi_{1}}-\overline{\psi_{2}})-k^{2}u\,(\overline{\psi_{1}}-\overline{\psi_{2}})\bigr{]}\,dx=0 for all ψjH01(Da)\psi_{j}\in H^{1}_{0}(D_{a}) with ψ1=ψ2\psi_{1}=\psi_{2} on CRC_{R}.

We finish this section by proving some mapping properties of Λ\Lambda.

Lemma 3.9.

The DtN operator Λ:H01/2(CR)H1/2(CR)\Lambda:H^{1/2}_{0}(C_{R})\rightarrow H^{-1/2}(C_{R}) is bounded. Moreover, the operator Λ-\Lambda can be decomposed into the sum of a coercive operator and a compact operator.

Proof. By the trace lemma and (33), the boundedness of Λ\Lambda follows from the estimate

ΛgH1/2(CR)=supψH01/2(CR)=1|Λg,ψ|cuH1(DaDR)cgH1/2(CR),\|\Lambda g\|_{H^{-1/2}(C_{R})}\ =\ \sup\limits_{\|\psi\|_{H^{1/2}_{0}(C_{R})}=1}|\langle\Lambda g,\psi\rangle|\ \leq\ c\,\|u\|_{H^{1}(D_{a}\setminus D_{R})}\ \leq\ c\,\|g\|_{H^{1/2}(C_{R})}\,,

where we have used the boundedness of the extension operator EE and the continuous dependence of uu from gg. Define Λi:H01/2(CR)H1/2(CR)\Lambda_{i}:H^{1/2}_{0}(C_{R})\rightarrow H^{-1/2}(C_{R}) as the DtN operator for the wave number k=ik=i; that is,

Λig,ψ=DaDR[vψ~¯+vψ~¯]𝑑x,ψH01/2(CR),\langle\Lambda_{i}g,\psi\rangle\ =\ -\int\limits_{D_{a}\setminus D_{R}}[\nabla v\cdot\nabla\overline{\tilde{\psi}}+v\,\overline{\tilde{\psi}}]\,dx\,,\quad\psi\in H^{1/2}_{0}(C_{R})\,,

where vH01(D)v\in H^{1}_{0}(D) solves (22) for φ:=Si1gH1/2(CR)\varphi:=S_{i}^{-1}g\in H^{-1/2}(C_{R}), and ψ~H01(Da)\tilde{\psi}\in H^{1}_{0}(D_{a}) is an extension of ψ\psi. The operator Λi-\Lambda_{i} is coercive over H01/2(CR)H_{0}^{1/2}(C_{R}). Indeed, choose ϕaC(2)\phi_{a}\in C^{\infty}(\mathbb{R}^{2}) with ϕa=1\phi_{a}=1 for |x|<R|x|<R and ϕa=0\phi_{a}=0 for |x|>a1|x|>a-1 and set ψ~=vϕa\tilde{\psi}=v\,\phi_{a} for a>R+1a>R+1. Then ψ~H01(Da)\tilde{\psi}\in H^{1}_{0}(D_{a}) and thus

Λig,g=vH1(Da1DR)2+DaDa1[v(vϕa)¯+ϕa|v|2]𝑑x.-\langle\Lambda_{i}g,g\rangle\ =\ \|v\|_{H^{1}(D_{a-1}\setminus D_{R})}^{2}\ +\ \int\limits_{D_{a}\setminus D_{a-1}}[\nabla v\cdot\nabla\overline{(v\phi_{a})}+\phi_{a}|v|^{2}]\,dx\,.

Now we let aa tend to infinity and use that vH1(D)v\in H^{1}(D). Therefore,

Λig,g=vH1(DDR)2cvH1/2(CR)2=cgH1/2(CR)2-\langle\Lambda_{i}g,g\rangle\ =\ \|v\|_{H^{1}(D\setminus D_{R})}^{2}\ \geq\ c\,\|v\|_{H^{1/2}(C_{R})}^{2}\ =\ c\,\|g\|_{H^{1/2}(C_{R})}^{2}

where we used the boundedness of the trace operator in the inequality. Furthermore, the operator ΛΛi\Lambda-\Lambda_{i} is compact. Indeed, this follows from

(ΛΛi)g,ψ=DaDR[(uv)(Eψ)¯(k2u+v)Eψ¯]𝑑x,ψH01/2(CR),\langle(\Lambda-\Lambda_{i})\,g,\psi\rangle\ =\ -\int\limits_{D_{a}\setminus D_{R}}\bigl{[}\nabla(u-v)\cdot\nabla\overline{(E\psi)}-(k^{2}u+v)\,\overline{E\psi}\bigr{]}\,dx\,,\quad\psi\in H^{1/2}_{0}(C_{R})\,,

and the compactness of the mapping g(uv)|DaDRg\mapsto(u-v)|_{D_{a}\setminus D_{R}} from H01/2(CR)H^{1/2}_{0}(C_{R}) into H1(DaDR)H^{1}(D_{a}\setminus D_{R}) (by the same arguments as in the proof of the compactness of SSiS-S_{i}) and the boundedness of gk2u+vg\mapsto k^{2}u+v from H01/2(CR)H^{1/2}_{0}(C_{R}) into H1(DaDR)H^{1}(D_{a}\setminus D_{R}) and the compact embedding of H1(DaDR)H^{1}(D_{a}\setminus D_{R}) into L2(DaDR)L^{2}(D_{a}\setminus D_{R}). ∎

4. Existence of Solutions of the Perturbed Problem

In this section we investigate well-posedness of time-harmonic scattering of an incoming wave uinu^{in} from a locally perturbed periodic curve Γ~=D~\tilde{\Gamma}=\partial\tilde{D} of Dirichlet kind; see Figure 2. We consider three kinds of incoming waves:

  • (i)

    Point source wave: uin(x):=Φ(x,y)=i4H0(1)(k|xy|)u^{in}(x):=\Phi(x,y)=\frac{i}{4}H_{0}^{(1)}(k|x-y|) with the source position yD~y\in\tilde{D}. Without loss of generality we suppose that yD~Ry\in\tilde{D}_{R}.

  • (ii)

    Plane wave: uin(x)=eikxθ^u^{in}(x)=e^{ikx\cdot\hat{\theta}} where θ^=(sinθ,cosθ)\hat{\theta}=(\sin\theta,-\cos\theta) is the incident direction with some incident angle θ(π/2,π/2)\theta\in(-\pi/2,\pi/2). In this case the incoming wave is incident onto Γ~\tilde{\Gamma} from above, and the parameter α:=ksinθ\alpha:=k\sin\theta is supposed to be not a propagative wavenumber (see Definition 2.1 (ii)).

  • (iii)

    uin(x)=ϕ^,j(x)u^{in}(x)=\hat{\phi}_{\ell,j}(x) is a right (resp. left) going surface wave at the propagative wavenumber α^j\hat{\alpha}_{j} for some jJj\in J which corresponds to the spectral problem (6) with the eigenvalue λ,j>0\lambda_{\ell,j}>0 (resp. λ,j<0\lambda_{\ell,j}<0).

Refer to caption
Figure 2. Illustration of wave scattering from perfectly reflecting periodic curves with a local perturbation.

We denote by uunpertscu^{sc}_{unpert} the unperturbed scattered field, defined in DD, which is caused by the unperturbed curve Γ\Gamma. In ΣR\Sigma_{R} the total field uu can be decomposed into u=uin+uunpertsc+upertscu=u^{in}+u^{sc}_{unpert}+u^{sc}_{pert}, and upertscu^{sc}_{pert} can be considered as the scattered part induced by the defect. The field upertscu^{sc}_{pert} is supposed to fulfill the open waveguide radiation condition of Definition 2.5 for all of the cases (i), (ii), (iii).

Define the spaces

YR\displaystyle Y_{R} :=\displaystyle:= {vH1(D~R):v=0 on Γ~D~R},\displaystyle\bigl{\{}v\in H^{1}(\tilde{D}_{R}):v=0\mbox{ on }\tilde{\Gamma}\cap\partial\tilde{D}_{R}\bigr{\}}\,,

where yD~y\in\tilde{D}. Well-posedness of our scattering problems will be stated separately for different incoming waves.

Theorem 4.1 (Well-posedness for point source waves).

Let uin:=Φ(,y)u^{in}:=\Phi(\cdot,y) be an incoming point source wave with yD~Ry\in\tilde{D}_{R}. Then the locally perturbed scattering problem admits a unique solution uu such that uuinHloc1(D~)u-u^{in}\in H^{1}_{loc}(\tilde{D}) and uu satisfies the open waveguide radiation conditions of Definitions 2.5 and 2.10.

In this theorem, the total field uu is required to satisfy the open waveguide radiation condition of Definition 2.5, because uu is nothing else but the Green’s function of the perturbed problem. We remark that in ΣR\Sigma_{R} the scattered field uuinu-u^{in} does not fulfill this radiation condition, since uin=Φ(,y)u^{in}=\Phi(\cdot,y) does not belong to H1(WhΣR)H^{1}(W_{h}\cap\Sigma_{R}) for any h>h0h>h_{0}.

If both Γ\Gamma and Γ~\tilde{{\Gamma}} can be represented as graphs of Lipschitz functions, the propagating part upropu_{prop} vanishes identically (see [4, 5]) and thus u=uradu=u_{rad}. In such a case, it was verified in [11, Theorem 2.2] that u=u(,y)Hρ1(WhΣR)u=u(\cdot,y)\in H^{1}_{\rho}(W_{h}\cap\Sigma_{R}) with R>|y1π|R>|y_{1}-\pi| for all ρ<1\rho<1 and that u(,y)Φ(,y)Hρ1(Wh)u(\cdot,y)-\Phi(\cdot,y)\in H^{1}_{\rho}(W_{h}) for all ρ<0\rho<0. In addition, both uu and uΦu-\Phi satisfy the Sommerfeld radiation condition of Definition 2.8. The above results of Theorem 4.1 have generalized those of [11] to non-graph curves where guided (propagating) waves may occur. On the other hand, the technical assumption made in [11, Section 2.3] that Γ\Gamma should contain at least one line segment in each period was removed in this paper by constructing a new form of the DtN operator; see subsection 3.2.

Proof of Theorem 4.1. Since the incident field is singular at yy, we transform our scattering problem to an equivalent source problem of the form (1). Introduce a smooth cut-off function χ:2\chi:\mathbb{R}^{2}\rightarrow\mathbb{R} such that χ(x)=1\chi(x)=1 for |xy|<ϵ/2|x-y|<\epsilon/2 and χ(x)=0\chi(x)=0 for |xy|ϵ|x-y|\geq\epsilon. Here ϵ>0\epsilon>0 is chosen to be less than the distance between yy and D~R\partial\tilde{D}_{R}. We make the ansatz on the total field uu as

u(x)=χ(x)Φ(x,y)+v(x,y),xD~,xy.\displaystyle u(x)\ =\ \chi(x)\,\Phi(x,y)\ +\ v(x,y)\,,\quad x\in\tilde{D},\quad x\neq y\,.

Then the scattering problem is equivalent of finding v(,y)H0,loc1(D~)v(\cdot,y)\in H_{0,loc}^{1}(\tilde{D}) such that

{Δxv(,y)+k2v(,y)=gy in D~,v(,y)=0 on Γ~,v(,y) satisfies the open waveguide radiation condition of Definition 2.5\displaystyle\left\{\begin{array}[]{l}\displaystyle\Delta_{x}v(\cdot,y)+k^{2}v(\cdot,y)\ =\ -g_{y}\mbox{ in }\tilde{D},\quad v(\cdot,y)=0\mbox{ on }\tilde{\Gamma},\\ \displaystyle\mbox{$v(\cdot,y)$ satisfies the open waveguide radiation condition of Definition~{}\ref{d-RC}}\end{array}\right.

with

gy:=ΔχΦ(,y)+2χxΦ(,y)L2(D~).\displaystyle g_{y}\ :=\ \Delta\chi\,\Phi(\cdot,y)+2\nabla\chi\cdot\nabla_{x}\Phi(\cdot,y)\in L^{2}(\tilde{D})\,.

Note that the source term gyg_{y} is compactly supported in D~R\tilde{D}_{R}. By the DtN operator Λ\Lambda, this problem can be reduced to an equivalent boundary value problem over the truncated domain D~R\tilde{D}_{R}. Consequently, we get the following variational formulation. Determine vYRv\in Y_{R} such that

(35) D~Rvψ¯k2vψ¯dxCRΛvψ¯𝑑s=D~Rgyψ¯𝑑sfor allψYR.\displaystyle\int\limits_{\tilde{D}_{R}}\nabla v\cdot\overline{\nabla\psi}-k^{2}v\,\overline{\psi}\,dx\ -\ \int\limits_{C_{R}}\Lambda v\,\overline{\psi}ds\ =\ \int\limits_{\tilde{D}_{R}}g_{y}\,\overline{\psi}\,ds\quad\mbox{for all}\quad\psi\in Y_{R}\,.

Here the integral over CRC_{R} is understood as the duality between H1/2(CR)H^{-1/2}(C_{R}) and H01/2(CR)H^{1/2}_{0}(C_{R}). In view of Lemma 3.9, the sesqui-linear form defined by the left hand side of (35) is strongly elliptic, leading to a Fredholm operator with index zero over YRY_{R}. By Theorem 2.16 we have uniqueness and thus also existence of v(,y)YRv(\cdot,y)\in Y_{R} by the Fredholm alternative. This solution can be extended to the exterior ΣR\Sigma_{R} by solving the problem of Theorem 3.2 with φ:=S1(v|CR)H1/2(CR)\varphi:=S^{-1}(v|_{C_{R}})\in H^{-1/2}(C_{R}).

Finally we note that uuin=v+(χ1)Φ(,y)Hloc1(D~)u-u^{in}=v+(\chi-1)\Phi(\cdot,y)\in H^{1}_{loc}(\tilde{D}) because χ\chi vanishes in a neighborhood of yy, and u=vu=v in ΣR\Sigma_{R} because χ\chi vanishes in ΣR\Sigma_{R}. This ends the proof. ∎


Remark 4.2.

If we decompose the field uu into u=uin+uunpertsc+upertscu=u^{in}+u^{sc}_{unpert}+u^{sc}_{pert} in ΣR\Sigma_{R} then we observe that also upertscu^{sc}_{pert} satisfies the open radiation condition because uu and uin+uunpertscu^{in}+u^{sc}_{unpert} do, the latter because it is the total field corresponding to the unperturbed problem.


We proceed with the scattering problem for plane waves.

Theorem 4.3 (Well-posedness for plane waves).

Let α:=ksinθ\alpha:=k\sin\theta be not a propagative wavenumber (see Definition 2.1 (ii)). Then the perturbed scattering problem for a plane wave incidence uin(x)=eikxθ^u^{in}(x)=e^{ikx\cdot\hat{\theta}} admits a unique solution u=uin+uscHloc,01(D~)u=u^{in}+u^{sc}\in H_{loc,0}^{1}(\tilde{D}) such that the scattered part uscu^{sc} has a decomposition in the form usc=uunpertsc+upertscu^{sc}=u^{sc}_{unpert}+u^{sc}_{pert} in the region ΣR\Sigma_{R} where uunpertscHα,loc1(D)u^{sc}_{unpert}\in H^{1}_{\alpha,loc}(D) is the scattered field corresponding to the unperturbed problem that satisfies the upward Rayleigh expansion (3b) with the quasi-periodic parameter α=ksinθ\alpha=k\sin\theta. The part upertscHloc1(ΣR)u^{sc}_{pert}\in H_{loc}^{1}(\Sigma_{R}) fulfills the open waveguide radiation conditions defined by Def. 2.5 and Def. 2.10.

Proof. In the unperturbed case, uniqueness and existence of the field uunpertscHα,loc1(D)u^{sc}_{unpert}\in H_{\alpha,loc}^{1}(D) can be justified using standard variational arguments in the truncated periodic cell QhQ_{h} (for some h>h0h>h_{0}) by enforcing the α\alpha-quasi-periodic DtN mapping on the artificial boundary Γh{\Gamma}_{h}. Uniqueness follows from the assumption that α=ksinθ\alpha=k\sin\theta is not a propagative wavenumber, and existence is a consequence of the Fredholm alternative.

Set u~in=uin+uunpertscHα,01(D)\tilde{u}^{in}=u^{in}+u^{sc}_{unpert}\in H_{\alpha,0}^{1}(D). This field is well defined in ΣR\Sigma_{R}. We make the ansatz for the perturbed problem in the form u=u~in+upertscu=\tilde{u}^{in}+u^{sc}_{pert} in ΣR\Sigma_{R} and u=uin+uscu=u^{in}+u^{sc} in D~R\tilde{D}_{R}. Since upertscHloc1(ΣR)u^{sc}_{pert}\in H^{1}_{loc}(\Sigma_{R}) is required to fulfill the open waveguide radiation condition and upertsc=0u^{sc}_{pert}=0 on ΓD~R{\Gamma}\setminus\tilde{D}_{R}, it has to satisfy

νupertsc|+=Λ(upertsc|+),thusνu|=νu~in|++Λ(u|u~in|+)on CR\displaystyle\partial_{\nu}u^{sc}_{pert}\bigr{|}_{+}\ =\ \Lambda(u^{sc}_{pert}|_{+})\,,\quad\mbox{thus}\quad\partial_{\nu}u\bigr{|}_{-}\ =\ \partial_{\nu}\tilde{u}^{in}\bigr{|}_{+}\ +\ \Lambda(u|_{-}-\tilde{u}^{in}|_{+})\quad\mbox{on }C_{R}

where |+|_{+} and ||_{-} denote the traces from ΣR\Sigma_{R} and D~R\tilde{D}_{R}, respectively. Therefore, we have to determine the total field uYRu\in Y_{R} such that

(36) D~Ruψ¯k2uψ¯dxCRΛuψ¯𝑑s=CR[νu~inΛu~in]ψ¯𝑑s\displaystyle\int\limits_{\tilde{D}_{R}}\nabla u\cdot\overline{\nabla\psi}-k^{2}u\overline{\psi}\,dx\ -\ \int\limits_{C_{R}}\Lambda u\,\overline{\psi}ds\ =\ \int\limits_{C_{R}}[\partial_{\nu}\tilde{u}^{in}-\Lambda\tilde{u}^{in}]\,\overline{\psi}\,ds

for all ψYR\psi\in Y_{R}. Application of Lemma 3.9, Theorem 2.16 and the Fredholm alternative yields the uniqueness and existence of uYRu\in Y_{R}. This also gives the scattered field usc=uuinH1(D~R)u^{sc}=u-u^{in}\in H^{1}(\tilde{D}_{R}) and the trace of the perturbed scattered field g:=upertsc|CR=(uscuunpertsc)|g:=u^{sc}_{pert}|_{C_{R}}=(u^{sc}-u^{sc}_{unpert})|_{-} on CRC_{R}. Finally, upertscu^{sc}_{pert} can be extended to ΣR\Sigma_{R} by solving the problem of Theorem 3.2 with φ=S1(g)\varphi=S^{-1}(g). ∎

Remark 4.4.

Suppose in Theorem 4.3 that ksinθ=α^jk\sin\theta=\hat{\alpha}_{j} is a propagative wavenumber for some fixed jJj\in J. Then it is well known that there exists still a α^j\hat{\alpha}_{j}-quasi-periodic solution uunpert,0=uin+uunpertscu_{unpert,0}=u^{in}+u^{sc}_{unpert} of the unperturbed problem. However, the solution is not unique, and the general solution is given by

(37) uunpert=uunpert,0+=1mjcϕ^,jin Du_{unpert}\ =\ u_{unpert,0}\ +\ \sum_{\ell=1}^{m_{j}}c_{\ell}\,\hat{\phi}_{\ell,j}\quad\mbox{in }D

where ϕ^,jXj\hat{\phi}_{\ell,j}\in X_{j} (see (4) and (6)) and cc_{\ell}\in\mathbb{C} are arbitrary. In our paper [12] we derive a new radiation condition based on the limiting absorption principe to prove uniqueness of the unperturbed scattering problem, even if ksinθk\sin\theta is a propagative wavenumber.


Now we consider the case that uin=ϕ^,ju^{in}=\hat{\phi}_{\ell,j} for some {1,,mj}\ell\in\{1,\ldots,m_{j}\} and jJj\in J is an incoming surface wave corresponding to the propagative wavenumber α^j\hat{\alpha}_{j}; that is,

{Δuin+k2uin=0inD,uin=0onΓ,uin is α^j-quasi-periodic in x1 and exponentially decays in the x2-direction.\displaystyle\left\{\begin{array}[]{lll}\Delta u^{in}+k^{2}u^{in}=0\quad\mbox{in}\quad D,\quad u^{in}=0\quad\mbox{on}\quad{\Gamma},\\ \mbox{$u^{in}$ is $\hat{\alpha}_{j}$-quasi-periodic in $x_{1}$ and exponentially decays in the $x_{2}$-direction.}\end{array}\right.

Since uinu^{in} vanishes already on Γ\Gamma and satisfies the radiation condition we conclude that the variational formulation for uYRu\in Y_{R} takes the same form as in (36) with u~in=uin\tilde{u}^{in}=u^{in}. Analogously to the proof of Theorem 4.3, we obtain

Theorem 4.5 (Well-posedness for incoming surface waves).

Given an incoming surface wave uin=ϕ^,ju^{in}=\hat{\phi}_{\ell,j} for some {1,,mj}\ell\in\{1,\ldots,m_{j}\} and jJj\in J, the perturbed scattering problem admits a unique solution u=uin+uscHloc,01(D~)u=u^{in}+u^{sc}\in H_{loc,0}^{1}(\tilde{D}) such that uscHloc1(D~)u^{sc}\in H_{loc}^{1}(\tilde{D}) fulfills the open waveguide radiation conditions of Def. 2.5 and Def. 2.10.

By Theorem 4.5, each surface wave ϕ^,j\hat{\phi}_{\ell,j} produces a non-trivial scattered field to the locally defected problem. Combining Theorems 4.3, 4.5 and Remark 4.4, we can get a general solution for plane wave incidence when ksinθk\sin\theta is a progagative wavenumber.

Corollary 4.6.

Let uinu^{in} be a plane wave and suppose that ksinθ=α^jk\sin\theta=\hat{\alpha}_{j} is a propagative wavenumber for some fixed jJj\in J. The general solution of the perturbed scattering problem for plane wave incidence takes the form

(39) u=uunpert,0+upertsc+=1mjcϕ^,j+=1mjcuscin ΣR.u\ =\ u_{unpert,0}\ +\ u^{sc}_{pert}\ +\ \sum_{\ell=1}^{m_{j}}c_{\ell}\,\hat{\phi}_{\ell,j}\ +\ \sum_{\ell=1}^{m_{j}}c_{\ell}\,u^{sc}_{\ell}\quad\mbox{in }\Sigma_{R}\,.

Here, upertscu_{pert}^{sc} is the open waveguide radiation solution determined in Theorem 4.3 excited by the incoming reference wave u~in=uunpert,0=uin+uunpertsc\tilde{u}^{in}=u_{unpert,0}=u^{in}+u^{sc}_{unpert}, and uscu^{sc}_{\ell} is the scattered field specified in Theorem 4.5 with uin:=ϕ^,ju^{in}:=\hat{\phi}_{\ell,j}.

5. Scattering by Neumann curves and by periodically arrayed obstacles

With slight changes our solvability results presented in Section 4 carry over to periodic and locally perturbed periodic curves of Neumann kind. Below we only remark the necessary modifications.

In the Neumann case, α[1/2,1/2]\alpha\in[-1/2,1/2] is called a propagative wave number if there exists a non-trivial ϕHα,loc1(D)\phi\in H^{1}_{\alpha,loc}(D) such that

Δϕ+k2ϕ= 0 in D,ϕν=0 on Γ,\displaystyle\Delta\phi+k^{2}\phi\ =\ 0\text{ in }D\,,\quad\frac{\partial\phi}{\partial\nu}=0\mbox{ on }\Gamma\,,

and ϕ\phi satisfies the Rayleigh expansion (3b). Here ν\nu denotes the normal direction at Γ\Gamma pointing into DD. Under the Assumption 2.2, one can still prove that there exist at most a finite number of propagative wavenumvers in the interval [1/2,1/2][-1/2,1/2]. The finite dimensional eigenspace XjX_{j} can be defined similarly to (4) but with the Neumann boundary condition on Γ\Gamma. The definition of the space H1(ΣR)H^{1}_{\ast}(\Sigma_{R}) should be replaced by

H1(ΣR):={uHloc1(ΣR):νu=0 on ΓΣR,uH1(WhΣR) for all h>h0}.H^{1}_{\ast}(\Sigma_{R})\ :=\ \bigl{\{}u\in H^{1}_{loc}(\Sigma_{R}):\partial_{\nu}u=0\mbox{ on }\Gamma\cap\partial\Sigma_{R},\ u\in H^{1}(W_{h}\cap\Sigma_{R})\mbox{ for all }h>h_{0}\bigr{\}}\,.

In this case, a bound state of the perturbed scattering problem is defined as a solution uH1(D~)u\in H^{1}(\tilde{D}) to the Helmholtz equation (Δ+k2)u=0(\Delta+k^{2})u=0 in D~\tilde{D} satisfying the Neumann boundary condition νu=0\partial_{\nu}u=0 on Γ~\tilde{\Gamma}. Assuming that there are no bound states in D~\tilde{D}, one can prove uniqueness to the perturbed scattering problem analogously to Theorem 2.16. To construct the DtN operator, we consider the problem of determining uHloc1(D)u\in H^{1}_{loc}(D) such that, for ϕH01/2(CR)\phi\in H_{0}^{-1/2}(C_{R}),

(40) Δu+k2u=0 in DCR,νu=0 on Γ,uνu+ν=φ on CR,\displaystyle\Delta u+k^{2}u=0\mbox{ in }D\setminus C_{R}\,,\quad\partial_{\nu}u=0\mbox{ on }\Gamma\,,\quad\frac{\partial u_{-}}{\partial\nu}-\frac{\partial u_{+}}{\partial\nu}=\varphi\mbox{ on }C_{R}\,,

and that uu satisfies the open waveguide radiation condition in ΣR\Sigma_{R}. The variational form of this transmission problem is to determine uHloc1(D)u\in H^{1}_{loc}(D) such that

(41) D[uψ¯k2uψ¯]𝑑x=CRφψ¯𝑑sfor all ψHc1(D),\displaystyle\int\limits_{D}[\nabla u\cdot\nabla\overline{\psi}-k^{2}u\,\overline{\psi}]\,dx\ =\ \int\limits_{C_{R}}\varphi\,\overline{\psi}\,ds\quad\mbox{for all }\psi\in H^{1}_{c}(D)\,,

together with the open waveguide radiation condition. Here,

Hc1(D):={ψH1(D):there exists a>0 with ϕ(x)=0 for all |x|>a}.\displaystyle H^{1}_{c}(D)\ :=\ \{\psi\in H^{1}(D):\mbox{there exists $a>0$ with $\phi(x)=0$ for all $|x|>a$}\}.

Note that the right hand side of (41) is understood as the duality between H01/2(CR)H_{0}^{-1/2}(C_{R}) and H1/2(CR)H^{1/2}(C_{R}). The mapping Sφ=u|CRS\varphi=u|_{C_{R}} defines the single layer operator under the Neumann boundary condition. Choose the open arc CRC_{R} such that the mixed boundary value problem

(Δ+k2)u=0 in DR,u=0 on CR,νu=0 on DRCR,\displaystyle(\Delta+k^{2})u=0\mbox{ in }D_{R}\,,\quad u=0\mbox{ on }C_{R}\,,\quad\partial_{\nu}u=0\mbox{ on }\partial D_{R}\setminus C_{R}\,,

admits the trivial solution only. We make the assumption that every solution uH1(ΣR)u\in H^{1}(\Sigma_{R}) to the exterior boundary value problem

(Δ+k2)u=0 in ΣR,u=0 on CR,νu=0 on ΓΣR,\displaystyle(\Delta+k^{2})u=0\mbox{ in }\Sigma_{R}\,,\quad u=0\mbox{ on }C_{R}\,,\quad\partial_{\nu}u=0\mbox{ on }\Gamma\cap\partial\Sigma_{R}\,,

must vanish identically, that is, there are no bound states to this special perturbation problem. The previous two conditions ensure that the single layer operator S:H01/2(CR)H1/2(CR)S:H^{-1/2}_{0}(C_{R})\rightarrow H^{1/2}(C_{R}) is injective and boundedly invertible. The DtN operator Λ\Lambda from H1/2(CR)H^{1/2}(C_{R}) into H01/2(CR)H^{-1/2}_{0}(C_{R}) takes the explicit form

Λg,ψ=DaDR[uψ~¯k2uψ~¯]𝑑x,ψH1/2(CR),\displaystyle\langle\Lambda g,\psi\rangle\ =\ -\int\limits_{D_{a}\setminus D_{R}}[\nabla u\cdot\nabla\overline{\tilde{\psi}}-k^{2}u\,\overline{\tilde{\psi}}]\,dx\,,\quad\psi\in H^{1/2}(C_{R})\,,

where ψ~=Eψ\tilde{\psi}=E\psi is a bounded extension operator from H1/2(CR)H^{1/2}(C_{R}) to H01(DaDR)H^{1}_{0}(D_{a}\setminus D_{R}) for some a>Ra>R. Here uu is the single layer potential with density φ:=S1gH01/2(CR)\varphi:=S^{-1}g\in H_{0}^{-1/2}(C_{R}); that is, the open waveguide radiation solution to the boundary value problem (40). Mapping properties of Λ\Lambda can be proved in the same way as Lemma 3.9. Finally, well-posedness results for scattering of point source waves, plane waves and surface waves from locally perturbed Neumann curves can be verified in the same manner as in the proofs of Theorems 4.1, 4.3 and 4.5,


Refer to caption
Figure 3. Illustration of the artificial boundary C:=KDC:=\partial K\subset D (in this case a circle) on which the DtN operator Λ\Lambda (see Definition 3.7) is defined for scattering by periodically arrayed obstacles with a local defect.

Let us now consider the TE and TM polarizations of time-harmonic electromagnetic scattering by periodically arrayed obstacles. Define the boundary conditions u:=u\mathcal{B}u:=u in the TE case and u:=νu\mathcal{B}u:=\partial_{\nu}u in the TM case. Let Ω×(H,H)\Omega\subset\mathbb{R}\times(-H,H) be a domain which is 2π2\pi-periodic with respect to x1x_{1} such that the exterior D:=2Ω¯D:=\mathbb{R}^{2}\setminus\overline{\Omega} is connected. Then α[1/2,1/2]\alpha\in[-1/2,1/2] is called a propagative wave number if there exists a non-trivial ϕHα,loc1(D)\phi\in H^{1}_{\alpha,loc}(D) such that

Δϕ+k2ϕ= 0 in D,ϕ=0 on D,\displaystyle\Delta\phi+k^{2}\phi\ =\ 0\text{ in }D\,,\quad\mathcal{B}\phi=0\mbox{ on }\partial D\,,

and ϕ\phi satisfies the Rayleigh expansions

ϕ(x)=ϕ±ei(+α)x1e±ik2(+α)2(x2H)for x2±H\displaystyle\phi(x)\ =\ \sum_{\ell\in\mathbb{Z}}\phi_{\ell}^{\pm}\,e^{i(\ell+\alpha)x_{1}}\,e^{\pm i\sqrt{k^{2}-(\ell+\alpha)^{2}}(x_{2}\mp H)}\quad\mbox{for }x_{2}\gtrless\pm H

for some ϕ±\phi^{\pm}_{\ell}\in\mathbb{C}. Then the spaces XjX_{j} of modes and their basis {ϕ^,j:=1,,mj}\{\hat{\phi}_{\ell,j}:\ell=1,\ldots,m_{j}\} are defined as in (4)–(6). Furthermore, let Ω\Omega be locally defected such that the periodic domain DD is replaced by a perturbed connected domain D~\tilde{D}. We assume that there exists a bounded Lipschitz domain KK which contains the defect (DD~)(D~D)(D\setminus\tilde{D})\cup(\tilde{D}\setminus D) and such that C:=KC:=\partial K is contained in DD. Defining Σ:=DK¯\Sigma:=D\setminus\overline{K} and the Sobolev space

H1(Σ):={uHloc1(Σ):u=0 on DΣ,uH1(WhΣ) for all h>H},H^{1}_{\ast}(\Sigma)\ :=\ \bigl{\{}u\in H^{1}_{loc}(\Sigma):\mathcal{B}u=0\mbox{ on }\partial D\cap\partial\Sigma\,,\ u\in H^{1}(W_{h}\cap\Sigma)\mbox{ for all }h>H\bigr{\}}\,,

where now Wh:=×(h,h)W_{h}:=\mathbb{R}\times(-h,h). Then the radiation conditions of Definitions 2.5, 2.8, and 2.12 carry over. CC, KK, and Σ\Sigma correspond to CRC_{R}, DRD_{R}, and ΣR\Sigma_{R}, respectively. A situation where CC can be chosen as a circle CRC_{R} is sketched in Figure 3. The Dirichlet-to-Neumann operator Λ\Lambda is again defined as Λg=νu|C\Lambda g=\partial_{\nu}u|_{C} where uHloc1(Σ)u\in H^{1}_{loc}(\Sigma) is the unique solution of

(43) Δu+k2u=0 in Σ,u=g on C,u=0 on DΣ,\displaystyle\Delta u+k^{2}u=0\mbox{ in }\Sigma\,,\quad u=g\mbox{ on }C\,,\quad\mathcal{B}u=0\mbox{ on }\partial D\cap\partial\Sigma\,,

together with the open waveguide radiation condition. We remark that the domain and range space of Λ\Lambda relies on the boundary condition under consideration. With proper assumptions on the domain KK, one can construct an invertible single layer operator Sφ=u|CS\varphi=u|_{C}, where uHloc1(D)u\in H^{1}_{loc}(D) is the radiating solution of the transmission problem

Δu+k2u=0 in DC,u=0 on D,uνu+ν=φ on C.\Delta u+k^{2}u=0\mbox{ in }D\setminus C\,,\quad\mathcal{B}u=0\mbox{ on }\partial D\,,\quad\frac{\partial u_{-}}{\partial\nu}-\frac{\partial u_{+}}{\partial\nu}=\varphi\mbox{ on }C\,.

Then one can define the DtN operator via Green’s formula, analogously to the scattering by Dirichlet and Neumann curves. The well-posedness results of Section 4 can be justified in the same manner.

Remark 5.1.

Exact boundary conditions (DtN maps) were also constructed for wave propagating in a closed periodic waveguide [13] and in a photonic crystal [7] containing a local perturbation. In comparision with [7], the DtN map defined by (43) applies to artificial boundary curves of arbitrary shape (although circular curves are used in this paper) and the medium is periodic in one direction only. The exact boundary condition of [7] is defined along square-shaped artificial boundaires, and the medium is periodic in two directions. In this paper the DtN map relies heavily on the open waveguide radiation condition of Def. 2.5.

Acknowledgements

The first author (G.H.) acknowledges the hospitality of the Institute for Applied and Numerical Mathematics, Karlsruhe Institute of Technology and the support of Alexander von Humboldt-Stiftung. The second author (A.K.) gratefully acknowledges the financial support by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Project-ID 258734477 – SFB 1173.

References

  • [1] A. S. Bonnet-Bendhia and P. Starling, Guided waves by electromagnetic gratings and non-uniqueness examples for the diffraction problem, Math. Meth. Appl. Sci., 17 (1994): 2305-338.
  • [2] S. N. Chandlea-Wilde, C.R. Ross, and B. Zhang, Scattering by rough surfaces, in Proceedings of the Fourth International Conference on Mathematical and Numerical Aspects of Wave Propagation (Golden, CO, 1998), SIAM, Philadelphia, PA, 1998, pp. 164-168.
  • [3] S.N. Chandlea-Wilde and B. Zhang, Electromagnetic scattering by an inhomogeneous conducting or dielectric layer on a perfectly conducting plate, Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences, 454 (1998): 519-542.
  • [4] S. N. Chandlea-Wilde and P. Monk, Existence, uniqueness, and variational methods for scattering by unbounded rough surfaces, SIAM J Math. Anal., 37 (2005): 598-618.
  • [5] S. N. Chandlea-Wilde and J. Elschner, Variational approach in weighted Sobolev spaces to scattering by unbounded rough surfaces, SIAM J. Math. Anal., 42 (2010): 2554-2580.
  • [6] J. Elschner and M. Yamamoto, An inverse problem in periodic diffractive optics: Reconstruction of Lipschitz grating profiles, Appl. Anal., 81 (2002): 1307-1328.
  • [7] S. Fliss and P. Joly, Exact boundary conditions for time-harmonic wave propagation in locally perturbed periodic media, Appl. Numer. Math., 59 (2009): 2155-2178.
  • [8] S. Fliss and P. Joly, Solutions of the time-harmonic wave equation in periodic waveguides: Asymptotic behavior and radiation condition, Arch. Ration. Mech. Anal., 219 (2016): 349-386.
  • [9] T. Furuya, Scattering by the local perturbation of an open periodic waveguide in the half plane, J. Math. Anal. Appl., 489 (2020): 124-149.
  • [10] G. Hu and A.Rathsfeld, Radiation conditions for the Helmholtz equation in a half plane filled by inhomogeneous periodic material, Journal of Differential Equations, 388 (2024): 215-252.
  • [11] G. Hu, W. Lu and A.Rathsfeld, Time-harmonic acoustic scattering from locally perturbed periodic curves, SIAM J. Appl. Math., 81 (2021): 2569-2595.
  • [12] G. Hu and A. Kirsch, Direct and inverse time-harmonic scattering by Dirichlet periodic curves with local perturbations.
  • [13] P. Joly, J.-R. Li and S. Fliss, Exact boundary conditions for periodic waveguides containing a local perturbation, Commun. Comput. Phys., 1 (2006): 945–973.
  • [14] I. V. Kamotski and S. A. Nazarov, The augmented scattering matrix and exponentially decaying solutions of an elliptic problem in a cylindrical domain, J. Math. Sci., 111 (2002): 3657-3666.
  • [15] A. Kirsch, Diffraction by periodic structures, in Proceedings of the Lapland Conference on Inverse Problems, L. Paivarinta and E. Summersalo, eds., Springer, Berlin, 1993, pp. 87- 102.
  • [16] A.Kirsch, Scattering by a periodic tube in 3\mathbb{R}^{3}: Part I. The limiting absorption principle, Inverse Problems, 35 (2019): 104004
  • [17] A. Kirsch, Scattering by a periodic tube in 3\mathbb{R}^{3}: Part II. A radiation condition, Inverse Problems, 35 (2019): 104005.
  • [18] A. Kirsch and A. Lechleiter, The limiting absorption principle and a radiation condition for the scattering by a periodic layer, SIAM J. Math. Anal., 50 (2018): 2536-2565.
  • [19] A. Kirsch, A scattering problem for a local perturbation of an open periodic waveguide, Math. Meth. Appl. Sci. 45 (2022): 5737-5773.
  • [20] A. Kirsch and A. Lechleiter, A radiation condition arising from the limiting absorption principle for a closed full- or half-waveguide problem, Math. Meth. Appl. Sci., 41 (2018): 3955-3975.
  • [21] J. W. S. Lord Rayleigh, On the dynamical theory of gratings, Proc. Roy. Soc. Lond. A, 79 (1907): 399-416.
  • [22] W. Mclean, Strongly Elliptic Systems and Boundary Integral Equations, Cambridge University Press, Cambridge, UK, 2010.
  • [23] R. Petit (ed.). Electromagnetic Theory of Gratings. Springer: Berlin, 1980.
  • [24] V. Yu. Gotlib, Solutions of the Helmholtz equation, concentrated near a plane periodic boundary, J. Math. Sci., 102 (2000): 4188-4194.