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

Invisibility enables super-visibility in electromagnetic imaging

Youzi He Department of Mathematics, The Chinese University of Hong Kong, Shatin, Hong Kong SAR, China [email protected] Hongjie Li Department of Mathematics, The Chinese University of Hong Kong, Shatin, Hong Kong SAR, China [email protected] Hongyu Liu Department of Mathematics, City University of Hong Kong, Kowloon, Hong Kong SAR, China [email protected]  and  Xianchao Wang School of Mathematics, Harbin Institute of Technology, Harbin, China [email protected]
Abstract.

This paper is concerned with the inverse electromagnetic scattering problem for anisotropic media. We use the interior resonant modes to develop an inverse scattering scheme for imaging the scatterer. The whole procedure consists of three phases. First, we determine the interior Maxwell transmission eigenvalues of the scatterer from a family of far-field data by the mechanism of the linear sampling method. Next, we determine the corresponding transmission eigenfunctions by solving a constrained optimization problem. Finally, based on both global and local geometric properties of the transmission eigenfunctions, we design an imaging functional which can be used to determine the shape of the medium scatterer. We provide rigorous theoretical basis for our method. Numerical experiments verify the effectiveness, better accuracy and super-resolution results of the proposed scheme.

Keywords:   inverse electromagnetic scattering, anisotropic media, transmission eigenfunctions, geometric structures, super-resolution

2010 Mathematics Subject Classification:  35Q60, 35P25, 35R30, 78A40

1. Introduction

1.1. Mathematical setup

We present the mathematical formulation of the forward and inverse electromagnetic scattering problems.

Let D3D\subset\mathbb{R}^{3} be the support of an anisotropic inhomogeneity, where DD is a bounded, simply connected Lipschitz domain with a piecewise smooth boundary. Next, we introduce the medium configuration, which is characterized by two material parameters including the electric permittivity 𝜺\bm{\varepsilon} and the magnetic permeability 𝝁\bm{\mu}. It is assumed that 𝝁(𝒙)μ0𝑰\bm{\mu}(\bm{x})\equiv\mu_{0}\bm{I}, 𝒙3\bm{x}\in\mathbb{R}^{3}, where μ0\mu_{0} is a positive constant and 𝑰\bm{I} is the 3×33\times 3 identity matrix. Let the electric permittivity 𝜺(L(3))3×3\bm{\varepsilon}\in(L^{\infty}(\mathbb{R}^{3}))^{3\times 3} be a uniformly symmetric-positive-definite bounded matrix-valued function. It is assumed that 𝜺(𝒙)=ε0𝑰\bm{\varepsilon}(\bm{x})=\varepsilon_{0}\bm{I} for 𝒙3\D¯\bm{x}\in\mathbb{R}^{3}\backslash\overline{D}, where ε0\varepsilon_{0} is a positive constant. Define the wavenumber kk to be

k2:=ε0μ0ω2,k^{2}:=\varepsilon_{0}\mu_{0}\omega^{2},

where ω+\omega\in\mathbb{R}_{+} signifies the temporal frequency of the electromagnetic waves. We define the matrix index of refraction of the anisotropic medium by

𝑵(𝒙):=𝜺(𝒙)ε0,𝒙D¯,\bm{N}(\bm{x}):=\frac{\bm{\varepsilon}(\bm{x})}{\varepsilon_{0}},\quad\bm{x}\in\overline{D}, (1.1)

and assume that (𝑵𝑰)1(\bm{N}-\bm{I})^{-1} is bounded and satisfies

((𝑵𝑰)1𝝃,𝝃)α|𝝃|2for any 𝝃3 a.e. in D((\bm{N}-\bm{I})^{-1}\bm{\xi},\bm{\xi})\geq\alpha|\bm{\xi}|^{2}\quad\text{for any }\bm{\xi}\in\mathbb{C}^{3}\text{ a.e. in }D (1.2)

for some constant α>0\alpha>0. Clearly, all the results of this paper also hold true in the particular case of isotropic media, i.e. 𝑵(𝒙)=n(x)𝑰\bm{N}(\bm{x})=n(x)\bm{I}, where n(x)n(x) is a positive piecewise smooth function in D¯\overline{D}. We take the time-harmonic electromagnetic incident fields 𝑬i\bm{E}^{i}, 𝑯i\bm{H}^{i} of the form:

𝑬i(𝒙,𝒅,𝒒)=ikcurl curl𝒒eik𝒙𝒅=ik(𝒅×𝒒)×𝒅eik𝒙𝒅,𝑯i(𝒙,𝒅,𝒒)=curl𝒒eik𝒙𝒅=ik𝒅×𝒒eik𝒙𝒅,\begin{split}\bm{E}^{i}(\bm{x},\bm{d},\bm{q})&=\frac{\mathrm{i}}{k}\text{curl curl}\,\bm{q}\mathrm{e}^{\mathrm{i}k\bm{x}\cdot\bm{d}}=\mathrm{i}k(\bm{d}\times\bm{q})\times\bm{d}\mathrm{e}^{\mathrm{i}k\bm{x}\cdot\bm{d}},\\ \bm{H}^{i}(\bm{x},\bm{d},\bm{q})&=\text{curl}\,\bm{q}\mathrm{e}^{\mathrm{i}k\bm{x}\cdot\bm{d}}=\mathrm{i}k\bm{d}\times\bm{q}\mathrm{e}^{\mathrm{i}k\bm{x}\cdot\bm{d}},\end{split} (1.3)

where i:=1\mathrm{i}:=\sqrt{-1} is the imaginary unit, 𝒅𝕊2:={x3:|𝒙|=1}\bm{d}\in\mathbb{S}^{2}:=\{x\in\mathbb{R}^{3}:|\bm{x}|=1\} is the direction of propagation, and 𝒒3\bm{q}\in\mathbb{R}^{3} is the polarization vector satisfying 𝒒𝒅\bm{q}\perp\bm{d}. The wave scattering of 𝑬i\bm{E}^{i}, 𝑯i\bm{H}^{i} by the medium scatterer (D,𝑵)(D,\bm{N}) leads to the following Maxwell equations for the interior electric and magnetic fields 𝑬,𝑯\bm{E},\bm{H}, and the scattered electric and magnetic fields 𝑬s,𝑯s\bm{E}^{s},\bm{H}^{s}:

{curl𝑬ik𝑯=0,curl𝑯+ik𝑵(𝒙)𝑬=0in D,curl𝑬sik𝑯s=0,curl𝑯s+ik𝑬s=0in 3D¯,𝝂×(𝑬s+𝑬i)𝝂×𝑬=0,𝝂×(𝑯s+𝑯i)𝝂×𝑯=0onD,lim|𝒙|(𝑯s×𝒙|𝒙|𝑬s)=0.\left\{\begin{aligned} &\text{curl}\,\bm{E}-\mathrm{i}k\bm{H}=0,\quad&&\text{curl}\,\bm{H}+\mathrm{i}k\bm{N}(\bm{x})\bm{E}=0\quad&&\text{in }D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ &\text{curl}\,\bm{E}^{s}-\mathrm{i}k\bm{H}^{s}=0,\quad&&\text{curl}\,\bm{H}^{s}+\mathrm{i}k\bm{E}^{s}=0\quad&&\text{in }\mathbb{R}^{3}\setminus\overline{D},\vskip 6.0pt plus 2.0pt minus 2.0pt\\ &\bm{\nu}\times(\bm{E}^{s}+\bm{E}^{i})-\bm{\nu}\times\bm{E}=0,\quad&&\bm{\nu}\times(\bm{H}^{s}+\bm{H}^{i})-\bm{\nu}\times\bm{H}=0\quad&&\text{on}\ \partial D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ &\lim\limits_{|\bm{x}|\to\infty}\left(\bm{H}^{s}\times\bm{x}-|\bm{x}|\bm{E}^{s}\right)=0.\end{aligned}\right. (1.4)

The last limit in (1.4) is known as the Silver-Müller radiation condition, which signifies that the scattered fields are outgoing. The well-posedness of the scattering problem (1.3)-(1.4) is established in [23] for the unique existence of a pair of solutions 𝑬,𝑯Hloc(curl;3)\bm{E},\bm{H}\in H_{loc}(\mathrm{curl};\mathbb{R}^{3}). If we eliminate the magnetic field in (1.4), the scattering system is reduced into:

{curl curl𝑬k2𝑵(𝒙)𝑬=0,in D,curl curl𝑬sk2𝑬s=0,in 3D¯,𝝂×(𝑬s+𝑬i)𝝂×𝑬=0,onD,𝝂×(curl𝑬s+curl𝑬i)𝝂×curl𝑬=0,onD,lim|𝒙|(curl𝑬s×𝒙ik|𝒙|𝑬s)=0.\left\{\begin{aligned} &\text{curl\,curl}\,\bm{E}-k^{2}\bm{N}(\bm{x})\bm{E}=0,\quad&&\text{in }D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ &\text{curl\,curl}\,\bm{E}^{s}-k^{2}\bm{E}^{s}=0,\quad&&\text{in }\mathbb{R}^{3}\setminus\overline{D},\vskip 6.0pt plus 2.0pt minus 2.0pt\\ &\bm{\nu}\times(\bm{E}^{s}+\bm{E}^{i})-\bm{\nu}\times\bm{E}=0,\quad&&\text{on}\ \partial D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ &\bm{\nu}\times(\text{curl}\,\bm{E}^{s}+\text{curl}\,\bm{E}^{i})-\bm{\nu}\times\text{curl}\,\bm{E}=0,\quad&&\text{on}\ \partial D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ &\lim\limits_{|\bm{x}|\to\infty}\left(\text{curl}\,\bm{E}^{s}\times\bm{x}-\mathrm{i}k|\bm{x}|\bm{E}^{s}\right)=0.\end{aligned}\right.

with the unit outward normal 𝝂\bm{\nu} on 𝕊2\mathbb{S}^{2}. The Silver-Müller radiation condition leads to the following asymptotic behavior of the scattered fields:

𝑸s(𝒙)=eik|𝒙||𝒙|{𝑸(𝒙^)+𝒪(1|𝒙|)},|𝒙|,𝑸:=𝑬,𝑯,\bm{Q}^{s}(\bm{x})=\frac{\mathrm{e}^{\mathrm{i}k|\bm{x}|}}{|\bm{x}|}\left\{\bm{Q}_{\infty}(\hat{\bm{x}})+\mathcal{O}\left(\frac{1}{|\bm{x}|}\right)\right\},\quad|\bm{x}|\to\infty,\quad\bm{Q}:=\bm{E},\bm{H},

uniformly in all directions 𝒙^=𝒙/|𝒙|𝕊2\hat{\bm{x}}=\bm{x}/|\bm{x}|\in\mathbb{S}^{2}. The vector fields 𝑬\bm{E}_{\infty} and 𝑯\bm{H}_{\infty} defined on the unit sphere 𝕊2\mathbb{S}^{2} are known as the electric far-field pattern and magnetic far-field pattern, respectively. They satisfy

𝑯=𝝂×𝑬and𝝂𝑬=𝝂𝑯=0.\bm{H}_{\infty}=\bm{\nu}\times\bm{E}_{\infty}\quad\text{and}\quad\bm{\nu}\cdot\bm{E}_{\infty}=\bm{\nu}\cdot\bm{H}_{\infty}=0.

The inverse scattering problem that we are concerned with is to recover the medium scatterer (D,𝑵)(D,\bm{N}) by knowledge of the associated far-field data, which can be formulated as the following nonlinear operator equation:

(D,𝑵)=𝑬(𝒙^,𝒅,𝒒;k),𝒙^𝕊2,𝒅𝕊2,𝒒3,kI:=[k0,k1],\mathcal{F}(D,\bm{N})=\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},\bm{q};k),\quad\hat{\bm{x}}\in\mathbb{S}^{2},\ \bm{d}\in\mathbb{S}^{2},\ \bm{q}\in\mathbb{R}^{3},\ k\in I:=[k_{0},k_{1}], (1.5)

where \mathcal{F} is defined by the Maxwell system (1.3)-(1.4). We shall mainly consider recovering the shape of the scatterer, namely DD, independent of its medium content 𝑵\bm{N}.

1.2. Overview of the proposed method and discussion

The inverse scattering problem (1.5) lies at the heart of many industrial applications including radar/sonar, medical imaging and nondestructive testing. Many methods have been developed for this practically important inverse problem and we refer to [2, 9, 16] and the references cited therein for the existing developments in the literature. In principle, most of the methods make use of the “visible patterns” for the reconstruction. Here, by visible patterns, we mean that the scattering data that one can observe in the far-field measurement. As a sharp contrast, we propose to reconstruct to scatterer by making use of “invisible patterns”, namely 𝑬=𝑯𝟎\bm{E}_{\infty}=\bm{H}_{\infty}\equiv\bm{0}.

If 𝑬=𝑯𝟎\bm{E}_{\infty}=\bm{H}_{\infty}\equiv\bm{0}, by Rellich’s theorem [16], one has 𝑬s=𝑯s=𝟎\bm{E}^{s}=\bm{H}^{s}=\bm{0} in 3\D¯\mathbb{R}^{3}\backslash\overline{D}. In such a case, one has by direct verifications that 𝑬|DHloc(curl,D)\bm{E}|_{D}\in H_{loc}(\mathrm{curl},D) and 𝑬0=𝑬i|DH(curl,D)\bm{E}_{0}=\bm{E}^{i}|_{D}\in H(\mathrm{curl},D) fulfill the following PDE system:

{curl curl𝑬k2𝑵(𝒙)𝑬=0inD,curl curl𝑬0k2𝑬0=0inD,𝝂×𝑬=𝝂×𝑬0onD,𝝂×curl𝑬=𝝂×curl𝑬0onD.\left\{\begin{array}[]{ll}\text{curl\,curl}\,\bm{E}-k^{2}\bm{N}(\bm{x})\bm{E}=0&\text{in}\ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \text{curl\,curl}\,\bm{E}_{0}-k^{2}\bm{E}_{0}=0&\text{in}\ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \bm{\nu}\times\bm{E}=\bm{\nu}\times\bm{E}_{0}&\text{on}\ \partial D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \bm{\nu}\times\text{curl}\,\bm{E}=\bm{\nu}\times\text{curl}\,\bm{E}_{0}&\text{on}\ \partial D.\end{array}\right. (1.6)

The system (1.6) is referred to as the Maxwell interior transmission eigenvalue problem (cf. [16]). That is, when invisibility occurs, the scattering patterns are trapped inside the scatterer DD (noting that outside the scatterer, 𝑬=𝑬i\bm{E}=\bm{E}^{i} and hence there is no scattering pattern to be observed). The proposed reconstruction scheme critically rely on the geometric properties of the transmission eigenfunctions to (1.6). Roughly speaking, we have the following global and local geometric patterns of the transmission eigenfunction 𝑬0\bm{E}_{0}: globally, its L2L^{2}-energy is localized around D\partial D; and locally, it is nearly vanishing around high-curvature point on D\partial D, and especially around boundary point where 𝝂\bm{\nu} is discontinuous which can be regarded as having infinite curvature. These global and local geometric patterns of the Maxwell transmission eigenfunctions have received considerable studies recently in the literature. Since it is not the focus of the current study, we refer to [8, 20, 21] for the theoretical and numerical justifications of those geometric patterns for the Maxwell transmission eigenfunctions, and we also refer to [3, 4, 5, 6, 7, 14, 18, 19, 24] for related studies on similar geometric patterns for the Helmholtz transmission eigenfunctions. Nevertheless, we would like to point out that all of the aforementioned geometric studies are concerned with transmission eigenfunctions associated with isotropic media. As an interesting byproduct of the current study, we numerically verify that the general geometric properties described above also hold for Maxwell transmission eigenfunctions associated with anisotropic media.

Based on the above discussion, the proposed method consists of three phases. First, we determine the interior Maxwell transmission eigenvalues of the unknown scatterer from the far-field data in (1.5) by the mechanism of the linear sampling method. Next, we determine the approximations of the corresponding transmission eigenfunctions by solving a constrained optimization problem. Finally, based on both global and local geometric properties of the transmission eigenfunctions, we design an imaging functional which can be used to determine the shape of the medium scatterer. Our study follows a similar spirit in a recent paper [14], where we developed an imaging scheme for the inverse acoustic scattering problem governed by the Helmholtz system. In this paper, we extend our study to the more complicated and challenging electromagnetic scattering problem and moreover associated with anisotropic media. In order to provide a rigorous basis for the proposed method, we establish several technical results that are of independent interest to the scattering theory of electromagnetic waves. In sharp contrast, the counterparts of these technical results required in [14] for the acoustic wave scattering were known in the literature.

We would like to mention that there are methods proposed for reconstructing 𝐍\mathbf{N} in (1.5) by making use of the transmission eigenvalues k2k^{2} associated with (1.6). Nevertheless, in the proposed method, we make use of the quantitative geometric properties of the transmission eigenfunctions for the reconstruction. This enables us to attain several salient features, in particular the super-resolution effect of the reconstructed images. Here, we recall the Abbe diffraction which states that in modern optics, the resolution limit is roughly around half of the operating wavelength. In our numerical study, it is shown that if DD possesses a thin layer of high refractive index, i.e. 𝑵\bm{N} is relatively large around D\partial D, then transmission eigenvalues to (1.6) can occur for even small kk, and moreover, the geometric patterns mentioned earlier are more evident for the corresponding transmission eigenfunctions. That means, in such a case, even if the (unknown) scatterer is of a size much smaller than the operating wavelength 2π/k2\pi/k (with kk being a transmission eigenvalue), the proposed method can still effectively reconstruct the shape of the scatterer. Hence, in practical applications, if one can coat the scattering object with a high refractive-index material (through indirect means), super-resolution imaging can be achieved via the proposed method. This is the main reason to suggest the title of our paper that invisibility enables super-visibility in electromagnetic imaging. In addition, the super-resolution effect can also be observed in the local reconstruction of geometrically singular places of D\partial D, say e.g. corners. Such a viewpoint has been advocated in [14] for the inverse acoustic imaging and we corroborate it in this paper for the inverse electromagnetic imaging.

The rest of the paper is organized as follows. In section 2, we discuss the determination of the Maxwell transmission eigenvalues by knowledge of the far-field data. In section 3, we determine the corresponding transmission eigenfunctions. Finally, in section 4, an imaging functional is designed and several numerical examples are presented.

2. Determination of transmission eigenvalues

In this section, we determine the transmission eigenvalues from a knowledge of a family of the electric far-field patterns. Although this problem has been solved in previous literature [10], for completeness and self-containedness, we briefly discuss the main process as well as the rationale behind the method.

For any given 𝒛3\bm{z}\in\mathbb{R}^{3}, let

𝑬e,(𝒙^,𝒛,𝒒;k)=ik4π(𝒙^×𝒒)×𝒙^eik𝒙^𝒛\bm{E}_{e,\infty}(\hat{\bm{x}},\bm{z},\bm{q};k)=\frac{\mathrm{i}k}{4\pi}(\hat{\bm{x}}\times\bm{q})\times\hat{\bm{x}}\mathrm{e}^{-\mathrm{i}k\hat{\bm{x}}\cdot\bm{z}}

be the far-field pattern of an electric dipole with source at 𝒛\bm{z} and polarization 𝒒\bm{q}. The determination of the Maxwell transmission eigenvalues is based on the so-called linear sampling method (LSM) [11], which is a qualitative method in inverse scattering theory. We now consider the following far-field equation

(Fk𝒈)(𝒙^)=𝑬e,(𝒙^,𝒛,𝒒;k),𝒛,𝒒3,kI,(F_{k}\bm{g})(\hat{\bm{x}})=\bm{E}_{e,\infty}(\hat{\bm{x}},\bm{z},\bm{q};k),\quad\bm{z},\bm{q}\in\mathbb{R}^{3},\ k\in I, (2.1)

where the far-field operator Fk:Lt2(𝕊2)Lt2(𝕊2)F_{k}:L_{t}^{2}(\mathbb{S}^{2})\to L_{t}^{2}(\mathbb{S}^{2}) is defined by

(Fk𝒈)(𝒙^):=𝕊2𝑬(𝒙^,𝒅,𝒈(𝒅);k)ds(𝒅),(F_{k}\bm{g})(\hat{\bm{x}}):=\int_{\mathbb{S}^{2}}\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},\bm{g}(\bm{d});k)\,\text{d}s(\bm{d}), (2.2)

with 𝑬(𝒙^,𝒅,𝒈(𝒅);k)\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},\bm{g}(\bm{d});k) the far-field pattern corresponding to the scattering problem (1.3)-(1.4). Since 𝒈(𝒅)\bm{g}(\bm{d}) is implicit in the far-field pattern 𝑬(𝒙^,𝒅,𝒈(𝒅);k)\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},\bm{g}(\bm{d});k), we can not solve the equation (2.1) directly.

Next, we present the numerical method to identify 𝒈(𝒅)\bm{g}(\bm{d}) in (2.1). For more details about the numerical calculation, we refer the reader to [11]. Let 𝒑3\bm{p}\in\mathbb{R}^{3} be a unit auxiliary vector such that 𝒑×𝒅0\bm{p}\times\bm{d}\neq 0 for any 𝒅𝕊2\bm{d}\in\mathbb{S}^{2}. Noting that 𝒅𝒈(𝒅)=0\bm{d}\cdot\bm{g}(\bm{d})=0, then the Herglotz kernel for each incident direction 𝒅\bm{d} can be expanded as

𝒈(𝒅)=gθ(𝒅)𝒑θ+gϕ(𝒅)𝒑ϕ,\bm{g}(\bm{d})=g^{\theta}(\bm{d})\,{\bm{p}}^{\theta}+g^{\phi}(\bm{d})\,{\bm{p}}^{\phi},

where gθ,gϕg^{\theta},g^{\phi}\in\mathbb{C} are undetermined coefficients and

𝒑θ=𝒑×𝒅|𝒑×𝒅|,𝒑ϕ=(𝒑×𝒅)×𝒅|(𝒑×𝒅)×𝒅|{\bm{p}}^{\theta}=\frac{\bm{p}\times\bm{d}}{|\bm{p}\times\bm{d}|},\quad{\bm{p}}^{\phi}=\frac{(\bm{p}\times\bm{d})\times\bm{d}}{|(\bm{p}\times\bm{d})\times\bm{d}|}

are two mutually orthogonal polarizations. Thus, the electric far-field pattern 𝑬\bm{E}_{\infty} can be represented by

𝑬(𝒙^,𝒅,𝒈(𝒅);k)=gθ𝑬(𝒙^,𝒅,𝒑θ;k)+gϕ𝑬(𝒙^,𝒅,𝒑ϕ;k).\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},\bm{g}(\bm{d});k)=g^{\theta}\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},{\bm{p}}^{\theta};k)+g^{\phi}\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},{\bm{p}}^{\phi};k).

Therefore, the far-field operator FkF_{k} can be rewritten as

(Fk𝒈)(𝒙^):=𝕊2(gθ(𝒅;𝒛,𝒒)𝑬(𝒙^,𝒅,𝒑θ;k)+gϕ(𝒅;𝒛,𝒒)𝑬(𝒙^,𝒅,𝒑ϕ;k))ds(𝒅).(F_{k}\bm{g})(\hat{\bm{x}}):=\int_{\mathbb{S}^{2}}\left(g^{\theta}(\bm{d};\bm{z},\bm{q})\,\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},{\bm{p}}^{\theta};k)+g^{\phi}(\bm{d};\bm{z},\bm{q})\,\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},{\bm{p}}^{\phi};k)\right)\,\mathrm{d}s(\bm{d}).

As the far-field equation (2.1) is ill posed, we usually adopt Tikhonov regularization [16] and instead solve

(δ𝑰+FkFk)𝒈δ=Fk𝑬e,(𝒙^,𝒛,𝒒;k),(\delta\bm{I}+F_{k}^{\ast}F_{k})\bm{g}_{\delta}=F_{k}^{\ast}\bm{E}_{e,\infty}(\hat{\bm{x}},\bm{z},\bm{q};k), (2.3)

where δ>0\delta>0 is the regularization parameter and FkF_{k}^{\ast} is the adjoint to FkF_{k}.

Now we introduce the rigorous mathematical justification of the LSM for determining the transmission eigenvalues. Define the Herglotz wave functions by

𝑬𝒈,k(𝒙):=𝕊2𝒈(𝒅)eik𝒙𝒅ds(𝒅),𝑯𝒈,k(𝒙):=1ikcurl𝑬𝒈,k(𝒙),\bm{E}_{\bm{g},k}(\bm{x}):=\int_{\mathbb{S}^{2}}\bm{g}(\bm{d})\mathrm{e}^{\mathrm{i}k\bm{x}\cdot\bm{d}}\text{d}s(\bm{d}),\quad\bm{H}_{\bm{g},k}(\bm{x}):=\frac{1}{\mathrm{i}k}\text{curl}\,\bm{E}_{\bm{g},k}(\bm{x}), (2.4)

where 𝒈Lt2(𝕊2)\bm{g}\in L_{t}^{2}(\mathbb{S}^{2}) is called the vector Herglotz kernel of the pair (𝑬𝒈,k,𝑯𝒈,k)(\bm{E}_{\bm{g},k},\bm{H}_{\bm{g},k}). The existence and discreteness of infinitely real transmission eigenvalues was established in [10]. Moreover, an effective discriminant method is proposed to distinguish whether kk is a Maxwell transmission eigenvalue or not in the following lemma[10].

Lemma 2.1.

Let 𝐠δ(,𝐳)Lt2(𝕊2)\bm{g}_{\delta}(\cdot,\bm{z})\in L_{t}^{2}(\mathbb{S}^{2}) be the Tikhonov regularized solution of the far-field equation, i.e., the solution of (2.3). Let 𝐄𝐠δ,k\bm{E}_{\bm{g}_{\delta},k} be the electric Herglotz wave function with kernel 𝐠δ\bm{g}_{\delta} defined in (2.4). Then, for any ball BDB\subset D, 𝐄𝐠δ,k(L2(D))3\|\bm{E}_{\bm{g}_{\delta},k}\|_{(L^{2}(D))^{3}} is bounded as δ0\delta\to 0 for a.e. 𝐳B\bm{z}\in B if and only if kk is not a transmission eigenvalue.

Remark 2.1.

There exists a constant M>0M>0 such that 𝑬𝒈δ,k(L2(D))3M𝒈δ(,𝒛)(L2(𝕊2))3\|\bm{E}_{\bm{g}_{\delta},k}\|_{(L^{2}(D))^{3}}\leq M\|\bm{g}_{\delta}(\cdot,\bm{z})\|_{(L^{2}(\mathbb{S}^{2}))^{3}}. Then, according to Lemma 2.1, we note that 𝒈δ(,𝒛)(L2(𝕊2))3\|\bm{g}_{\delta}(\cdot,\bm{z})\|_{(L^{2}(\mathbb{S}^{2}))^{3}} behaves quite differently whether kk is a Maxwell transmission eigenvalue or not. Although 𝑬𝒈δ,k(L2(D))3\|\bm{E}_{\bm{g}_{\delta},k}\|_{(L^{2}(D))^{3}} also behaves differently when kk is a transmission eigenvalue or not, as DD is unknown, we therefore use 𝒈δ(,𝒛)(L2(𝕊2))3\|\bm{g}_{\delta}(\cdot,\bm{z})\|_{(L^{2}(\mathbb{S}^{2}))^{3}} as an indicator to determine whether or not kk is a Maxwell transmission eigenvalue.

As discussed above, we formulate the following scheme and summarize the procedure of determining transmission eigenvalues.

Algorithm I: Transmission eigenvalues determination scheme
Step 1 Find a unit auxiliary vector 𝒑\bm{p} and collect a pair of electric far-field data 𝑬(𝒙^,𝒅,𝒑θ;k)\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},{\bm{p}}^{\theta};k) and 𝑬(𝒙^,𝒅,𝒑ϕ;k)\bm{E}_{\infty}(\hat{\bm{x}},\bm{d},{\bm{p}}^{\phi};k) for (𝒙^,𝒅,k)𝕊2×𝕊2×I(\hat{\bm{x}},\bm{d},k)\in\mathbb{S}^{2}\times\mathbb{S}^{2}\times I, where II is a finite interval in +\mathbb{R}_{+}.
Step 2 Pick a point 𝒛D\bm{z}\in D (a priori information), and take three independent artificial polarization vectors 𝒒1=[1,0,0],𝒒2=[0,1,0]\bm{q}_{1}=[1,0,0]^{\top},\bm{q}_{2}=[0,1,0]^{\top} and 𝒒3=[0,0,1]\bm{q}_{3}=[0,0,1]^{\top}. Then, for each kIk\in I, solve (2.3) to obtain the solutions 𝒈δ(,𝒛,𝒒1;k)\bm{g}_{\delta}(\cdot,\bm{z},\bm{q}_{1};k), 𝒈δ(,𝒛,𝒒2;k)\bm{g}_{\delta}(\cdot,\bm{z},\bm{q}_{2};k) and 𝒈δ(,𝒛,𝒒3;k)\bm{g}_{\delta}(\cdot,\bm{z},\bm{q}_{3};k).
Step 3 Plot =13𝒈δ(,𝒛,𝒒;k)(L2(𝕊2))32\sum_{\ell=1}^{3}\|\bm{g}_{\delta}(\cdot,\bm{z},\bm{q}_{\ell};k)\|_{(L^{2}(\mathbb{S}^{2}))^{3}}^{2} against kIk\in I and find the transmission eigenvalues where peaks appear in the graph.

3. Determination of transmission eigenfunctions

In section 2, we have determined a number of Maxwell transmission eigenvalues within the interval II by knowledge of a family of the electric far-field patterns in (1.5). Next, we are devoted to determining the corresponding transmission eigenfunctions. To that end, we first show the following denseness result of the Herglotz wave in a Sobolev space.

Lemma 3.1.

(see [11, 17]). The set of Herglotz wave functions 𝐄𝐠,k\bm{E}_{\bm{g},k} defined by (2.4) is dense in M(D)¯\overline{M(D)} with respect to the H(curl,D)H(\mathrm{curl},D) norm, where

M(D):={𝑬C2(D)C1(D¯):curl curl𝑬=k2𝑬inD}.M(D):=\left\{\bm{E}\in C^{2}(D)\cap C^{1}(\overline{D}):\text{curl\,curl}\,\bm{E}=k^{2}\bm{E}\ \ \mathrm{in}\ D\right\}.

The following theorem states that if k+k\in\mathbb{R}_{+} is a Maxwell transmission eigenvalue, then there exists a Herglotz wave function 𝑬𝒈ϵ,k\bm{E}_{\bm{g}_{\epsilon},k} such that the scattered field corresponding to this 𝑬𝒈ϵ,k\bm{E}_{\bm{g}_{\epsilon},k} as the incident field is nearly vanishing.

Theorem 3.1.

Suppose that k+k\in\mathbb{R}_{+} is a Maxwell transmission eigenvalue in DD. For any sufficiently small ϵ+\epsilon\in\mathbb{R}_{+}, there exists 𝐠ϵLt2(𝕊2)\bm{g}_{\epsilon}\in L_{t}^{2}(\mathbb{S}^{2}) such that

Fk𝒈ϵ(L2(𝕊2))3=𝒪(ϵ)and𝑬𝒈ϵ,kH(curl,D)=1+𝒪(ϵ),\|F_{k}\bm{g}_{\epsilon}\|_{(L^{2}(\mathbb{S}^{2}))^{3}}=\mathcal{O}(\epsilon)\ \ \text{and}\ \ \|\bm{E}_{\bm{g}_{\epsilon},k}\|_{H(\mathrm{curl},D)}=1+\mathcal{O}(\epsilon),

where FkF_{k} is the far-field operator defined by (2.2) and 𝐄𝐠ϵ,k\bm{E}_{\bm{g}_{\epsilon},k} is the Herglotz wave function defined by (2.4) with the kernel 𝐠ϵ\bm{g}_{\epsilon}.

Proof.

Let 𝑬0,k\bm{E}_{0,k} be a normalized Maxwell transmission eigenfunction in DD associated with the Maxwell transmission eigenvalue kk, which means that 𝑬0,k\bm{E}_{0,k} with 𝑬0,kH(curl,D)=1\|\bm{E}_{0,k}\|_{H(\mathrm{curl},D)}=1 is a solution of

curl curl𝑬0,kk2𝑬0,k=0in D.\text{curl curl}\,\bm{E}_{0,k}-k^{2}\bm{E}_{0,k}=0\quad\text{in }D.

By Lemma 3.1, for any sufficiently small ϵ>0\epsilon>0, there exists 𝒈ϵLt2(𝕊2)\bm{g}_{\epsilon}\in L_{t}^{2}(\mathbb{S}^{2}) such that

𝑬𝒈ϵ,k𝑬0,kH(curl,D)<ϵ,\|\bm{E}_{\bm{g}_{\epsilon},k}-\bm{E}_{0,k}\|_{H(\mathrm{curl},D)}<\epsilon,

where 𝑬𝒈ϵ,k\bm{E}_{\bm{g}_{\epsilon},k} is the Herglotz wave function with the kernel 𝒈ϵ\bm{g}_{\epsilon}. Then, by the triangle inequality,

𝑬𝒈ϵ,kH(curl,D)𝑬𝒈ϵ,k𝑬0,kH(curl,D)+𝑬0,kH(curl,D)<ϵ+𝑬0,kH(curl,D),\|\bm{E}_{\bm{g}_{\epsilon},k}\|_{H(\mathrm{curl},D)}\leq\|\bm{E}_{\bm{g}_{\epsilon},k}-\bm{E}_{0,k}\|_{H(\mathrm{curl},D)}+\|\bm{E}_{0,k}\|_{H(\mathrm{curl},D)}<\epsilon+\|\bm{E}_{0,k}\|_{H(\mathrm{curl},D)},

and

𝑬𝒈ϵ,kH(curl,D)𝑬0,kH(curl,D)𝑬𝒈ϵ,k𝑬0,kH(curl,D)>𝑬0,kH(curl,D)ϵ.\|\bm{E}_{\bm{g}_{\epsilon},k}\|_{H(\mathrm{curl},D)}\geq\|\bm{E}_{0,k}\|_{H(\mathrm{curl},D)}-\|\bm{E}_{\bm{g}_{\epsilon},k}-\bm{E}_{0,k}\|_{H(\mathrm{curl},D)}>\|\bm{E}_{0,k}\|_{H(\mathrm{curl},D)}-\epsilon.

So, one has that 𝑬𝒈ϵ,kH(curl,D)=1+𝒪(ϵ)\|\bm{E}_{\bm{g}_{\epsilon},k}\|_{H(\mathrm{curl},D)}=1+\mathcal{O}(\epsilon).

Besides, from the definition of the far-field operator, Fk𝒈ϵF_{k}\bm{g}_{\epsilon} is the far-field pattern produced by the incident wave 𝑬𝒈ϵ,k\bm{E}_{\bm{g}_{\epsilon},k}. According to Theorem 2.5 in [26], one obtains that

Fk𝒈ϵ(L2(𝕊2))3<Cϵ,\|F_{k}\bm{g}_{\epsilon}\|_{(L^{2}(\mathbb{S}^{2}))^{3}}<C\epsilon,

where C=C(D,k)C=C(D,k) is a positive constant.

The proof is complete. ∎

On the basis of Theorem 3.1 and normalization if necessary, we can say that the following optimization problem:

min𝒈Lt2(𝕊2)Fk𝒈(L2(𝕊2))3 s.t. 𝑬𝒈,kH(curl,D)=1\min\limits_{\bm{g}\in L_{t}^{2}(\mathbb{S}^{2})}\|F_{k}\bm{g}\|_{(L^{2}(\mathbb{S}^{2}))^{3}}\quad\text{ s.t. }\|\bm{E}_{\bm{g},k}\|_{H(\mathrm{curl},D)}=1

exists at least one solution 𝒈0Lt2(𝕊2)\bm{g}_{0}\in L_{t}^{2}(\mathbb{S}^{2}) when k+k\in\mathbb{R}_{+} is a Maxwell transmission eigenvalue in DD. Unfortunately, since DD is unknown, the constraint 𝑬𝒈,kH(curl,D)=1\|\bm{E}_{\bm{g},k}\|_{H(\mathrm{curl},D)}=1 in the problem above is unpractical. However, it is reasonable to address this issue by considering an alternative optimization problem:

min𝒈Lt2(𝕊2)Fk𝒈(L2(𝕊2))3 s.t. 𝑬𝒈,kH(curl,Ω)=1,\min\limits_{\bm{g}\in L_{t}^{2}(\mathbb{S}^{2})}\|F_{k}\bm{g}\|_{(L^{2}(\mathbb{S}^{2}))^{3}}\quad\text{ s.t. }\|\bm{E}_{\bm{g},k}\|_{H(\mathrm{curl},\Omega)}=1, (3.1)

where Ω\Omega is an a priori ball containing DD.

Let 𝒈0\bm{g}_{0} be a “satisfactory” solution to the optimization problem (3.1). Next, we illustrate that the Herglotz wave 𝑬𝒈0,k\bm{E}_{\bm{g}_{0},k} with the Herglotz kernel 𝒈0\bm{g}_{0} is indeed an approximation to the Maxwell transmission eigenfunction 𝑬0,k\bm{E}_{0,k} associated with the transmission eigenvalue kk. Before further discussion, we give a quantitative Rellich’s result for the electromagnetic medium scattering, which is the generalization of the argument in the acoustic scattering case [5].

Lemma 3.2.

Let R>1,ρ>0R>1,\rho>0, and 𝐄i,𝐇iWloc1,3+ρ(div;3)Hloc(curl;3)\bm{E}^{i},\bm{H}^{i}\in W^{1,3+\rho}_{loc}(\mathrm{div};\mathbb{R}^{3})\cap H_{loc}(\mathrm{curl};\mathbb{R}^{3}) be the incident fields, satisfying

curl𝑬iik𝑯i=0,curl𝑯i+ik𝑬i=0,\text{curl}\,\bm{E}^{i}-\mathrm{i}k\bm{H}^{i}=0,\ \ \text{curl}\,\bm{H}^{i}+\mathrm{i}k\bm{E}^{i}=0,

with 𝐄iW1,3+ρ(div;B2R)\|\bm{E}^{i}\|_{W^{1,3+\rho}(\mathrm{div};B_{2R})}\leq\mathcal{I}. Besides, let 𝐄,𝐇Wloc1,3+ρ(div;3)Hloc(curl;3)\bm{E},\bm{H}\in W^{1,3+\rho}_{loc}(\mathrm{div};\mathbb{R}^{3})\cap H_{loc}(\mathrm{curl};\mathbb{R}^{3}) be the total electromagnetic fields satisfying

curl𝑬ik𝑯=0,curl𝑯+ik𝑵(𝒙)𝑬=0,\text{curl}\,\bm{E}-\mathrm{i}k\bm{H}=0,\ \ \text{curl}\,\bm{H}+\mathrm{i}k\bm{N}(\bm{x})\bm{E}=0,

where the matrix index of refraction 𝐍(𝐱)\bm{N}(\bm{x}) is defined in (1.1). The scattered electric and magnetic fields 𝐄s=𝐄𝐄i,𝐇s=𝐇𝐇i\bm{E}^{s}=\bm{E}-\bm{E}^{i},\bm{H}^{s}=\bm{H}-\bm{H}^{i} satisfy the Silver–Müller radiation condition. Let 𝐄,𝐇\bm{E}_{\infty},\bm{H}_{\infty} be their far-field patterns. We further assume that 𝐄s(W1,3+ρ(B2R))3𝒯\|\bm{E}^{s}\|_{(W^{1,3+\rho}(B_{2R}))^{3}}\leq\mathcal{T}. There is ϵ0>0\epsilon_{0}>0 such that if

𝑬(L2(𝕊2))3ϵ0,\|\bm{E}_{\infty}\|_{(L^{2}(\mathbb{S}^{2}))^{3}}\leq\epsilon_{0},

then for some positive constant βρ3+ρ\beta\leq\frac{\rho}{3+\rho} depending on R,𝐍R,\bm{N}, there is

supD|𝑬s|𝒞(lnln(𝒯𝑬(L2(𝕊2))31))β\sup_{\partial D}|\bm{E}^{s}|\leq\mathcal{C}\left(\mathrm{ln\,ln}(\mathcal{T}\|\bm{E}_{\infty}\|_{(L^{2}(\mathbb{S}^{2}))^{3}}^{-1})\right)^{-\beta}

for some 𝒞=C(𝐍,ω,R)(𝒯+)\mathcal{C}=C(\bm{N},\omega,R)(\mathcal{T}+\mathcal{I}).

Proof.

In what follows, we assume that DBRD\subset B_{R}, where BRB_{R} is a central ball of radius R+R\in\mathbb{R}_{+}. Clearly, the scattered electric field 𝑬s\bm{E}^{s} satisfies the vector Helmholtz equation outside DD. That is to say, the Cartesian components of 𝑬s\bm{E}^{s} satisfy the scalar Helmholtz equation outside DD. According to Sobolev embedding and the far-field to near-field estimate for the acoustic scattering (see e.g. Corollary 5.3 in [5]), there are constants C,c>0C,c>0 depending on k,R,Bk,R,B such that

EjsL(B)EjsW1,3+ρ(B)C max(E,jL2(𝕊2),𝒯ecln(𝒯/E,jL2(𝕊2))),\|E^{s}_{j}\|_{L^{\infty}(B)}\leq\|E^{s}_{j}\|_{W^{1,3+\rho}(B)}\leq C\mbox{ max}\left(\|E_{\infty,j}\|_{L^{2}(\mathbb{S}^{2})},\mathcal{T}\mathrm{e}^{-c\sqrt{\mbox{ln}\left(\mathcal{T}/\|E_{\infty,j}\|_{L^{2}(\mathbb{S}^{2})}\right)}}\right), (3.2)

where the elements with subscripts j=1,2,3j=1,2,3 stand for the Cartesian components of the corresponding vectors, and BB is a domain such that B¯B2RB¯R\overline{B}\subset B_{2R}\setminus\overline{B}_{R}. One of our requirements on ϵ0\epsilon_{0} is that the maximum picks the number on the right side of (3.2). Therefore we must have ϵ0𝒯ec2\epsilon_{0}\leq\mathcal{T}\mathrm{e}^{-c^{2}}.

Besides, we assume that 𝑬s(L(B(2λ)RB(1+λ)R))3γ1\|\bm{E}^{s}\|_{(L^{\infty}(B_{(2-\lambda)R}\setminus B_{(1+\lambda)R}))^{3}}\leq\gamma\leq 1. Then according to Proposition 5.7 in [5],

|𝑬s(𝒙)|C𝒯γc2(2+λ)R/r+2|\bm{E}^{s}(\bm{x})|\leq C\mathcal{T}\gamma^{c_{2}^{(2+\lambda)R/r+2}} (3.3)

for 𝒙B2RB(D,4r)\bm{x}\in B_{2R}\setminus B(D,4r), and some positive constants C,c2<1/4,1/4<λ<1/2,r<(12λ)R/2C,c_{2}<1/4,1/4<\lambda<1/2,r<(1-2\lambda)R/2. Here, B(D,4r)=𝒙DB(𝒙,4r)B(D,4r)=\cup_{\bm{x}\in D}B(\bm{x},4r). Then, when we take 𝒙B3R/2B(D,4r)\bm{x}^{\prime}\in B_{3R/2}\setminus B(D,4r), by the Hölder continuity of 𝑬s\bm{E}^{s} (see e.g. Theorem 4 in [1]) and (3.3), there exists 0<βρ3+ρ0<\beta\leq\frac{\rho}{3+\rho} depending on R,𝑵R,\bm{N} such that

|𝑬s(𝒙)|𝑬sC0,β(B¯3R/2)|𝒙𝒙|β+|𝑬s(𝒙)|C8βrβ(𝑬s(L2(B3R/2))3+𝑬s(W1,3+ρ(B3R/2))3+𝑬iW1,3+ρ(div;B3R/2))+C𝒯γc2(2+λ)R/r+2C(𝒯+)8βrβ+C𝒯γc2(2+λ)R/r+2\begin{split}|\bm{E}^{s}(\bm{x})|\leq&\|\bm{E}^{s}\|_{C^{0,\beta}(\overline{B}_{3R/2})}|\bm{x}-\bm{x}^{\prime}|^{\beta}+|\bm{E}^{s}(\bm{x}^{\prime})|\\ \leq&C8^{\beta}r^{\beta}\left(\|\bm{E}^{s}\|_{(L^{2}(B_{3R/2}))^{3}}+\|\bm{E}^{s}\|_{(W^{1,3+\rho}(B_{3R/2}))^{3}}+\|\bm{E}^{i}\|_{W^{1,3+\rho}(\mathrm{div};B_{3R/2})}\right)\\ &+C\mathcal{T}\gamma^{c_{2}^{(2+\lambda)R/r+2}}\\ \leq&C(\mathcal{T}+\mathcal{I})8^{\beta}r^{\beta}+C\mathcal{T}\gamma^{c_{2}^{(2+\lambda)R/r+2}}\end{split} (3.4)

for d(𝒙,D)4rd(\bm{x},\partial D)\leq 4r. The constant CC depends on R,𝑵,ωR,\bm{N},\omega.

Next, we require

γ=C𝒯ecln(𝒯/ϵ0)<1/exp exp(C(λ,k,R))\gamma=C\mathcal{T}\mathrm{e}^{-c\sqrt{\mbox{ln}(\mathcal{T}/\epsilon_{0})}}<1/\text{exp exp}\,(C(\lambda,k,R))

and

r=r((ln|lnγ|)1,λ,k,R).r=r\left((\mbox{ln}|\mbox{ln}\gamma|)^{-1},\lambda,k,R\right).

If we further choose λ=λ(R)\lambda=\lambda(R) and let ϵ0\epsilon_{0} small enough, the claim follows from (3.4).

Remark 3.1.

Lemma  3.2 shows that the smallness of the far-field pattern can ensure that the scattered field on the boundary are also small.

Remark 3.2.

We note that Lemma  3.2 is analyzed in terms of electric fields. If we do the same analysis for the magnetic fields, we can get similar results, that is |𝑯s|ψ(ϵ0)|\bm{H}^{s}|\leq\psi(\epsilon_{0}) on D\partial D if 𝑬(L2(𝕊2))3ϵ0\|\bm{E}_{\infty}\|_{(L^{2}(\mathbb{S}^{2}))^{3}}\leq\epsilon_{0}. Here, ψ\psi is a double logarithmic type function and satisfies ψ(ϵ0)0\psi(\epsilon_{0})\to 0 as ϵ0+0\epsilon_{0}\to+0.

Remark 3.3.

We would like to emphasize that the quantitative Rellich’s result for the electromagnetic case is verified here for the first time. Actually, there is a higher requirement for the regularity of the scatterer [5]. However, we are not overly concerned about this issue, which is not the focus of this paper. Nevertheless, we believe that this regularity assumption can be relaxed. In fact, the subsequent numerical examples demonstrate the effectiveness of our method even if D\partial D is only Lipschitz continuous.

In the following theorem, we demonstrate that the Herglotz wave 𝑬𝒈0,k\bm{E}_{\bm{g}_{0},k} is indeed generically an approximation to the Maxwell transmission eigenfunction 𝑬0,k\bm{E}_{0,k}.

Theorem 3.2.

Suppose that the assumptions in Lemma  3.2 hold. Additionally, we suppose that k+k\in\mathbb{R}_{+} is a Maxwell transmission eigenvalue in DD and 𝐠0\bm{g}_{0} is a solution to the optimization problem (3.1) satisfying

Fk𝒈0(L2(𝕊2))3ϵ1.\|F_{k}\bm{g}_{0}\|_{(L^{2}(\mathbb{S}^{2}))^{3}}\leq\epsilon\ll 1. (3.5)

Then the Herglotz wave 𝐄𝐠0,k\bm{E}_{\bm{g}_{0},k} is an approximation to a transmission eigenfunction 𝐄0,k\bm{E}_{0,k} associated with the transmission eigenvalue kk in (L2(D))3(L^{2}(D))^{3}-norm.

Proof.

Consider the scattering system (1.3)-(1.4). Let 𝑬i=𝑬𝒈0,k\bm{E}^{i}=\bm{E}_{\bm{g}_{0},k}, 𝑬s\bm{E}^{s} and 𝑬t\bm{E}^{t} be, respectively, the incident, scattered and total electric fields. After expressing the magnetic fields in terms of the electric fields, one has

{curl curl𝑬tk2𝑵(𝒙)𝑬t=0inD,curl curl𝑬𝒈0,kk2𝑬𝒈0,k=0inD,𝝂×(𝑬𝒈0,k+𝑬s)=𝝂×𝑬tonD,𝝂×(curl𝑬𝒈0,k+curl𝑬s)=𝝂×curl𝑬tonD.\left\{\begin{array}[]{ll}\text{curl\,curl}\,\bm{E}^{t}-k^{2}\bm{N}(\bm{x})\bm{E}^{t}=0&\text{in}\ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \text{curl\,curl}\,\bm{E}_{\bm{g}_{0},k}-k^{2}\bm{E}_{\bm{g}_{0},k}=0&\text{in}\ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \bm{\nu}\times(\bm{E}_{\bm{g}_{0},k}+\bm{E}^{s})=\bm{\nu}\times\bm{E}^{t}&\text{on}\ \partial D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \bm{\nu}\times(\text{curl}\,\bm{E}_{\bm{g}_{0},k}+\text{curl}\,\bm{E}^{s})=\bm{\nu}\times\text{curl}\,\bm{E}^{t}&\text{on}\ \partial D.\end{array}\right. (3.6)

According to our earlier discussion, Fk𝒈0F_{k}\bm{g}_{0} is the far-field pattern of 𝑬s\bm{E}^{s}. By virtue of Lemma  3.2 and (3.5), there is

|𝝂×𝑬s|ψ(ϵ) and |𝝂×curl𝑬s|ψ(ϵ)on D,|\bm{\nu}\times\bm{E}^{s}|\leq\psi(\epsilon)\text{ and }|\bm{\nu}\times\text{curl}\,\bm{E}^{s}|\leq\psi(\epsilon)\quad\text{on }\partial D, (3.7)

where ψ\psi is the stability function, which is of double logarithmic type and satisfies ψ(ϵ)0\psi(\epsilon)\to 0 as ϵ+0\epsilon\to+0. For more details about the quantitative Rellich’s theorem and the stability function ψ\psi, we recommend that readers refer to [5].

Recall the Maxwell interior transmission eigenvalue problem (1.6). Setting

𝒖=𝑬𝑬0and𝒗=𝑵(𝒙)𝑬𝑬0,\bm{u}=\bm{E}-\bm{E}_{0}\quad\text{and}\quad\bm{v}=\bm{N}(\bm{x})\bm{E}-\bm{E}_{0}, (3.8)

one has the following fourth order PDE of 𝒖\bm{u}:

(curl curlk2𝑵)(𝑵𝑰)1(curl curl𝒖k2𝒖)=0inD.(\text{curl curl}-k^{2}\bm{N})(\bm{N}-\bm{I})^{-1}(\text{curl curl}\,\bm{u}-k^{2}\bm{u})=0\quad\text{in}\ D.

After using a variational framework, the system (1.6) becomes

𝒜k(𝒖,𝒖)k2(𝒖,𝒖)=0, for any 𝒖H02(curl,D),\mathcal{A}_{k}(\bm{u},\bm{u}^{\prime})-k^{2}\mathcal{B}(\bm{u},\bm{u}^{\prime})=0,\text{ for any }\bm{u}^{\prime}\in H_{0}^{2}(\mathrm{curl},D), (3.9)

where 𝒜k\mathcal{A}_{k} and \mathcal{B} are continuous sesquilinear forms on H2(curl,D)×H2(curl,D)H^{2}(\mathrm{curl},D)\times H^{2}(\mathrm{curl},D), defined by

𝒜k(𝒖,𝒖)=((𝑵𝑰)1(curl curl𝒖k2𝒖),curl curl𝒖k2𝒖)D+k4(𝒖,𝒖)D,\mathcal{A}_{k}(\bm{u},\bm{u}^{\prime})=\left((\bm{N}-\bm{I})^{-1}(\text{curl curl}\,\bm{u}-k^{2}\bm{u}),\text{curl curl}\,\bm{u}^{\prime}-k^{2}\bm{u}^{\prime}\right)_{D}+k^{4}(\bm{u},\bm{u}^{\prime})_{D},

and

(𝒖,𝒖)=(curl𝒖,curl𝒖)D.\mathcal{B}(\bm{u},\bm{u}^{\prime})=(\text{curl}\,\bm{u},\text{curl}\,\bm{u}^{\prime})_{D}.

With the help of the Riesz representation theorem, we define two bounded linear operators Ak:H02(curl,D)H02(curl,D)A_{k}:H_{0}^{2}(\mathrm{curl},D)\to H_{0}^{2}(\mathrm{curl},D) and B:H02(curl,D)H02(curl,D)B:H_{0}^{2}(\mathrm{curl},D)\to H_{0}^{2}(\mathrm{curl},D) by

(Ak𝒖,𝒖)H2(curl,D):=𝒜k(𝒖,𝒖)and(B𝒖,𝒖)H2(curl,D):=(𝒖,𝒖).(A_{k}\bm{u},\bm{u}^{\prime})_{H^{2}(\mathrm{curl},D)}:=\mathcal{A}_{k}(\bm{u},\bm{u}^{\prime})\quad\text{and}\quad(B\bm{u},\bm{u}^{\prime})_{H^{2}(\mathrm{curl},D)}:=\mathcal{B}(\bm{u},\bm{u}^{\prime}).

Next, we consider the scattering system (3.6), by (3.7) and some analytical regularization techniques (i.e., mollifiers), we could take 𝜻H2(curl,D)\bm{\zeta}\in H^{2}(\mathrm{curl},D) such that

𝝂×𝜻=𝝂×𝑬s,𝝂×curl𝜻=𝝂×curl𝑬son D,\bm{\nu}\times\bm{\zeta}=\bm{\nu}\times\bm{E}^{s},\quad\bm{\nu}\times\text{curl}\,\bm{\zeta}=\bm{\nu}\times\text{curl}\,\bm{E}^{s}\quad\text{on }\partial D, (3.10)

and

𝜻H2(curl,D)Cψ(ϵ),\|\bm{\zeta}\|_{H^{2}(\mathrm{curl},D)}\leq C\psi(\epsilon), (3.11)

where CC is a constant depending on DD. Setting 𝑬~:=𝑬t𝜻H2(curl,D)\widetilde{\bm{E}}:=\bm{E}^{t}-\bm{\zeta}\in H^{2}(\mathrm{curl},D), the system (3.6) could be rewritten as

{curl curl𝑬~k2𝑵(𝒙)𝑬~=(curl curlk2𝑵(𝒙))𝜻inD,curl curl𝑬𝒈0,kk2𝑬𝒈0,k=0inD,𝝂×𝑬~=𝝂×𝑬𝒈0,konD,𝝂×curl𝑬~=𝝂×curl𝑬𝒈0,konD.\left\{\begin{array}[]{ll}\text{curl curl}\,\widetilde{\bm{E}}-k^{2}\bm{N}(\bm{x})\widetilde{\bm{E}}=-(\text{curl curl}-k^{2}\bm{N}(\bm{x}))\bm{\zeta}&\text{in}\ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \text{curl curl}\,\bm{E}_{\bm{g}_{0},k}-k^{2}\bm{E}_{\bm{g}_{0},k}=0&\text{in}\ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \bm{\nu}\times\widetilde{\bm{E}}=\bm{\nu}\times\bm{E}_{\bm{g}_{0},k}&\text{on}\ \partial D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \bm{\nu}\times\text{curl}\,\widetilde{\bm{E}}=\bm{\nu}\times\text{curl}\,\bm{E}_{\bm{g}_{0},k}&\text{on}\ \partial D.\end{array}\right. (3.12)

Then, we take 𝒖~:=𝑬~𝑬𝒈0,kH02(curl,D)\widetilde{\bm{u}}:=\widetilde{\bm{E}}-\bm{E}_{\bm{g}_{0},k}\in H_{0}^{2}(\mathrm{curl},D). According to the variational formulae, the PDE system (3.12) can be rewritten as

(Ak𝒖~,𝒖)k2(B𝒖~,𝒖)=𝔣(𝒖),for any 𝒖H02(curl,D),(A_{k}\widetilde{\bm{u}},\bm{u}^{\prime})-k^{2}(B\widetilde{\bm{u}},\bm{u}^{\prime})=\mathfrak{f}(\bm{u}^{\prime}),\quad\text{for any }\bm{u}^{\prime}\in H_{0}^{2}(\mathrm{curl},D), (3.13)

where 𝔣(𝒖)\mathfrak{f}(\bm{u}^{\prime}) is an antilinear form on H2(curl,D)H^{2}(\mathrm{curl},D), defined by

𝔣(𝒖)=((𝑵𝑰)1(curl curlk2𝑵(𝒙))𝜻,(curl curlk2)𝒖)D.\mathfrak{f}(\bm{u}^{\prime})=(-(\bm{N}-\bm{I})^{-1}(\text{curl curl}-k^{2}\bm{N}(\bm{x}))\bm{\zeta},(\text{curl curl}-k^{2})\bm{u}^{\prime})_{D}. (3.14)

Since we have already known that AkA_{k} is an isomorphism, and BB is compact (see e.g. Chapter 4 in [11]), the Fredholm alternative could be applied to equations (3.9) and (3.13). Now, we need to prove that the operator equation (3.13) is solvable in the quotient space H02(curl,D)/𝕌H_{0}^{2}(\mathrm{curl},D)/\mathbb{U}, where 𝕌\mathbb{U} is the eigenspace of the system (3.9). We denote by 𝒇H02(curl,D)\bm{f}\in H_{0}^{2}(\mathrm{curl},D) the Riesz representation of 𝔣\mathfrak{f} in H02(curl,D)H_{0}^{2}(\mathrm{curl},D). Next, we illustrate that

𝒇ker[(Akk2B)]\bm{f}\in\mathrm{ker}[(A_{k}-k^{2}B)^{*}]^{\bot}

where (Akk2B)(A_{k}-k^{2}B)^{*} is adjoint to Akk2BA_{k}-k^{2}B. If 𝒗ker(Akk2B)\bm{v}\in\mathrm{ker}(A_{k}-k^{2}B)^{*}, then for any 𝒘H2(curl,D)\bm{w}\in H^{2}(\mathrm{curl},D), there is

(𝒘,(Akk2B)𝒗)H2(curl,D)=0,\left(\bm{w},(A_{k}-k^{2}B)^{*}\bm{v}\right)_{H^{2}(\mathrm{curl},D)}=0,

which means

((𝑵𝑰)1(curl curlk2)𝒘,(curl curlk2)𝒗)D+k4(𝒘,𝒗)Dk2(curl𝒘,curl𝒗)D=0.\left((\bm{N}-\bm{I})^{-1}(\text{curl curl}-k^{2})\bm{w},(\text{curl curl}-k^{2})\bm{v}\right)_{D}+k^{4}(\bm{w},\bm{v})_{D}-k^{2}(\text{curl}\,\bm{w},\text{curl}\,\bm{v})_{D}=0. (3.15)

Then we shall show that (𝒗,𝒇)H2(curl,D)=0(\bm{v},\bm{f})_{H^{2}(\mathrm{curl},D)}=0. Since 𝒇\bm{f} is the Riesz representation of 𝔣\mathfrak{f}, applying (3.15), one has

(𝒗,𝒇)H2(curl,D)=𝔣(𝒗)=((𝑵𝑰)1(curl curlk2𝑵(𝒙))𝜻,(curl curlk2)𝒗)D=((𝑵𝑰)1(curl curlk2)𝜻(𝑵𝑰)1k2(𝑵𝑰)𝜻,(curl curlk2)𝒗)D=((𝑵𝑰)1(curl curlk2)𝜻,(curl curlk2)𝒗)D+(k2𝜻,(curl curlk2)𝒗)D=((𝑵𝑰)1(curl curlk2)𝜻,(curl curlk2)𝒗)D+k2(curl𝜻,curl𝒗)Dk4(𝜻,𝒗)D=0,\begin{split}(\bm{v},&\bm{f})_{H^{2}(\mathrm{curl},D)}\\ &=\mathfrak{f}(\bm{v})\\ &=\left(-(\bm{N}-\bm{I})^{-1}(\text{curl curl}-k^{2}\bm{N}(\bm{x}))\bm{\zeta},(\text{curl curl}-k^{2})\bm{v}\right)_{D}\\ &=-\left((\bm{N}-\bm{I})^{-1}(\text{curl curl}-k^{2})\bm{\zeta}-(\bm{N}-\bm{I})^{-1}k^{2}(\bm{N}-\bm{I})\bm{\zeta},(\text{curl curl}-k^{2})\bm{v}\right)_{D}\\ &=-\left((\bm{N}-\bm{I})^{-1}(\text{curl curl}-k^{2})\bm{\zeta},(\text{curl curl}-k^{2})\bm{v}\right)_{D}+(k^{2}\bm{\zeta},(\text{curl curl}-k^{2})\bm{v})_{D}\\ &=-\left((\bm{N}-\bm{I})^{-1}(\text{curl curl}-k^{2})\bm{\zeta},(\text{curl curl}-k^{2})\bm{v}\right)_{D}+k^{2}(\text{curl}\,\bm{\zeta},\text{curl}\,\bm{v})_{D}-k^{4}(\bm{\zeta},\bm{v})_{D}\\ &=0,\end{split}

where in the second last equality, Green’s vector theorem and Gauss’s divergence theorem have been used to get

(𝜻,curl curl𝒗)D=(curl𝜻,curl𝒗)D.(\bm{\zeta},\text{curl curl}\,\bm{v})_{D}=(\text{curl}\,\bm{\zeta},\text{curl}\,\bm{v})_{D}.

Therefore, the equation (3.13) is solvable in the quotient space H02(curl,D)/𝕌H_{0}^{2}(\mathrm{curl},D)/\mathbb{U}. Set

𝒖^=(Akk2B)1𝒇in H02(curl,D)/𝕌.\hat{\bm{u}}=(A_{k}-k^{2}B)^{-1}\bm{f}\quad\text{in }H_{0}^{2}(\mathrm{curl},D)/\mathbb{U}. (3.16)

Noting (3.11) and (3.14), one obtains

𝒇(L2(D))3Cψ(ϵ),\|\bm{f}\|_{(L^{2}(D))^{3}}\leq C\psi(\epsilon), (3.17)

where CC is a constant depending on 𝑵\bm{N}, kk and DD. Combining (3.11), (3.16) and (3.17), one has

𝑬0,k𝑬𝒈0,k+𝑬t𝑬k(L2(D))3𝒖~𝒖(L2(D))3+𝜻(L2(D))3Cψ(ϵ)0asϵ+0.\begin{split}\|\bm{E}_{0,k}-\bm{E}_{\bm{g}_{0},k}+\bm{E}^{t}-\bm{E}_{k}\|_{(L^{2}(D))^{3}}\leq&\|\widetilde{\bm{u}}-\bm{u}\|_{(L^{2}(D))^{3}}+\|\bm{\zeta}\|_{(L^{2}(D))^{3}}\\ \leq&C\psi(\epsilon)\rightarrow 0\ \ \mbox{as}\ \ \epsilon\rightarrow+0.\end{split} (3.18)

Now, we analyze the interior transmission eigenvalue problem by studying the function 𝒗\bm{v} instead of 𝒖\bm{u} defined in (3.8). Taking the difference between the first two equations in (1.6), with the help of (3.8), one obtains

curl curl𝒖=k2𝒗inD.\text{curl curl}\,\bm{u}=k^{2}\bm{v}\quad\text{in}\ D. (3.19)

Next, we define the spaces

Si:={𝒇Hi(curl,D)|div𝒇=0},i=0,1,S_{i}:=\{\bm{f}\in H^{i}(\mathrm{curl},D)|\mbox{div}\bm{f}=0\},\ i=0,1,

and the mapping

curl1:S0S1/S\mathrm{curl}^{-1}:S_{0}\rightarrow S_{1}/S^{\prime}

that maps a vector field to its vector potential. Here, S:={𝒇H(curl,D)|curl𝒇=0,div𝒇=0}S^{\prime}:=\{\bm{f}\in H(\mathrm{curl},D)|\mbox{curl}\bm{f}=0,\ \mbox{div}\bm{f}=0\}. Besides, we define another mapping that maps a vector field to its vector potential,

curl~1:S1/SH2(curl,D)/S′′,\widetilde{\mathrm{curl}}^{-1}:S_{1}/S^{\prime}\rightarrow H^{2}(\mathrm{curl},D)/S^{\prime\prime},

where S′′:={𝒇H2(curl,D)|curl𝒇=0}S^{\prime\prime}:=\{\bm{f}\in H^{2}(\mathrm{curl},D)|\mbox{curl}\bm{f}=0\}. We then represent these two mappings indiscriminately when the space is clear. Accordingly, we define the inner product space

V0:={𝒇S0|curl1curl1𝒇H02(curl,D)}V_{0}:=\{\bm{f}\in S_{0}|\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{f}\in H_{0}^{2}(\mathrm{curl},D)\}

equipped with the (L2(D))3(L^{2}(D))^{3} scalar product. Therefore, the existence and uniqueness of a strong solution to (1.6) is equivalent to the existence and uniqueness of 𝒗V0\bm{v}\in V_{0} and 𝒖H02(curl,D)\bm{u}\in H_{0}^{2}(\mathrm{curl},D) satisfying (3.19) and

(curl curlk2𝑵)(𝑵𝑰)1(𝒗k2curl1curl1𝒗)=0in D.(\text{curl curl}-k^{2}\bm{N})(\bm{N}-\bm{I})^{-1}(\bm{v}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v})=0\quad\text{in }D. (3.20)

After using the variational formulation, one easily see that 𝒗V0\bm{v}\in V_{0} satisfies (3.20) if and only if

((𝑵𝑰)1(𝒗k2curl1curl1𝒗),𝒗k2𝑵curl1curl1𝒗)D=0\left((\bm{N}-\bm{I})^{-1}(\bm{v}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}),\bm{v}^{\prime}-k^{2}\bm{N}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}^{\prime}\right)_{D}=0 (3.21)

for all 𝒗V0\bm{v}^{\prime}\in V_{0}.

Then, we set

𝒢k(𝒗,𝒗)=((𝑵𝑰)1(𝒗k2curl1curl1𝒗),𝒗k2curl1curl1𝒗)D+(k2curl1curl1𝒗,k2curl1curl1𝒗)D\begin{split}\mathcal{G}_{k}(\bm{v},\bm{v}^{\prime})=&\left((\bm{N}-\bm{I})^{-1}(\bm{v}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}),\bm{v}^{\prime}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}^{\prime}\right)_{D}\\ &+\left(k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v},k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}^{\prime}\right)_{D}\end{split} (3.22)

and

𝒥(𝒗,𝒗)=(curl1𝒗,curl1𝒗)D,\mathcal{J}(\bm{v},\bm{v}^{\prime})=\left(\mathrm{curl}^{-1}\bm{v},\mathrm{curl}^{-1}\bm{v}^{\prime}\right)_{D}, (3.23)

which define two sesquilinear forms on V×VV\times V. Using the identities

𝑵(𝑵𝑰)1=𝑰+(𝑵𝑰)1\bm{N}(\bm{N}-\bm{I})^{-1}=\bm{I}+(\bm{N}-\bm{I})^{-1}

and

(𝒗,curl1curl1𝒗)D=(curl1𝒗,curl1𝒗)D\left(\bm{v},\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}^{\prime}\right)_{D}=\left(\mathrm{curl}^{-1}\bm{v},\mathrm{curl}^{-1}\bm{v}^{\prime}\right)_{D}

for all (𝒗,𝒗)V×V0(\bm{v},\bm{v}^{\prime})\in V\times V_{0}, the equation (3.21) can be rewritten as

𝒢k(𝒗,𝒗)k2𝒥(𝒗,𝒗)=0\mathcal{G}_{k}(\bm{v},\bm{v}^{\prime})-k^{2}\mathcal{J}(\bm{v},\bm{v}^{\prime})=0 (3.24)

for all 𝒗V0\bm{v}^{\prime}\in V_{0}. With the help of Poincaré-type inequality [28], we know that the sesquilinear forms 𝒢k\mathcal{G}_{k} and 𝒥\mathcal{J} are bounded. Based on the Riesz representation theorem, we could define the continuous operators Gk:V0V0G_{k}:V_{0}\to V_{0} such that

(Gk𝒗0,𝒗)D=𝒢k(𝒗0,𝒗)(G_{k}\bm{v}_{0},\bm{v}^{\prime})_{D}=\mathcal{G}_{k}(\bm{v}_{0},\bm{v}^{\prime})

for all 𝒗V0\bm{v}^{\prime}\in V_{0}, and J:V0V0J:V_{0}\to V_{0} such that

(J𝒗0,𝒗)D=𝒥(𝒗0,𝒗)(J\bm{v}_{0},\bm{v}^{\prime})_{D}=\mathcal{J}(\bm{v}_{0},\bm{v}^{\prime})

for all 𝒗V0\bm{v}^{\prime}\in V_{0}.

Next, we show that 𝒢k\mathcal{G}_{k} is a coercive sesquilinear form on V0×V0V_{0}\times V_{0}. According to (1.2), there is

𝒢k(𝒗0,𝒗0)α𝒗0k2curl1curl1𝒗0(L2(D))32+k4curl1curl1𝒗0(L2(D))32.\mathcal{G}_{k}(\bm{v}_{0},\bm{v}_{0})\geq\alpha\|\bm{v}_{0}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}_{0}\|^{2}_{(L^{2}(D))^{3}}+k^{4}\|\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}_{0}\|^{2}_{(L^{2}(D))^{3}}. (3.25)

Setting a=𝒗0(L2(D))3a=\|\bm{v}_{0}\|_{(L^{2}(D))^{3}} and b=k2curl1curl1𝒗0(L2(D))3b=k^{2}\|\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}_{0}\|_{(L^{2}(D))^{3}}, we have

𝒗0k2curl1curl1𝒗0(L2(D))32a2+b22ab.\|\bm{v}_{0}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}_{0}\|^{2}_{(L^{2}(D))^{3}}\geq a^{2}+b^{2}-2ab. (3.26)

Combining (3.25) and (3.26), one concludes that

𝒢k(𝒗0,𝒗0)αa22αab+(1+α)b2α1+2α(a2+b2)α1+2α𝒗0(L2(D))32.\mathcal{G}_{k}(\bm{v}_{0},\bm{v}_{0})\geq\alpha a^{2}-2\alpha ab+(1+\alpha)b^{2}\geq\frac{\alpha}{1+2\alpha}(a^{2}+b^{2})\geq\frac{\alpha}{1+2\alpha}\|\bm{v}_{0}\|^{2}_{(L^{2}(D))^{3}}.

Since 𝒢k\mathcal{G}_{k} is a coercive sesquilinear form on V0×V0V_{0}\times V_{0}, by the Lax–Milgram theorem, we deduce that Gk:V0V0G_{k}:V_{0}\to V_{0} is a bijective operator.

Regarding JJ, we would like to illustrate that it is compact. Let {𝒗n}\{\bm{v}_{n}\} be a bounded sequence in V0V_{0}, by the Bolzano-Weierstrass theorem, there exists a subsequence, denoted by {𝒗ni}\{\bm{v}_{n_{i}}\}, that converges weakly to some 𝒗0\bm{v}_{0} in V0V_{0}. Provided D\partial D is sufficiently smooth, the space

{𝒖H0(curl,D):div𝒖=0inD}\{\bm{u}\in H_{0}(\mathrm{curl},D):\text{div}\,\bm{u}=0\ \mathrm{in}\ D\}

is continuously embedded into H1(D)H^{1}(D). So, the sequence {curl1𝒗ni}\{\mathrm{curl}^{-1}\bm{v}_{n_{i}}\} is bounded in H1(D)H^{1}(D). By the Rellich compact embedding theorem, one has that {curl1𝒗ni}\{\mathrm{curl}^{-1}\bm{v}_{n_{i}}\} converges strongly to curl1𝒗0\mathrm{curl}^{-1}\bm{v}_{0} in (L2(D))3(L^{2}(D))^{3}. According to the definition of JJ and the Cauchy-Schwarz inequality, there is

J(𝒗ni𝒗0)(L2(D))32curl1(𝒗ni𝒗0)(L2(D))3curl1(J(𝒗ni𝒗0))(L2(D))3.\|J(\bm{v}_{n_{i}}-\bm{v}_{0})\|^{2}_{(L^{2}(D))^{3}}\leq\|\mathrm{curl}^{-1}(\bm{v}_{n_{i}}-\bm{v}_{0})\|_{(L^{2}(D))^{3}}\|\mathrm{curl}^{-1}(J(\bm{v}_{n_{i}}-\bm{v}_{0}))\|_{(L^{2}(D))^{3}}.

Besides, using Poincaré-type inequality, we deduce that

J(𝒗ni𝒗0)(L2(D))3curl1(J(𝒗ni𝒗0))(L2(D))3.\|J(\bm{v}_{n_{i}}-\bm{v}_{0})\|_{(L^{2}(D))^{3}}\geq\|\mathrm{curl}^{-1}(J(\bm{v}_{n_{i}}-\bm{v}_{0}))\|_{(L^{2}(D))^{3}}.

Combining the above two formulas, we get

J(𝒗ni𝒗0)(L2(D))3curl1(𝒗ni𝒗0)(L2(D))3.\|J(\bm{v}_{n_{i}}-\bm{v}_{0})\|_{(L^{2}(D))^{3}}\leq\|\mathrm{curl}^{-1}(\bm{v}_{n_{i}}-\bm{v}_{0})\|_{(L^{2}(D))^{3}}.

Therefore, the sequence {J𝒗ni}\{J\bm{v}_{n_{i}}\} converges strongly to J𝒗0J\bm{v}_{0} in V0V_{0}.

We now consider the scattering system (3.6) again. We take 𝜻H2(curl,D)\bm{\zeta}\in H^{2}(\mathrm{curl},D) satisfying (3.10) and (3.11), and define 𝜾:=curl curl𝜻V\bm{\iota}:=\text{curl curl}\,\bm{\zeta}\in V. Setting 𝑬¯:=𝑬t𝜻H2(curl,D)\bar{\bm{E}}:=\bm{E}^{t}-\bm{\zeta}\in H^{2}(\mathrm{curl},D), the system (3.6) becomes

{curl curl𝑬¯k2𝑵(𝒙)𝑬¯=(curl curlk2𝑵(𝒙))𝜻inD,curl curl𝑬𝒈0,kk2𝑬𝒈0,k=0inD,𝝂×𝑬¯=𝝂×𝑬𝒈0,konD,𝝂×curl𝑬¯=𝝂×curl𝑬𝒈0,konD.\left\{\begin{array}[]{ll}\text{curl curl}\,\bar{\bm{E}}-k^{2}\bm{N}(\bm{x})\bar{\bm{E}}=-(\text{curl curl}-k^{2}\bm{N}(\bm{x}))\bm{\zeta}&\text{in}\ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \text{curl curl}\,\bm{E}_{\bm{g}_{0},k}-k^{2}\bm{E}_{\bm{g}_{0},k}=0&\text{in}\ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \bm{\nu}\times\bar{\bm{E}}=\bm{\nu}\times\bm{E}_{\bm{g}_{0},k}&\text{on}\ \partial D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \bm{\nu}\times\text{curl}\,\bar{\bm{E}}=\bm{\nu}\times\text{curl}\,\bm{E}_{\bm{g}_{0},k}&\text{on}\ \partial D.\end{array}\right. (3.27)

Let 𝒖¯:=𝑬¯𝑬𝒈0,kH02(curl,D)\bar{\bm{u}}:=\bar{\bm{E}}-\bm{E}_{\bm{g}_{0},k}\in H_{0}^{2}(\mathrm{curl},D), 𝒗¯:=𝑵(𝒙)𝑬¯𝑬𝒈0,kk2curl curl𝜻+𝑵(𝒙)𝜻V0\bar{\bm{v}}:=\bm{N}(\bm{x})\bar{\bm{E}}-\bm{E}_{\bm{g}_{0},k}-k^{-2}\text{curl curl}\,\bm{\zeta}+\bm{N}(\bm{x})\bm{\zeta}\in V_{0}. Then curl curl𝒖¯=k2𝒗¯\text{curl curl}\,\bar{\bm{u}}=k^{2}\bar{\bm{v}}, and the equations in (3.27) can be rewritten as

(curl curlk2𝑵)(𝑵𝑰)1(𝒗¯k2curl1curl1𝒗¯+k2𝜾𝑵curl1curl1𝜾)+𝜾k2𝑵curl1curl1𝜾=0in D.\begin{split}(\text{curl curl}-k^{2}\bm{N})(\bm{N}-\bm{I})^{-1}(\bar{\bm{v}}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bar{\bm{v}}+&k^{-2}\bm{\iota}-\bm{N}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\iota})\\ &+\bm{\iota}-k^{2}\bm{N}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\iota}=0\quad\text{in }D.\end{split} (3.28)

Using a variational framework, we can see that 𝒗¯V0\bar{\bm{v}}\in V_{0} satisfying (3.28) is equivalent to

𝒢k(𝒗¯,𝒗)k2𝒥(𝒗¯,𝒗)=𝔮(𝒗)for all 𝒗V0,\mathcal{G}_{k}(\bar{\bm{v}},\bm{v}^{\prime})-k^{2}\mathcal{J}(\bar{\bm{v}},\bm{v}^{\prime})=\mathfrak{q}(\bm{v}^{\prime})\quad\text{for all }\bm{v}^{\prime}\in V_{0}, (3.29)

where 𝒢k\mathcal{G}_{k} and 𝒥\mathcal{J} are sesquilinear forms defined in (3.22) and (3.23). And 𝔮(𝒗)\mathfrak{q}(\bm{v}^{\prime}) is an antilinear form defined by

𝔮(𝒗):=(curl1curl1𝜾+(𝑵𝑰)1curl1curl1𝜾k2(𝑵𝑰)1𝜾,𝒗)D+((𝑵𝑰)1𝜾k2curl1curl1𝜾k2(𝑵𝑰)1curl1curl1𝜾,curl1curl1𝒗)D.\begin{split}\mathfrak{q}(\bm{v}^{\prime}):=&\left(\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\iota}+(\bm{N}-\bm{I})^{-1}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\iota}-k^{-2}(\bm{N}-\bm{I})^{-1}\bm{\iota},\bm{v}^{\prime}\right)_{D}\\ &+\left((\bm{N}-\bm{I})^{-1}\bm{\iota}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\iota}-k^{2}(\bm{N}-\bm{I})^{-1}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\iota},\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{v}^{\prime}\right)_{D}.\end{split} (3.30)

As GkG_{k} is an isomorphism, and JJ is compact, the Fredholm alternative can be applied to the systems (3.24) and (3.29). Now, we show that (3.29) is solvable in the quotient space V0/𝕍V_{0}/\mathbb{V}, where 𝕍\mathbb{V} is the eigenspace of (3.24). We denote by 𝒒~\bm{\tilde{q}} the Riesz representation of 𝔮\mathfrak{q} in V0V_{0}. Next, we verify that

𝒒~ker[(Gkk2J)].\bm{\tilde{q}}\in\text{ker}[(G_{k}-k^{2}J)^{*}]^{\bot}.

If the function 𝝇V0\bm{\varsigma}\in V_{0} belongs to the kernel of (Gkk2J)(G_{k}-k^{2}J)^{*}, then for any 𝝉V\bm{\tau}\in V, one has

(𝝉,(Gkk2J)𝝇)D=0,\left(\bm{\tau},(G_{k}-k^{2}J)^{*}\bm{\varsigma}\right)_{D}=0,

that is

((𝑵𝑰)1(𝝉k2curl1curl1𝝉),𝝇k2curl1curl1𝝇)D+(k2curl1curl1𝝉,k2curl1curl1𝝇)Dk2(curl1𝝉,curl1𝝇)D=0.\begin{split}&\left((\bm{N}-\bm{I})^{-1}(\bm{\tau}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\tau}),\bm{\varsigma}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\varsigma}\right)_{D}\\ &\quad+(k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\tau},k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\varsigma})_{D}-k^{2}(\mathrm{curl}^{-1}\bm{\tau},\mathrm{curl}^{-1}\bm{\varsigma})_{D}=0.\end{split} (3.31)

As 𝒒~\bm{\tilde{q}} is the Riesz representation of 𝔮\mathfrak{q}, with the help of (3.31), there is

(𝝇,𝒒~)D=𝔮(𝝇)=(curl1curl1𝜾+(𝑵𝑰)1curl1curl1𝜾k2(𝑵𝑰)1𝜾,𝝇)D+((𝑵𝑰)1𝜾k2curl1curl1𝜾k2(𝑵𝑰)1curl1curl1𝜾,curl1curl1𝝇)D=0.\begin{split}(\bm{\varsigma},\bm{\tilde{q}}&)_{D}\\ =&\mathfrak{q}(\bm{\varsigma})\\ =&\left(\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\iota}+(\bm{N}-\bm{I})^{-1}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\iota}-k^{-2}(\bm{N}-\bm{I})^{-1}\bm{\iota},\bm{\varsigma}\right)_{D}\\ &+\left((\bm{N}-\bm{I})^{-1}\bm{\iota}-k^{2}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\iota}-k^{2}(\bm{N}-\bm{I})^{-1}\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\iota},\mathrm{curl}^{-1}\mathrm{curl}^{-1}\bm{\varsigma}\right)_{D}\\ =&0.\end{split}

Up to now, we have verified that (3.29) is solvable in the quotient space V0/𝕍V_{0}/\mathbb{V}. By (3.11) and (3.30), we observes that

𝒒~(L2(D))3Cψ(ϵ),\|\bm{\tilde{q}}\|_{(L^{2}(D))^{3}}\leq C\psi(\epsilon), (3.32)

where CC is a constant depending on 𝑵,k\bm{N},k and DD. Setting

𝒗^=(Gkk2J)1𝒒~in V0/𝕍,\hat{\bm{v}}=(G_{k}-k^{2}J)^{-1}\bm{\tilde{q}}\quad\text{in }V_{0}/\mathbb{V},

and using (3.11), (3.32), one has

𝑵𝑬t𝑵𝑬k𝑬𝒈0,k+𝑬0,k(L2(D))3𝒗^(L2(D))3+k2curl curl𝜻(L2(D))3Cψ(ϵ)0 as ϵ+0.\begin{split}\|\bm{N}\bm{E}^{t}-\bm{N}\bm{E}_{k}-\bm{E}_{\bm{g}_{0},k}+\bm{E}_{0,k}\|_{(L^{2}(D))^{3}}\leq&\|\hat{\bm{v}}\|_{(L^{2}(D))^{3}}+k^{-2}\|\text{curl curl}\,\bm{\zeta}\|_{(L^{2}(D))^{3}}\\ \leq&C\psi(\epsilon)\to 0\text{ as }\epsilon\to+0.\end{split} (3.33)

Finally, by combining (3.18) and (3.33), one obtains

𝑬𝒈0,k𝑬0,k(L2(D))3Cψ(ϵ)0 as ϵ+0.\|\bm{E}_{\bm{g}_{0},k}-\bm{E}_{0,k}\|_{(L^{2}(D))^{3}}\leq C\psi(\epsilon)\to 0\text{ as }\epsilon\to+0.

The proof is complete. ∎

Remark 3.4.

It is remarked that Lemma  3.2 is used in the proof of Theorem  3.2. So, the higher requirement for the regularity of the scatterer is needed technically. However, we believe that this regularity assumption can be relaxed, which is showed in the following numerical examples. The approximation result Theorem  3.2 is an extension of the similar result in the acoustic scattering problem, but this study is more technical and challenging. Meanwhile, we believe that this result is not only of independent interest to the scattering theory of electromagnetic waves, but also provides a theoretical basis for practical applications, such as super-resolution imaging and invisibility cloaking.

Remark 3.5.

Theorem 3.1 and 3.2 indicate that if the Herglotz kernel 𝒈0\bm{g}_{0} is a solution to the optimization problem (3.1), then the Herglotz wave function 𝑬𝒈0,k\bm{E}_{\bm{g}_{0},k} (2.4) is approximated to the transmission eigenfunction 𝑬0,k\bm{E}_{0,k} associated with the correspondingly eigenvalue kk. Thus, in order to recover the transmission eigenfunction, it is critical for solving the constrained optimization problem (3.1) and obtain a satisfactory Herglotz kernel 𝒈0\bm{g}_{0}. In what follows, we will use the Fourier-total-least-square (FTLS) method and Gradient-total-least-square (GTLS) method [25] to compute 𝒈0\bm{g}_{0}, respectively.

4. Imaging of the scatterer

According to sections 2 and 3, we can determine the transmission eigenvalues and the corresponding transmission eigenfunctions from the electric far-field data in (1.5). In this section, we show that the “invisible patterns”, the transmission eigenfunctions, can be used to qualitatively image the shape of the electromagnetic medium scatterer DD, independent of 𝑵\bm{N}.

In our recent studies, we find that the Maxwell transmission eigenfunctions 𝑬0,k\bm{E}_{0,k} possess the following interesting global and local geometric properties. From a global perspective, there exists a sequence of surface-localized transmission eigenfunctions associated with sufficient large transmission eigenvalues, that is, the L2L^{2}-energy of the eigenfunction concentrates on the boundary D\partial D [20]. We would like to point out that these surface-localized eigenstates occur very frequently through theoretical and numerical observations. From a local perspective, the transmission eigenfunctions are nearly vanishing around the high-curvature points on D\partial D, especially around the corner points of D\partial D [8]. Meanwhile, the rigorous mathematical justifications in section 3 show that the Herglotz wave 𝑬𝒈0,k\bm{E}_{\bm{g}_{0},k} is an approximation of the transmission eigenfunction 𝑬0,k\bm{E}_{0,k}. So the Herglotz wave has the same geometric attributes as the transmission eigenfunction. Hence, based on the global and local properties of the Herglotz wave, we introduce the following indicator function for identifying the shape of the scatterer DD:

IkRes(𝒛):=ln𝑬𝒈0,k(𝒛),I_{k}^{\text{Res}}(\bm{z}):=-\text{ln}\|\bm{E}_{\bm{g}_{0},k}(\bm{z})\|,

where 𝒛D~\bm{z}\in\widetilde{D} is the sampling point with DD~3D\subset\widetilde{D}\subset\mathbb{R}^{3}. It is worth noting that the wave field 𝑬𝒈0,k\bm{E}_{\bm{g}_{0},k} is nearly vanishing in the interior of DD and it is vanishing around the high-curvature points. Thus, the indicator function IkRes(𝒛)I_{k}^{\text{Res}}(\bm{z}) possesses a relative large value if 𝒛\bm{z} belongs to the interior of DD or locates at the corner/edge/highly-curved place on D\partial D [22], whereas it possesses a relatively small value if 𝒛\bm{z} locates in the other places around D\partial D and outside of DD.

Noting that the multiple transmission eigenfunctions possess the same geometric properties, we superimpose the imaging effects by introducing the following multi-frequency imaging functional:

I𝕂LRes(𝒛):=lnk𝕂L|𝑬𝒈0,k(𝒛)|,I_{\mathbb{K}_{L}}^{\text{Res}}(\bm{z}):=-\text{ln}\sum\limits_{k\in{\mathbb{K}_{L}}}|\bm{E}_{\bm{g}_{0},k}(\bm{z})|, (4.1)

where 𝕂L={k1,k2,,kL}\mathbb{K}_{L}=\{k_{1},k_{2},\cdots,k_{L}\} denotes the set of LL distinct transmission eigenvalues. Based on the imaging functional (4.1), we then propose the following imaging scheme, which is referred to as imaging by interior resonant modes.

Algorithm II: Imaging by interior resonant modes
Step 1 For each resonant wavenumber kk found in Algorithm I, solve the optimization problem (3.1) by the FTLS method or the GTLS method [25] to obtain the vector Herglotz kernel 𝒈0\bm{g}_{0}.
Step 2 Calculate the electric Herglotz wave function 𝑬𝒈0,k\bm{E}_{\bm{g}_{0},k} with the Herglotz kernel 𝒈0\bm{g}_{0} by the definition (2.4).
Step 3 Plot the indicator function (4.1) in a proper domain containing the electromagnetic medium scatterer DD and identify the interior and corners (two dimension) or edges (three dimension) as bright points, and other boundary places as dark points in the graph to obtain the shape of the scatterer DD.

4.1. Transmission eigenvalues reconstruction

In this part, we present a numerical experiment to verify the validity of Algorithm I for Maxwell transmission eigenvalues reconstruction based on the LSM. We take the regularization parameter δ=105\delta=10^{-5}. To begin with, we consider a unit ball, see Figure 1(a). In order to reduce the calculation expense, we use a normal mesh with mesh size h0.15h\approx 0.15. To avoid the inverse crime, we use the finite element method (FEM) to compute a pair of electric far-field pattern 𝑬(𝒙^i,𝒅j,𝒑jθ;ks)\bm{E}_{\infty}(\hat{\bm{x}}_{i},\bm{d}_{j},{\bm{p}}_{j}^{\theta};k_{s}) and 𝑬(𝒙^i,𝒅j,𝒑jϕ;ks)\bm{E}_{\infty}(\hat{\bm{x}}_{i},\bm{d}_{j},{\bm{p}}_{j}^{\phi};k_{s}), where i=1,2,,Mi=1,2,\cdots,M, j=1,2,,Nj=1,2,\cdots,N and s=1,2,,Ss=1,2,\cdots,S. Here, the observation and incident directions are pseudo uniformly distributed on the unit spherical surface. The artificial far-field data are computed at 100100 observation directions (M=100M=100), 100100 incident directions (N=100N=100) and 8181 equally distributed wave numbers (S=81S=81) in [1.1,1.5][1.1,1.5]. Here the refractive index is given by

𝑵(𝒙)=(160001600016).\bm{N}(\bm{x})=\left(\begin{array}[]{ccc}16&0&0\\ 0&16&0\\ 0&0&16\\ \end{array}\right).

In Figure 1(b), the solid blue line shows the value of =13𝒈δ(,𝒛0,𝒒;ks)(L2(𝕊2))32\sum_{\ell=1}^{3}\|\bm{g}_{\delta}(\cdot,\bm{z}_{0},\bm{q}_{\ell};k_{s})\|_{(L^{2}(\mathbb{S}^{2}))^{3}}^{2} against ks[1.1,1.5]k_{s}\in[1.1,1.5] for a fixed test point 𝒛0=(0.1,0.3,0.4)\bm{z}_{0}=(0.1,0.3,0.4) with 1%1\% noise. As expected, it can be seen that the solid blue line has clear spikes, which indicate the locations of the Maxwell transmission eigenvalues. To show the accuracy of the reconstruction, we list the exact eigenvalues, the reconstructed eigenvalues by the mixed FEM [29] and the LSM, in Table 1. From the results of the comparison, we can see that Algorithm I is valid for identifying the Maxwell transmission eigenvalues.

Refer to caption
(a) a unit ball centered at (0,0,0)(0,0,0)
Refer to caption
(b) the indicator function vs. kk
Figure 1. (a) DD: a unit ball; (b) plot of the indicator function =13𝒈δ(,𝒛0,𝒒;ks)(L2(𝕊2))32\sum_{\ell=1}^{3}\|\bm{g}_{\delta}(\cdot,\bm{z}_{0},\bm{q}_{\ell};k_{s})\|_{(L^{2}(\mathbb{S}^{2}))^{3}}^{2} for a unit ball centered at (0,0,0)(0,0,0), where ks[1.1,1.5]k_{s}\in[1.1,1.5].
Table 1. The first three real Maxwell transmission eigenvalues of the spherical domain.
Index of eigenvalue 11 22 33
Exact: 1.1651.165 1.4611.461 1.4751.475
FEM: 1.1661.166 1.4621.462 1.4761.476
LSM: 1.1701.170 1.4651.465 1.4901.490

4.2. Scatterer reconstruction

In this part, we give several numerical examples to show that the proposed imaging scheme can effectively reconstruct the scatterer DD.

Refer to caption
(a) Gourd
Refer to caption
(b) Kite
Refer to caption
(c) Square
Figure 2. Shapes considered in the numerical examples. (a) gourd-shaped domain in 3D; (b) layered kite-shaped domain in 2D; (c) square domain in 2D.

4.2.1. Gourd-shaped domain

In the first example, we utilize the global properties of transmission eigenfunctions to identify the anisotropic inhomogeneity in the 33D setting. Let the geometry be a gourd-shaped domain, see Figure 2(a). The refractive index is given by

𝑵(𝒙)=(161011600014).\bm{N}(\bm{x})=\left(\begin{array}[]{ccc}16&1&0\\ 1&16&0\\ 0&0&14\\ \end{array}\right).

The artificial far-field data are computed at 100100 observation directions and 100100 incident directions. To begin with, we use the mixed FEM to calculate the transmission eigenfunctions associated with different eigenvalues. Figure 3 shows the surface plots and multiple slices plots of the eigenfunctions |𝑬0,k||\bm{E}_{0,k}| for three different eigenvalues. It is clear to see that the transmission eigenfunctions are surface-localized around the boundary. In this case, the medium is anisotropic. The surface localization result here extends the theoretical statement of isotropic cases [20] to the anisotropic cases numerically. Besides, it is noted that |𝑬0,k||\bm{E}_{0,k}| is nearly vanishing near the top of the gourd-shaped domain. This observation is consistent with the theoretical result that the transmission eigenfunctions must be nearly vanishing near the points of large curvature [7].

Next, we test the proposed reconstruction scheme with 1%1\% noise. Here we use the GTLS method with regularization parameter β=104\beta=10^{-4}. Figure 4 presents iso-surface plots and slice plots of the multi-frequency indicator function, respectively. Here we use five different frequencies in interval [1.5,1.6][1.5,1.6]. In the slice-view plots, the white dashed lined are the boundary of the exact scatterer. One can find that the proposed method could approximately identify the boundary of the scatterer.

4.2.2. TE mode in 2D

Refer to caption
(a) k=1.5141k=1.5141
Refer to caption
(b) k=1.5149k=1.5149
Refer to caption
(c) k=1.5338k=1.5338
Refer to caption
(d) k=1.5141k=1.5141
Refer to caption
(e) k=1.5149k=1.5149
Refer to caption
(f) k=1.5338k=1.5338
Figure 3. Surface plots and multiple slices plots of the transmission eigenfunctions |𝑬0,k||\bm{E}_{0,k}| with three different wavenumbers.
Refer to caption
(a) isovalue 2020
Refer to caption
(b) isovalue 3030
Refer to caption
(c) isovalue 4040
Refer to caption
(d) slice at x=0x=0
Refer to caption
(e) slice at y=0y=0
Refer to caption
(f) slice at z=0z=0
Figure 4. Iso-surface plots and slice plots of multi-frequency indicator function I𝕂5ResI_{\mathbb{K}_{5}}^{Res}.

In this example, we consider the case that the scatterer coats a thin layer of high refractive-index material. It is worth mentioning that it requires heavy costs in computing the three-dimensional far-field data with high refractive index. Due to the limited computational resources, we could only afford to the two-dimensional calculation. Thus, we test the transverse electric (TE) mode in 2D. The transmission eigenvalue problem (1.6) for TE mode is reduced to the following system:

{¯(𝐄)k2𝑵𝐄=0,inD,¯(𝐄0)k2𝐄0=0,inD,𝝂𝐄=𝝂𝐄0,𝝂¯(𝐄)=𝝂¯(𝐄0)onD,\begin{cases}\underline{\nabla}\wedge(\nabla\wedge\mathbf{E})-k^{2}\bm{N}\mathbf{E}=0,\quad&\mbox{in}\ \ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \underline{\nabla}\wedge(\nabla\wedge\mathbf{E}_{0})-k^{2}\mathbf{E}_{0}=0,\quad&\mbox{in}\ \ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \bm{\nu}\wedge\mathbf{E}=\bm{\nu}\wedge\mathbf{E}_{0},\ \underline{\bm{\nu}}\wedge(\nabla\wedge\mathbf{E})=\underline{\bm{\nu}}\wedge(\nabla\wedge\mathbf{E}_{0})\quad&\mbox{on}\ \ \partial D,\end{cases}

where

𝑬:=(E(1),E(2))and𝑬0:=(E0(1),E0(2)),\bm{E}:=\left(E^{(1)},\,E^{(2)}\right)^{\top}\quad\text{and}\quad\bm{E}_{0}:=\left(E_{0}^{(1)},\,E_{0}^{(2)}\right)^{\top},

and

𝐅:=xF(2)yF(1),¯E:=(yExE),𝝂𝐅:=ν1F(2)ν2F(1),𝝂¯E:=(ν2Eν1E).\nabla\wedge\mathbf{F}:=\partial_{x}F^{(2)}-\partial_{y}F^{(1)},\ \underline{\nabla}\wedge E:=\left(\begin{array}[]{c}\partial_{y}E\\ -\partial_{x}E\end{array}\right),\ \bm{\nu}\wedge\mathbf{F}:=\nu_{1}F^{(2)}-\nu_{2}F^{(1)},\ \underline{\bm{\nu}}\wedge E:=\left(\begin{array}[]{c}\nu_{2}E\\ -\nu_{1}E\end{array}\right).

Here we consider a kite-shaped domain with a thin layer, see Figure 2(b). The refractive indexes 𝑵\bm{N} of the outside layer and the inside domain are given by

𝑵out(𝒙)=(2561616256),𝑵in(𝒙)=(2002).\bm{N}_{out}(\bm{x})=\left(\begin{array}[]{ccc}256&16\\ 16&256\\ \end{array}\right),\quad\bm{N}_{in}(\bm{x})=\left(\begin{array}[]{ccc}2&0\\ 0&2\\ \end{array}\right).

Figure 5 presents the exact transmission eigenfunctions in the predetermined domain. We can see from the refractive index 𝑵\bm{N} that this problem is associated with anisotropic media. In this case both transmission eigenfunctions 𝑬\bm{E} and 𝑬0\bm{E}_{0} are localized around the boundary of the domain. This example also numerically expands the theoretical result of isotropic cases [20]. The synthetic far-field data are computed within the interval [1,2][1,2] with additional 1%1\% noise. First, we use Algorithm I to determine eight transmission eigenvalues. Next, we use the GTLS method with regularization parameter β=106\beta=10^{-6} to recover the Herglotz wave functions E𝒈0,kE_{\bm{g}_{0},k}. We present the reconstruction results in Figure 6 by using two, five and eight eigenmodes, respectively. One readily sees that the kite is already finely reconstructed with eight interior resonant modes. It is clear that the scale of the kite-shaped domain is much smaller than the underlying wavelength, 2π/k42\pi/k\approx 4. This result shows that super-resolution reconstruction can be realized by the proposed imaging scheme. This is unobjectionably expected since we make use of the interior resonant modes for the reconstruction. Here, we want to point out that the contrast 𝑵\bm{N} just needed to be high around D\partial D. Then small transmission eigenvalues can occur no matter the refractive index is large or normal inside the object. That means in real life, for a regular refractive inhomogeneity, one may first coat the object via indirect means with a relatively thin layer of high-contrast medium. For example, one could spray high-contrast material onto the surface of the scatterer. Then super-resolution imaging can be achieved by the same reconstruction procedure as above.

Remark 4.1.

From Figure 4 and 6, we can observe that the concave part of the scatterer can be reconstructed well. This is physically reasonable since the proposed method makes use of the interior resonant modes, which is equivalent to “looking” the scatterer from its inside. At this point, the concave part of the scatterer observed from the exterior becomes convex.

4.2.3. TM mode in 2D

Refer to caption
(a) E(1)E^{(1)}
Refer to caption
(b) E(2)E^{(2)}
Refer to caption
(c) E0(1)E_{0}^{(1)}
Refer to caption
(d) E0(2)E_{0}^{(2)}
Figure 5. Imagesc plots of the transmission eigenfunctions with k=1.5098k=1.5098.
Refer to caption
(a) L=2L=2
Refer to caption
(b) L=5L=5
Refer to caption
(c) L=8L=8
Figure 6. Imagesc plots of multi-frequency indicator function I𝕂LResI_{\mathbb{K}_{L}}^{Res} with L=2, 5, 8L=2,\,5,\,8, respectively.

In the final example, we use the local geometric properties of transmission eigenfunctions to image the scatterer. To reduce the computational cost, we consider the transverse magnetic (TM) mode in 2D. The transmission eigenvalue problem (1.6) for TM mode is reduced to the following system:

{(¯E)k2n(𝒙)E=0,inD,(¯E0)k2E0=0,inD,𝝂¯E=𝝂¯E0,𝝂(¯E)=𝝂(¯E0)onD,\begin{cases}\nabla\wedge(\underline{\nabla}\wedge E)-k^{2}n(\bm{x})E=0,\quad&\mbox{in}\ \ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \nabla\wedge(\underline{\nabla}\wedge E_{0})-k^{2}E_{0}=0,\quad&\mbox{in}\ \ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \underline{\bm{\nu}}\wedge E=\underline{\bm{\nu}}\wedge E_{0},\ \bm{\nu}\wedge(\underline{\nabla}\wedge E)=\bm{\nu}\wedge(\underline{\nabla}\wedge E_{0})\quad&\mbox{on}\ \ \partial D,\end{cases} (4.2)

where the refractive index n(𝒙)=1/4n(\bm{x})=1/4. In particular, the system (4.2) can be rewritten as the following transmission eigenvalue problem associated with the scalar Helmholtz equation:

{ΔE+k2n(𝒙)E=0inD,ΔE0+k2E0=0inD,E=E0,E𝝂=E0𝝂onD.\left\{\begin{array}[]{ll}\Delta E+k^{2}n(\bm{x})E=0&\text{in}\ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \Delta E_{0}+k^{2}E_{0}=0&\text{in}\ D,\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \displaystyle{E=E_{0},\ \ \frac{\partial E}{\partial\bm{\nu}}=\frac{\partial E_{0}}{\partial\bm{\nu}}}&\text{on}\ \partial D.\end{array}\right.

Here, we let DD be a square defined by (x,y)[1,1]×[1,1](x,y)\in[-1,1]\times[-1,1]. The exact domain is shown in Figure 2(c).

Refer to caption
(a) k1=5.48k_{1}=5.48
Refer to caption
(b) k2=6.10k_{2}=6.10
Refer to caption
(c) k3=6.65k_{3}=6.65
Refer to caption
(d) k1=5.40k_{1}=5.40
Refer to caption
(e) k2=6.11k_{2}=6.11
Refer to caption
(f) k3=6.67k_{3}=6.67
Refer to caption
(g) k1=5.40k_{1}=5.40
Refer to caption
(h) k2=6.11k_{2}=6.11
Refer to caption
(i) k3=6.67k_{3}=6.67
Figure 7. The top row: transmission eigenfunctions E0,kE_{0,k} with the first three kk; the middle row: the Herglotz wave recovered by FTLS method; the bottom row: the Herglotz wave recovered by GTLS method.
Refer to caption
(a) L=2L=2
Refer to caption
(b) L=4L=4
Refer to caption
(c) L=6L=6
Refer to caption
(d) L=2L=2
Refer to caption
(e) L=4L=4
Refer to caption
(f) L=6L=6
Figure 8. Imagesc plots of multi-frequency indicator function I𝕂LResI_{\mathbb{K}_{L}}^{Res} with L=2, 4, 6L=2,\,4,\,6, respectively. The top rows: recovered by the FTLS method ; the bottom rows: recovered by the GTLS method.

To begin with, we proceed to determine the transmission eigenfunctions E0,kE_{0,k}. We consider the square domain as discussed above and in order to avoid inverse crimes, extra 5%5\% noise is added to the far-field data associated with those eigenvalues. For comparison, we use the FEM to compute the exact eigenfunction of E0,kE_{0,k} from the exact domain, see the top row of Figure 7. In addition, the middle and bottom rows of Figure 7 present the Herglotz wave functions E𝒈0,kE_{{\bm{g}}_{0},k} recovered by the FTLS and GTLS methods, respectively. For the FTLS method, we take the Fourier truncation order by Nt=6N_{t}=6. For the GTLS method, the regularization parameter is chosen as β=101\beta=10^{-1}. The dashed red lines denote the exact support of the square scatterer. It is clear that the reconstructed Herglotz wave functions by both methods are close to the exact transmission eigenfunctions E0,kE_{0,k} inside the scatterer. This numerically verifies the conclusion in section 3. Moreover, from the top row of Figure 7, one can find that the eigenfunctions are vanishing near the corners, which verifies the previous theoretical study [3, 6, 8]. In particular, we can observe that the reconstructed eigenfunctions are also nearly vanishing near the corner points rather than at the corners. This is because that recovering the Herglotz kernel function 𝒈0\bm{g}_{0} from the far-field data is very ill-posed.

Although the nodal lines of eigenfunctions appear in different locations for different eigenvalues (cf. [12, 13]), the nodal lines always go through the corners. So, the corners of the domain will stand out if we superimpose the indicator function with multi-frequency far-field data. Figure 8 presents the multi-frequency indicator function (4.1) with the first two, four and six eigenvalues with 5%5\% noise. In this case, the truncated Fourier order is given by Nt=8N_{t}=8 for the FTLS method and the regularization parameter is chosen as β=103\beta=10^{-3} for the GTLS method. One readily sees that the corner points are finely reconstructed with four interior resonant modes. If further a priori information is available on the shape, say, it is a polygon, then one can actually recover the scatterer. This example also shows that the proposed imaging scheme can break the Abbe resolution limit in recovering the fine details of D\partial D, even the singular points.

Remark 4.2.

It is worth to mention that the proposed method is valid for reconstructing an electromagnetic medium scatterer no matter if the medium is isotropic or anisotropic. In addition, the aforementioned numerical experiments demonstrate that super-resolution reconstruction can be realized by the proposed imaging scheme.

Acknowledgments. The authors would like to thank Professor Wei Wu of Jilin University for providing many helpful discussions during the completion of this paper. The work was supported by the Hong Kong RGC General Research Fund (projects 12301420, 12302919, 11300821) and NSFC/RGC Joint Research Grant (project N_CityU101/21).

References

  • [1] G. S. Alberti and Y. Capdeboscq, Elliptic regularity theory applied to time harmonic anisotropic Maxwell’s equations with less than Lipschitz complex coefficients, SIAM Journal on Mathematical Analysis, 46 (2014), no. 1, 998–1016.
  • [2] H. Ammari, J. Garnier, W. Jing, H. Kang, M. Lim, K. S\varnothinglna and H. Wang, Mathematical and Statistical Methods for Multistatic Imaging, Springer, New York, 2013
  • [3] E. Blåsten, Nonradiating sources and transmission eigenfunctions vanish at corners and edges, SIAM J. Math. Anal., 50 (2018), no. 6, 6255–6270.
  • [4] E. Blåsten, X. Li, H. Liu and Y. Wang, On vanishing and localization near cusps of transmission eigenfunctions: a numerical study, Inverse Problems, 33 (2017), 105001.
  • [5] E. Blåsten and H. Liu, On corners scattering stably and stable shape determination by a single far-field pattern, Indiana Univ. Math. J., 70 (2021), 907–947.
  • [6] E. Blåsten and H. Liu, On vanishing near corners of transmission eigenfunctions, J. Funct. Anal., 273 (2017), 3616–3632. Addendum available from https://arxiv.org/abs/1710.08089.
  • [7] E. Blåsten and H. Liu, Scattering by curvatures, radiationless sources, transmission eigenfunctions, and inverse scattering problems, SIAM J. Math. Anal., 53 (2021), 3801–3837.
  • [8] E. Blåsten, H. Liu and J. Xiao, On an electromagnetic problem in a corner and its applications, Anal. PDE, 14 (2021), no. 7, 2207–2224.
  • [9] F. Cakoni and D. Colton, Qualitative Methods in Inverse Scattering Theory, Springer, Berlin, 2006.
  • [10] F. Cakoni, D. Colton and H. Haddar, Inverse Scattering Theory and Transmission Eigenvalues, SIAM, Philadelphia, 2016.
  • [11] F. Cakoni, D. Colton and P. Monk, The Linear Sampling Method in Inverse Electromagnetic Scattering, SIAM, Philadelphia, 2011.
  • [12] X. Cao, H. Diao, H. Liu and J. Zou, On novel geometric structures of Laplacian eigenfunctions in 3\mathbb{R}^{3} and applications to inverse problems, SIAM J. Math. Anal., 53 (2021), 1263–1294.
  • [13] X. Cao, H. Diao, H. Liu and J. Zou, On nodal and generalized singular structures of Laplacian eigenfunctions and applications to inverse scattering problems, J. Math. Pures Appl., 143 (2020), 116–161.
  • [14] Y.-T. Chow, Y. Deng, Y. He, H. Liu and X. Wang, Surface-localized transmission eigenstates, super-resolution imaging and pseudo surface plasmon modes, SIAM J. Imaging Sci., 14 (2021), 946–975.
  • [15] Y.-T. Chow, Y. Deng, H. Liu and M. Sunkula, Surface concentration of transmission eigenfunctions, arXiv:2109.14361.
  • [16] D. Colton and R. Kress, Inverse Acoustic and Electromagnetic Scattering Theory, 4th Edition, Springer, New York, 2019.
  • [17] D. Colton and R. Kress, On the denseness of Herglotz wave functions and electromagnetic Herglotz pairs in Sobolev spaces, Math. Meth. Appl. Sci, 24 (2001), no. 16, 1289–1303.
  • [18] Y. Deng, C. Duan and H. Liu, On vanishing near corners of conductive transmission eigenfunctions, Res. Math. Sci., 9 (2022), 2.
  • [19] Y. Deng, Y. Jiang, H. Liu and K. Zhang, On new surface-localized transmission eigenmodes, Inverse Problems and Imaging, (2021), https://doi.org/10.3934/ipi.2021063
  • [20] Y. Deng, H. Liu, X. Wang, and W. Wu, On geometric properties of electromagnetic transmission eigenfunctions and artificial mirage, SIAM J. Appl. Math., (2021), https://doi.org/10.1137/21M1413547
  • [21] H. Diao, X. Cao and H. Liu, On the geometric structures of transmission eigenfunctions with a conductive boundary condition and applications, Comm. Partial Differential Equations, 46 (2021), 630–679.
  • [22] H. Diao, H. Liu, X. Wang and K. Yang, On vanishing and localizing around corners of electromagnetic transmission resonances, Partial Differential Equations and Applications, 2 (2021), 1–20.
  • [23] A. Kirsch and P. Monk, A finite element/spectral method for approximating the time-harmonic Maxwell system in 3\mathbb{R}^{3}, SIAM J. Appl. Math., 55 (1995), 1324–1344.
  • [24] H. Liu, On local and global structures of transmission eigenfunctions and beyond, J. Inverse Ill-Posed Probl., (2020), https://doi.org/10.1515/jiip-2020-0099.
  • [25] H. Liu, X. Liu, X. Wang and Y. Wang, On a novel inverse scattering scheme using resonant modes with enhanced imaging resolution, Inverse Problems, 35 (2019), 125012.
  • [26] H. Liu, Y. Wang, and S. Zhong, Nearly non-scattering electromagnetic wave set and its application, Z. Angew. Math. Phys, 68 (2017), no. 2, 35.
  • [27] J. Marschall, The trace of Sobolev-Slobodeckij spaces on Lipschitz domains, Manuscripta Math., 58 (1987), no. 1, 47–65.
  • [28] P. Monk, Finite Element Methods for Maxwell’s Equations, Oxford University Press, New York, 2003.
  • [29] P. Monk and J. Sun, Finite element methods for Maxwell’s transmission eigenvalues, SIAM J. Sci. Comput., 34 (2012), no. 3, B247–B264.