Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Examples of non-scattering inhomogeneities

Lucas Chesnel1, Houssem Haddar1, Hongjie Li2, Jingni Xiao3
1 IDEFIX, Ensta Paris, Institut Polytechnique de Paris, Palaiseau, France;
2 Yau Mathematical Sciences Center, Tsinghua University, Beijing, China;
3 Department of Mathematics, Drexel University, Philadelphia, Pennsylvania, USA.
E-mails: lucas.chesnel@inria.fr, houssem.haddar@inria.fr, hongjieli@tsinghua.edu.cn, jingni.xiao@drexel.edu.
(June 25, 2024)

Abstract. We consider the scattering of waves by a penetrable inclusion embedded in some reference medium. We exhibit examples of materials and geometries for which non-scattering frequencies exist, i.e., for which at some frequencies there are incident fields which produce null scattered fields outside of the inhomogeneity. We show in particular that certain domains with corners or even cusps can support non-scattering frequencies. We relate the latter, for some inclusions, to resonance frequencies for Dirichlet or Neumann cavities. We also find situations where incident non-scattering fields solve the Helmholtz equation in a neighborhood of the inhomogeneity and not in the whole space. Finally, in relation with invisibility, we give examples of inclusions of anisotropic materials which are non-scattering for all real frequencies. We prove that corresponding material indices must have a special structure on the boundary.
Key words. Non-scattering frequencies, transmission eigenvalues, interior transmission problem.
Mathematics Subject Classification. 35J05, 35P25, 35N25

1 Introduction

The so-called transmission eigenvalues play an important role in the study of inverse scattering from inhomogeneous media. They can be helpful in addressing theoretical questions such as uniqueness of the perturbation or in the justification of some reconstruction methods such as the Linear Sampling Method [12, 6]. These special frequencies can also be exploited in imaging algorithms for highly cluttered media [1, 2]. In general, transmission eigenvalues are associated with eigenfunctions that cannot be exactly represented as solutions to the Helmholtz equation in the whole space. In that case, exact non-scattering does not occur. Proving that incident fields always scatter in certain configurations has been the subject of several works and is usually associated with the presence of some geometric singularities such as corners in the boundary of the inhomogeneity. The techniques and results differ according to the considered model for the propagation of waves inside the inclusion, namely

Δu+k2qu=0Δ𝑢superscript𝑘2𝑞𝑢0\Delta u+k^{2}qu=0roman_Δ italic_u + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_q italic_u = 0 (1)

or

div(Au)+k2qu=0.div𝐴𝑢superscript𝑘2𝑞𝑢0\mathrm{div}(A\nabla u)+k^{2}qu=0.roman_div ( italic_A ∇ italic_u ) + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_q italic_u = 0 . (2)

Here k𝑘kitalic_k denotes the frequency and A𝐴Aitalic_A, q𝑞qitalic_q are functions characterizing the physical properties of the materials. We refer the reader respectively to [3, 25, 16, 9, 28, 19] and [4, 11, 29, 10] for studies concerning cases (1) and (2). In particular, it is proven for model (1) that if there is a corner in the boundary of the support of q1𝑞1q-1italic_q - 1, then at any frequency an incident wave will always produce a non zero scattered field outside the inhomogeneity, provided that q𝑞qitalic_q is sufficiently regular and different from the background coefficient 1111 near the corner. However, things can be different for model (2) especially when A𝐴Aitalic_A is different from the background coefficient IdId\mathrm{Id}roman_Id near the boundary of the support of the inhomogeneity. The main goal of the present article is to provide examples for which non-scattering occur for model (2). We show that this can happen even when the domain has singularities like corners or cusps. This analysis can be helpful in better understanding the optimality of some results concerning the absence of non-scattering frequencies in the literature, and provide some insight on the difference between cases (1) and (2). We also give examples of situations where non-scattering incident fields have some singularities outside the inhomogeneity. This motivates the definition of non-scattering frequencies adopted below.

Let us first describe the scattering problem we are considering. It is associated with the scalar acoustic wave equation in dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, d2𝑑2d\geq 2italic_d ≥ 2. We assume that some inclusion is located in a bounded domain ΩΩ\Omegaroman_Ω with Lipschitz boundary ΩΩ\partial\Omega∂ roman_Ω such that dΩ¯superscript𝑑¯Ω\mathbb{R}^{d}\setminus\overline{\Omega}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ∖ over¯ start_ARG roman_Ω end_ARG is connected. In accordance with (2), the inclusion is characterized by some material properties AL(Ω,d×d)𝐴superscriptLΩsuperscript𝑑𝑑A\in\mathrm{L}^{\infty}(\Omega,\mathbb{R}^{d\times d})italic_A ∈ roman_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( roman_Ω , blackboard_R start_POSTSUPERSCRIPT italic_d × italic_d end_POSTSUPERSCRIPT ), qL(Ω,)𝑞superscriptLΩq\in\mathrm{L}^{\infty}(\Omega,\mathbb{R})italic_q ∈ roman_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( roman_Ω , blackboard_R ) such that

A=Id,q=1 in dΩ¯;ess inf xΩinf ξd,|ξ|=1(ξA(x)ξ¯)>0;ess inf xΩq(x)>0.formulae-sequence𝐴Idformulae-sequence𝑞1 in superscript𝑑¯Ωformulae-sequencexΩess inf formulae-sequence𝜉superscript𝑑𝜉1inf 𝜉𝐴x¯𝜉0xΩess inf 𝑞x0A=\mathrm{Id},\ q=1\mbox{ in }\mathbb{R}^{d}\setminus\overline{\Omega};\qquad% \underset{\mathrm{x}\in\Omega}{\mbox{ess inf }}\underset{\xi\in\mathbb{C}^{d},% |\xi|=1}{\mbox{inf }}(\xi\cdot A(\mathrm{x})\overline{\xi})>0;\qquad\underset{% \mathrm{x}\in\Omega}{\mbox{ess inf }}q(\mathrm{x})>0.italic_A = roman_Id , italic_q = 1 in blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ∖ over¯ start_ARG roman_Ω end_ARG ; start_UNDERACCENT roman_x ∈ roman_Ω end_UNDERACCENT start_ARG ess inf end_ARG start_UNDERACCENT italic_ξ ∈ blackboard_C start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT , | italic_ξ | = 1 end_UNDERACCENT start_ARG inf end_ARG ( italic_ξ ⋅ italic_A ( roman_x ) over¯ start_ARG italic_ξ end_ARG ) > 0 ; start_UNDERACCENT roman_x ∈ roman_Ω end_UNDERACCENT start_ARG ess inf end_ARG italic_q ( roman_x ) > 0 . (3)

Below, we refer without distinction to (A,q;Ω)𝐴𝑞Ω(A,q;\Omega)( italic_A , italic_q ; roman_Ω ) as the inclusion or the inhomogeneity. Consider some incident field uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT satisfying

Δui+k2ui=0 in Ω~,Δsubscript𝑢𝑖superscript𝑘2subscript𝑢𝑖0 in ~Ω\Delta u_{i}+k^{2}u_{i}=0\quad\mbox{ in }\tilde{\Omega},roman_Δ italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 in over~ start_ARG roman_Ω end_ARG , (4)

where Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG is a neighborhood of ΩΩ\Omegaroman_Ω, more precisely a domain such that Ω¯Ω~¯Ω~Ω\overline{\Omega}\subset\tilde{\Omega}over¯ start_ARG roman_Ω end_ARG ⊂ over~ start_ARG roman_Ω end_ARG. Note that we do not assume uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT be defined and smooth in the whole dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT. This allows us to take into account the possibility of illuminating the inclusion for example by incident fields due to point sources. The scattering of uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT due to the inhomogeneity is governed by the following time-harmonic problem where ussubscript𝑢𝑠u_{s}italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT denotes the scattered field,

div(Aus)+k2qus=div((AId)ui)k2(q1)ui in dlimR+|x|=R|usrikus|2𝑑s(x)=0.div𝐴subscript𝑢𝑠superscript𝑘2𝑞subscript𝑢𝑠div𝐴Idsubscript𝑢𝑖superscript𝑘2𝑞1subscript𝑢𝑖 in superscript𝑑subscript𝑅subscript𝑥𝑅superscriptsubscript𝑢𝑠𝑟𝑖𝑘subscript𝑢𝑠2differential-d𝑠x0\begin{array}[]{|rcll}\mathrm{div}(A\nabla u_{s})+k^{2}qu_{s}&=&-\mathrm{div}(% (A-\mathrm{Id})\nabla u_{i})-k^{2}(q-1)u_{i}&\mbox{ in }\mathbb{R}^{d}\\[5.0pt% ] \vrule\lx@intercol\hfil\displaystyle{\displaystyle\lim_{R\to+\infty}\int_{|x|=% R}\left|\frac{\partial u_{s}}{\partial r}-iku_{s}\right|^{2}ds(\mathrm{x})=0.}% \hfil\lx@intercol\end{array}start_ARRAY start_ROW start_CELL roman_div ( italic_A ∇ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_q italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_CELL start_CELL = end_CELL start_CELL - roman_div ( ( italic_A - roman_Id ) ∇ italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_q - 1 ) italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_CELL start_CELL in blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL roman_lim start_POSTSUBSCRIPT italic_R → + ∞ end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT | italic_x | = italic_R end_POSTSUBSCRIPT | divide start_ARG ∂ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_r end_ARG - italic_i italic_k italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d italic_s ( roman_x ) = 0 . end_CELL end_ROW end_ARRAY (5)

The derivatives in the first line of (5) are to be understood in the weak sense where uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is extended by zero outside Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG. The last line of (5), in which r=|x|𝑟xr=|\mathrm{x}|italic_r = | roman_x |, is the so-called Sommerfeld radiation condition. The function u:=ui+usassign𝑢subscript𝑢𝑖subscript𝑢𝑠u:=u_{i}+u_{s}italic_u := italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is usually referred to as the total field. For all k>0𝑘0k>0italic_k > 0, Problem (5) has a unique solution ussubscript𝑢𝑠u_{s}italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT belonging to H1(𝒪)superscriptH1𝒪\mathrm{H}^{1}(\mathcal{O})roman_H start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( caligraphic_O ) for all bounded domains 𝒪d𝒪superscript𝑑\mathcal{O}\subset\mathbb{R}^{d}caligraphic_O ⊂ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT if d=2𝑑2d=2italic_d = 2. When d=3𝑑3d=3italic_d = 3, the existence of solutions can be established under the same assumptions. The uniqueness for d=3𝑑3d=3italic_d = 3 has been proven with the additional condition that A𝐴Aitalic_A is Lipschitz continuous in Ω¯¯Ω\overline{\Omega}over¯ start_ARG roman_Ω end_ARG, by using the unique continuation principle, see e.g. [6].
We define in the following the so-called non-scattering frequencies.

Definition 1.1.

We say that k>0𝑘0k>0italic_k > 0 is a non-scattering frequency if there is Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG such that Ω¯Ω~¯Ω~Ω\overline{\Omega}\subset\tilde{\Omega}over¯ start_ARG roman_Ω end_ARG ⊂ over~ start_ARG roman_Ω end_ARG and uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT solving (4) for which the corresponding scattered field satisfies us=0subscript𝑢𝑠0u_{s}=0italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 0 in dΩsuperscript𝑑Ω\mathbb{R}^{d}\setminus\Omegablackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ∖ roman_Ω. Such uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is referred to as a non-scattering incident field. We denote by 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT the set of non-scattering frequencies of (5).

If uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is a non-scattering incident field, then by setting v=ui𝑣subscript𝑢𝑖v=u_{i}italic_v = italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, we see that (u,v)H1(Ω)×H1(Ω)𝑢𝑣superscriptH1ΩsuperscriptH1Ω(u,v)\in\mathrm{H}^{1}(\Omega)\times\mathrm{H}^{1}(\Omega)( italic_u , italic_v ) ∈ roman_H start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( roman_Ω ) × roman_H start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( roman_Ω ) solves the interior transmission eigenvalue problem

div(Au)+k2qu=0 in ΩΔv+k2v=0 in Ωuv=0 on ΩνAuνv=0 on Ω.div𝐴𝑢superscript𝑘2𝑞𝑢0 in Ωmissing-subexpressionΔ𝑣superscript𝑘2𝑣0 in Ωmissing-subexpression𝑢𝑣0 on Ωmissing-subexpression𝜈𝐴𝑢𝜈𝑣0 on Ωmissing-subexpression\begin{array}[]{|rclll}\mathrm{div}(A\nabla u)+k^{2}qu&=&0&\mbox{ in }\Omega\\% [3.0pt] \Delta v+k^{2}v&=&0&\mbox{ in }\Omega\\[3.0pt] u-v&=&0&\mbox{ on }\partial\Omega\\[3.0pt] \nu\cdot A\nabla u-\nu\cdot\nabla v&=&0&\mbox{ on }\partial\Omega.\end{array}start_ARRAY start_ROW start_CELL roman_div ( italic_A ∇ italic_u ) + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_q italic_u end_CELL start_CELL = end_CELL start_CELL 0 end_CELL start_CELL in roman_Ω end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL roman_Δ italic_v + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v end_CELL start_CELL = end_CELL start_CELL 0 end_CELL start_CELL in roman_Ω end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_u - italic_v end_CELL start_CELL = end_CELL start_CELL 0 end_CELL start_CELL on ∂ roman_Ω end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_ν ⋅ italic_A ∇ italic_u - italic_ν ⋅ ∇ italic_v end_CELL start_CELL = end_CELL start_CELL 0 end_CELL start_CELL on ∂ roman_Ω . end_CELL start_CELL end_CELL end_ROW end_ARRAY (6)

Here ν𝜈\nuitalic_ν stands for the unit outward normal vector to ΩΩ\partial\Omega∂ roman_Ω.

Definition 1.2.

We say that k>0𝑘0k>0italic_k > 0 is a transmission eigenvalue if there is a non-zero (u,v)𝑢𝑣(u,v)( italic_u , italic_v ) which satisfies (6). We denote by 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT the set of transmission eigenvalues.

The above discussion shows that 𝒮NS𝒮TEsubscript𝒮NSsubscript𝒮TE\mathscr{S}_{\mathrm{NS}}\subset\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT ⊂ script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT. However the converse does not hold in general. More precisely, we have the following characterization result whose proof is straightforward.

Proposition 1.3.

A real frequency k𝒮TE𝑘subscript𝒮TEk\in\mathscr{S}_{\mathrm{TE}}italic_k ∈ script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT is also in 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT if and only if there is a non-zero eigenpair (u,v)𝑢𝑣(u,v)( italic_u , italic_v ) of (6) associated with k𝑘kitalic_k such that v𝑣vitalic_v can be extended in Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG, a neighborhood of ΩΩ\Omegaroman_Ω, as a function v~~𝑣\tilde{v}over~ start_ARG italic_v end_ARG satisfying

Δv~+k2v~=0 in Ω~.Δ~𝑣superscript𝑘2~𝑣0 in ~Ω\Delta\tilde{v}+k^{2}\tilde{v}=0\mbox{ in }\tilde{\Omega}.roman_Δ over~ start_ARG italic_v end_ARG + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over~ start_ARG italic_v end_ARG = 0 in over~ start_ARG roman_Ω end_ARG . (7)

Concerning (6), it has first been shown that 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT is discrete for rather mild conditions on the coefficients A𝐴Aitalic_A and n𝑛nitalic_n [6]. Under a bit more restrictive hypothesis on A𝐴Aitalic_A and q𝑞qitalic_q, people have been able to prove that 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT contains a countably infinite sequence of eigenvalues accumulating only at ++\infty+ ∞ (see [26, 17, 8, 7, 27]). In contrast, the set 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT has been less studied. Certain simple examples of geometries (rectangle, balls) where 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}\neq\emptysetscript_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT ≠ ∅ have been given in [21, 10]. On the other hand, it has been established that for certain scatterers with non-smooth geometries (having corners in 2D or conical tips/edges in 3D), there holds 𝒮NS=subscript𝒮NS\mathscr{S}_{\mathrm{NS}}=\emptysetscript_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT = ∅ (see [3, 25, 16]).

As indicated above, the main goal of the present article is to provide examples of configurations where 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}\neq\emptysetscript_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT ≠ ∅. We first consider the case where the material properties have the special form A=aId𝐴𝑎IdA=a\,\mathrm{Id}italic_A = italic_a roman_Id, q=a𝑞𝑎q=aitalic_q = italic_a, with a>0𝑎0a>0italic_a > 0 (Section 2). This configuration is interesting for several reasons. For instance, one can completely characterize in this case the set 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT as the union of Dirichlet and Neumann cavity eigenvalues. One can therefore exhibit non-scattering frequencies by considering either Dirichlet or Neumann eigenvalues. We show in particular that some polygonal domains (convex or not) possess non-scattering frequencies. We also exhibit examples, some inspired by and some extracted from [15], where non-scattering incident fields are not entire solutions to the Helmholtz equation (i.e. Ω~d~Ωsuperscript𝑑\tilde{\Omega}\neq\mathbb{R}^{d}over~ start_ARG roman_Ω end_ARG ≠ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT). We consider separately Dirichlet and Neumann eigenvalues. While examples for the first case are already present in the literature, the examples for Neumann eigenvalues are harder to find analytically and seem less known. We propose a method to construct such domains based on analysing the characteristics for the gradient of the solutions to the Helmholtz equation. This allows for instance to prove existence of non-scattering frequencies for certain domains with corners of arbitrary angles and even cusps. We emphasize that on the contrary, non-scattering domains for the Dirichlet eigenvalues can only have corners of particular apertures. In Section 3, we work with anisotropic materials and exhibit configurations where non-scattering occurs at any frequencies. The first examples are inspired by the concept of invisibility by diffeomorphisms [18]. For associated anisotropies we prove in particular that the matrix A𝐴Aitalic_A has a special structure at irregular points of the boundary as it must satisfy Aν=ν𝐴𝜈𝜈A\nu=\nuitalic_A italic_ν = italic_ν (see also [10]). We also provide other examples of anisotropies (not associated with diffeomorphism transformations) for which non-scattering occur at any frequency in the case of regular and non-regular domains.

Note: we say that a function is an entire solution of the Helmholtz equation if it solves the homogeneous Helmholtz equation in dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT.

2 Non-scattering in the case A=aId𝐴𝑎IdA=a\,\mathrm{Id}italic_A = italic_a roman_Id, q=a𝑞𝑎q=aitalic_q = italic_a, with a>0𝑎0a>0italic_a > 0

In this section, we make the assumption that there holds

A=aIdandq=a in Ω,formulae-sequence𝐴𝑎Idand𝑞𝑎 in ΩA=a\,\mathrm{Id}\quad\text{and}\quad q=a\qquad\mbox{ in }\Omega,italic_A = italic_a roman_Id and italic_q = italic_a in roman_Ω , (8)

where a>0𝑎0a>0italic_a > 0 is a constant such that a1𝑎1a\neq 1italic_a ≠ 1. In this particular situation, we first show that 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT is formed by the union of the eigenvalues of the Dirichlet and Neumann Laplacians in ΩΩ\Omegaroman_Ω. More precisely, introduce the problems

ΔwD+k2wD=0 in ΩwD=0 on Ω and ΔwN+k2wN=0 in ΩνwN=0 on Ω.Δsubscript𝑤𝐷superscript𝑘2subscript𝑤𝐷0 in Ωsubscript𝑤𝐷0 on Ω and Δsubscript𝑤𝑁superscript𝑘2subscript𝑤𝑁0 in Ωsubscript𝜈subscript𝑤𝑁0 on Ω\begin{array}[]{|rl}\Delta w_{D}+k^{2}w_{D}=0&\mbox{ in }\Omega\\[3.0pt] w_{D}=0&\mbox{ on }\partial\Omega\end{array}\qquad\mbox{ and }\qquad\begin{% array}[]{|rl}\Delta w_{N}+k^{2}w_{N}=0&\mbox{ in }\Omega\\[3.0pt] \partial_{\nu}w_{N}=0&\mbox{ on }\partial\Omega.\end{array}start_ARRAY start_ROW start_CELL roman_Δ italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT = 0 end_CELL start_CELL in roman_Ω end_CELL end_ROW start_ROW start_CELL italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT = 0 end_CELL start_CELL on ∂ roman_Ω end_CELL end_ROW end_ARRAY and start_ARRAY start_ROW start_CELL roman_Δ italic_w start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = 0 end_CELL start_CELL in roman_Ω end_CELL end_ROW start_ROW start_CELL ∂ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = 0 end_CELL start_CELL on ∂ roman_Ω . end_CELL end_ROW end_ARRAY (9)

We denote by 𝒮Dsubscript𝒮D\mathscr{S}_{\mathrm{D}}script_S start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT (resp. 𝒮Nsubscript𝒮N\mathscr{S}_{\mathrm{N}}script_S start_POSTSUBSCRIPT roman_N end_POSTSUBSCRIPT) the values of k>0𝑘0k>0italic_k > 0 such that the Dirichlet (resp. Neumann) problem (9) admits non-zero solutions in H1(Ω)superscriptH1Ω\mathrm{H}^{1}(\Omega)roman_H start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( roman_Ω ). We have the following result.

Lemma 2.1.

When A𝐴Aitalic_A, q𝑞qitalic_q satisfy (8), we have 𝒮TE=𝒮D𝒮Nsubscript𝒮TEsubscript𝒮Dsubscript𝒮N\mathscr{S}_{\mathrm{TE}}=\mathscr{S}_{\mathrm{D}}\cup\mathscr{S}_{\mathrm{N}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT = script_S start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT ∪ script_S start_POSTSUBSCRIPT roman_N end_POSTSUBSCRIPT.

Proof.

Suppose that (u,v)𝑢𝑣(u,v)( italic_u , italic_v ) is a transmission eigenpair solving (6). Define wD:=uvassignsubscript𝑤𝐷𝑢𝑣w_{D}:=u-vitalic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT := italic_u - italic_v and wN:=auvassignsubscript𝑤𝑁𝑎𝑢𝑣w_{N}:=au-vitalic_w start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT := italic_a italic_u - italic_v. Then one readily checks that wDsubscript𝑤𝐷w_{D}italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT, wNsubscript𝑤𝑁w_{N}italic_w start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT solve respectively the Dirichlet and Neumann eigenvalue problems (9). Moreover, at least one of the functions wDsubscript𝑤𝐷w_{D}italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT and wNsubscript𝑤𝑁w_{N}italic_w start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT must be non-trivial. Hence, we have 𝒮TE(𝒮D𝒮N)subscript𝒮TEsubscript𝒮Dsubscript𝒮N\mathscr{S}_{\mathrm{TE}}\subset(\mathscr{S}_{\mathrm{D}}\cup\mathscr{S}_{% \mathrm{N}})script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT ⊂ ( script_S start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT ∪ script_S start_POSTSUBSCRIPT roman_N end_POSTSUBSCRIPT ).
Conversely, assume that wDsubscript𝑤𝐷w_{D}italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT is a non-zero solution of the Dirichlet problem appearing in (9). Then (u,v)=(awD,wD)𝑢𝑣𝑎subscript𝑤𝐷subscript𝑤𝐷(u,v)=(aw_{D},w_{D})( italic_u , italic_v ) = ( italic_a italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT ) is a (non-trivial) eigenpair of (6). Similarly, if wNsubscript𝑤𝑁w_{N}italic_w start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT is a non-zero solution of the Neumann problem appearing in (9), then (u,v)=(wN,wN)𝑢𝑣subscript𝑤𝑁subscript𝑤𝑁(u,v)=(w_{N},w_{N})( italic_u , italic_v ) = ( italic_w start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) solves (6). This guarantees that (𝒮D𝒮N)𝒮TEsubscript𝒮Dsubscript𝒮Nsubscript𝒮TE(\mathscr{S}_{\mathrm{D}}\cup\mathscr{S}_{\mathrm{N}})\subset\mathscr{S}_{% \mathrm{TE}}( script_S start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT ∪ script_S start_POSTSUBSCRIPT roman_N end_POSTSUBSCRIPT ) ⊂ script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT. ∎

Remark 2.2.

Note that the case where A𝐴Aitalic_A, q𝑞qitalic_q satisfy (8) is a situation where all the k2superscript𝑘2k^{2}\in\mathbb{C}italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∈ blackboard_C such that (6) admits a non-zero solution are real. It is not known if this can happen with other pairs A𝐴Aitalic_A, q𝑞qitalic_q.

From Lemma 2.1, when A𝐴Aitalic_A, q𝑞qitalic_q satisfy (8), the question of finding non-scattering eigenvalues can be reformulated as “are there eigenfunctions of the Dirichlet or Neumann Laplacian in ΩΩ\Omegaroman_Ω that can be extended as solutions to the Helmholtz equation in a neighborhood of Ω¯¯Ω\overline{\Omega}over¯ start_ARG roman_Ω end_ARG?”. Some positive answers can be given in three situations:
- For certain geometries ΩΩ\Omegaroman_Ω, one can compute analytically the eigenfunctions of the Dirichlet/Neumann Laplacians and observe that they are defined in larger domains than ΩΩ\Omegaroman_Ω;
- For certain ΩΩ\Omegaroman_Ω, one can use reflections to extend the eigenfunctions of the Dirichlet/Neumann Laplacians in ΩΩ\Omegaroman_Ω to larger domains;
- Given a function uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT solving the Helmholtz equation (4) in some given domain Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG, one can look for ΩΩ\Omegaroman_Ω, with Ω¯Ω~¯Ω~Ω\overline{\Omega}\subset\tilde{\Omega}over¯ start_ARG roman_Ω end_ARG ⊂ over~ start_ARG roman_Ω end_ARG, for which uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is an eigenfunction of the Dirichlet or Neumann Laplacian in ΩΩ\Omegaroman_Ω.
We present corresponding results in the following subsections.

2.1 Domains where analytic expressions can be obtained for the eigenfunctions

Assume here that ΩΩ\Omegaroman_Ω coincides with the unit square and consider k𝒮D𝑘subscript𝒮Dk\in\mathscr{S}_{\mathrm{D}}italic_k ∈ script_S start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT as well as wDsubscript𝑤𝐷w_{D}italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT a corresponding eigenfunction of the Dirichlet problem appearing in (9). In an appropriate set of coordinates, wDsubscript𝑤𝐷w_{D}italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT writes as a linear combination of the functions

sin(mπx)sin(nπy),m,n:={1,2,3,}.𝑚𝜋𝑥𝑛𝜋𝑦𝑚𝑛superscriptassign123\sin(m\pi x)\sin(n\pi y),\qquad m,n\in\mathbb{N}^{\ast}:=\{1,2,3,\dots\}.roman_sin ( italic_m italic_π italic_x ) roman_sin ( italic_n italic_π italic_y ) , italic_m , italic_n ∈ blackboard_N start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT := { 1 , 2 , 3 , … } .

In the proof of Lemma 2.1, we have seen that (u,v)=(awD,wD)𝑢𝑣𝑎subscript𝑤𝐷subscript𝑤𝐷(u,v)=(aw_{D},w_{D})( italic_u , italic_v ) = ( italic_a italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT ) constitutes an eigenpair of (6). Clearly v𝑣vitalic_v extends as a function solving the homogeneous Helmholtz equation (7) in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. From Proposition 1.3, we infer that k𝒮NS𝑘subscript𝒮NSk\in\mathscr{S}_{\mathrm{NS}}italic_k ∈ script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT. Interestingly, as explained in [3, 10], in that case one can exhibit non-scattering incident fields uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT which are simple combinations of plane waves. More precisely, uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT such that

ui(x,y)=4sin(mπx)sin(nπy)=eiπ(mxny)+eiπ(mxny)eiπ(mx+ny)eiπ(mx+ny)subscript𝑢𝑖𝑥𝑦4𝑚𝜋𝑥𝑛𝜋𝑦missing-subexpressionsuperscript𝑒𝑖𝜋𝑚𝑥𝑛𝑦superscript𝑒𝑖𝜋𝑚𝑥𝑛𝑦superscript𝑒𝑖𝜋𝑚𝑥𝑛𝑦superscript𝑒𝑖𝜋𝑚𝑥𝑛𝑦\begin{array}[]{rcl}u_{i}(x,y)&=&4\sin(m\pi x)\sin(n\pi y)\\[3.0pt] &=&e^{i\pi(mx-ny)}+e^{-i\pi(mx-ny)}-e^{i\pi(mx+ny)}-e^{-i\pi(mx+ny)}\end{array}start_ARRAY start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x , italic_y ) end_CELL start_CELL = end_CELL start_CELL 4 roman_sin ( italic_m italic_π italic_x ) roman_sin ( italic_n italic_π italic_y ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = end_CELL start_CELL italic_e start_POSTSUPERSCRIPT italic_i italic_π ( italic_m italic_x - italic_n italic_y ) end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT - italic_i italic_π ( italic_m italic_x - italic_n italic_y ) end_POSTSUPERSCRIPT - italic_e start_POSTSUPERSCRIPT italic_i italic_π ( italic_m italic_x + italic_n italic_y ) end_POSTSUPERSCRIPT - italic_e start_POSTSUPERSCRIPT - italic_i italic_π ( italic_m italic_x + italic_n italic_y ) end_POSTSUPERSCRIPT end_CELL end_ROW end_ARRAY

produces a scattered field which is exactly zero outside of ΩΩ\Omegaroman_Ω. Similarly we establish that 𝒮N𝒮NSsubscript𝒮Nsubscript𝒮NS\mathscr{S}_{\mathrm{N}}\subset\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_N end_POSTSUBSCRIPT ⊂ script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT.

This reasoning can be adapted to deal with other simple domains ΩΩ\Omegaroman_Ω where we can use separation of variables. This allows one to state the following result.

Proposition 2.3.

Assume that Ω2Ωsuperscript2\Omega\subset\mathbb{R}^{2}roman_Ω ⊂ blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is a rectangle, a disk, or an elliptical domain (whose boundary is an ellipse). Then when A𝐴Aitalic_A, q𝑞qitalic_q satisfy (8), we have 𝒮NS=𝒮TE=𝒮D𝒮Nsubscript𝒮NSsubscript𝒮TEsubscript𝒮Dsubscript𝒮N\mathscr{S}_{\mathrm{NS}}=\mathscr{S}_{\mathrm{TE}}=\mathscr{S}_{\mathrm{D}}% \cup\mathscr{S}_{\mathrm{N}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT = script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT = script_S start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT ∪ script_S start_POSTSUBSCRIPT roman_N end_POSTSUBSCRIPT.

Note that to consider the case of elliptic domains, we work with the elliptic coordinates (μ,ϕ)𝜇italic-ϕ(\mu,\phi)( italic_μ , italic_ϕ ) such that, after a rigid change of variables,

x=αcoshμcosϕ,y=αsinhμsinϕ,formulae-sequence𝑥𝛼𝜇italic-ϕ𝑦𝛼𝜇italic-ϕx=\alpha\cosh\mu\cos\phi,\qquad y=\alpha\sinh\mu\sin\phi,italic_x = italic_α roman_cosh italic_μ roman_cos italic_ϕ , italic_y = italic_α roman_sinh italic_μ roman_sin italic_ϕ ,

where μ0𝜇0\mu\geq 0italic_μ ≥ 0, ϕ/(2π)italic-ϕ2𝜋\phi\in\mathbb{R}/(2\pi)italic_ϕ ∈ blackboard_R / ( 2 italic_π ) and (±α,0)plus-or-minus𝛼0(\pm\alpha,0)( ± italic_α , 0 ) stand for the foci of the ellipse. In particular, the curves μ=constant>0𝜇constant0\mu=\mathrm{constant}>0italic_μ = roman_constant > 0 are confocal ellipses. Solutions of (9) can be decomposed on functions with separate variables in μ𝜇\muitalic_μ, ϕitalic-ϕ\phiitalic_ϕ. For the latter functions, we find that the dependences in μ𝜇\muitalic_μ, ϕitalic-ϕ\phiitalic_ϕ satisfy respectively a modified Mathieu’s equation and a Mathieu’s equation with periodic boundary conditions, which are both Sturm-Liouville problems. Moreover it is known that the solutions of the Mathieu’s equation which are regular at zero can be extended in \mathbb{R}blackboard_R so that all eigenfunctions of the Dirichlet and Neumann Laplacians in a bounded elliptical domain can be extended as solutions of an Helmholtz equation in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. For more details, we refer the readers to [14, 23].
Similar to Proposition 2.3, we have the following statement in 3superscript3\mathbb{R}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT.

Proposition 2.4.

Assume that Ω3Ωsuperscript3\Omega\subset\mathbb{R}^{3}roman_Ω ⊂ blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT is a ball, an ellipsoid, or, in an appropriate system of coordinates, there holds Ω=I×ωΩ𝐼𝜔\Omega=I\times\omegaroman_Ω = italic_I × italic_ω where I𝐼Iitalic_I is a bounded open interval and ω2𝜔superscript2\omega\subset\mathbb{R}^{2}italic_ω ⊂ blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is a rectangle, a disk, or an elliptical domain. Then when A𝐴Aitalic_A, q𝑞qitalic_q satisfy (8), we have 𝒮NS=𝒮TE=𝒮D𝒮Nsubscript𝒮NSsubscript𝒮TEsubscript𝒮Dsubscript𝒮N\mathscr{S}_{\mathrm{NS}}=\mathscr{S}_{\mathrm{TE}}=\mathscr{S}_{\mathrm{D}}% \cup\mathscr{S}_{\mathrm{N}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT = script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT = script_S start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT ∪ script_S start_POSTSUBSCRIPT roman_N end_POSTSUBSCRIPT.

Refer to caption Refer to captionRefer to caption

Figure 1: Sector (left), spherical wedge (center) and cylindrical wedge (right).

We now turn our attention to other geometries. Consider some constants α(0,2π)𝛼02𝜋\alpha\in(0,2\pi)italic_α ∈ ( 0 , 2 italic_π ) and >00\ell>0roman_ℓ > 0. Define in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT the sector

Ω={(rcosϕ,rsinϕ)|r(0,),ϕ(0,α)}Ωconditional-set𝑟italic-ϕ𝑟italic-ϕformulae-sequence𝑟0italic-ϕ0𝛼\Omega=\{(r\cos\phi,r\sin\phi)\,|\,r\in(0,\ell),\,\phi\in(0,\alpha)\}roman_Ω = { ( italic_r roman_cos italic_ϕ , italic_r roman_sin italic_ϕ ) | italic_r ∈ ( 0 , roman_ℓ ) , italic_ϕ ∈ ( 0 , italic_α ) }

and in 3superscript3\mathbb{R}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, the spherical wedge

Ω={(rcosϕsinθ,rsinϕsinθ,rcosθ)|r(0,),θ(0,π),ϕ(0,α)}Ωconditional-set𝑟italic-ϕ𝜃𝑟italic-ϕ𝜃𝑟𝜃formulae-sequence𝑟0formulae-sequence𝜃0𝜋italic-ϕ0𝛼\Omega=\{(r\cos\phi\sin\theta,r\sin\phi\sin\theta,r\cos\theta)\,|\,r\in(0,\ell% ),\theta\in(0,\pi),\phi\in(0,\alpha)\}roman_Ω = { ( italic_r roman_cos italic_ϕ roman_sin italic_θ , italic_r roman_sin italic_ϕ roman_sin italic_θ , italic_r roman_cos italic_θ ) | italic_r ∈ ( 0 , roman_ℓ ) , italic_θ ∈ ( 0 , italic_π ) , italic_ϕ ∈ ( 0 , italic_α ) }

as well as the cylindrical wedge

Ω={(rcosϕ,rsinϕ,z)|r(0,),ϕ(0,α),zI}Ωconditional-set𝑟italic-ϕ𝑟italic-ϕ𝑧formulae-sequence𝑟0formulae-sequenceitalic-ϕ0𝛼𝑧𝐼\Omega=\{(r\cos\phi,r\sin\phi,z)\,|\,r\in(0,\ell),\,\phi\in(0,\alpha),z\in I\}roman_Ω = { ( italic_r roman_cos italic_ϕ , italic_r roman_sin italic_ϕ , italic_z ) | italic_r ∈ ( 0 , roman_ℓ ) , italic_ϕ ∈ ( 0 , italic_α ) , italic_z ∈ italic_I }

(see Figure 1 for illustrations). Here I𝐼Iitalic_I is an open bounded interval. For theses domains, depending on the opening angle, there may exist zero or infinitely many non-scattering frequencies.

Proposition 2.5.

With the notation above, assume that ΩΩ\Omegaroman_Ω is either a sector in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, a spherical wedge in 3superscript3\mathbb{R}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT or a cylindrical wedge in 3superscript3\mathbb{R}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, of angle α𝛼\alphaitalic_α. When A𝐴Aitalic_A, q𝑞qitalic_q satisfy (8), we have:
- if α(0,2π)π𝛼02𝜋𝜋\alpha\in(0,2\pi)\cap\mathbb{Q}\piitalic_α ∈ ( 0 , 2 italic_π ) ∩ blackboard_Q italic_π, then 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT contains an unbounded sequence which accumulates only at ++\infty+ ∞;
- if α(0,2π)(ππ)𝛼02𝜋𝜋𝜋\alpha\in(0,2\pi)\cap(\mathbb{R}\pi\setminus\mathbb{Q}\pi)italic_α ∈ ( 0 , 2 italic_π ) ∩ ( blackboard_R italic_π ∖ blackboard_Q italic_π ), then 𝒮NS=subscript𝒮NS\mathscr{S}_{\mathrm{NS}}=\emptysetscript_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT = ∅.

Remark 2.6.

In fact, we show in the proof that 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT coincides exactly with 𝒮TE=𝒮D𝒮Nsubscript𝒮TEsubscript𝒮Dsubscript𝒮N\mathscr{S}_{\mathrm{TE}}=\mathscr{S}_{\mathrm{D}}\cup\mathscr{S}_{\mathrm{N}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT = script_S start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT ∪ script_S start_POSTSUBSCRIPT roman_N end_POSTSUBSCRIPT if and only if α(0,2π)(π/)𝛼02𝜋𝜋superscript\alpha\in(0,2\pi)\cap(\pi/\mathbb{N}^{\ast})italic_α ∈ ( 0 , 2 italic_π ) ∩ ( italic_π / blackboard_N start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ).

Remark 2.7.

In the second item of Proposition 2.5, the fact that 𝒮NS=subscript𝒮NS\mathscr{S}_{\mathrm{NS}}=\emptysetscript_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT = ∅ in that domains ΩΩ\Omegaroman_Ω is not only due to the value of the opening angle α𝛼\alphaitalic_α but also to the shape of the whole ΩΩ\Omegaroman_Ω. Indeed in §2.3.2, we will exhibit examples of geometries in the plane supporting non-scattering frequencies with corners of arbitrary α(0,2π)𝛼02𝜋\alpha\in(0,2\pi)italic_α ∈ ( 0 , 2 italic_π ).

Proof.

Let ΩΩ\Omegaroman_Ω be a sector in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The eigenfunctions of the Dirichlet Laplacian in ΩΩ\Omegaroman_Ω coincide with the functions Jmπ/α(μmr)sin(mπϕ/α)subscript𝐽𝑚𝜋𝛼subscript𝜇𝑚𝑟𝑚𝜋italic-ϕ𝛼J_{m\pi/\alpha}(\mu_{m}r)\sin(m\pi\phi/\alpha)italic_J start_POSTSUBSCRIPT italic_m italic_π / italic_α end_POSTSUBSCRIPT ( italic_μ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_r ) roman_sin ( italic_m italic_π italic_ϕ / italic_α ), with m𝑚superscriptm\in\mathbb{N}^{\ast}italic_m ∈ blackboard_N start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and μm>0subscript𝜇𝑚0\mu_{m}>0italic_μ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT > 0 such that Jmπ/α(μm)=0subscript𝐽𝑚𝜋𝛼subscript𝜇𝑚0J_{m\pi/\alpha}(\mu_{m}\ell)=0italic_J start_POSTSUBSCRIPT italic_m italic_π / italic_α end_POSTSUBSCRIPT ( italic_μ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT roman_ℓ ) = 0. Here Jmπ/αsubscript𝐽𝑚𝜋𝛼J_{m\pi/\alpha}italic_J start_POSTSUBSCRIPT italic_m italic_π / italic_α end_POSTSUBSCRIPT is the Bessel function of the first kind with order mπ/α𝑚𝜋𝛼m\pi/\alphaitalic_m italic_π / italic_α. As a consequence, by looking at the singularity of Bessel functions at the origin and periodicity of the sin function, we infer that an eigenfunction can be extended as an entire solution of the Helmholtz equation if and only if α(0,2π)π/𝛼02𝜋𝜋superscript\alpha\in(0,2\pi)\cap\pi/\mathbb{N}^{\ast}italic_α ∈ ( 0 , 2 italic_π ) ∩ italic_π / blackboard_N start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. The analysis is similar for eigenfunctions of the Neumann Laplacian.
The cases of spherical and cylindrical wedges in 3superscript3\mathbb{R}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT can be dealt similarly by working, respectively, with spherical and cylindrical coordinates. ∎

With this approach, we could also consider 3D conical tip, i.e. domains of the form

Ω={(rcosθcosϕ,rcosθsinϕ,rsinθ)| 0<r<,0<θ<θ0,0ϕ<2π}Ωconditional-set𝑟𝜃italic-ϕ𝑟𝜃italic-ϕ𝑟𝜃formulae-sequence 0𝑟0𝜃subscript𝜃00italic-ϕ2𝜋\Omega=\{(r\cos\theta\cos\phi,r\cos\theta\sin\phi,r\sin\theta)\,|\,0<r<\ell,0<% \theta<\theta_{0},0\leq\phi<2\pi\}roman_Ω = { ( italic_r roman_cos italic_θ roman_cos italic_ϕ , italic_r roman_cos italic_θ roman_sin italic_ϕ , italic_r roman_sin italic_θ ) | 0 < italic_r < roman_ℓ , 0 < italic_θ < italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , 0 ≤ italic_ϕ < 2 italic_π }

with θ0>0subscript𝜃00\theta_{0}>0italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0. In this case, the possibility of extending eigenfunctions of the Dirichlet/Neumann Laplacians in ΩΩ\Omegaroman_Ω depends on whether cosθ0subscript𝜃0\cos\theta_{0}roman_cos italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is a zero of the Legendre polynomial Pnmsubscriptsuperscript𝑃𝑚𝑛P^{m}_{n}italic_P start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT for some integers n𝑛nitalic_n and m[n,n]𝑚𝑛𝑛m\in[-n,n]italic_m ∈ [ - italic_n , italic_n ]. We choose not to elaborate much on this direction.

Let us mention that some of the discussions above for 2D cases can also be found in [20].

2.2 Domains where eigenfunctions can be extended by reflections

Now we present other geometries where some or all the Dirichlet/Neumann eigenfunctions of (9) can be extended, this time by working with reflections. We start with a definition.

Definition 2.8.

Let ΩΩ\Omegaroman_Ω be a polygon of 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (resp. a polyhedron of 3superscript3\mathbb{R}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT). We say that ΩΩ\Omegaroman_Ω is a proper paving unit if the following two conditions hold:
- One can find a combination of successive reflections of Ω¯¯Ω\overline{\Omega}over¯ start_ARG roman_Ω end_ARG with respect to the edges (resp. faces) that allows one to cover a domain Ω~Ω¯¯Ω~Ω\tilde{\Omega}\supset\overline{\Omega}over~ start_ARG roman_Ω end_ARG ⊃ over¯ start_ARG roman_Ω end_ARG with the overlaps only on the skeleton 𝒮𝒮\mathcal{S}caligraphic_S consisting of the edges (resp. faces) of ΩΩ\partial\Omega∂ roman_Ω and their images by the reflections.
- If one assigns a different colour to each of the edges (resp. faces) of ΩΩ\partial\Omega∂ roman_Ω, then each element of 𝒮𝒮\mathcal{S}caligraphic_S may inherit two or more colours due to reflections. We require that every element of 𝒮𝒮\mathcal{S}caligraphic_S has only one colour under reflections.

In Figure 2, we present an example of domain ΩΩ\Omegaroman_Ω and covering which does not satisfy the second item of Definition 2.8.

A𝐴Aitalic_AB𝐵Bitalic_BC𝐶Citalic_C
(a) Domain ΩΩ\Omegaroman_Ω. The vertices of the triangle are such that A=2π/5𝐴2𝜋5\angle A=2\pi/5∠ italic_A = 2 italic_π / 5 and B=C=3π/10𝐵𝐶3𝜋10\angle B=\angle C=3\pi/10∠ italic_B = ∠ italic_C = 3 italic_π / 10.
A𝐴Aitalic_AB𝐵Bitalic_BC𝐶Citalic_C
(b) By reflecting Ω¯¯Ω\overline{\Omega}over¯ start_ARG roman_Ω end_ARG, we can fill the above pentagon.
A𝐴Aitalic_AB𝐵Bitalic_BC𝐶Citalic_CB2subscript𝐵2B_{2}italic_B start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPTC3subscript𝐶3C_{3}italic_C start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPTB4subscript𝐵4B_{4}italic_B start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPTC5subscript𝐶5C_{5}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT
(c) The second item of Definition 2.8 is not satisfied because we have two different colours on [AB]delimited-[]𝐴𝐵[AB][ italic_A italic_B ].
Figure 2: A domain ΩΩ\Omegaroman_Ω that is not a proper paving unit.
Proposition 2.9.

Assume that ΩdΩsuperscript𝑑\Omega\subset\mathbb{R}^{d}roman_Ω ⊂ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, d=2,3𝑑23d=2,3italic_d = 2 , 3, is a proper paving unit. Then when A𝐴Aitalic_A, q𝑞qitalic_q satisfy (8), we have 𝒮NS=𝒮TE=𝒮D𝒮Nsubscript𝒮NSsubscript𝒮TEsubscript𝒮Dsubscript𝒮N\mathscr{S}_{\mathrm{NS}}=\mathscr{S}_{\mathrm{TE}}=\mathscr{S}_{\mathrm{D}}% \cup\mathscr{S}_{\mathrm{N}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT = script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT = script_S start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT ∪ script_S start_POSTSUBSCRIPT roman_N end_POSTSUBSCRIPT.

Proof.

Let ΩΩ\Omegaroman_Ω be a proper paving unit. Consider one particular edge or face of ΩΩ\partial\Omega∂ roman_Ω and introduce some system of coordinates x,xdsuperscript𝑥subscript𝑥𝑑x^{\prime},x_{d}italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT with xd1superscript𝑥superscript𝑑1x^{\prime}\in\mathbb{R}^{d-1}italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT, such that this edge or face lies in the hypersurface {xd=0}subscript𝑥𝑑0\{x_{d}=0\}{ italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 0 }. Let Ω1subscriptΩ1\Omega_{1}roman_Ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT denote the reflection of ΩΩ\Omegaroman_Ω with respect to the line or plane {xd=0}subscript𝑥𝑑0\{x_{d}=0\}{ italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 0 }. Consider v𝑣vitalic_v an eigenfunction of the Dirichlet Laplacian in ΩΩ\Omegaroman_Ω. Classically we can extend it to Ω1subscriptΩ1\Omega_{1}roman_Ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT by defining v(x,xd)=v(x,xd)𝑣superscript𝑥subscript𝑥𝑑𝑣superscript𝑥subscript𝑥𝑑v(x^{\prime},x_{d})=-v(x^{\prime},-x_{d})italic_v ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) = - italic_v ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , - italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) for all (x,xd)Ω1superscript𝑥subscript𝑥𝑑subscriptΩ1(x^{\prime},x_{d})\in\Omega_{1}( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) ∈ roman_Ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. Similarly, Neumann eigenfunctions can be extended in Ω1subscriptΩ1\Omega_{1}roman_Ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT by defining v(x,xd)=v(x,xd)𝑣superscript𝑥subscript𝑥𝑑𝑣superscript𝑥subscript𝑥𝑑v(x^{\prime},x_{d})=v(x^{\prime},-x_{d})italic_v ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) = italic_v ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , - italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ). It can be verified straightforwardly that such extended function v𝑣vitalic_v satisfies Δv+k2v=0Δ𝑣superscript𝑘2𝑣0\Delta v+k^{2}v=0roman_Δ italic_v + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v = 0 in the whole interior of ΩΩ1¯¯ΩsubscriptΩ1\overline{\Omega\cup\Omega_{1}}over¯ start_ARG roman_Ω ∪ roman_Ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG. Repeating this reflection argument successively, we can extend Dirichlet/Neumann eigenfunctions to Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG as solutions to the homogeneous Helmholtz equation. The proof is completed. ∎

From classical results of interior regularity, this shows that in proper paving unit, Dirichlet and Neumann eigenfunctions are smooth up to the boundary because their extensions are real-analytic in Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG.

One can check that triangles that are equilateral (60°-60°-60°), hemiequilateral (30°-60°-90°) or isosceles right (45°-45°-90°) are all proper paving unit of 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. In fact, for these three types of triangles the Dirichlet/Neumann eigenfunctions can be expressed with trigonometric functions as discovered by G. Lamé (see [22]). Besides, observe that if ω𝜔\omegaitalic_ω is a proper paving unit of 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT then Ω=I×ωΩ𝐼𝜔\Omega=I\times\omegaroman_Ω = italic_I × italic_ω is a proper paving unit of 3superscript3\mathbb{R}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT for any bounded interval I𝐼Iitalic_I.

Finally, let us mention that for domains ω𝜔\omegaitalic_ω, with ω¯Ω~¯𝜔~Ω\overline{\omega}\subset\tilde{\Omega}over¯ start_ARG italic_ω end_ARG ⊂ over~ start_ARG roman_Ω end_ARG, obtained by “properly” reflecting a given proper paving unit ΩΩ\Omegaroman_Ω a finite number of times (see an illustration in 2D with Figure 3), 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT contains an unbounded sequence. Indeed, any eigenfunction of the Dirichlet/Neumann Laplacian in ΩΩ\Omegaroman_Ω extended to Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG according to the process described above is clearly an eigenfunction of the Dirichlet/Neumann Laplacian in ω𝜔\omegaitalic_ω.

Refer to caption
Refer to caption
Figure 3: Examples of domains for which 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT contains an unbounded sequence.

2.3 Combining solutions to the homogeneous Helmholtz equation

In the previous subsections, we presented canonical examples of geometries ΩΩ\Omegaroman_Ω where eigenfunctions of the Dirichlet/Neumann Laplacians can be extended as smooth solutions of the homogeneous Helmholtz equation in a neighborhood of ΩΩ\Omegaroman_Ω. Here, by working directly with solutions to the homogeneous Helmholtz equation, we provide examples of more general domains such that non-scattering frequencies exist. To keep notations simple, we work in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT only. The ideas can be applied to generate examples in 3superscript3\mathbb{R}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT.

2.3.1 Extendable Dirichlet eigenfunctions

Consider the function v0subscript𝑣0v_{0}italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT such that

v0(x,y)=sinxsiny.subscript𝑣0𝑥𝑦𝑥𝑦v_{0}(x,y)=\sin x\sin y.italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x , italic_y ) = roman_sin italic_x roman_sin italic_y .

It is a particular solution of the equation

Δv+2v=0 in 2.Δ𝑣2𝑣0 in superscript2\Delta v+2v=0\quad\mbox{ in }\mathbb{R}^{2}.roman_Δ italic_v + 2 italic_v = 0 in blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (10)

Now let us, roughly speaking, rotate this function by an angle α(0,2π)𝛼02𝜋\alpha\in(0,2\pi)italic_α ∈ ( 0 , 2 italic_π ) by defining vαsubscript𝑣𝛼v_{\alpha}italic_v start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT such that

vα(x,y)=v0(cos(α)xsin(α)y,sin(α)x+cos(α)y).subscript𝑣𝛼𝑥𝑦subscript𝑣0𝛼𝑥𝛼𝑦𝛼𝑥𝛼𝑦v_{\alpha}(x,y)=v_{0}(\cos(\alpha)x-\sin(\alpha)y,\sin(\alpha)x+\cos(\alpha)y).italic_v start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_x , italic_y ) = italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( roman_cos ( italic_α ) italic_x - roman_sin ( italic_α ) italic_y , roman_sin ( italic_α ) italic_x + roman_cos ( italic_α ) italic_y ) .

Note that vαsubscript𝑣𝛼v_{\alpha}italic_v start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT also solves (10). Then for λ𝜆\lambda\in\mathbb{R}italic_λ ∈ blackboard_R, set

vαλ=v0+λvαsubscriptsuperscript𝑣𝜆𝛼subscript𝑣0𝜆subscript𝑣𝛼v^{\lambda}_{\alpha}=v_{0}+\lambda\,v_{\alpha}italic_v start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_λ italic_v start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT

and define the nodal set 𝒩αλ:={(x,y)2|vαλ(x,y)=0}assignsubscriptsuperscript𝒩𝜆𝛼conditional-set𝑥𝑦superscript2subscriptsuperscript𝑣𝜆𝛼𝑥𝑦0\mathscr{N}^{\lambda}_{\alpha}:=\{(x,y)\in\mathbb{R}^{2}\,|\,v^{\lambda}_{% \alpha}(x,y)=0\}script_N start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT := { ( italic_x , italic_y ) ∈ blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_v start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_x , italic_y ) = 0 }. As classical in literature (see e.g. [13]), we call nodal domains of vαλsubscriptsuperscript𝑣𝜆𝛼v^{\lambda}_{\alpha}italic_v start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT the maximally connected subsets of 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for which vαλsubscriptsuperscript𝑣𝜆𝛼v^{\lambda}_{\alpha}italic_v start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT does not change sign. From Proposition 1.3 and Lemma 2.1, we obtain the following statement.

Proposition 2.10.

Any bounded nodal domain of vαλsubscriptsuperscript𝑣𝜆𝛼v^{\lambda}_{\alpha}italic_v start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is a domain where 2𝒮NS2subscript𝒮NS2\in\mathscr{S}_{\mathrm{NS}}2 ∈ script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT.

Let us make a few comments concerning this approach. First the choice k2=2superscript𝑘22k^{2}=2italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 2 is arbitrary here and we could consider any other k2>0superscript𝑘20k^{2}>0italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT > 0. On the other hand, we combined only two particular solutions of (10). We could have worked similarly with any other linear combinations of functions satisfying (10). To exhibit different solutions of (10), we can proceed to rotations of v0subscript𝑣0v_{0}italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT as we did. We can also translate v0subscript𝑣0v_{0}italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT or its rotated versions. This a priori offers a large variety of functions. A natural question then is “how rich is the family of corresponding bounded nodal domains?”. Finally, observe that we prove only the existence of one element in 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT and not that 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT contains an unbounded sequence as in the statements of the previous subsections.

In Figure 4, we display the nodal sets 𝒩αλsubscriptsuperscript𝒩𝜆𝛼\mathscr{N}^{\lambda}_{\alpha}script_N start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT for α=π/4𝛼𝜋4\alpha=\pi/4italic_α = italic_π / 4 and λ{1,0.2,0.5,2}𝜆10.20.52\lambda\in\{-1,-0.2,0.5,2\}italic_λ ∈ { - 1 , - 0.2 , 0.5 , 2 }. In each situation we observe that there exist bounded nodal domains. The question of proving the existence of bounded nodal domains would deserve to be studied in more details.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: Nodal sets 𝒩αλsubscriptsuperscript𝒩𝜆𝛼\mathscr{N}^{\lambda}_{\alpha}script_N start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT for α=π/4𝛼𝜋4\alpha=\pi/4italic_α = italic_π / 4 and, from left to right, λ=1,0.2,0.5,2𝜆10.20.52\lambda=-1,-0.2,0.5,2italic_λ = - 1 , - 0.2 , 0.5 , 2.

Interestingly, this technique can be exploited to exhibit domains ΩΩ\Omegaroman_Ω such that eigenfunctions of the Dirichlet Laplacian in ΩΩ\Omegaroman_Ω can be extended only to an open neighborhood of Ω¯¯Ω\overline{\Omega}over¯ start_ARG roman_Ω end_ARG but not to the whole 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. This has been found by J. Eckmann and C. Pillet in [15] and we reproduce their work here.

Given L𝐿superscriptL\in\mathbb{N}^{\ast}italic_L ∈ blackboard_N start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and μ+:=(0,+)𝜇subscriptassign0\mu\in\mathbb{R}_{+}:=(0,+\infty)italic_μ ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT := ( 0 , + ∞ ), we define the function v𝑣vitalic_v such that

v(r,θ)=l=0L1Jμ(kρl)cos(μψl).𝑣𝑟𝜃superscriptsubscript𝑙0𝐿1subscript𝐽𝜇𝑘subscript𝜌𝑙𝜇subscript𝜓𝑙v(r,\theta)=\sum_{l=0}^{L-1}J_{\mu}(k\rho_{l})\cos(\mu\psi_{l}).italic_v ( italic_r , italic_θ ) = ∑ start_POSTSUBSCRIPT italic_l = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L - 1 end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_k italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) roman_cos ( italic_μ italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) . (11)

Here (r,θ)𝑟𝜃(r,\theta)( italic_r , italic_θ ) stands for the polar coordinates with r>0𝑟0r>0italic_r > 0, θ(π,π]𝜃𝜋𝜋\theta\in(-\pi,\pi]italic_θ ∈ ( - italic_π , italic_π ], Jμsubscript𝐽𝜇J_{\mu}italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT is again the Bessel function of the first kind with order μ𝜇\muitalic_μ and ρl>0subscript𝜌𝑙0\rho_{l}>0italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT > 0, ψl(π,π]subscript𝜓𝑙𝜋𝜋\psi_{l}\in(-\pi,\pi]italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ∈ ( - italic_π , italic_π ] satisfy the relations

ρlcosψl=a+rcos(θ+2πl/L),ρlsinψl=rsin(θ+2πl/L),formulae-sequencesubscript𝜌𝑙subscript𝜓𝑙𝑎𝑟𝜃2𝜋𝑙𝐿subscript𝜌𝑙subscript𝜓𝑙𝑟𝜃2𝜋𝑙𝐿\rho_{l}\cos\psi_{l}=a+r\cos(\theta+2\pi l/L),\qquad\qquad\rho_{l}\sin\psi_{l}% =r\sin(\theta+2\pi l/L),italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT roman_cos italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = italic_a + italic_r roman_cos ( italic_θ + 2 italic_π italic_l / italic_L ) , italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT roman_sin italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = italic_r roman_sin ( italic_θ + 2 italic_π italic_l / italic_L ) ,

with a𝑎a\in\mathbb{R}italic_a ∈ blackboard_R. In particular, the term for l=0𝑙0l=0italic_l = 0 in (11) corresponds to the spherical wave function Jμ(kr)cos(μθ)subscript𝐽𝜇𝑘𝑟𝜇𝜃J_{\mu}(kr)\cos(\mu\theta)italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_k italic_r ) roman_cos ( italic_μ italic_θ ) with the center translated from (0,0)00(0,0)( 0 , 0 ) to (a,0)𝑎0(-a,0)( - italic_a , 0 ). If we further rotate the initial spherical wave function Jμ(kr)cos(μθ)subscript𝐽𝜇𝑘𝑟𝜇𝜃J_{\mu}(kr)\cos(\mu\theta)italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_k italic_r ) roman_cos ( italic_μ italic_θ ) clockwise by 2π/L2𝜋𝐿2\pi/L2 italic_π / italic_L (resp. 2πl/L2𝜋𝑙𝐿2\pi l/L2 italic_π italic_l / italic_L), we obtain the term in (11) for l=1𝑙1l=1italic_l = 1 (resp. l𝑙litalic_l in general). More generally, we can consider functions v𝑣vitalic_v of the form

v(r,θ)=l=0L1blJμl(kρl)cos(μl(ψlϕl)),𝑣𝑟𝜃superscriptsubscript𝑙0𝐿1subscript𝑏𝑙subscript𝐽subscript𝜇𝑙𝑘subscript𝜌𝑙subscript𝜇𝑙subscript𝜓𝑙subscriptitalic-ϕ𝑙v(r,\theta)=\sum_{l=0}^{L-1}b_{l}J_{\mu_{l}}(k\rho_{l})\cos\big{(}\mu_{l}(\psi% _{l}-\phi_{l})\big{)},italic_v ( italic_r , italic_θ ) = ∑ start_POSTSUBSCRIPT italic_l = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L - 1 end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_k italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) roman_cos ( italic_μ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) ) , (12)

with different Bessel orders μlsubscript𝜇𝑙\mu_{l}italic_μ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, rotational parameters ϕlsubscriptitalic-ϕ𝑙\phi_{l}italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, and weights blsubscript𝑏𝑙b_{l}italic_b start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, for each l=0,,L1𝑙0𝐿1l=0,\ldots,L-1italic_l = 0 , … , italic_L - 1. We can also impose different translational and rotational effects to each term by defining

ρlcosψl=al+rcos(θ+θl)andρlsinψl=rsin(θ+θl)formulae-sequencesubscript𝜌𝑙subscript𝜓𝑙subscript𝑎𝑙𝑟𝜃subscript𝜃𝑙andsubscript𝜌𝑙subscript𝜓𝑙𝑟𝜃subscript𝜃𝑙\rho_{l}\cos\psi_{l}=a_{l}+r\cos(\theta+\theta_{l})\qquad\text{and}\qquad\rho_% {l}\sin\psi_{l}=r\sin(\theta+\theta_{l})italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT roman_cos italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT + italic_r roman_cos ( italic_θ + italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) and italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT roman_sin italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = italic_r roman_sin ( italic_θ + italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT )

with different alsubscript𝑎𝑙a_{l}italic_a start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and θlsubscript𝜃𝑙\theta_{l}italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT for each l=0,,L1𝑙0𝐿1l=0,\ldots,L-1italic_l = 0 , … , italic_L - 1. Throughout this subsection, we shall let k𝑘kitalic_k be the first positive zero of Jμsubscript𝐽𝜇J_{\mu}italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT.

The following example, taken from [15], gives a bounded and simple connected domain ΩΩ\Omegaroman_Ω with analytic boundary where the Dirichlet Laplacian has an eigenfunction which is extendable in an open neighborhood of Ω¯¯Ω\overline{\Omega}over¯ start_ARG roman_Ω end_ARG but not to the whole 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

Example 2.11.

Let us consider the function v𝑣vitalic_v in (11) with μ=3/2𝜇32\mu=3/2italic_μ = 3 / 2, L=3𝐿3L=3italic_L = 3, a=0.6𝑎0.6a=0.6italic_a = 0.6. Since μ=3/2𝜇32\mu=3/2italic_μ = 3 / 2, the function θcos(μθ)maps-to𝜃𝜇𝜃\theta\mapsto\cos(\mu\theta)italic_θ ↦ roman_cos ( italic_μ italic_θ ) has a derivative which is not continuous at θ=π𝜃𝜋\theta=\piitalic_θ = italic_π and then v𝑣vitalic_v is real analytic only in 2Σsuperscript2Σ\mathbb{R}^{2}\setminus\Sigmablackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∖ roman_Σ with Σ:={(rcosθ,rsinθ)|r[a,),θ=π+2lπ/L,l=0,,2}assignΣconditional-set𝑟𝜃𝑟𝜃formulae-sequence𝑟𝑎formulae-sequence𝜃𝜋2𝑙𝜋𝐿𝑙02\Sigma:=\{(r\cos\theta,r\sin\theta)\,|\,r\in[a,\infty),\,\theta=-\pi+2l\pi/L,% \,l=0,\ldots,2\}roman_Σ := { ( italic_r roman_cos italic_θ , italic_r roman_sin italic_θ ) | italic_r ∈ [ italic_a , ∞ ) , italic_θ = - italic_π + 2 italic_l italic_π / italic_L , italic_l = 0 , … , 2 }. In particular, v𝑣vitalic_v is real analytic in the open ball Basubscript𝐵𝑎B_{a}italic_B start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT centred at the origin and of radius a𝑎aitalic_a. Let us explain how to prove that there is a bounded and simple-connected domain ΩΩ\Omegaroman_Ω with real-analytic boundary such that v𝑣vitalic_v is real-analytic in (an open neighborhood of) ΩΩ\Omegaroman_Ω and v=0𝑣0v=0italic_v = 0 on ΩΩ\partial\Omega∂ roman_Ω.
First notice that (x,y)J3/2(kr)cos(3θ/2)maps-to𝑥𝑦subscript𝐽32𝑘𝑟3𝜃2(x,y)\mapsto J_{3/2}(kr)\cos(3\theta/2)( italic_x , italic_y ) ↦ italic_J start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) roman_cos ( 3 italic_θ / 2 ) is positive in D1={(rcosθ,rsinθ)|r(0,1),θ(π/3,π/3)}subscript𝐷1conditional-set𝑟𝜃𝑟𝜃formulae-sequence𝑟01𝜃𝜋3𝜋3D_{1}=\{(r\cos\theta,r\sin\theta)\,|\,r\in(0,1),\theta\in(-\pi/3,\pi/3)\}italic_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = { ( italic_r roman_cos italic_θ , italic_r roman_sin italic_θ ) | italic_r ∈ ( 0 , 1 ) , italic_θ ∈ ( - italic_π / 3 , italic_π / 3 ) } and negative in B1D1¯subscript𝐵1¯subscript𝐷1B_{1}\setminus\overline{D_{1}}italic_B start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∖ over¯ start_ARG italic_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG (see Figure 5(a)). Using this property, we display in Figure 5(b) a yellow region where the three terms in the sum defining v𝑣vitalic_v are all positive and a blue region where these three terms are all negative. We note by Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG the green region in which the sign of v𝑣vitalic_v is a priori not clear. Looking at every direction θ(π,π]𝜃𝜋𝜋\theta\in(-\pi,\pi]italic_θ ∈ ( - italic_π , italic_π ], with the intermediate value theorem, we can show that there is (xθ,yθ)subscript𝑥𝜃subscript𝑦𝜃(x_{\theta},y_{\theta})( italic_x start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ) in Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG such that v(xθ,yθ)=0𝑣subscript𝑥𝜃subscript𝑦𝜃0v(x_{\theta},y_{\theta})=0italic_v ( italic_x start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ) = 0. Moreover, by exploiting that v𝑣vitalic_v is real analytic in Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG, we infer that there is a closed analytic curve ΓΩ~Γ~Ω\Gamma\subset\tilde{\Omega}roman_Γ ⊂ over~ start_ARG roman_Ω end_ARG such that v=0𝑣0v=0italic_v = 0 on ΓΓ\Gammaroman_Γ (see Figure 5(c) for a plot of ΓΓ\Gammaroman_Γ). Denote by ΩΩ\Omegaroman_Ω the bounded open set surrounded by ΓΓ\Gammaroman_Γ.
Then v𝑣vitalic_v is an eigenfunction of the Dirichlet Laplacian in ΩΩ\Omegaroman_Ω which can be analytically extended to Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG (actually to 2Σsuperscript2Σ\mathbb{R}^{2}\setminus\Sigmablackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∖ roman_Σ) as a solution to the Helmholtz equation. Because of this property, we conclude that we have 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}\neq\emptysetscript_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT ≠ ∅ for this ΩΩ\Omegaroman_Ω. Let us emphasize however that v𝑣vitalic_v can not be extended in the whole 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT analytically because its derivative with respect to θ𝜃\thetaitalic_θ is not continuous on ΣΣ\Sigmaroman_Σ.

Refer to caption
(a) Properties of (x,y)J3/2(kr)cos(3θ/2)maps-to𝑥𝑦subscript𝐽32𝑘𝑟3𝜃2(x,y)\mapsto J_{3/2}(kr)\cos(3\theta/2)( italic_x , italic_y ) ↦ italic_J start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) roman_cos ( 3 italic_θ / 2 ): positive in the yellow regions and negative in the blue regions.
Refer to caption
(b) Properties of v𝑣vitalic_v: in the yellow (resp. blue) region, the three terms in the sum defining v𝑣vitalic_v are positive (resp. negative).
Refer to caption
(c) Curves of constant value of the function v𝑣vitalic_v.
l
l
Figure 5: Illustration of the properties of the functions involved in Example 2.11.

We can modify Example 2.11 by taking different parameters in the definition of the function v𝑣vitalic_v appearing in (11). This allows us to exhibit other analytic domains ΩΩ\Omegaroman_Ω where the Dirichlet Laplacian has eigenfunctions which are extendable only in a neighborhood of ΩΩ\Omegaroman_Ω. In Figure 6, we display certain of these domains.

Refer to caption
(a) L=2𝐿2L=2italic_L = 2, μ=5/2𝜇52\mu=5/2italic_μ = 5 / 2, a=0.58𝑎0.58a=0.58italic_a = 0.58.
Refer to caption
(b) L=3𝐿3L=3italic_L = 3, μ=5/2𝜇52\mu=5/2italic_μ = 5 / 2, a=0.6𝑎0.6a=0.6italic_a = 0.6.
Refer to caption
(c) L=4𝐿4L=4italic_L = 4, μ=5/2𝜇52\mu=5/2italic_μ = 5 / 2, a=0.5𝑎0.5a=0.5italic_a = 0.5
Figure 6: Curves of constant value of the function v𝑣vitalic_v for different parameters in (11). In the analytic domains enclosed by the curve v=0𝑣0v=0italic_v = 0, the Dirichlet Laplacian has eigenfunctions which are only locally extendable.

Note that if we take the Bessel order μ𝜇\muitalic_μ in (11) in superscript\mathbb{N}^{\ast}blackboard_N start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, we can find domains ΩΩ\Omegaroman_Ω for which the eigenfunctions of the Dirichlet Laplacian can be extended in the whole 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. We present such examples in Figure 7. Let us emphasize that the domains ΩΩ\Omegaroman_Ω may have analytic boundaries as in Figure 7(a), or Lipschitz boundaries as in Figures 27(c). The latter situation appears when two or more nodal curves of the function v𝑣vitalic_v intersect. For example, in Figure 2, both v𝑣vitalic_v and v𝑣\nabla v∇ italic_v vanishes at the two points ±(a+1,0)plus-or-minus𝑎10\pm(a+1,0)± ( italic_a + 1 , 0 ), but v𝑣\mathcal{H}vcaligraphic_H italic_v, the Hessian of v𝑣vitalic_v, is different from zero at ±(a+1,0)plus-or-minus𝑎10\pm(a+1,0)± ( italic_a + 1 , 0 ). Hence there are two nodal curves intersecting at ±(a+1,0)plus-or-minus𝑎10\pm(a+1,0)± ( italic_a + 1 , 0 ). In Figure 7(c), at (0,2/2)022(0,-\sqrt{2}/2)( 0 , - square-root start_ARG 2 end_ARG / 2 ) we have αv=0superscript𝛼𝑣0\partial^{\alpha}v=0∂ start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_v = 0 for all multi-indices α2𝛼superscript2\alpha\in\mathbb{N}^{2}italic_α ∈ blackboard_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT with 0|α|20𝛼20\leq|\alpha|\leq 20 ≤ | italic_α | ≤ 2 but x3v0superscriptsubscript𝑥3𝑣0\partial_{x}^{3}v\neq 0∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_v ≠ 0. This explains why there are three nodal curves passing through (0,2/2)022(0,-\sqrt{2}/2)( 0 , - square-root start_ARG 2 end_ARG / 2 ).

Refer to caption
(a) L=2𝐿2L=2italic_L = 2, μ=1𝜇1\mu=1italic_μ = 1, a=0.55𝑎0.55a=0.55italic_a = 0.55.
22\sqrt{2}square-root start_ARG 2 end_ARG
Refer to caption
(b) L=2𝐿2L=2italic_L = 2, μ=1𝜇1\mu=1italic_μ = 1,
a=k2k2k𝑎subscript𝑘2𝑘2𝑘a=\frac{{k_{2}}-{k}}{2k}italic_a = divide start_ARG italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_k end_ARG start_ARG 2 italic_k end_ARG222Here k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT denotes the second positive zero of Jμsubscript𝐽𝜇J_{\mu}italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT., ϕ0=π/2=ϕ1subscriptitalic-ϕ0𝜋2subscriptitalic-ϕ1\phi_{0}=\pi/2=-\phi_{1}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_π / 2 = - italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.
Refer to caption
(c) L=2𝐿2L=2italic_L = 2, μ=1𝜇1\mu=1italic_μ = 1,
a=2/2𝑎22a=\sqrt{2}/2italic_a = square-root start_ARG 2 end_ARG / 2, ϕ0=π/4subscriptitalic-ϕ0𝜋4\phi_{0}=\pi/4italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_π / 4, ϕ1=3π/4subscriptitalic-ϕ13𝜋4\phi_{1}=3\pi/4italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 3 italic_π / 4.
Figure 7: Curves of constant value of the function v𝑣vitalic_v for different parameters in (11). Here we get analytic and Lipschitz domains where the Dirichlet Laplacian has eigenfunctions which extend to solutions of the Helmholtz equation in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

Similarly, we can find Lipschitz domains ΩΩ\Omegaroman_Ω for which there exist eigenfunctions are the Dirichlet Laplacian which can be extended in a neighborhood of Ω¯¯Ω\overline{\Omega}over¯ start_ARG roman_Ω end_ARG but not to the whole 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (see Figure 8). We refer the reader to the Appendix for the justification of existence of such nodal curves as shown in Figures 68.

Refer to caption
(a) L=2𝐿2L=2italic_L = 2, μ=7/2𝜇72\mu=7/2italic_μ = 7 / 2, a=cos(π/7)𝑎𝜋7a=\cos(\pi/7)italic_a = roman_cos ( italic_π / 7 ), b0=1=b1subscript𝑏01subscript𝑏1b_{0}=1=-b_{1}italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 = - italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.
Refer to caption
(b) L=2𝐿2L=2italic_L = 2, μ=5/2𝜇52\mu=5/2italic_μ = 5 / 2, a=1𝑎1a=1italic_a = 1, ϕ0=π/5=ϕ1subscriptitalic-ϕ0𝜋5subscriptitalic-ϕ1\phi_{0}=\pi/5=-\phi_{1}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_π / 5 = - italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.
Refer to caption
(c) L=2𝐿2L=2italic_L = 2, μ=3/2𝜇32\mu=3/2italic_μ = 3 / 2, a=2/2𝑎22a=\sqrt{2}/2italic_a = square-root start_ARG 2 end_ARG / 2, ϕ0=π/12=ϕ1subscriptitalic-ϕ0𝜋12subscriptitalic-ϕ1\phi_{0}=\pi/{12}=-\phi_{1}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_π / 12 = - italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.
Figure 8: Curves of constant value of the function v𝑣vitalic_v for different parameters in (11). Here we get Lipschitz domains where the Dirichlet Laplacian has eigenfunctions which are only locally extendable.

We conclude this paragraph with an important result which generalizes the constraint we have in Proposition 2.5 concerning the angle of a sector so that non-scattering frequencies can exist.

Let Ω be a domain such that the Dirichlet Laplacian has an eigenfunction v thatcan be extended into a function v~ satisfying Δv~+k2v~=0 in a neighborhood of Ω¯.Let Ω be a domain such that the Dirichlet Laplacian has an eigenfunction v thatcan be extended into a function v~ satisfying Δv~+k2v~=0 in a neighborhood of Ω¯.\begin{array}[]{|l}\mbox{Let $\Omega$ be a domain such that the Dirichlet % Laplacian has an eigenfunction $v$ that}\\ \mbox{can be extended into a function $\tilde{v}$ satisfying $\Delta\tilde{v}+% k^{2}\tilde{v}=0$ in a neighborhood of $\overline{\Omega}$.}\end{array}start_ARRAY start_ROW start_CELL Let roman_Ω be a domain such that the Dirichlet Laplacian has an eigenfunction italic_v that end_CELL end_ROW start_ROW start_CELL can be extended into a function over~ start_ARG italic_v end_ARG satisfying roman_Δ over~ start_ARG italic_v end_ARG + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over~ start_ARG italic_v end_ARG = 0 in a neighborhood of over¯ start_ARG roman_Ω end_ARG . end_CELL end_ROW end_ARRAY (13)

First, for ΩΩ\Omegaroman_Ω as in (13), due to the implicit functions theorem, ΩΩ\partial\Omega∂ roman_Ω must be piecewise analytic. Additionally, we have the following statement (see also [13, Chapter V, §16]).

Proposition 2.12.

If PΩ𝑃ΩP\in\partial\Omegaitalic_P ∈ ∂ roman_Ω is a corner point of a domain ΩΩ\Omegaroman_Ω satisfying (13) with opening angle α𝛼\alphaitalic_α, then necessarily α(0,2π)π𝛼02𝜋𝜋\alpha\in(0,2\pi)\cap\mathbb{Q}\piitalic_α ∈ ( 0 , 2 italic_π ) ∩ blackboard_Q italic_π.

Proof.

We start by observing that ΩΩ\partial\Omega∂ roman_Ω belongs to the nodal set of v𝑣vitalic_v, an eigenfunction of the Dirichlet Laplacian in ΩΩ\Omegaroman_Ω. In an adapted system of coordinates, we can assume that P𝑃Pitalic_P coincides with the origin O𝑂Oitalic_O. By the assumption that v𝑣vitalic_v satisfies the Helmholtz equation in an open ball around O𝑂Oitalic_O, we obtain that v𝑣vitalic_v is of the form

v(x,y)=cNJN(kr)cos(Nθθ0)+fN+1(x,y),𝑣𝑥𝑦subscript𝑐𝑁subscript𝐽𝑁𝑘𝑟𝑁𝜃subscript𝜃0subscript𝑓𝑁1𝑥𝑦v(x,y)=c_{N}J_{N}(kr)\cos(N\theta-\theta_{0})+f_{N+1}(x,y),italic_v ( italic_x , italic_y ) = italic_c start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_k italic_r ) roman_cos ( italic_N italic_θ - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + italic_f start_POSTSUBSCRIPT italic_N + 1 end_POSTSUBSCRIPT ( italic_x , italic_y ) ,

near O𝑂Oitalic_O, where (r,θ)𝑟𝜃(r,\theta)( italic_r , italic_θ ) are the polar coordinates of (x,y)𝑥𝑦(x,y)( italic_x , italic_y ). Here cN\{0}subscript𝑐𝑁\0c_{N}\in\mathbb{R}\backslash\{0\}italic_c start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ∈ blackboard_R \ { 0 } and θ0[0,2π)subscript𝜃002𝜋\theta_{0}\in[0,2\pi)italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ [ 0 , 2 italic_π ) are constants depending on v𝑣vitalic_v, N𝑁superscriptN\in\mathbb{N}^{\ast}italic_N ∈ blackboard_N start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT is the vanishing order of v𝑣vitalic_v at P𝑃Pitalic_P and fN+1=O(rN+1)subscript𝑓𝑁1𝑂superscript𝑟𝑁1f_{N+1}=O(r^{N+1})italic_f start_POSTSUBSCRIPT italic_N + 1 end_POSTSUBSCRIPT = italic_O ( italic_r start_POSTSUPERSCRIPT italic_N + 1 end_POSTSUPERSCRIPT ) is a real-analytic function in (x,y)𝑥𝑦(x,y)( italic_x , italic_y ). As a consequence, the Taylor series of v𝑣vitalic_v near O𝑂Oitalic_O is of the form

v(x,y)=c~NrNcos(Nθθ0)+f~N+1(x,y),𝑣𝑥𝑦subscript~𝑐𝑁superscript𝑟𝑁𝑁𝜃subscript𝜃0subscript~𝑓𝑁1𝑥𝑦v(x,y)=\tilde{c}_{N}r^{N}\cos(N\theta-\theta_{0})+\tilde{f}_{N+1}(x,y),italic_v ( italic_x , italic_y ) = over~ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_cos ( italic_N italic_θ - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + over~ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_N + 1 end_POSTSUBSCRIPT ( italic_x , italic_y ) ,

with c~N0subscript~𝑐𝑁0\tilde{c}_{N}\neq 0over~ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ≠ 0 and f~N+1subscript~𝑓𝑁1\tilde{f}_{N+1}over~ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_N + 1 end_POSTSUBSCRIPT real-analytic. Hence, there are exactly N𝑁Nitalic_N analytic nodal curves of v𝑣vitalic_v intersecting at P𝑃Pitalic_P, with tangential directions along angles in the set {(π/2+κπ+θ0)/N|κ}conditional-set𝜋2𝜅𝜋subscript𝜃0𝑁𝜅\{(\pi/2+\kappa\pi+\theta_{0})/N\,|\,\kappa\in\mathbb{Z}\}{ ( italic_π / 2 + italic_κ italic_π + italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_N | italic_κ ∈ blackboard_Z }. Therefore we see that the angle between two tangential directions is necessarily equal to κπ/Nsuperscript𝜅𝜋𝑁\kappa^{\prime}\pi/Nitalic_κ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_π / italic_N for some κsuperscript𝜅\kappa^{\prime}\in\mathbb{N}italic_κ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ blackboard_N. ∎

2.3.2 Extendable Neumann eigenfunctions

What we did in §2.3.1 to find domains ΩΩ\Omegaroman_Ω where the Dirichlet Laplacian admits extendable eigenfunctions can not be directly applied for Neumann boundary conditions. Instead of looking for nodal curves, we should consider the characteristics associated with the gradient of solutions to the Helmholtz equation. Let us explain in the following.

Start with some non-zero real valued function v𝑣vitalic_v satisfying Δv+k2v=0Δ𝑣superscript𝑘2𝑣0\Delta v+k^{2}v=0roman_Δ italic_v + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v = 0 in some domain, say 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for simplicity, for a given k>0𝑘0k>0italic_k > 0. Pick some point (x0,y0)subscript𝑥0subscript𝑦0(x_{0},y_{0})( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) such that v(x0,y0)0𝑣subscript𝑥0subscript𝑦00\nabla v(x_{0},y_{0})\neq 0∇ italic_v ( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ≠ 0. Then consider the gradient system

X(t)=v(X(t))X(0)=(x0,y0).superscript𝑋𝑡𝑣𝑋𝑡𝑋0subscript𝑥0subscript𝑦0\begin{array}[]{|rcl}X^{\prime}(t)&=&\nabla v(X(t))\\[3.0pt] X(0)&=&(x_{0},y_{0}).\end{array}start_ARRAY start_ROW start_CELL italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) end_CELL start_CELL = end_CELL start_CELL ∇ italic_v ( italic_X ( italic_t ) ) end_CELL end_ROW start_ROW start_CELL italic_X ( 0 ) end_CELL start_CELL = end_CELL start_CELL ( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) . end_CELL end_ROW end_ARRAY (14)

The Cauchy-Lipschitz theorem ensures that (14) admits a unique maximal solution X𝑋Xitalic_X defined on an interval I𝐼I\subset\mathbb{R}italic_I ⊂ blackboard_R containing 00. Set Γ:={X(t)|tI}2assignΓconditional-set𝑋𝑡𝑡𝐼superscript2\Gamma:=\{X(t)\,|\,t\in I\}\subset\mathbb{R}^{2}roman_Γ := { italic_X ( italic_t ) | italic_t ∈ italic_I } ⊂ blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and introduce ν𝜈\nuitalic_ν a unit normal vector to the smooth curve ΓΓ\Gammaroman_Γ whose orientation is arbitrarily fixed. For tI𝑡𝐼t\in Iitalic_t ∈ italic_I, we have

νv(X(t))=ν(X(t))v(X(t))=ν(X(t))X(t)=0subscript𝜈𝑣𝑋𝑡𝜈𝑋𝑡𝑣𝑋𝑡𝜈𝑋𝑡superscript𝑋𝑡0\partial_{\nu}v(X(t))=\nu(X(t))\cdot\nabla v(X(t))=\nu(X(t))\cdot X^{\prime}(t% )=0∂ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_v ( italic_X ( italic_t ) ) = italic_ν ( italic_X ( italic_t ) ) ⋅ ∇ italic_v ( italic_X ( italic_t ) ) = italic_ν ( italic_X ( italic_t ) ) ⋅ italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) = 0

because X(t)superscript𝑋𝑡X^{\prime}(t)italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) and ν(X(t))𝜈𝑋𝑡\nu(X(t))italic_ν ( italic_X ( italic_t ) ) are respectively tangent and normal to the orbit tX(t)maps-to𝑡𝑋𝑡t\mapsto X(t)italic_t ↦ italic_X ( italic_t ). Thus v𝑣vitalic_v satisfies an homogeneous Neumann boundary condition on ΓΓ\Gammaroman_Γ. Therefore if we are able to find a bounded domain ΩΩ\Omegaroman_Ω whose boundary coincides with the union of orbits of Problem (14) for different initial conditions (x0,y0)subscript𝑥0subscript𝑦0(x_{0},y_{0})( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), then this shows that k𝑘kitalic_k belongs to the set 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT associated with ΩΩ\Omegaroman_Ω. Let us illustrate the approach with some specific examples.

Example 2.13.

Consider the function

v(x,y)=(cosx+cosy)𝑣𝑥𝑦𝑥𝑦v(x,y)=-(\cos x+\cos y)italic_v ( italic_x , italic_y ) = - ( roman_cos italic_x + roman_cos italic_y ) (15)

which satisfies Δv+2v=0Δ𝑣2𝑣0\Delta v+2v=0roman_Δ italic_v + 2 italic_v = 0 in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. For this v𝑣vitalic_v, (14) reads

x(t)=sinx(t)y(t)=siny(t)(x(0),y(0))=(x0,y0).superscript𝑥𝑡𝑥𝑡superscript𝑦𝑡𝑦𝑡𝑥0𝑦0subscript𝑥0subscript𝑦0\begin{array}[]{|rcl}x^{\prime}(t)&=&\sin x(t)\\[3.0pt] y^{\prime}(t)&=&\sin y(t)\\[3.0pt] (x(0),y(0))&=&(x_{0},y_{0}).\end{array}start_ARRAY start_ROW start_CELL italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) end_CELL start_CELL = end_CELL start_CELL roman_sin italic_x ( italic_t ) end_CELL end_ROW start_ROW start_CELL italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) end_CELL start_CELL = end_CELL start_CELL roman_sin italic_y ( italic_t ) end_CELL end_ROW start_ROW start_CELL ( italic_x ( 0 ) , italic_y ( 0 ) ) end_CELL start_CELL = end_CELL start_CELL ( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) . end_CELL end_ROW end_ARRAY (16)

The stationary points of this system of ordinary differential equations are the (mπ,nπ)𝑚𝜋𝑛𝜋(m\pi,n\pi)( italic_m italic_π , italic_n italic_π ), m,n𝑚𝑛m,n\in\mathbb{Z}italic_m , italic_n ∈ blackboard_Z. These are constant orbits. Moreover, one can check that

{0}×(0,π),{π}×(0,π),(0,π)×{0},(0,π)×{π}00𝜋𝜋0𝜋0𝜋00𝜋𝜋\{0\}\times(0,\pi),\quad\{\pi\}\times(0,\pi),\quad(0,\pi)\times\{0\},\quad(0,% \pi)\times\{\pi\}{ 0 } × ( 0 , italic_π ) , { italic_π } × ( 0 , italic_π ) , ( 0 , italic_π ) × { 0 } , ( 0 , italic_π ) × { italic_π }

are also orbits of (16). Now pick some (x0,y0)(0,π)2subscript𝑥0subscript𝑦0superscript0𝜋2(x_{0},y_{0})\in(0,\pi)^{2}( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ∈ ( 0 , italic_π ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Since different orbits cannot cross, we know that the one passing through (x0,y0)subscript𝑥0subscript𝑦0(x_{0},y_{0})( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) stays inside (0,π)2superscript0𝜋2(0,\pi)^{2}( 0 , italic_π ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and so the associated trajectory is global, i.e. defined for all t𝑡t\in\mathbb{R}italic_t ∈ blackboard_R. It turns out that we can compute it explicitly and we find

x(t)=2arctan(ettanx02),y(t)=2arctan(ettany02).formulae-sequence𝑥𝑡2superscript𝑒𝑡subscript𝑥02𝑦𝑡2superscript𝑒𝑡subscript𝑦02x(t)=2\arctan(e^{t}\tan\frac{x_{0}}{2}),\qquad\quad y(t)=2\arctan(e^{t}\tan% \frac{y_{0}}{2}).italic_x ( italic_t ) = 2 roman_arctan ( italic_e start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT roman_tan divide start_ARG italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) , italic_y ( italic_t ) = 2 roman_arctan ( italic_e start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT roman_tan divide start_ARG italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) .

After a few operations, we obtain that the corresponding orbit coincides with the curve

Γ(δ):={(x,y(x))=2arctan(δtan(x/2)))|x(0,π)}\Gamma(\delta):=\{(x,y(x))=2\arctan(\delta\tan(x/2)))\,|\,x\in(0,\pi)\}roman_Γ ( italic_δ ) := { ( italic_x , italic_y ( italic_x ) ) = 2 roman_arctan ( italic_δ roman_tan ( italic_x / 2 ) ) ) | italic_x ∈ ( 0 , italic_π ) }

where δ:=tan(y0/2)tan(x0/2)>0assign𝛿subscript𝑦02subscript𝑥020\delta:=\tan(y_{0}/2)\tan(x_{0}/2)>0italic_δ := roman_tan ( italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 ) roman_tan ( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 ) > 0 (see Figure 9 left for representations of different Γ(δ)Γ𝛿\Gamma(\delta)roman_Γ ( italic_δ )). For any 0<δ1<δ20subscript𝛿1subscript𝛿20<\delta_{1}<\delta_{2}0 < italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, denote Ω(δ1,δ2)Ωsubscript𝛿1subscript𝛿2\Omega(\delta_{1},\delta_{2})roman_Ω ( italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) the domain located between Γ(δ1)Γsubscript𝛿1\Gamma(\delta_{1})roman_Γ ( italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) and Γ(δ2)Γsubscript𝛿2\Gamma(\delta_{2})roman_Γ ( italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ). Then v𝑣vitalic_v is an eigenfunction of the Neumann Laplacian in Ω(δ1,δ2)Ωsubscript𝛿1subscript𝛿2\Omega(\delta_{1},\delta_{2})roman_Ω ( italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (see Figure 9 right for pictures of some Ω(δ1,δ2)Ωsubscript𝛿1subscript𝛿2\Omega(\delta_{1},\delta_{2})roman_Ω ( italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT )). Since v𝑣vitalic_v satisfies Δv+2v=0Δ𝑣2𝑣0\Delta v+2v=0roman_Δ italic_v + 2 italic_v = 0 in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, we deduce that 2222 belongs to 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT for this domain.
Note that we also have 2𝒮NS2subscript𝒮NS2\in\mathscr{S}_{\mathrm{NS}}2 ∈ script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT for the domain delimited for example by the curves

{0}×(0,π),{π}×(0,π),Γ(δ),for any δ>0.00𝜋𝜋0𝜋Γ𝛿for any δ>0\{0\}\times(0,\pi),\qquad\{\pi\}\times(0,\pi),\qquad\Gamma(\delta),\qquad\mbox% {for any $\delta>0$}.{ 0 } × ( 0 , italic_π ) , { italic_π } × ( 0 , italic_π ) , roman_Γ ( italic_δ ) , for any italic_δ > 0 .
Refer to caption

Refer to caption Refer to caption Refer to caption Refer to caption

Figure 9: Left: curves Γ(δ)Γ𝛿\Gamma(\delta)roman_Γ ( italic_δ ) for δ=0.5,1,4𝛿0.514\delta=0.5,1,4italic_δ = 0.5 , 1 , 4. Right: function v𝑣vitalic_v defined in (15) in a few domains where it is an eigenfunction of the Neumann Laplacian associated with the eigenvalue 2222.

At this stage, let us make a few comments. First, when one enlarges a bounded domain, due to the min-max principle, the eigenvalues of the Dirichlet Laplacian decrease. This is not true for the eigenvalues of the Neumann Laplacian, roughly speaking because one cannot extend them by zero. This is well illustrated by the above example. We have a family of continously deformed shapes which all have 2222 among the eigenvalues of the Neumann Laplacian.

Second, this example shows that one can find domains where the Neumann Laplacian has eigenfunctions which extend as solutions to the Helmholtz solution in the whole space and which admit corners with arbitrary apertures in (0,2π)02𝜋(0,2\pi)( 0 , 2 italic_π ). This is very different from the Dirichlet case where in order to be able to extend the eigenfunctions of the Laplace operator, the aperture of the corners must be in (0,2π)π02𝜋𝜋(0,2\pi)\cap\mathbb{Q}\pi( 0 , 2 italic_π ) ∩ blackboard_Q italic_π (see Proposition 2.12).

The reason we can have such domains ΩΩ\Omegaroman_Ω with corners of arbitrary aperture in Example 2.13 is that v(O)𝑣𝑂\mathcal{H}v(O)caligraphic_H italic_v ( italic_O ), the Hessian of v𝑣vitalic_v at the stationary point O=(0,0)𝑂00O=(0,0)italic_O = ( 0 , 0 ), is equal to the identity matrix. As a consequence, the linearized problem of (16) at O𝑂Oitalic_O simply writes U(t)=U(t)superscript𝑈𝑡𝑈𝑡U^{\prime}(t)=U(t)italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) = italic_U ( italic_t ) with U=(Ux,Uy)𝑈superscriptsubscript𝑈𝑥subscript𝑈𝑦topU=(U_{x},U_{y})^{\top}italic_U = ( italic_U start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_U start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT. Solving it, we get Ux(t)=axetsubscript𝑈𝑥𝑡subscript𝑎𝑥superscript𝑒𝑡U_{x}(t)=a_{x}\,e^{t}italic_U start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_t ) = italic_a start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT, Uy(t)=ayetsubscript𝑈𝑦𝑡subscript𝑎𝑦superscript𝑒𝑡U_{y}(t)=a_{y}\,e^{t}italic_U start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ( italic_t ) = italic_a start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT for some constants axsubscript𝑎𝑥a_{x}italic_a start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT, aysubscript𝑎𝑦a_{y}italic_a start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT. Thus we see that the curve {(Ux(t),Uy(t)),t(;0)}subscript𝑈𝑥𝑡subscript𝑈𝑦𝑡𝑡0\{(U_{x}(t),U_{y}(t)),\,t\in(-\infty;0)\}{ ( italic_U start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_t ) , italic_U start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ( italic_t ) ) , italic_t ∈ ( - ∞ ; 0 ) } is a line which can take any direction by choosing properly axsubscript𝑎𝑥a_{x}italic_a start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT, aysubscript𝑎𝑦a_{y}italic_a start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT. This explains why the orbits of (16) can leave from O𝑂Oitalic_O asymptotically with any angle.

Let us consider a second example.

Example 2.14.

Start from the function v𝑣vitalic_v such that v(x,y)=cosxcos2y𝑣𝑥𝑦𝑥2𝑦v(x,y)=\cos x\cos 2yitalic_v ( italic_x , italic_y ) = roman_cos italic_x roman_cos 2 italic_y. It satisfies Δv+5v=0Δ𝑣5𝑣0\Delta v+5v=0roman_Δ italic_v + 5 italic_v = 0 in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The set of stationary points associated with (14) is {(mπ,nπ/2),(π/2+mπ,π/4+nπ/2)|m,n}conditional-set𝑚𝜋𝑛𝜋2𝜋2𝑚𝜋𝜋4𝑛𝜋2𝑚𝑛\{(m\pi,n\pi/2),(\pi/2+m\pi,\pi/4+n\pi/2)\,|\,m,n\in\mathbb{Z}\}{ ( italic_m italic_π , italic_n italic_π / 2 ) , ( italic_π / 2 + italic_m italic_π , italic_π / 4 + italic_n italic_π / 2 ) | italic_m , italic_n ∈ blackboard_Z }. By computing the Hessian v𝑣\mathcal{H}vcaligraphic_H italic_v, one finds that the (π/2+mπ,π/4+nπ/2)𝜋2𝑚𝜋𝜋4𝑛𝜋2(\pi/2+m\pi,\pi/4+n\pi/2)( italic_π / 2 + italic_m italic_π , italic_π / 4 + italic_n italic_π / 2 ) are saddle points. At saddle points, orbits of (14) can meet the equilibrium point at only two particular directions (see Figure 10(b)). On the other hand, for m,n𝑚𝑛m,n\in\mathbb{Z}italic_m , italic_n ∈ blackboard_Z, we have v(mπ,nπ/2)=(1)m+n1diag(1,4)𝑣𝑚𝜋𝑛𝜋2superscript1𝑚𝑛1diag14\mathcal{H}v(m\pi,n\pi/2)=(-1)^{m+n-1}\mathrm{diag}(1,4)caligraphic_H italic_v ( italic_m italic_π , italic_n italic_π / 2 ) = ( - 1 ) start_POSTSUPERSCRIPT italic_m + italic_n - 1 end_POSTSUPERSCRIPT roman_diag ( 1 , 4 ). Hence the orbits of (14) are all tangential to the horizontal line at (0,0)00(0,0)( 0 , 0 ) except one that is tangential to the vertical line. Indeed, for example for m=n=0𝑚𝑛0m=n=0italic_m = italic_n = 0, the linearized problem associated with (14) writes U(t)=diag(1,4)U(t)superscript𝑈𝑡diag14𝑈𝑡U^{\prime}(t)=-\mathrm{diag}(1,4)\cdot U(t)italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) = - roman_diag ( 1 , 4 ) ⋅ italic_U ( italic_t ) so that its solution is given by Ux(t)=axetsubscript𝑈𝑥𝑡subscript𝑎𝑥superscript𝑒𝑡U_{x}(t)=a_{x}\,e^{-t}italic_U start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_t ) = italic_a start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_t end_POSTSUPERSCRIPT, Uy(t)=aye4tsubscript𝑈𝑦𝑡subscript𝑎𝑦superscript𝑒4𝑡U_{y}(t)=a_{y}\,e^{-4t}italic_U start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ( italic_t ) = italic_a start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - 4 italic_t end_POSTSUPERSCRIPT for some constants axsubscript𝑎𝑥a_{x}italic_a start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT, aysubscript𝑎𝑦a_{y}italic_a start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT. Thus we see that except when ax=0subscript𝑎𝑥0a_{x}=0italic_a start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = 0, the curve {(Ux(t),Uy(t)),t(0;+)}subscript𝑈𝑥𝑡subscript𝑈𝑦𝑡𝑡0\{(U_{x}(t),U_{y}(t)),\,t\in(0;+\infty)\}{ ( italic_U start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_t ) , italic_U start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ( italic_t ) ) , italic_t ∈ ( 0 ; + ∞ ) } is tangential to the horizontal line when t+𝑡t\to+\inftyitalic_t → + ∞.
We can make this more explicitly by computing the orbits. First we find that

(0,0),(π,0),(π/2,0),(π/2,π),(0,π)×{0},{π}×(0,π/2),(0,π)×{π/2},{0}×(0,π/2)00𝜋0𝜋20𝜋2𝜋0𝜋0𝜋0𝜋20𝜋𝜋200𝜋2(0,0),\quad(\pi,0),\quad(\pi/2,0),\quad(\pi/2,\pi),\quad(0,\pi)\times\{0\},% \quad\{\pi\}\times(0,\pi/2),\quad(0,\pi)\times\{\pi/2\},\quad\{0\}\times(0,\pi% /2)( 0 , 0 ) , ( italic_π , 0 ) , ( italic_π / 2 , 0 ) , ( italic_π / 2 , italic_π ) , ( 0 , italic_π ) × { 0 } , { italic_π } × ( 0 , italic_π / 2 ) , ( 0 , italic_π ) × { italic_π / 2 } , { 0 } × ( 0 , italic_π / 2 )

are particular orbits of (14). Then we obtain that the (global) trajectory (x(t),y(t))𝑥𝑡𝑦𝑡(x(t),y(t))( italic_x ( italic_t ) , italic_y ( italic_t ) ) passing through (x0,y0)(0,π)×(0,π/2)subscript𝑥0subscript𝑦00𝜋0𝜋2(x_{0},y_{0})\in(0,\pi)\times(0,\pi/2)( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ∈ ( 0 , italic_π ) × ( 0 , italic_π / 2 ) is such that

sin(2y(t))=δsin4x(t)2𝑦𝑡𝛿superscript4𝑥𝑡\sin(2y(t))=\delta\sin^{4}x(t)roman_sin ( 2 italic_y ( italic_t ) ) = italic_δ roman_sin start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_x ( italic_t )

for δ>0𝛿0\delta>0italic_δ > 0. We show in Figure 10 the normalized gradient flow and some of the “Neumann curves”. The latter enclose domains ΩΩ\Omegaroman_Ω for which v𝑣vitalic_v is a Neumann eigenfunction so that 5555 belongs to 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT. We get that ΩΩ\Omegaroman_Ω can be of class 𝒞1superscript𝒞1\mathscr{C}^{1}script_C start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, Lipschitz but may also possess cusps.

Refer to caption
(a) Normalized gradient v/|v|𝑣𝑣\nabla v/|\nabla v|∇ italic_v / | ∇ italic_v |.

Refer to caption

(b) Neumann curves.
Figure 10: Case v(x,y)=cosxcos2y𝑣𝑥𝑦𝑥2𝑦v(x,y)=\cos x\cos 2yitalic_v ( italic_x , italic_y ) = roman_cos italic_x roman_cos 2 italic_y.
Example 2.15 (Neumann eigenfunctions extendable in a neighborhood of Ω¯¯Ω\overline{\Omega}over¯ start_ARG roman_Ω end_ARG but not to the whole dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT).

Let us defined v𝑣vitalic_v as in (12) with L=2𝐿2L=2italic_L = 2, μ=7/2𝜇72\mu=7/2italic_μ = 7 / 2, a=cos(π/7)𝑎𝜋7a=\cos(\pi/7)italic_a = roman_cos ( italic_π / 7 ) and b0=1=b1subscript𝑏01subscript𝑏1b_{0}=1=-b_{1}italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 = - italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, with k𝑘kitalic_k being the first positive zero of Jμsubscript𝐽𝜇J_{\mu}italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT (see Figure 8(a)). We can calculate the gradient v𝑣\nabla v∇ italic_v explicitly, say, when |x1|<asubscript𝑥1𝑎|x_{1}|<a| italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | < italic_a, as

v(x,y)=l=01(1)l(kJμ(kρl)cos(μψl)ρlμJμ(kρl)sin(μψl)ψl),𝑣𝑥𝑦superscriptsubscript𝑙01superscript1𝑙𝑘subscriptsuperscript𝐽𝜇𝑘subscript𝜌𝑙𝜇subscript𝜓𝑙subscript𝜌𝑙𝜇subscript𝐽𝜇𝑘subscript𝜌𝑙𝜇subscript𝜓𝑙subscript𝜓𝑙\nabla v(x,y)=\sum_{l=0}^{1}(-1)^{l}\Big{(}kJ^{\prime}_{\mu}(k\rho_{l})\cos(% \mu\psi_{l})\nabla\rho_{l}-\mu J_{\mu}(k\rho_{l})\sin(\mu\psi_{l})\nabla\psi_{% l}\Big{)},∇ italic_v ( italic_x , italic_y ) = ∑ start_POSTSUBSCRIPT italic_l = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT ( italic_k italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_k italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) roman_cos ( italic_μ italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) ∇ italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_μ italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_k italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) roman_sin ( italic_μ italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) ∇ italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) ,

with

ρl=(a+(1)lx1)2+x22,andψl=(1)larcsinx2ρl,l=0,1.formulae-sequencesubscript𝜌𝑙superscript𝑎superscript1𝑙subscript𝑥12superscriptsubscript𝑥22andformulae-sequencesubscript𝜓𝑙superscript1𝑙subscript𝑥2subscript𝜌𝑙𝑙01\rho_{l}=\sqrt{\left(a+(-1)^{l}x_{1}\right)^{2}+x_{2}^{2}},\qquad\text{and}% \qquad\psi_{l}=(-1)^{l}\arcsin\frac{x_{2}}{\rho_{l}},\qquad l=0,1.italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = square-root start_ARG ( italic_a + ( - 1 ) start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , and italic_ψ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = ( - 1 ) start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT roman_arcsin divide start_ARG italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_ARG , italic_l = 0 , 1 .

It can be shown that v=0𝑣0\nabla v=0∇ italic_v = 0 at (0,±sin(π/7))0plus-or-minus𝜋7(0,\pm\sin(\pi/7))( 0 , ± roman_sin ( italic_π / 7 ) ) and (±t0,0)plus-or-minussubscript𝑡00(\pm t_{0},0)( ± italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , 0 ), where t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the (only) solution to Jμ(k(a+t0))+Jμ(k(at0))=0superscriptsubscript𝐽𝜇𝑘𝑎subscript𝑡0superscriptsubscript𝐽𝜇𝑘𝑎subscript𝑡00J_{\mu}^{\prime}\left(k(a+t_{0})\right)+J_{\mu}^{\prime}\left(k(a-t_{0})\right% )=0italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k ( italic_a + italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ) + italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k ( italic_a - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ) = 0 in the interval (0,a)0𝑎(0,a)( 0 , italic_a ). The normalized gradient flow is shown in Figure 11(a). Moreover, by (numerically) solving the system of ODEs (14), we can construct curves on which νv=0subscript𝜈𝑣0\partial_{\nu}v=0∂ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_v = 0, and some of those curves enclose bounded 𝒞1superscript𝒞1\mathscr{C}^{1}script_C start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT or Lipschitz domains (with corners of opening angle π/2𝜋2\pi/2italic_π / 2), or domains with cusps. We display some of these domains in Figure 11, where the starting points X(0)=(x0,y0)𝑋0subscript𝑥0subscript𝑦0X(0)=(x_{0},y_{0})italic_X ( 0 ) = ( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) satisfy x0=0subscript𝑥00x_{0}=0italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 and y0{αsin(π/7)||α|<1}subscript𝑦0conditional-set𝛼𝜋7𝛼1y_{0}\in\{\alpha\sin(\pi/7)\,|\,|\alpha|<1\}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ { italic_α roman_sin ( italic_π / 7 ) | | italic_α | < 1 }. In the numerics, we take ±α{0,0.1,0.5,0.8,0.9,0.9999}plus-or-minus𝛼00.10.50.80.90.9999\pm\alpha\in\{0,0.1,0.5,0.8,0.9,0.9999\}± italic_α ∈ { 0 , 0.1 , 0.5 , 0.8 , 0.9 , 0.9999 }. The curves on the left side of the starting points are obtained by solving (14) and letting t+𝑡t\to+\inftyitalic_t → + ∞. We get limt+X(t)=(t0,0)subscript𝑡𝑋𝑡subscript𝑡00\lim_{t\to+\infty}X(t)=(-t_{0},0)roman_lim start_POSTSUBSCRIPT italic_t → + ∞ end_POSTSUBSCRIPT italic_X ( italic_t ) = ( - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , 0 ). Those on the right side are obtained by solving the “backward” system X(t)=v(X(t))superscript𝑋𝑡𝑣𝑋𝑡X^{\prime}(t)=-\nabla v(X(t))italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) = - ∇ italic_v ( italic_X ( italic_t ) ), in which case limt+X(t)=(t0,0)subscript𝑡𝑋𝑡subscript𝑡00\lim_{t\to+\infty}X(t)=(t_{0},0)roman_lim start_POSTSUBSCRIPT italic_t → + ∞ end_POSTSUBSCRIPT italic_X ( italic_t ) = ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , 0 ). An explicit computation gives v(±t0,0)=±diag(λ1,λ2)𝑣plus-or-minussubscript𝑡00plus-or-minusdiagsubscript𝜆1subscript𝜆2\mathcal{H}v(\pm t_{0},0)=\pm\text{diag}\,(\lambda_{1},\lambda_{2})caligraphic_H italic_v ( ± italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , 0 ) = ± diag ( italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) with 0<λ2<λ10subscript𝜆2subscript𝜆10<\lambda_{2}<\lambda_{1}0 < italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. As a consequence, though this does not appear clearly in Figure 11(b) because here we do not zoom at (±t0,0)plus-or-minussubscript𝑡00(\pm t_{0},0)( ± italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , 0 ), all the orbits except the horizontal one, are tangential to the vertical axis at (±t0,0)plus-or-minussubscript𝑡00(\pm t_{0},0)( ± italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , 0 ).

Refer to caption
(a) Normalized gradient v/|v|𝑣𝑣\nabla v/|\nabla v|∇ italic_v / | ∇ italic_v |.
Refer to caption
(b) Neumann curves.
Figure 11: Case v𝑣vitalic_v as in (12) with L=2𝐿2L=2italic_L = 2, μ=7/2𝜇72\mu=7/2italic_μ = 7 / 2, a=cos(π/7)𝑎𝜋7a=\cos(\pi/7)italic_a = roman_cos ( italic_π / 7 ), b0=1=b1subscript𝑏01subscript𝑏1b_{0}=1=-b_{1}italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 = - italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.

3 Non-scattering with anisotropic materials

In this section, we provide examples of anisotropic materials, more precisely materials for which A𝐴Aitalic_A in (2) is a matrix valued function, that support non-scattering frequencies. For some of them, non-scattering occurs at all k>0𝑘0k>0italic_k > 0 so that in particular the set of transmission eigenvalues 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT contains (0,+)0(0,+\infty)( 0 , + ∞ ).

3.1 Non-scattering via diffeomorphism

We first look for non-scattering waves and media by using diffeomorphisms following for example [18] (see also [10]). Let v𝑣vitalic_v be an entire solution to the Helmholtz equation, i.e. such that

Δv+k2v=0in d.Δ𝑣superscript𝑘2𝑣0in d\Delta v+k^{2}v=0\qquad\mbox{in $\mathbb{R}^{d}$}.roman_Δ italic_v + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v = 0 in blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT .

Given a bounded Lipschitz domain ΩΩ\Omegaroman_Ω in dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, let ΨΨ\Psiroman_Ψ be a 𝒞m(Ω¯)superscript𝒞𝑚¯Ω\mathscr{C}^{m}(\overline{\Omega})script_C start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( over¯ start_ARG roman_Ω end_ARG ), m2𝑚2m\geq 2italic_m ≥ 2, diffeomorphism satisfying Ψ=IdΨId\Psi=\operatorname{Id}roman_Ψ = roman_Id on ΩΩ\partial\Omega∂ roman_Ω. Define

A=DΨDTΨ|detDΨ|Ψ1andq=1|detDΨ|Ψ1in Ω,formulae-sequence𝐴𝐷Ψsuperscript𝐷𝑇Ψ𝐷ΨsuperscriptΨ1and𝑞1𝐷ΨsuperscriptΨ1in Ω{A}=\frac{D\Psi D^{T}\Psi}{\left\lvert\det D\Psi\right\rvert}\circ\Psi^{-1}% \quad\text{and}\quad{q}=\frac{1}{\left\lvert\det D\Psi\right\rvert}\circ\Psi^{% -1}\qquad\mbox{in $\Omega$},italic_A = divide start_ARG italic_D roman_Ψ italic_D start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT roman_Ψ end_ARG start_ARG | roman_det italic_D roman_Ψ | end_ARG ∘ roman_Ψ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and italic_q = divide start_ARG 1 end_ARG start_ARG | roman_det italic_D roman_Ψ | end_ARG ∘ roman_Ψ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT in roman_Ω , (17)

where DΨ𝐷ΨD\Psiitalic_D roman_Ψ is the Jacobian matrix of ΨΨ\Psiroman_Ψ. Note in particular that A𝐴Aitalic_A is symmetric. Then the function u:=vΨ1assign𝑢𝑣superscriptΨ1u:=v\circ\Psi^{-1}italic_u := italic_v ∘ roman_Ψ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT satisfies u=v𝑢𝑣u=vitalic_u = italic_v on ΩΩ\partial\Omega∂ roman_Ω and

div(Au)+k2qu=0in Ω,νAu=νvon Ω,formulae-sequencediv𝐴𝑢superscript𝑘2𝑞𝑢0in Ωsuperscriptsubscript𝜈𝐴𝑢subscript𝜈𝑣on Ω,\begin{split}\mathrm{div}(A\nabla u)+k^{2}qu=0\quad\mbox{in $\Omega$},&\qquad% \partial_{\nu}^{{A}}u=\partial_{\nu}v\quad\mbox{on $\partial\Omega$,}\end{split}start_ROW start_CELL roman_div ( italic_A ∇ italic_u ) + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_q italic_u = 0 in roman_Ω , end_CELL start_CELL ∂ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT italic_u = ∂ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_v on ∂ roman_Ω , end_CELL end_ROW (18)

with νA:=νA\partial_{\nu}^{A}\cdot:=\nu\cdot A\nabla\cdot∂ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ⋅ := italic_ν ⋅ italic_A ∇ ⋅. As a consequence, we have the following lemma.

Lemma 3.1.

Let ΩΩ\Omegaroman_Ω be a bounded Lipschitz domain in dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT and let ΨΨ\Psiroman_Ψ be a 𝒞2(Ω¯)superscript𝒞2¯Ω\mathscr{C}^{2}(\overline{\Omega})script_C start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( over¯ start_ARG roman_Ω end_ARG ) diffeomorphism on Ω¯¯Ω\overline{\Omega}over¯ start_ARG roman_Ω end_ARG satisfying Ψ=IdΨId\Psi=\operatorname{Id}roman_Ψ = roman_Id on ΩΩ\partial\Omega∂ roman_Ω. Define A𝐴Aitalic_A, q𝑞qitalic_q by (17). Then A𝐴Aitalic_A, q𝑞qitalic_q satisfy assumptions (3) and (A,q;Ω)𝐴𝑞Ω(A,q;\Omega)( italic_A , italic_q ; roman_Ω ) is non-scattering for any incident waves at any frequencies.

Remark 3.2.

Notice for A𝐴Aitalic_A defined in (17) we always have detA=|detDΨ|2dΨ1𝐴superscript𝐷Ψ2𝑑superscriptΨ1\det A=\left\lvert\det D\Psi\right\rvert^{2-d}\circ\Psi^{-1}roman_det italic_A = | roman_det italic_D roman_Ψ | start_POSTSUPERSCRIPT 2 - italic_d end_POSTSUPERSCRIPT ∘ roman_Ψ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. In particular detA=1𝐴1\det A=1roman_det italic_A = 1 in dimension two and detA=q𝐴𝑞\det A=qroman_det italic_A = italic_q in dimension three. This may shed light to the verification of optimality for results on the discreteness of transmission eigenvalues; See also [21, 24].

As observed in [10], given any bounded Lipschitz domain ΩdΩsuperscript𝑑\Omega\subset\mathbb{R}^{d}roman_Ω ⊂ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, one can construct infinitely many diffeomorphisms that satisfy the conditions in Lemma 3.1. For instance, let Φ:dd:Φsuperscript𝑑superscript𝑑\Phi:\mathbb{R}^{d}\to\mathbb{R}^{d}roman_Φ : blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT → blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT be a 𝒞2superscript𝒞2\mathscr{C}^{2}script_C start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT map such that Φ=0Φ0\Phi=0roman_Φ = 0 in dΩsuperscript𝑑Ω\mathbb{R}^{d}\setminus\Omegablackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ∖ roman_Ω. Then the mapping ΨΨ\Psiroman_Ψ defined as

Ψ=Id+εΦ.ΨId𝜀Φ\Psi=\mathrm{Id}+\varepsilon\Phi.roman_Ψ = roman_Id + italic_ε roman_Φ . (19)

is a 𝒞2(Ω¯)superscript𝒞2¯Ω\mathscr{C}^{2}(\overline{\Omega})script_C start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( over¯ start_ARG roman_Ω end_ARG ) diffeomorphism for ε>0𝜀0\varepsilon>0italic_ε > 0 sufficiently small that satisfies Ψ=IdΨId\Psi=\operatorname{Id}roman_Ψ = roman_Id on ΩΩ\partial\Omega∂ roman_Ω.

Next, we give two explicit non-scattering examples (where ε𝜀\varepsilonitalic_ε in (19) does not have to be small) with ΩΩ\Omegaroman_Ω being a square and a disk, respectively.

Example 3.3 (Geometry with corners).

Let Ω=(1,1)22Ωsuperscript112superscript2\Omega=(-1,1)^{2}\subset\mathbb{R}^{2}roman_Ω = ( - 1 , 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⊂ blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. For a fixed constant α(1/2,1/2)𝛼1212\alpha\in(-1/2,1/2)italic_α ∈ ( - 1 / 2 , 1 / 2 ), define

Ψ(x,y)=(x+α(1x2)(1y2),y).Ψ𝑥𝑦𝑥𝛼1superscript𝑥21superscript𝑦2𝑦\Psi(x,y)=\left(x+\alpha(1-x^{2})(1-y^{2})\,,\,y\right).roman_Ψ ( italic_x , italic_y ) = ( italic_x + italic_α ( 1 - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( 1 - italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , italic_y ) .

It can be verified that Ψ:Ω¯Ω¯:Ψ¯Ω¯Ω\Psi:\overline{\Omega}\to\overline{\Omega}roman_Ψ : over¯ start_ARG roman_Ω end_ARG → over¯ start_ARG roman_Ω end_ARG is a 𝒞superscript𝒞\mathscr{C}^{\infty}script_C start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT diffeomorphism with Ψ|Ω=Idevaluated-atΨΩId\Psi|_{\partial\Omega}=\operatorname{Id}roman_Ψ | start_POSTSUBSCRIPT ∂ roman_Ω end_POSTSUBSCRIPT = roman_Id. Then the medium (A,q;(1,1)2)𝐴𝑞superscript112(A,q;(-1,1)^{2})( italic_A , italic_q ; ( - 1 , 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) as in (17) is non-scattering for all incident waves at all frequencies. We can express A𝐴Aitalic_A and q𝑞qitalic_q in a more explicit form as

AΨ(x,y)=112αx(1y2)((12αx(1y2))2+4α2y2(1x2)22αy(1x2)2αy(1x2)1)𝐴Ψ𝑥𝑦112𝛼𝑥1superscript𝑦2superscript12𝛼𝑥1superscript𝑦224superscript𝛼2superscript𝑦2superscript1superscript𝑥222𝛼𝑦1superscript𝑥22𝛼𝑦1superscript𝑥21A\circ\Psi(x,y)=\frac{1}{1-2\alpha x(1-y^{2})}\left(\begin{array}[]{cc}\left(1% -2\alpha x(1-y^{2})\right)^{2}+4\alpha^{2}y^{2}(1-x^{2})^{2}&-2\alpha y(1-x^{2% })\\ -2\alpha y(1-x^{2})&1\end{array}\right)italic_A ∘ roman_Ψ ( italic_x , italic_y ) = divide start_ARG 1 end_ARG start_ARG 1 - 2 italic_α italic_x ( 1 - italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG ( start_ARRAY start_ROW start_CELL ( 1 - 2 italic_α italic_x ( 1 - italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL - 2 italic_α italic_y ( 1 - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_CELL end_ROW start_ROW start_CELL - 2 italic_α italic_y ( 1 - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_CELL start_CELL 1 end_CELL end_ROW end_ARRAY )

and

qΨ=112αx(1y2).𝑞Ψ112𝛼𝑥1superscript𝑦2q\circ\Psi=\frac{1}{1-2\alpha x(1-y^{2})}.italic_q ∘ roman_Ψ = divide start_ARG 1 end_ARG start_ARG 1 - 2 italic_α italic_x ( 1 - italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG .
Example 3.4 (Non-spherically stratified disk).

Let f𝒞02(+)𝑓subscriptsuperscript𝒞20superscriptf\in\mathscr{C}^{2}_{0}(\mathbb{R}^{+})italic_f ∈ script_C start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ) satisfy f(1)=0𝑓10f(1)=0italic_f ( 1 ) = 0. Define ΨΨ\Psiroman_Ψ, in polar coordinates, as

Ψ(r,θ)=(r,θ+f(r)).Ψ𝑟𝜃𝑟𝜃𝑓𝑟\Psi(r,\theta)=(r,\theta+f(r)).roman_Ψ ( italic_r , italic_θ ) = ( italic_r , italic_θ + italic_f ( italic_r ) ) .

Then Ψ:B1¯B1¯:Ψ¯subscript𝐵1¯subscript𝐵1\Psi:\overline{B_{1}}\to\overline{B_{1}}roman_Ψ : over¯ start_ARG italic_B start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG → over¯ start_ARG italic_B start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG is a 𝒞2superscript𝒞2\mathscr{C}^{2}script_C start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT diffeomorphism with Ψ|Ω=Idevaluated-atΨΩId\Psi|_{\partial\Omega}=\operatorname{Id}roman_Ψ | start_POSTSUBSCRIPT ∂ roman_Ω end_POSTSUBSCRIPT = roman_Id (this latter property comes from the constraint f(1)=0𝑓10f(1)=0italic_f ( 1 ) = 0). Moreover, ΨΨ\Psiroman_Ψ is orientation- and area-preserving with detDΨ=1𝐷Ψ1\det D\Psi=1roman_det italic_D roman_Ψ = 1. As a consequence, the inhomogeneity (A,1;B1)𝐴1subscript𝐵1(A,1;B_{1})( italic_A , 1 ; italic_B start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) with A𝐴Aitalic_A as in (17) is non-scattering for all incident waves at all frequencies. Note that we find

A(x,y)=Idf(r)r(2xyy2x2y2x22xy)+(f(r))2(y2xyxyx2),r=(x2+y2)1/2.formulae-sequence𝐴𝑥𝑦Idsuperscript𝑓𝑟𝑟2𝑥𝑦superscript𝑦2superscript𝑥2superscript𝑦2superscript𝑥22𝑥𝑦superscriptsuperscript𝑓𝑟2superscript𝑦2𝑥𝑦𝑥𝑦superscript𝑥2𝑟superscriptsuperscript𝑥2superscript𝑦212A(x,y)=\mathrm{Id}-\frac{f^{\prime}(r)}{r}\left(\begin{array}[]{cc}2xy&y^{2}-x% ^{2}\\ y^{2}-x^{2}&-2xy\end{array}\right)+\left(f^{\prime}(r)\right)^{2}\left(\begin{% array}[]{cc}y^{2}&-xy\\ -xy&x^{2}\end{array}\right),\quad r=\left(x^{2}+y^{2}\right)^{1/2}.italic_A ( italic_x , italic_y ) = roman_Id - divide start_ARG italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_r ) end_ARG start_ARG italic_r end_ARG ( start_ARRAY start_ROW start_CELL 2 italic_x italic_y end_CELL start_CELL italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL - 2 italic_x italic_y end_CELL end_ROW end_ARRAY ) + ( italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_r ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( start_ARRAY start_ROW start_CELL italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL - italic_x italic_y end_CELL end_ROW start_ROW start_CELL - italic_x italic_y end_CELL start_CELL italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW end_ARRAY ) , italic_r = ( italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT .
On the structure of the anisotropy induced by diffemorphism transforms.

The following two results reveal an interesting geometry property of diffeomorphisms satisfying the conditions in Lemma 3.1.

Lemma 3.5.

Let ΩΩ\Omegaroman_Ω and ΨΨ\Psiroman_Ψ be as in Lemma 3.1. Then we must have

(DTΨ|detDΨ|Id)ν=0on Ω.superscript𝐷𝑇Ψ𝐷ΨId𝜈0on Ω\Big{(}\frac{D^{T}\Psi}{\left\lvert\det D\Psi\right\rvert}-\mathrm{Id}\Big{)}% \nu=0\qquad\mbox{on $\partial\Omega$}.( divide start_ARG italic_D start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT roman_Ψ end_ARG start_ARG | roman_det italic_D roman_Ψ | end_ARG - roman_Id ) italic_ν = 0 on ∂ roman_Ω .
Proof.

Let j{1,,d}𝑗1𝑑j\in\{1,\ldots,d\}italic_j ∈ { 1 , … , italic_d } be fixed. Let v(x)=xj𝑣xsubscript𝑥𝑗v(\mathrm{x})=x_{j}italic_v ( roman_x ) = italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and let u=vΨ1𝑢𝑣superscriptΨ1u=v\circ\Psi^{-1}italic_u = italic_v ∘ roman_Ψ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Then u𝑢uitalic_u, v𝑣vitalic_v satisfy

div(Au)=0andΔv=0in Ω,u=vandνAu=νvon Ω,formulae-sequencediv𝐴𝑢0andformulae-sequenceΔ𝑣0in Ωformulae-sequence𝑢𝑣andsuperscriptsubscript𝜈𝐴𝑢subscript𝜈𝑣on Ω\mathrm{div}(A\nabla u)=0\quad\text{and}\quad\Delta v=0\quad\mbox{in $\Omega$}% ,\qquad u=v\quad\text{and}\quad\partial_{\nu}^{A}u=\partial_{\nu}v\quad\mbox{% on $\partial\Omega$},roman_div ( italic_A ∇ italic_u ) = 0 and roman_Δ italic_v = 0 in roman_Ω , italic_u = italic_v and ∂ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT italic_u = ∂ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_v on ∂ roman_Ω ,

with A𝐴Aitalic_A defined in (17). Furthermore, by straightforward calculations we have

νv=νTejandνAu=νTDΨDTΨ|detDΨ|(DTΨ1)ej=νTDΨ|detDΨ|ejon Ω.formulae-sequenceformulae-sequencesubscript𝜈𝑣superscript𝜈𝑇subscript𝑒𝑗andsuperscriptsubscript𝜈𝐴𝑢superscript𝜈𝑇𝐷Ψsuperscript𝐷𝑇Ψ𝐷Ψsuperscript𝐷𝑇superscriptΨ1subscript𝑒𝑗superscript𝜈𝑇𝐷Ψ𝐷Ψsubscript𝑒𝑗on Ω\partial_{\nu}v=\nu^{T}\vec{e}_{j}\qquad\text{and}\qquad\partial_{\nu}^{A}u=% \nu^{T}\frac{D\Psi D^{T}\Psi}{\left\lvert\det D\Psi\right\rvert}(D^{T}\Psi^{-1% })\,\vec{e}_{j}=\nu^{T}\frac{D\Psi}{\left\lvert\det D\Psi\right\rvert}\,\vec{e% }_{j}\qquad\mbox{on $\partial\Omega$}.∂ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_v = italic_ν start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT over→ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and ∂ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT italic_u = italic_ν start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT divide start_ARG italic_D roman_Ψ italic_D start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT roman_Ψ end_ARG start_ARG | roman_det italic_D roman_Ψ | end_ARG ( italic_D start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT roman_Ψ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) over→ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_ν start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT divide start_ARG italic_D roman_Ψ end_ARG start_ARG | roman_det italic_D roman_Ψ | end_ARG over→ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT on ∂ roman_Ω .

The proof is complete by taking all j{1,,d}𝑗1𝑑j\in\{1,\ldots,d\}italic_j ∈ { 1 , … , italic_d }. ∎

Combining this result with [10, Theorem 2.1], we obtain the following statement when ΩΩ\Omegaroman_Ω is not smooth.

Proposition 3.6.

Let ΩΩ\Omegaroman_Ω be a 𝒞1,αsuperscript𝒞1𝛼\mathscr{C}^{1,\alpha}script_C start_POSTSUPERSCRIPT 1 , italic_α end_POSTSUPERSCRIPT bounded domain in dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, and let ΨΨ\Psiroman_Ψ be a 𝒞m+2superscript𝒞𝑚2\mathscr{C}^{m+2}script_C start_POSTSUPERSCRIPT italic_m + 2 end_POSTSUPERSCRIPT (resp. 𝒞superscript𝒞\mathscr{C}^{\infty}script_C start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT or (real) analytic) diffeomorphism on Ω¯¯Ω\overline{\Omega}over¯ start_ARG roman_Ω end_ARG with m1𝑚1m\geq 1italic_m ≥ 1 satisfying Ψ|Ω=Idevaluated-atΨΩId\Psi|_{\partial\Omega}=\operatorname{Id}roman_Ψ | start_POSTSUBSCRIPT ∂ roman_Ω end_POSTSUBSCRIPT = roman_Id. Then

(DΨId)ν=0for x0Ω,𝐷ΨId𝜈0for x0Ω\left(D\Psi-\mathrm{Id}\right)\nu=0\qquad\mbox{for\quad$\mathrm{x}_{0}\in% \partial\Omega$},( italic_D roman_Ψ - roman_Id ) italic_ν = 0 for roman_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ ∂ roman_Ω , (20)

where x0Ωsubscriptx0Ω\mathrm{x}_{0}\in\partial\Omegaroman_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ ∂ roman_Ω is such that ΩΩ\partial\Omega∂ roman_Ω is not 𝒞m,αsuperscript𝒞𝑚𝛼\mathscr{C}^{m,\alpha}script_C start_POSTSUPERSCRIPT italic_m , italic_α end_POSTSUPERSCRIPT (resp. 𝒞superscript𝒞\mathscr{C}^{\infty}script_C start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT or (real) analytic) in any neighborhood of x0subscriptx0\mathrm{x}_{0}roman_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

Proof.

Let x0Ωsubscriptx0Ω\mathrm{x}_{0}\in\partial\Omegaroman_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ ∂ roman_Ω be such that ΩΩ\partial\Omega∂ roman_Ω is not 𝒞m,αsuperscript𝒞𝑚𝛼\mathscr{C}^{m,\alpha}script_C start_POSTSUPERSCRIPT italic_m , italic_α end_POSTSUPERSCRIPT (resp. 𝒞superscript𝒞\mathscr{C}^{\infty}script_C start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT or (real) analytic) in any neighborhood of x0subscriptx0\mathrm{x}_{0}roman_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Assume by contradiction that (DΨId)ν0𝐷ΨId𝜈0(D\Psi-\mathrm{Id})\nu\neq 0( italic_D roman_Ψ - roman_Id ) italic_ν ≠ 0 at x0subscriptx0\mathrm{x}_{0}roman_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Let

A=DΨDTΨ|detDΨ| in Ω¯.𝐴𝐷Ψsuperscript𝐷𝑇Ψ𝐷Ψ in Ω¯A=\frac{D\Psi D^{T}\Psi}{\left\lvert\det D\Psi\right\rvert}\quad\mbox{ in $% \overline{\Omega}$}.italic_A = divide start_ARG italic_D roman_Ψ italic_D start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT roman_Ψ end_ARG start_ARG | roman_det italic_D roman_Ψ | end_ARG in over¯ start_ARG roman_Ω end_ARG .

Then by Lemma 3.5 we have

(AId)ν=(DΨId)νon Ω,𝐴Id𝜈𝐷ΨId𝜈on Ω(A-\mathrm{Id})\,\nu=(D\Psi-\mathrm{Id})\,\nu\qquad\mbox{on $\partial\Omega$},( italic_A - roman_Id ) italic_ν = ( italic_D roman_Ψ - roman_Id ) italic_ν on ∂ roman_Ω ,

and so (AId)ν0𝐴Id𝜈0(A-\mathrm{Id})\nu\neq 0( italic_A - roman_Id ) italic_ν ≠ 0 at x0Ωsubscriptx0Ω\mathrm{x}_{0}\in\partial\Omegaroman_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ ∂ roman_Ω. Exploiting this, we can find a solution v𝑣vitalic_v to the homogeneous Helmholtz equation in dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT such that νT(AId)v0superscript𝜈𝑇𝐴Id𝑣0\nu^{T}(A-\mathrm{Id})\nabla v\neq 0italic_ν start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ( italic_A - roman_Id ) ∇ italic_v ≠ 0 at x0subscriptx0\mathrm{x}_{0}roman_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. But from Lemma 3.1 we know that v𝑣vitalic_v is non-scattering for (A,q;Ω)𝐴𝑞Ω(A,q;\Omega)( italic_A , italic_q ; roman_Ω ) with A𝐴Aitalic_A, q𝑞qitalic_q given in (17). Hence by [10, Theorem 2.1] we must have that ΩΩ\partial\Omega∂ roman_Ω is 𝒞m,αsuperscript𝒞𝑚𝛼\mathscr{C}^{m,\alpha}script_C start_POSTSUPERSCRIPT italic_m , italic_α end_POSTSUPERSCRIPT (resp. 𝒞superscript𝒞\mathscr{C}^{\infty}script_C start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT or (real) analytic) near x0subscriptx0\mathrm{x}_{0}roman_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, which leads to a contradiction. ∎

Remark 3.7.

In particular, we show in the proof that the matrix A𝐴Aitalic_A defined in (17) must satisfy

Aν=ν𝐴𝜈𝜈A\,\nu=\nuitalic_A italic_ν = italic_ν

at the considered “non-smooth” point x0subscriptx0\mathrm{x}_{0}roman_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT of ΩΩ\partial\Omega∂ roman_Ω

3.2 Other explicit non-scattering examples

In this section, we are focused on non-scattering examples where the total and incident fields are identical in the whole dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT.

Example 3.8.

Let Ω=(0,π)2Ωsuperscript0𝜋2\Omega=(0,\pi)^{2}roman_Ω = ( 0 , italic_π ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and

A=(a100a2),q=q0in Ω,formulae-sequence𝐴matrixsubscript𝑎100subscript𝑎2𝑞subscript𝑞0in Ω,A=\left(\begin{matrix}a_{1}&0\\ 0&a_{2}\end{matrix}\right),\qquad\quad q=q_{0}\qquad\mbox{in $\Omega$,}italic_A = ( start_ARG start_ROW start_CELL italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) , italic_q = italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in roman_Ω , (21)

with a1,a2,q0+subscript𝑎1subscript𝑎2subscript𝑞0subscripta_{1},a_{2},q_{0}\in\mathbb{R}_{+}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT. Assume that these coefficients satisfy one of the following assumptions:
\bullet a11subscript𝑎11a_{1}\neq 1italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≠ 1, a21subscript𝑎21a_{2}\neq 1italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≠ 1 and there exist m,n𝑚𝑛m,n\in\mathbb{N}italic_m , italic_n ∈ blackboard_N, with (m,n)(0,0(m,n)\neq(0,0( italic_m , italic_n ) ≠ ( 0 , 0), such that m2(q0a1)=n2(a2q0)superscript𝑚2subscript𝑞0subscript𝑎1superscript𝑛2subscript𝑎2subscript𝑞0m^{2}(q_{0}-a_{1})=n^{2}(a_{2}-q_{0})italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT );
\bullet a1=1subscript𝑎11a_{1}=1italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 1, a21subscript𝑎21a_{2}\neq 1italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≠ 1 and (q01)(a21)>0subscript𝑞01subscript𝑎210(q_{0}-1)(a_{2}-1)>0( italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) ( italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - 1 ) > 0;
\bullet a11subscript𝑎11a_{1}\neq 1italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≠ 1, a2=1subscript𝑎21a_{2}=1italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1 and (q01)(a11)>0subscript𝑞01subscript𝑎110(q_{0}-1)(a_{1}-1)>0( italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) ( italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - 1 ) > 0.
Then 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT contains an unbounded sequence which accumulates at ++\infty+ ∞.

Remark 3.9.

It is known for the last two cases in Example 3.8 that 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT is at most countable, and hence such (A,q;Ω)𝐴𝑞Ω(A,q;\Omega)( italic_A , italic_q ; roman_Ω ) cannot be non-scattering for all frequencies. The same is true for the first case provided a1q01subscript𝑎1subscript𝑞01a_{1}q_{0}\neq 1italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≠ 1 or a2q01subscript𝑎2subscript𝑞01a_{2}q_{0}\neq 1italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≠ 1 (see [21, 24]); Otherwise if a1q0=1subscript𝑎1subscript𝑞01a_{1}q_{0}=1italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 or a2q0=1subscript𝑎2subscript𝑞01a_{2}q_{0}=1italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 but q01subscript𝑞01q_{0}\neq 1italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≠ 1, the discreteness of 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT is to our best knowledge still an open question.

Proof.

Suppose that a1,a2,q0subscript𝑎1subscript𝑎2subscript𝑞0a_{1},a_{2},q_{0}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT satisfy the first set of assumptions. Then it can be verified straightforwardly that for any κ𝜅superscript\kappa\in\mathbb{N}^{\ast}italic_κ ∈ blackboard_N start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, the functions u𝑢uitalic_u, v𝑣vitalic_v such that

u(x)=v(x)=cos(κmx)cos(κny)in Ωformulae-sequence𝑢x𝑣x𝜅𝑚𝑥𝜅𝑛𝑦in Ωu(\mathrm{x})=v(\mathrm{x})=\cos(\kappa mx)\cos(\kappa ny)\quad\mbox{in $% \Omega$}italic_u ( roman_x ) = italic_v ( roman_x ) = roman_cos ( italic_κ italic_m italic_x ) roman_cos ( italic_κ italic_n italic_y ) in roman_Ω

solve the interior transmission eigenvalue problem (ITEP) (6) with k=κm2+n2𝑘𝜅superscript𝑚2superscript𝑛2k=\kappa\sqrt{m^{2}+n^{2}}italic_k = italic_κ square-root start_ARG italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG.
Now assume that a1,a2,q0subscript𝑎1subscript𝑎2subscript𝑞0a_{1},a_{2},q_{0}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT satisfy the second set of assumptions (the third one can be dealt with similarly). Then one can verify that for any m𝑚superscriptm\in\mathbb{N}^{\ast}italic_m ∈ blackboard_N start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, the functions u𝑢uitalic_u, v𝑣vitalic_v such that

u(x)=v(x)=(c1cos(bx)+c2sin(bx))cos(my)in Ω,formulae-sequence𝑢x𝑣xsubscript𝑐1𝑏𝑥subscript𝑐2𝑏𝑥𝑚𝑦in Ωu(\mathrm{x})=v(\mathrm{x})=\big{(}c_{1}\cos(bx)+c_{2}\sin(bx)\big{)}\cos(my)% \quad\mbox{in $\Omega$},italic_u ( roman_x ) = italic_v ( roman_x ) = ( italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_cos ( italic_b italic_x ) + italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_sin ( italic_b italic_x ) ) roman_cos ( italic_m italic_y ) in roman_Ω ,

with b=k2m2+i+𝑏superscript𝑘2superscript𝑚2subscript𝑖subscriptb=\sqrt{k^{2}-m^{2}}\in\mathbb{R}_{+}\cup i\mathbb{R}_{+}italic_b = square-root start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ∪ italic_i blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT and where c1subscript𝑐1c_{1}italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, c2subscript𝑐2c_{2}italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are non zero constants, solve (6) for k=m(a21)/(q01)𝑘𝑚subscript𝑎21subscript𝑞01k=m\sqrt{(a_{2}-1)/(q_{0}-1)}italic_k = italic_m square-root start_ARG ( italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - 1 ) / ( italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) end_ARG. ∎

Remark 3.10.

By scaling and/or rigid change of coordinates, we can adapt Example 3.8 to the case of any rectangle, with properly modified form of A(x)𝐴xA(\mathrm{x})italic_A ( roman_x ) (not necessarily diagonal) and conditions between A𝐴Aitalic_A and q𝑞qitalic_q. Analogous arguments and results apply when ΩΩ\Omegaroman_Ω is a cuboid in 3superscript3\mathbb{R}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT.

The case in Example 3.8 but for q=aj=1𝑞subscript𝑎𝑗1q=a_{j}=1italic_q = italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 1 for j=1𝑗1j=1italic_j = 1 or 2222 can be generalized to the following situation where q1𝑞1q\equiv 1italic_q ≡ 1 and AId𝐴IdA-\mathrm{Id}italic_A - roman_Id is rank deficient. In this situation we lost the discreteness of the sets 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT and 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT. A particular case of this example appears in [21, p.1170].

Example 3.11.

Given any bounded Lipschitz domain ΩΩ\Omegaroman_Ω in dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, suppose that A=A(x)𝐴𝐴xA=A(\mathrm{x})italic_A = italic_A ( roman_x ) satisfies

A(x)=U(100An1(x))UT,xΩ¯,formulae-sequence𝐴x𝑈100subscript𝐴𝑛1xsuperscript𝑈𝑇x¯ΩA(\mathrm{x})=U\left(\begin{array}[]{cc}1&0\\ 0&A_{n-1}(\mathrm{x})\end{array}\right)U^{T},\qquad\mathrm{x}\in\overline{% \Omega},italic_A ( roman_x ) = italic_U ( start_ARRAY start_ROW start_CELL 1 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_A start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ( roman_x ) end_CELL end_ROW end_ARRAY ) italic_U start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , roman_x ∈ over¯ start_ARG roman_Ω end_ARG ,

with some (n1)×(n1)𝑛1𝑛1(n-1)\times(n-1)( italic_n - 1 ) × ( italic_n - 1 ) symmetric positive definite matrix An1subscript𝐴𝑛1A_{n-1}italic_A start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT and a constant orthogonal matrix U𝑈Uitalic_U. Importantly, note that in this case 1111 is an eigenvalue of A(x)𝐴xA(\mathrm{x})italic_A ( roman_x ) for all xΩ¯x¯Ω\mathrm{x}\in\overline{\Omega}roman_x ∈ over¯ start_ARG roman_Ω end_ARG. Then any k>0𝑘0k>0italic_k > 0 is a non-scattering frequency for (A,1;Ω)𝐴1Ω(A,1;\Omega)( italic_A , 1 ; roman_Ω ).

Proof.

For any non-zero constants c1subscript𝑐1c_{1}italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, c2subscript𝑐2c_{2}italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, define v0(x)=v0(x1,x)=c1cos(kx1)+c2sin(kx1)subscript𝑣0xsubscript𝑣0subscript𝑥1superscript𝑥subscript𝑐1𝑘subscript𝑥1subscript𝑐2𝑘subscript𝑥1v_{0}(\mathrm{x})=v_{0}(x_{1},x^{\prime})=c_{1}\cos(kx_{1})+c_{2}\sin(kx_{1})italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( roman_x ) = italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_cos ( italic_k italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) + italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_sin ( italic_k italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ). It can be verified straightforwardly that for any bounded Lipschitz domain Ω~~Ω\tilde{\Omega}over~ start_ARG roman_Ω end_ARG and any fixed k>0𝑘0k>0italic_k > 0, (u0,v0)=(v0,v0)subscript𝑢0subscript𝑣0subscript𝑣0subscript𝑣0(u_{0},v_{0})=(v_{0},v_{0})( italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = ( italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) is a pair of eigenfunctions to the ITEP (6) for (A~,1;Ω~)~𝐴1~Ω(\tilde{A},1;\tilde{\Omega})( over~ start_ARG italic_A end_ARG , 1 ; over~ start_ARG roman_Ω end_ARG ) with A~=diag(1,An1)~𝐴diag1subscript𝐴𝑛1\tilde{A}=\mathrm{diag}(1,A_{n-1})over~ start_ARG italic_A end_ARG = roman_diag ( 1 , italic_A start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ) (note in particular that (A~I)v00~𝐴𝐼subscript𝑣00(\tilde{A}-I)\nabla v_{0}\equiv 0( over~ start_ARG italic_A end_ARG - italic_I ) ∇ italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≡ 0). Then consider the change of coordinates y=Uxy𝑈x\mathrm{y}=U\mathrm{x}roman_y = italic_U roman_x for xdxsuperscript𝑑\mathrm{x}\in\mathbb{R}^{d}roman_x ∈ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT and define v𝑣vitalic_v such that v(y):=v0(U1y)assign𝑣ysubscript𝑣0superscript𝑈1yv(\mathrm{y}):=v_{0}\big{(}U^{-1}\mathrm{y}\big{)}italic_v ( roman_y ) := italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_U start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_y ). The pair (u,v)=(v,v)𝑢𝑣𝑣𝑣(u,v)=(v,v)( italic_u , italic_v ) = ( italic_v , italic_v ) satisfies (6) for (A,1;Ω)𝐴1Ω(A,1;\Omega)( italic_A , 1 ; roman_Ω ) with Ω={y=Ux|xΩ~}Ωconditional-sety𝑈xx~Ω\Omega=\{\mathrm{y}=U\mathrm{x}\,|\,\mathrm{x}\in\tilde{\Omega}\}roman_Ω = { roman_y = italic_U roman_x | roman_x ∈ over~ start_ARG roman_Ω end_ARG }. Moreover, v𝑣vitalic_v is an entire solution to the Helmholtz equation. The proof is complete. ∎

The following can be viewed as a generalization of Example 3.8 when q=aj1𝑞subscript𝑎𝑗1q=a_{j}\neq 1italic_q = italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≠ 1 for j=1𝑗1j=1italic_j = 1 or 2222.

Example 3.12.

Let Ω=(b1,b2)×DdΩsubscript𝑏1subscript𝑏2𝐷superscript𝑑\Omega=(b_{1},b_{2})\times D\subset\mathbb{R}^{d}roman_Ω = ( italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) × italic_D ⊂ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT with D𝐷Ditalic_D an open interval if d=2𝑑2d=2italic_d = 2 or a bounded Lipschitz domain in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT if d=3𝑑3d=3italic_d = 3. Then there are infinitely many non-scattering frequencies for (A,q;Ω)𝐴𝑞Ω(A,q;\Omega)( italic_A , italic_q ; roman_Ω ) with

A(x)=(a000An1(x))andq(x)=a0,xΩ,formulae-sequence𝐴xsubscript𝑎000subscript𝐴𝑛1xandformulae-sequence𝑞xsubscript𝑎0xΩA(\mathrm{x})=\left(\begin{array}[]{cc}a_{0}&0\\ 0&A_{n-1}(\mathrm{x})\end{array}\right)\qquad\text{and}\qquad q(\mathrm{x})=a_% {0},\qquad\mathrm{x}\in\Omega,italic_A ( roman_x ) = ( start_ARRAY start_ROW start_CELL italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_A start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ( roman_x ) end_CELL end_ROW end_ARRAY ) and italic_q ( roman_x ) = italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , roman_x ∈ roman_Ω ,

for some positive constant a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and some (n1)×(n1)𝑛1𝑛1(n-1)\times(n-1)( italic_n - 1 ) × ( italic_n - 1 ) symmetric positive definite matrix An1subscript𝐴𝑛1A_{n-1}italic_A start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT.

Proof.

It can be verified straightforwardly that v(x)=cos(k(x1b1))𝑣x𝑘subscript𝑥1subscript𝑏1v(\mathrm{x})=\cos(k(x_{1}-b_{1}))italic_v ( roman_x ) = roman_cos ( italic_k ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ) is a non-scattering incident field with u=v𝑢𝑣u=vitalic_u = italic_v being the total field for k=mπ/|b2b1|𝑘𝑚𝜋subscript𝑏2subscript𝑏1k=m\pi/|b_{2}-b_{1}|italic_k = italic_m italic_π / | italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT |, m𝑚superscriptm\in\mathbb{N}^{\ast}italic_m ∈ blackboard_N start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. ∎

Remark 3.13.

We notice from Example 3.11 that if a0=1subscript𝑎01a_{0}=1italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 in Example 3.12, the sets 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT and hence 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT coincide with +superscript\mathbb{R}^{+}blackboard_R start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. However, in some other cases, for example when both a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and the smallest eigenvalue of An1(x)subscript𝐴𝑛1xA_{n-1}(\mathrm{x})italic_A start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ( roman_x ) are larger than 1111 for all xx\mathrm{x}roman_x in a small neighborhood of ΩΩ\partial\Omega∂ roman_Ω, the sets 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT and hence 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT are discrete [5, 24]. The discreteness of 𝒮TEsubscript𝒮TE\mathscr{S}_{\mathrm{TE}}script_S start_POSTSUBSCRIPT roman_TE end_POSTSUBSCRIPT or 𝒮NSsubscript𝒮NS\mathscr{S}_{\mathrm{NS}}script_S start_POSTSUBSCRIPT roman_NS end_POSTSUBSCRIPT in general is unknown for Example 3.12 especially if detA=1𝐴1\det A=1roman_det italic_A = 1 on ΩΩ\partial\Omega∂ roman_Ω.

Acknowledgements

The work of J. Xiao was supported in part by the National Science Foundation under Grant DMS-2307737.

Appendix

The following figures illustrate how to justify the existence of analytic nodal curves as shown in Figures 68. Below the corresponding functions are positive in the yellow regions and negative in the blue regions.

As an example of such justification, let us look at the nodal set of Figure 6(a). That is, when L=2𝐿2L=2italic_L = 2, μ=5/2𝜇52\mu=5/2italic_μ = 5 / 2 and a=0.58𝑎0.58a=0.58italic_a = 0.58 in (11). Based on the “generating function” Jμ(kr)cos(μθ)subscript𝐽𝜇𝑘𝑟𝜇𝜃J_{\mu}(kr)\cos(\mu\theta)italic_J start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_k italic_r ) roman_cos ( italic_μ italic_θ ), we can identify the positive and negative regions of each terms l=0,,L1𝑙0𝐿1l=0,\ldots,L-1italic_l = 0 , … , italic_L - 1 in (11). See Figure 12(a) for an example for the term l=0𝑙0l=0italic_l = 0 in (11), where the dashed lines are the nodal curves of the terms l=1,,L1𝑙1𝐿1l=1,\ldots,L-1italic_l = 1 , … , italic_L - 1 in (11). Then we can find certain regions where v𝑣vitalic_v in (11) has to be either positive/negative, say, the domains where every term is of the same sign (see Figure 12(b)). Therefore, there must be a nodal set in between the strictly positive and the strictly negative regions. The justification for Figures 6(b)6(c), and 7(a) is similar.

Refer to caption
(a) The term l=0𝑙0l=0italic_l = 0 in (11).
Refer to caption
(b) v𝑣vitalic_v in (11).
Refer to caption
(c) The nodal curves of v𝑣vitalic_v.
Figure 12: Justification for Figure 6(a): L=2𝐿2L=2italic_L = 2, μ=5/2𝜇52\mu=5/2italic_μ = 5 / 2, a=0.58𝑎0.58a=0.58italic_a = 0.58.
Refer to caption
(a) The term l=0𝑙0l=0italic_l = 0 in (11).
Refer to caption
(b) v𝑣vitalic_v in (11).
Refer to caption
(c) The nodal curves of v𝑣vitalic_v.
Figure 13: Justification for Figure 6(b): L=3𝐿3L=3italic_L = 3, μ=5/2𝜇52\mu=5/2italic_μ = 5 / 2, a=0.6𝑎0.6a=0.6italic_a = 0.6.
Refer to caption
(a) The term l=0𝑙0l=0italic_l = 0 in (11).
Refer to caption
(b) v𝑣vitalic_v in (11).
Refer to caption
(c) The nodal curves of v𝑣vitalic_v.
Figure 14: Justification for Figure 6(c): L=4𝐿4L=4italic_L = 4, μ=5/2𝜇52\mu=5/2italic_μ = 5 / 2, a=0.5𝑎0.5a=0.5italic_a = 0.5.
Refer to caption
(a) The term l=0𝑙0l=0italic_l = 0 in (11).
Refer to caption
(b) v𝑣vitalic_v in (11).
Refer to caption
(c) The nodal curves of v𝑣vitalic_v.
Figure 15: Justification for Figure 7(a): L=2𝐿2L=2italic_L = 2, μ=1𝜇1\mu=1italic_μ = 1, a=0.55𝑎0.55a=0.55italic_a = 0.55.

The situation for the case of Figure 2 is slightly different. First, we can identify the positive and negative regions for each term in (12) as before, and hence some one-sign regions for v𝑣vitalic_v in (12) (see Figures 16(a) and 16(b)). In particular, thanks to the symmetry of the two terms in (12), we have that v=0𝑣0v=0italic_v = 0 on the axis {y=0}𝑦0\{y=0\}{ italic_y = 0 }. Additionally, we observe that v=0𝑣0v=0italic_v = 0 at the two points (0,±1a2)0plus-or-minus1superscript𝑎2(0,\pm\sqrt{1-a^{2}})( 0 , ± square-root start_ARG 1 - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ), where in fact both terms in (12) are zero. Moreover, it can be verified straightforwardly that v0𝑣0\nabla v\neq 0∇ italic_v ≠ 0 at (0,±1a2)0plus-or-minus1superscript𝑎2(0,\pm\sqrt{1-a^{2}})( 0 , ± square-root start_ARG 1 - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ). So there is exactly one (real-analytic) nodal curve of v𝑣vitalic_v passing through the point (0,1a2)01superscript𝑎2(0,\sqrt{1-a^{2}})( 0 , square-root start_ARG 1 - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ). This curve will continue, say, to the right side of (0,1a2)01superscript𝑎2(0,\sqrt{1-a^{2}})( 0 , square-root start_ARG 1 - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) until it hits some point (x1,0)subscript𝑥10(x_{1},0)( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , 0 ) for some x1[1a,1+a]subscript𝑥11𝑎1𝑎x_{1}\in[1-a,1+a]italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∈ [ 1 - italic_a , 1 + italic_a ]. On the other hand, using the fact that a=(k2/k1)/2𝑎subscript𝑘2𝑘12a=(k_{2}/k-1)/2italic_a = ( italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_k - 1 ) / 2 with k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT the second positive zero of J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, it can be verified that v=0𝑣0\nabla v=0∇ italic_v = 0 at the two points ±(a+1,0)plus-or-minus𝑎10\pm(a+1,0)± ( italic_a + 1 , 0 ) and v0𝑣0\nabla v\neq 0∇ italic_v ≠ 0 at any (x1,0)subscript𝑥10(x_{1},0)( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , 0 ) with x1(a1,a+1)subscript𝑥1𝑎1𝑎1x_{1}\in(-a-1,a+1)italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∈ ( - italic_a - 1 , italic_a + 1 ). Therefore, the nodal curve of v𝑣vitalic_v generated from (0,1a2)01superscript𝑎2(0,\sqrt{1-a^{2}})( 0 , square-root start_ARG 1 - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) must also pass through (a+1,0)𝑎10(a+1,0)( italic_a + 1 , 0 ). In fact, one can also verify that the Hessian of v𝑣vitalic_v satisfies v0𝑣0\mathcal{H}v\neq 0caligraphic_H italic_v ≠ 0 at ±(a+1,0)plus-or-minus𝑎10\pm(a+1,0)± ( italic_a + 1 , 0 ). Hence there are exactly two (real-analytic) nodal curves passing at (a+1,0)𝑎10(a+1,0)( italic_a + 1 , 0 ): one is the line {y=0}𝑦0\{y=0\}{ italic_y = 0 } and the other is the one coming from (0,1a2)01superscript𝑎2(0,\sqrt{1-a^{2}})( 0 , square-root start_ARG 1 - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ). Applying analogous arguments for the other three quadrant we can then complete the justification of a closed (real-analytic) nodal curve of v𝑣vitalic_v.

Refer to caption
(a) The term l=0𝑙0l=0italic_l = 0 in (12).
Refer to caption
(b) v𝑣vitalic_v in (12).
Refer to caption
(c) The nodal curves of v𝑣vitalic_v.
Figure 16: Justification for Figure 2: L=2𝐿2L=2italic_L = 2, μ=1𝜇1\mu=1italic_μ = 1, a=k2k2k𝑎subscript𝑘2𝑘2𝑘a=\frac{{k_{2}}-{k}}{2k}italic_a = divide start_ARG italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_k end_ARG start_ARG 2 italic_k end_ARG, ϕ1=π/2=ϕ2subscriptitalic-ϕ1𝜋2subscriptitalic-ϕ2\phi_{1}=\pi/2=-\phi_{2}italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_π / 2 = - italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT.

In Figure 7(c), we make use of the function v𝑣vitalic_v in (12) with μ=1𝜇1\mu=1italic_μ = 1, L=2𝐿2L=2italic_L = 2, a=2/2𝑎22a=\sqrt{2}/2italic_a = square-root start_ARG 2 end_ARG / 2, ϕ1=π/4subscriptitalic-ϕ1𝜋4\phi_{1}=\pi/4italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_π / 4 and ϕ2=3π/4subscriptitalic-ϕ23𝜋4\phi_{2}=3\pi/4italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 3 italic_π / 4. The choice of ϕ1subscriptitalic-ϕ1\phi_{1}italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ϕ2subscriptitalic-ϕ2\phi_{2}italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ensures the symmetry of v𝑣vitalic_v with respect to the y𝑦yitalic_y-axis. In particular, v=0𝑣0v=0italic_v = 0 on {x=0}𝑥0\{x=0\}{ italic_x = 0 }. In addition, thanks to the value of a𝑎aitalic_a, one can verify that at the point P0=(0,2/2)subscript𝑃0022P_{0}=(0,-\sqrt{2}/2)italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ( 0 , - square-root start_ARG 2 end_ARG / 2 ) there holds v=xv=yv=ij2v=0𝑣subscript𝑥𝑣subscript𝑦𝑣subscriptsuperscript2𝑖𝑗𝑣0v=\partial_{x}v=\partial_{y}v=\partial^{2}_{ij}v=0italic_v = ∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_v = ∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_v = ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_v = 0 for all i,j{x,y}𝑖𝑗𝑥𝑦i,j\in\{x,y\}italic_i , italic_j ∈ { italic_x , italic_y } but x3v0superscriptsubscript𝑥3𝑣0\partial_{x}^{3}v\neq 0∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_v ≠ 0. Hence, there are exactly three nodal curves passing through (0,2/2)022(0,-\sqrt{2}/2)( 0 , - square-root start_ARG 2 end_ARG / 2 ). Furthermore, one can verify straightforwardly that v=0𝑣0v=0italic_v = 0 while v0𝑣0\nabla v\neq 0∇ italic_v ≠ 0 at the points Pjsubscript𝑃𝑗P_{j}italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, j=1,,9𝑗19j=1,\ldots,9italic_j = 1 , … , 9, where the coordinates of each point can be calculated explicitly in terms of a𝑎aitalic_a, k𝑘kitalic_k and k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. Finally, based on the locations of strictly positive and negative regions we can conclude to the existence of real-analytic nodal curves for v𝑣vitalic_v that enclose some bounded Lipschitz domains, as shown Figure 17(c).

Refer to caption
(a) The term l=0𝑙0l=0italic_l = 0 in (12).
Refer to caption
(b) v𝑣vitalic_v in (12).
Refer to caption
(c) The nodal curves of v𝑣vitalic_v.
Figure 17: Justification for Figure 7(c): L=2𝐿2L=2italic_L = 2, μ=1𝜇1\mu=1italic_μ = 1, a=2/2𝑎22a=\sqrt{2}/2italic_a = square-root start_ARG 2 end_ARG / 2, ϕ1=π/4subscriptitalic-ϕ1𝜋4\phi_{1}=\pi/4italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_π / 4, ϕ2=3π/4subscriptitalic-ϕ23𝜋4\phi_{2}=3\pi/4italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 3 italic_π / 4.

The justification of the results appearing in Figure 8 is similar.

Refer to caption
(a) The term l=0𝑙0l=0italic_l = 0 in (12).
Refer to caption
(b) v𝑣vitalic_v in (12)
Refer to caption
(c) The nodal curves of v𝑣vitalic_v.
Figure 18: Justification for Figure 8(a): L=2𝐿2L=2italic_L = 2, μ=7/2𝜇72\mu=7/2italic_μ = 7 / 2, a=cos(π/7)𝑎𝜋7a=\cos(\pi/7)italic_a = roman_cos ( italic_π / 7 ), b0=1=b1subscript𝑏01subscript𝑏1b_{0}=1=-b_{1}italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 = - italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.
Refer to caption
(a) The term l=0𝑙0l=0italic_l = 0 in (12).
Refer to caption
(b) v𝑣vitalic_v in (12).
Refer to caption
(c) The nodal curves of v𝑣vitalic_v.
Figure 19: Justification for Figure 8(b): L=2𝐿2L=2italic_L = 2, μ=5/2𝜇52\mu=5/2italic_μ = 5 / 2, a=1𝑎1a=1italic_a = 1, ϕ0=π/5=ϕ1subscriptitalic-ϕ0𝜋5subscriptitalic-ϕ1\phi_{0}=\pi/5=-\phi_{1}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_π / 5 = - italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.
Refer to caption
(a) The term l=0𝑙0l=0italic_l = 0 in (12).
Refer to caption
(b) v𝑣vitalic_v in (12).
Refer to caption
(c) The nodal curves of v𝑣vitalic_v.
Figure 20: Justification for Figure 8(c): L=2𝐿2L=2italic_L = 2, μ=3/2𝜇32\mu=3/2italic_μ = 3 / 2, a=2/2𝑎22a=\sqrt{2}/2italic_a = square-root start_ARG 2 end_ARG / 2, ϕ0=π/12=ϕ1subscriptitalic-ϕ0𝜋12subscriptitalic-ϕ1\phi_{0}=\pi/12=-\phi_{1}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_π / 12 = - italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.

References

  • [1] L. Audibert, L. Chesnel, H. Haddar, and K. Napal. Qualitative indicator functions for imaging crack networks using acoustic waves. SIAM J. Sci. Comput., 43(2):B271–B297, 2021.
  • [2] L. Audibert, H. Haddar, and F. Pourre. Reconstruction of averaging indicators for highly heterogeneous media. Inverse Probl., 40(4):045028, 2024.
  • [3] E. Blåsten, L. Päivärinta, and J. Sylvester. Corners always scatter. Commun. Math. Phys., 331(2):725–753, 2014.
  • [4] E. Blåsten, H. Liu, and J. Xiao. On an electromagnetic problem in a corner and its applications. Analysis & PDE, 14(7):2207–2224, 2021.
  • [5] A.-S. Bonnet-Ben Dhia, L. Chesnel, and H. Haddar. On the use of T𝑇Titalic_T-coercivity to study the Interior Transmission Eigenvalue Problem. C. R. Acad. Sci., Ser. I, 340:647–651, 2011.
  • [6] F. Cakoni, D. Colton, and H. Haddar. Inverse scattering theory and transmission eigenvalues, volume 98 of CBMS-NSF Regional Conference Series in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, second edition, 2023.
  • [7] F. Cakoni, D. Gintides, and H. Haddar. The existence of an infinite discrete set of transmission eigenvalues. SIAM J. Math. Anal., 42(1):237–255, 2010.
  • [8] F. Cakoni and H. Haddar. On the existence of transmission eigenvalues in an inhomogeneous medium. Appl. Anal., 88(4):475–493, 2009.
  • [9] F. Cakoni and M. Vogelius. Singularities almost always scatter: regularity results for non-scattering inhomogeneities. Comm. Pure Appl. Math., 76(12):4022–4047, 2023.
  • [10] F. Cakoni, M. Vogelius, and J. Xiao. On the regularity of non-scattering anisotropic inhomogeneities. Arch. Ration. Mech. Anal., 247(3):31, 2023.
  • [11] F. Cakoni and J. Xiao. On corner scattering for operators of divergence form and applications to inverse scattering. Commun. Partial Differ. Equ., 46(3):413–441, 2021.
  • [12] D. Colton and A. Kirsch. A simple method for solving inverse scattering problems in the resonance region. Inverse probl., 12(4):383, 1996.
  • [13] R. Courant and D. Hilbert. Methods of mathematical physics, volume I. Wiley-Interscience, 1989.
  • [14] S. D. Daymond. The principal frequencies of vibrating systems with elliptic boundaries. Q. J. Mech. Appl. Math., 8(3):361–372, 1955.
  • [15] J. Eckmann and C. Pillet. Spectral duality for planar billiards. Comm. Math. Phys., 170(2):283–313, 1995.
  • [16] J. Elschner and G. Hu. Acoustic scattering from corners, edges and circular cones. Arch. Ration. Mech. Anal., 228(2):653–690, 2018.
  • [17] A. Kirsch. On the existence of transmission eigenvalues. Inverse problems and imaging, 3(2):155–172, 2009.
  • [18] R. Kohn and M. Vogelius. Identification of an unknown conductivity by means of measurements at the boundary, volume 14 of SIAM-AMS Proc., pages 113–123. Amer. Math. Soc., Providence, RI, 1984.
  • [19] P. Kow, S. Larson, M. Salo, and H. Shahgholian. Quadrature domains for the Helmholtz equation with applications to non-scattering phenomena. Potential Anal., 60(1):387–424, 2024.
  • [20] J. R. Kuttler and V. G. Sigillito. Eigenvalues of the Laplacian in two dimensions. SIAM Review, 26(2):163–193, 1984.
  • [21] E. Lakshtanov and B. Vainberg. Ellipticity in the interior transmission problem in anisotropic media. SIAM J. Math. Anal., 44(2):1165–1174, 2012.
  • [22] B. McCartin. Laplacian eigenstructure of the equilateral triangle. Hikari Ltd., Ruse, 2011.
  • [23] N. W. McLachlan. Theory and application of Mathieu functions. Dover Publications, Inc., New York, 1964.
  • [24] H. Nguyen and Q. Nguyen. Discreteness of interior transmission eigenvalues revisited. Calc. Var. Partial Differ., 56(2):51, 2017.
  • [25] L. Päivärinta, M. Salo, and E. Vesalainen. Strictly convex corners scatter. Rev. Mat. Iberoam., 33(4):1369–1396, 2017.
  • [26] L. Päivärinta and J. Sylvester. Transmission eigenvalues. SIAM J. Math. Anal., 40(2):738–753, 2008.
  • [27] L. Robbiano. Spectral analysis of the interior transmission eigenvalue problem. Inverse Probl., 29(10):104001, 2013.
  • [28] M. Salo and H. Shahgholian. Free boundary methods and non-scattering phenomena. Res. Math. Sci., 8(4):58, 2021.
  • [29] J. Xiao. A new type of CGO solutions and its applications in corner scattering. Inverse Probl., 38(3):Paper No. 034001, 23, 2022.