Publication Cover
Applicable Analysis
An International Journal
Volume 98, 2019 - Issue 4
1,556
Views
9
CrossRef citations to date
0
Altmetric
Articles

The inverse electromagnetic scattering problem by a penetrable cylinder at oblique incidence

&
Pages 781-798 | Received 07 Jan 2017, Accepted 19 Jun 2017, Published online: 17 Nov 2017

ABSTRACT

In this work, we consider the method of non-linear boundary integral equation for solving numerically the inverse scattering problem of obliquely incident electromagnetic waves by a penetrable homogeneous cylinder in three dimensions. We consider the indirect method and simple representations for the electric and the magnetic fields in order to derive a system of five integral equations, four on the boundary of the cylinder and one on the unit circle where we measure the far-field pattern of the scattered wave. We solve the system iteratively by linearizing only the far-field equation. Numerical results illustrate the feasibility of the proposed scheme.

AMS SUBJECT CLASSIFICATIONS:

1. Introduction

The inverse obstacle scattering problem is to image the scattering object, i.e. find its shape and location, from the knowledge of the far-field pattern of the scattered wave. The medium is illuminated by light at given direction and polarization. Then, Maxwell’s equations are used to model the propagation of the light through the medium, see [Citation1,Citation2] for an overview. This problem is of great interest because of its applications in many areas of physics and engineering (non-destructive testing, biomedical imaging, remote sensing, and target identification). We refer to [Citation3Citation7] for some recent applications.

Due to the complexity of the combined system of equations for the electric and the magnetic fields, it is common to impose additional assumptions on the incident illumination and the nature of the scatterer. We consider time-harmonic incident electromagnetic plane wave that due to the linearity of the problem will result to a time-independent system of equations. In addition, the penetrable object is considered as an infinitely long homogeneous cylinder. Then, it is characterized by constant permittivity and permeability. The problem is further simplified if we impose normal incidence for the incident wave. However, in this work, we consider the more complicated case of oblique incidence.

The three-dimensional scattering problem modeled by Maxwell’s equations is then equivalent to a pair of two-dimensional Helmholtz equations for two scalar fields (the third components of the electric and the magnetic fields). This approach reduces the difficulty of the problem but results to more complicated boundary conditions. The transmission conditions now contain also the tangential derivatives of the electric and magnetic fields. In [Citation8], we showed that the corresponding direct problem is well-posed and we constructed a unique solution using the direct integral equation method. A similar problem has been considered for an impedance cylinder embedded in a homogeneous [Citation9] and in an inhomogeneous medium [Citation10]. Different numerical solutions of the direct problem have been also proposed using finite difference/element methods [Citation11,Citation12], the Galerkin method [Citation13], the Nyström method [Citation14], the Green’s tensor method [Citation15], the method of auxiliary sources [Citation16,Citation17], the generalized Debye method [Citation18], and the separation of variables method [Citation19Citation21].

On the other hand, the inverse problem is non-linear and ill-posed. The non-linearity is due to the dependence of the solution of the scattering problem on the unknown boundary curve. The smoothness of the mapping from the boundary to the far-field pattern reflects the ill-posedness of the inverse problem. The unique solvability of the inverse problem is still an open problem. The first and only, to our knowledge, uniqueness result was presented recently in [Citation22] for the case of an impedance cylinder using the Lax–Phillips method.

In this work, we solve the inverse problem by formulating an equivalent system of non-linear integral equations that is solved using a regularized iterative scheme. This method was introduced by Kress and Rundell [Citation23] and then considered in many different problems, in acoustic scattering problems [Citation24,Citation25], in elasticity [Citation26,Citation27] and in electrical impedance problem [Citation28]. Our iterative scheme is based on the idea of Johansson and Sleeman [Citation29] first applied to the inverse acoustic scattering problem for a sound soft object. See [Citation26,Citation30], for applications of the method in different problems. We assume integral representations for the solutions that result to a system consisting of four integral equations on the unknown boundary (considering the transmission conditions) and one on the unit circle (taking into account the asymptotic expansion of the solutions). In our case, compared to [Citation29,Citation30] where only smooth and weakly singular integral operators are present in the systems of equations, appears also a singular operator (the tangential derivative of the single layer) due to the much more involved transmission conditions.

We solve the system of equations in two steps. First, given an initial guess for the boundary curve, we solve the well-posed subsystem (equations on the boundary) to obtain the corresponding densities and then we solve the linearized (with respect to the boundary) ill-posed far-field equation to update the initial approximation of the radial function. We consider Tikhonov regularization and the normal equations are solved by the conjugate gradient method. To improve the reconstructions, we take also into account measurements for few incident waves.

This work can be seen as a first step for solving the problem in the more complicated anisotropic case. There, one has to treat the three-dimensional problem differently and the integral equation method will result to a more complicated system of equations. The simplification due to the symmetry of the problem is also questionable and the unique solvability even for the direct problem is still an open problem. We refer to [Citation31] for a numerical solution using a subspace-based optimization method.

The paper is organized as follows: in Section 2, we present the direct scattering problem, the elastic potentials, and the equivalent system of integral equations that provide us with the far-field data. The inverse problem is stated in Section 3, where we construct an equivalent system of integral equation using the indirect integral equation method. In Section 4, the two-step method for the parametrized form of the system and the necessary Fréchet derivative of the integral operators are presented. The numerical examples give satisfactory results and justify the applicability of the proposed iterative scheme.

2. The direct problem

We consider the scattering of an electromagnetic wave by a penetrable cylinder in R3. Let x=(x,y,z)R3. We denote by Ωint={x:(x,y)Ω,zR} the cylinder, where Ω is a bounded domain in R2 with smooth boundary Γ. The cylinder Ωint is oriented parallel to the z-axis and Ω is its horizontal cross section. We assume constant permittivity ϵ0 and permeability μ0 for the exterior domain Ωext:=R3\Ω¯int. The interior domain Ωint is also characterized by constant parameters ϵ1 and μ1.

We define the exterior magnetic Hext(x,t) and electric field Eext(x,t) for xΩext,tR and the interior fields Hint(x,t) and Eint(x,t) for xΩint,tR, that satisfy the Maxwell’s equations(1) ×Eext+μ0Hextt=0,×Hext-ϵ0Eextt=0,xΩext,×Eint+μ1Hintt=0,×Hint-ϵ1Eintt=0,xΩint.(1)

and the transmission conditions(2) n^×Eint=n^×Eext,n^×Hint=n^×Hext,xΓ,(2)

where n^ is the outward normal vector, directed into Ωext.

We illuminate the cylinder with an incident electromagnetic plane wave at oblique incidence, meaning transverse magnetic (TM) polarized wave. We define by θ the incident angle with respect to the negative z axis and by ϕ the polar angle of the incident direction d^ (in spherical coordinates), see Figure . Then, d^=(sinθcosϕ,sinθsinϕ,-cosθ) and the polarization vector is given by p^=(cosθcosϕ,cosθsinϕ,sinθ), satisfying d^p^, for θ(0,π). The upcoming analysis can also be carried out to the case of transverse electric polarized incident plane wave.

Figure 1. The geometry of the scattering problem.

Figure 1. The geometry of the scattering problem.

In the following, due to the linearity of the problem, we suppress the time-dependence of the fields and because of the cylindrical symmetry of the medium we express the incident fields as separable functions of x:=(x,y) and z.

Let ω>0 be the frequency and k0=ωμ0ϵ0 the wave number in Ωext. We define β=k0cosθ and κ0=k02-β2=k0sinθ and it follows that the incident fields can be decomposed to [Citation8](3) Einc(x;d^,p^)=einc(x)e-iβz,Hinc(x;d^,p^)=hinc(x)e-iβz,(3)

whereeinc(x)=1ϵ0p^eiκ0(xcosϕ+ysinϕ),hinc(x)=1μ0(sinϕ,-cosϕ,0)eiκ0(xcosϕ+ysinϕ).

After some calculations, we can reformulate Maxwell’s euqations (Equation1) as a system of equations only for the z-component of the electric and magnetic fields [Citation8]. The interior fields e3int(x) and h3int(x),xΩ1:=Ω and the exterior fields e3ext(x) and h3ext(x), xΩ0:=R2\Ω satisfy the Helmholtz equations(4) Δe3int+κ12e3int=0,Δh3int+κ12h3int=0,xΩ1,Δe3ext+κ02e3ext=0,Δh3ext+κ02h3ext=0,xΩ0,(4)

where κ12=μ1ϵ1ω2-β2. Here, we assume μ1ϵ1>μ0ϵ0cos2θ in order to have κ12>0.

The transmission conditions (Equation2) can also be written only for the z-component of the fields. Let (n^,τ^) be a local coordinate system, where n^=(n1,n2) is the outward normal vector and τ^=(-n2,n1) the outward tangent vector on Γ. We define n=n^·t, τ=τ^·t, where t=e1x+e2y and e1,e2 denote the unit vectors in R2. Then, we rewrite the boundary conditions as [Citation8](5) e3int=e3ext,xΓ,μ~1ωh3intn+β1e3intτ=μ~0ωh3extn+β0e3extτ,xΓ,h3int=h3ext,xΓ,ϵ~1ωe3intn-β1h3intτ=ϵ~0ωe3extn-β0h3extτ,xΓ,(5)

where μ~j=μj/κj2,ϵ~j=ϵj/κj2,βj=β/κj2, for j=0,1. The exterior fields are decomposed to e3ext=e3sc+e3inc and h3ext=h3sc+h3inc, where e3sc and h3sc denote the scattered electric and magnetic field, respectively. From (Equation3), we see that(6) e3inc(x)=1ϵ0sinθeiκ0(xcosϕ+ysinϕ),h3inc(x)=0.(6)

To ensure that the scattered fields are outgoing, we impose in addition the radiation conditions in R2: (7) limrre3scr-iκ0e3sc=0,limrrh3scr-iκ0h3sc=0,(7)

where r=|x|, uniformly over all directions.

Now we are in position to formulate the direct transmission problem for oblique incident wave: find the fields h3int,h3sc,e3int, and e3sc that satisfy the Helmholtz equations (Equation4), the transmission conditions (Equation5) and the radiation conditions (Equation7).

Theorem 2.1:

If κ12 is not an interior Dirichlet eigenvalue and κ02 is not an interior Dirichlet and Neumann eigenvalue, then the direct transmission problem (Equation4)–(Equation7) admits a unique solution.

Proof The proof is based on the integral representation of the solution resulting to a Fredholm type system of boundary integral equations. For more details, see [Citation8, Theorem 3.2].

In the following, j=0,1 counts for the exterior (xΩ0) and interior domain (xΩ1), respectively. We introduce the single- and double-layer potentials defined by:(8) (Sjf)(x)=ΓΦj(x,y)f(y)ds(y),xΩj,(Djf)(x)=ΓΦjn(y)(x,y)f(y)ds(y),xΩj,(8)

where Φj is the fundamental solution of the Helmholtz equation in R2: (9) Φj(x,y)=i4H0(1)(κj|x-y|),x,yΩj,xy,(9)

and H0(1) is the Hankel function of the first kind and zero order. We define also the integral operators(10) (Sjf)(x)=ΓΦj(x,y)f(y)ds(y),xΓ,(Djf)(x)=ΓΦjn(y)(x,y)f(y)ds(y),xΓ,(NSjf)(x)=ΓΦjn(x)(x,y)f(y)ds(y),xΓ,(NDjf)(x)=Γ2Φjn(x)n(y)(x,y)f(y)ds(y),xΓ,(TSjf)(x)=ΓΦjτ(x)(x,y)f(y)ds(y),xΓ,(TDjf)(x)=Γ2Φjτ(x)n(y)(x,y)f(y)ds(y),xΓ.(10)

The following theorem was proven in [Citation8].

Theorem 2.2:

Let the assumptions of Theorem 2.1 still hold. Then, the potentials(11) e3int(x)=-(D1ϕ1)(x)+(S1η1)(x),xΩ1,h3int(x)=-(D1ψ1)(x)+(S1ξ1)(x),xΩ1,e3ext(x)=(D0ϕ0)(x)-(S0η0)(x),xΩ0,h3ext(x)=(D0ψ0)(x)-(S0ξ0)(x),xΩ0,(11)

solve the direct transmission problem (Equation4)–(Equation7) provided that the densities ϕ0H1/2(Γ) and ψ0H1/2(Γ) satisfy the system of integral equations(12) (D0+K0)ϕ0ψ0=b0,(12)

whereD0=D0-12I00D0-12I,K0=-ϵ~1ϵ~0S0K1-1ϵ~0ωS0(β1L1+β0L0)1μ~0ωS0(β1L1+β0L0)-μ~1μ~0S0K1,b0=-S0η+ϵ~1ϵ~0S0K1-1μ~0ωS0(β0τ+β1L1)e3inc,

and Kj:=NSj±12I-1NDj,Lj:=2(TDj-TSjKj). The rest of the densities satisfy ϕ1=ϕ0+e3inc,ψ1=ψ0,ηj=Kjϕj and ξj=Kjψj.

The solutions e3sc and h3sc of (Equation4)–(Equation7) have the asymptotic behavior(13) e3sc(x)=eiκ0rre(x^)+O(r-3/2),h3sc(x)=eiκ0rrh(x^)+O(r-3/2),(13)

where x^=x/|x|. The pair (e,h) is called the far-field pattern corresponding to the scattering problem (Equation4)–(Equation7). Its knowledge is essential for the inverse problem and using (Equation11) we can compute it by: (14) e(x^)=(Dϕ0)(x^)-(Sη0)(x^),x^S,h(x^)=(Dψ0)(x^)-(Sξ0)(x^),x^S,(14)

where S is the unit ball. The far-field operators are given by: (15) (Sf)(x^)=ΓΦ(x^,y)f(y)ds(y),x^S,(Df)(x^)=ΓΦn(y)(x^,y)f(y)ds(y),x^S,(15)

where Φ is the far-field of the Green function Φ, given by:Φ(x^,y)=eiπ/48πκ0e-iκ0x^·y.

3. The inverse problem

The inverse scattering problem, we address here, reads: find the shape and the position of the inclusion Ω, meaning reconstruct its boundary Γ, given the far-field patterns (e(x^),h(x^)), for all x^S, for one or few incident fields (Equation6).

3.1. The integral equation method

To solve the inverse problem we apply the method of non-linear boundary integral equations, which in our case, results to a system of four integral equations on the unknown boundary and one on the unit circle where the far-field data are defined. This method was first introduced in [Citation23] and further considered in various inverse problems, see for instance, [Citation25,Citation27,Citation32,Citation33]. Since the direct problem was solved with the direct method (Green’s formulas), in order to obtain our numerical data, here we adopt a different approach based on the indirect integral equation method, using simple representations for the fields.

We assume a double-layer representation for the interior fields and a single-layer representation for the exterior fields. Thus, we set(16) e3int(x)=1ϵ~1(D1ϕe)(x),h3int(x)=1μ~1(D1ϕh)(x),xΩ1,e3sc(x)=1ϵ~0(S0ψe)(x),h3sc(x)=1μ~0(S0ψh)(x),xΩ0.(16)

Substituting the above representations in the transmission conditions (Equation5) and considering the well-known jump relations, we get the system of integral equations on Γ (17) 1ϵ~1D1-12ϕe-1ϵ~0S0ψe=e3inc,ωND1ϕh+β1ϵ~1TD1-12τϕe-ωNS0-12ψh-β0ϵ~0TS0ψe=β0e3incτ,1μ~1D1-12ϕh-1μ~0S0ψh=0,ωND1ϕe-β1μ~1TD1-12τϕh-ωNS0-12ψe+β0μ~0TS0ψh=ϵ~0ωe3incn.(17)

In addition, given the far-field operators (Equation15) and the representations (Equation16) of the exterior fields, we see that the unknown boundary Γ and the densities ψe and ψh satisfy also the far-field equations (18a) 1ϵ~0Sψe=e,onS,(18a) (18b) 1μ~0Sψh=h,onS,(18b) where the right-hand sides are the known far-field patterns from the direct problem. The Equation (Equation17) in matrix form reads(19) (T+K)ϕ=b,(19)

whereT=ω2β12μ~1τ00[10pt]0-12μ~10000ω2-β12ϵ~1τ[10pt]000-12ϵ~1,K=-ωNS0-β1μ~1TD1β0μ~0TS0ωND1[10pt]01μ~1D1-1μ~0S00-β0ϵ~0TS0ωND1-ωNS0β1ϵ~1TD1[10pt]-1ϵ~0S0001ϵ~1D1,ϕ=ψeϕhψhϕe,b=ϵ~0ωn0β0τ1e3inc.

The matrix T due to its special form and the boundness of τ:H1/2(Γ)H-1/2(Γ) has a bounded inverse given by:(20) T-1=2ω2β1ωτ00[10pt]0-2μ~100[6pt]002ω-2β1ωτ[10pt]000-2ϵ~1.(20)

Then, Equation (Equation19) takes the form(21) (I+C)ϕ=g,(21)

where now I is the identity matrix andC=T-1K=-2NS002ωμ~0(β0-β1)TS02ND10-2D12μ~1μ~0S00-2ωϵ~0(β0-β1)TS02ND1-2NS002ϵ~1ϵ~0S000-2D1,g=T-1b=2ϵ~0n02ω(β0-β1)τ-2ϵ~1e3inc.

Using the mapping properties of the integral operators [Citation34], we see that the operator C:(H-1/2(Γ)×H1/2(Γ))2(H-3/2(Γ)×H-1/2(Γ))2 is compact.

We observe that we have six equations (Equation21) and () for the five unknowns: Γ and the four densities. Thus, we consider the linear combination ϵ~0·(Equation18a) + μ~0·(Equation18b) as a replacement for the far-field equations in order to state the following theorem as a formulation of the inverse problem.

Theorem 3.1:

Given the incident field (Equation6) and the far-field patterns (e(x^),h(x^)), for all x^S, if the boundary Γ and the densities ψe,ϕh,ψh and ϕe satisfy the system of equations (22a) ψe-2NS0ψe+2ωμ~0(β0-β1)TS0ψh+2ND1ϕe=2ϵ~0ne3inc,(22a) (22b) ϕh-2D1ϕh+2μ~1μ~0S0ψh=0,(22b) (22c) -2ωϵ~0(β0-β1)TS0ψe+2ND1ϕh+ψh-2NS0ψh=2ω(β0-β1)τe3inc,(22c) (22d) 2ϵ~1ϵ~0S0ψe+ϕe-2D1ϕe=-2ϵ~1e3inc,(22d) (22e) Sψe+Sψh=ϵ~0e+μ~0h,(22e) then, Γ solves the inverse problem.

The integral operators in () are linear with respect to the densities but non-linear with respect to the unknown boundary Γ. The smoothness of the kernels in the far-field Equation (Equation22e) reflects the ill-posedness of the inverse problem.

To solve the above system of equations, we consider the method first introduced in [Citation29] and then applied in different problems, see for instance [Citation26,Citation30,Citation35]. More precisely, given an initial approximation for the boundary Γ, we solve the subsystem (Equation22a)–(Equation22d) for the densities ψe,ϕh,ψh and ϕe. Then, keeping the densities ψe and ψh fixed we linearize the far-field Equation (Equation22e) with respect to the boundary. The linearized equation is solved to obtain the update for the boundary. The linearization is performed using Fréchet derivatives of the operators and we also regularize the ill-posed last equation.

To present the proposed method in detail, we consider the following parametrization for the boundaryΓ={z(t)=r(t)(cost,sint):t[0,2π]},

where z:RR2 is a C2-smooth, 2π-periodic, injective in [0,2π), meaning that z(t)0, for all t[0,2π]. The non-negative function r represents the radial distance of Γ from the origin. Then, we defineζe(t)=ψe(z(t)),ζh(t)=ψh(z(t)),t[0,2π]ξe(t)=ϕe(z(t)),ξh(t)=ϕh(z(t)),t[0,2π]

and the parametrized form of () is given by:(23) A1A2A3A4A5(r;ζe)+B1B2B3B4B5(r;ξh)+C1C2C3C4C5(r;ζh)+D1D2D3D4D5(r;ξe)=F1F2F3F4F5,(23)

with the parametrized operators(A1(r;ζ))(t)=(C3(r;ζ))(t)=ζ(t)-202πMNS0(t,s)ζ(s)ds,(A3(r;ζ))(t)=-μ~0ϵ~0(C1(r;ζ))(t)=-2ωϵ~0(β0-β1)02πMTS0(t,s)ζ(s)ds,(A4(r;ζ))(t)=μ~0ϵ~1μ~1ϵ~0(C2(r;ζ))(t)=2ϵ~1ϵ~002πMS0(t,s)ζ(s)ds,(A5(r;ζ))(t)=(C5(r;ζ))(t)=02πΦ(z^(t),z(s))ζ(s)|z(s)|ds,(B2(r;ξ))(t)=(D4(r;ξ))(t)=ξ(t)-202πMD1(t,s)ξ(s)ds,(B3(r;ξ))(t)=(D1(r;ξ))(t)=202πMND1(t,s)ξ(s)ds,

and the right-hand side(F1(r))(t)=2ϵ~0ne3inc(z(t)),(F3(r))(t)=2ω(β0-β1)τe3inc(z(t)),(F4(r))(t)=-2ϵ~1e3inc(z(t)),(F5)(t)=ϵ~0e(z^(t))+μ~0h(z^(t)).

In addition, we set A2=B1=B4=B5=C4=D2=D3=D5=F2=0. The matrix MKj denotes the discretized kernel of the operator Kj. The explicit forms of the kernels can be found for example, in [Citation8, Equation 4.3]. The operators Ak,Bk,Ck,Dk, k=1,2,3,4,5 act on the densities and the first variable r shows the dependence on the unknown parametrization of the boundary. Only F5 is independent of the radial function.

Let the function q stand for the radial function of the perturbed boundaryΓq={q(t)=q(t)(cost,sint):t[0,2π]}.

Then the iterative method reads:

Iterative Scheme 1:

Let r(0) be an initial approximation of the radial function. Then, in the kth iteration step:

(i)

We assume that we know r(k-1) and we solve the subsystem (24) A1A2A3A4(r(k-1);ζe)+B1B2B3B4(r(k-1);ξh)+C1C2C3C4(r(k-1);ζh)+D1D2D3D4(r(k-1);ξe)=F1F2F3F4,(24) to obtain the densities ζe(k),ξh(k),ζh(k), and ξe(k).

(ii)

Keeping the densities ζe and ζh fixed, we linearize the fifth equation of (Equation23), namely (25) A5(r(k-1);ζe(k))+(A5(r(k-1);ζe(k)))(q)+C5(r(k-1);ζh(k))+(C5(r(k-1);ζh(k)))(q)=F5.(25) We solve this equation for q and we update the radial function r(k)=r(k-1)+q.

The iteration stops when a suitable stopping criterion is satisfied.

Remark 1:

In order to take advantage of the available measurement data, we can also keep the overdetermined system (Equation17) and () instead of (Equation22e) and replace Equation (Equation25) with(26) A5(r(k-1);ζe(k))A5(r(k-1);ζh(k))q=FeFh-A5(r(k-1);ζe(k))A5(r(k-1);ζh(k)),(26)

where now Fe=ϵ~0e and Fh=μ~0h.

The Fréchet derivatives of the operators are calculated by formally differentiating their kernels with respect to r (27) ((A5(r;ζ))(q))(t)=eiπ/48πκ002πe-iκ0z^(t)·z(s)(-iκ0z^(t)·q(s)|z(s)|+z(s)·q(s)|z(s)|)ζ(s)ds.(27)

Recall that A5=C5=S. If κ02 is not an interior Neumann eigenvalue, then the operator A5 is injective [Citation25]. Using similar arguments as in [Citation30,Citation36], we can relate the above iterative scheme to the classical Newton’s method.

The iterative scheme 3.1 can also be generalized to the case of multiple illuminations elinc,l=1,,L.

Iterative Scheme 2:

[Multiple illuminations] Let r(0) be an initial approximation of the radial function. Then, in the kth iteration step:

(i)

We assume that we know r(k-1) and we solve the L subsystems (28) A1A2A3A4(r(k-1);ζe,l)+B1B2B3B4(r(k-1);ξh,l)+C1C2C3C4(r(k-1);ζh,l)+D1D2D3D4(r(k-1);ξe,l)=F1,lF2,lF3,lF4,l,l=1,,L(28) to obtain the densities ζe,l(k),ξh,l(k),ζh,l(k) and ξe,l(k).

(ii)

Then, keeping the densities fixed, we solve the overdetermined version of the linearized fifth equation of (Equation23) A5(r(k-1);ζe,1(k)+ζh,1(k))A5(r(k-1);ζe,2(k)+ζh,2(k))A5(r(k-1);ζe,l(k)+ζh,l(k))q=F5,1-A5(r(k-1);ζe,1(k)+ζh,1(k))F5,2-A5(r(k-1);ζe,2(k)+ζh,2(k))F5,L-A5(r(k-1);ζe,l(k)+ζh,l(k)) for q and we update the radial function r(k)=r(k-1)+q.

The iteration stops when a suitable stopping criterion is satisfied.

4. Numerical implementation

In this section, we present numerical examples that illustrate the applicability of the proposed method. We use quadrature rules for integrating the singularities considering trigonometric interpolation. The convergence and error analysis are given in [Citation37,Citation38]. Then, the system of integral equations is solved using the Nyström method. The parametrized forms of the integral operators are presented in [Citation8, Section 4]. We approximate the smooth kernels with the trapezoidal rule and the singular ones with the well-known quadratures rules [Citation38].

In the following examples, we consider two different boundary curves. A peanut-shaped and an apple-shaped boundary with radial functionr(t)=(0.5cos2t+0.15sin2t)1/2,t[0,2π],

andr(t)=0.45+0.3cost-0.1sin2t1+0.7cost,t[0,2π],

respectively.

To avoid an inverse crime, we construct the simulated far-field data using the numerical scheme (Equation12) and considering double amount of quadrature points compared to the inverse problem. We approximate the radial function q by a trigonometric polynomial of the formq(t)k=0makcoskt+k=1mbksinkt,t[0,2π],

and we consider 2n equidistant points tj=jπ/n,j=0,,2n-1. The well-posed subsystem (Equation24) does not require any special treatment. The ill-posed linearized far-field Equation (Equation25) is solved by Tikhonov regularization. We rewrite (Equation25) as:(29) (A5(r(k-1);ζ(k)))(q)=F5-A5(r(k-1);ζ(k)),(29)

for ζ(k):=ζe(k)+ζh(k), and we decompose (Equation27) as:(30) ((A5(r;ζ))(q))(t)=((G1(r;ζ))(q))(t)+((G2(r;ζ))(q))(t),(30)

where((G1(r;ζ))(q))(t):=eiπ/48πκ002πe-iκ0z^(t)·z(s)-iκ0z^(t)·(coss,sins)|z(s)|+z(s)·(-sins,coss)|z(s)|ζ(s)q(s)ds,((G2(r;ζ))(q))(t):=eiπ/48πκ002πe-iκ0z^(t)·z(s)z(s)·(coss,sins)|z(s)|ζ(s)q(s)ds.

We replace the derivative of q by the derivative of the trigonometric interpolation polynomialq(t)j=02n-1Q(t,tj)q(tj),

with weightQ(tk,tj)=12(-1)k-jcottk-tj2,kj,k=0,,2n-1.

Then, at the kth step we minimize the Tikhonov functional of the discretized equationATx-b22+λxpp,λ>0,

where xR(2m+1)×1 is the vector with the unknowns coefficients a0,,am, b1,,bm of the radial function, and AC2n×2n,bC2n×1 are given by:Akj=MG1(tk,tj)+MG2(tk,tj)Q(tk,tj),bk=F5(tk)-(MA5ζ)(tk),

for k,j=0,,2n-1. The multiplication matrix TR2n×(2m+1) stands for the trigonometric functions of the approximated radial function and is given by:Tkj=coskjπn,k=0,,2n-1,j=0,,msink(j-m)πn,k=0,,2n-1,j=m+1,,2m

Here, p0 defines the corresponding Sobolev norm. Since q is real valued, we solve the following regularized equation:(31) TR(A)R(A)+I(A)I(A)T+λkIpx=TR(A)R(b)+I(A)I(b),(31)

on the kth step, where the matrix IpR(2m+1)×(2m+1) corresponds to the Sobolev Hp penalty term. We solve (Equation31) using the conjugate gradient method. We update the regularization parameter in each iteration step k by:λk=λ023k-1,k=1,2,

for some given initial parameter λ0>0. To test the stability of the iterative method against noisy data, we add also noise to the far-field patterns with respect to the L2-normeδ=e+δ1e2u2u,hδ=h+δ2h2v2v,

for some given noise levels, δ1,δ2 where u=u1+iu2,v=v1+iv2, for u1,u2,v1,v2R normally distributed random variables.

Figure 2. Reconstruction of a peanut-shaped boundary for two incident fields, frequency ω=2.5, for exact data (left) and data with 5% noise (right).

Figure 2. Reconstruction of a peanut-shaped boundary for two incident fields, frequency ω=2.5, for exact data (left) and data with 5% noise (right).

Already in simpler cases [Citation30], the knowledge of the far-field patterns for one incident wave is not enough to produce satisfactory reconstructions. Thus, we will also use multiple incident directions. To do so, we have to consider different values of the polar angle ϕ since in R2, as we see from (Equation6), corresponds to the incident direction d=(cosϕ,sinϕ). We setdl=cosϕl,sinϕl,whereϕl=2πlL,forl=1,,L.

4.1. Numerical results

We present reconstructions for different boundary curves, different number of incident directions and initial guesses for exact and perturbed far-field data. In all figures the initial guess is a circle with radius r0, a green solid line, the exact curve is represented by a dashed red line and the reconstructed by a solid blue line. The arrows denote the directions of the incoming incident fields.

We use n=64 collocation points for the direct problem and n=32 for the inverse. In the first five examples, we set the exterior parameters (ϵ0,μ0)=(1,1) and the interior (ϵ1,μ1)=(2,2). We set θ=π/3 and λ0[0.5,0.8] as the initial regularization parameter.

Figure 3. Reconstruction of a peanut-shaped boundary for four incident fields, frequency ω=2.5, noisy data (5% noise), with initial guess r0=0.6 (left) and r0=1 (right).

Figure 3. Reconstruction of a peanut-shaped boundary for four incident fields, frequency ω=2.5, noisy data (5% noise), with initial guess r0=0.6 (left) and r0=1 (right).

Figure 4. Reconstruction of a peanut-shaped boundary for four incident fields, m=5 coefficients, frequency ω=2, for exact data (left) and data with 3% noise (right).

Figure 4. Reconstruction of a peanut-shaped boundary for four incident fields, m=5 coefficients, frequency ω=2, for exact data (left) and data with 3% noise (right).

In the first three examples, we consider the peanut-shaped boundary. In the first example, the regularized Equation (Equation31) is solved with L2 penalty term, meaning p=0 and m=3 coefficients. We solve Equation (Equation26) for different incident directions. The reconstructions for ω=2.5 and r0=0.6 are presented in Figure for two incident fields with directions dl+1/2. On the left picture, we see the reconstructed curve for exact data and 9 iterations and on the right picture for noisy data with δ1=δ2=5% and 14 iterations. In the second example, we consider Equation (Equation25), four incident fields, noisy data δ1=δ2=5% and we keep all the parameters as before. The reconstructions for r0=0.6 and 14 iterations are shown in the left picture of Figure , and for r0=1 and 20 iterations in the right one. We set m=5 and p=1 (H1 penalty term) in the third example. The results for r0=1 and four incident fields are shown in Figure . Here, ω=2 and we use Equation (Equation26). We need 26 iterations for the exact data and 30 iterations for the noisy data (δ1=δ2=3%).

Figure 5. Reconstruction of an apple-shaped boundary for four incident fields, frequency ω=3, exact data, with initial guess r0=0.5 (left) and r0=1 (right).

Figure 5. Reconstruction of an apple-shaped boundary for four incident fields, frequency ω=3, exact data, with initial guess r0=0.5 (left) and r0=1 (right).

Figure 6. Reconstruction of an apple-shaped boundary for four incident fields, frequency ω=3, data with 3% noise, for three (left) and four (right) incident fields.

Figure 6. Reconstruction of an apple-shaped boundary for four incident fields, frequency ω=3, data with 3% noise, for three (left) and four (right) incident fields.

In the next two examples we consider the apple-shaped boundary, H1 penalty term, ω=3 and m=3 coefficients. In the fourth example, we consider Equation (Equation25), noise-free data and four incident fields in order to examine the dependence of the iterative scheme on the initial radial guess. On the left picture of Figure , we see the reconstructed curve for r0=0.5 after 13 iterations and on the right picture for r0=1 after 20 iterations. In the fifth example, we consider δ1=δ2=3% noise and r0=0.6. Figure shows the improvement of the reconstruction for more incident fields. On the left picture we see the results for three incident fields, Equation (Equation26) and seven iterations and the reconstructed curve for four incident fields, Equation (Equation25) and 15 iterations is shown on the right picture.

In the last example we consider an electrically larger object, meaning we set (ϵ0,μ0)=(3,3) and ω=7, resulting to a seven times larger (electrically) scatterer compared to the previous examples. We choose (ϵ1,μ1)=(9,1) such that the condition μ1ϵ1>μ0ϵ0cos2θ is satisfied. To account also for different oblique directions we set θ=π/10. For m=3 and data with 3% noise, we present the reconstructions in Figure . We consider for both boundary curves the same initial guess r0=0.6. The results on the left picture are for two incident fields, considering Equation (Equation26) and 35 iterations. On the right picture we present the reconstruction for four incident fields, Equation (Equation25) and 13 iterations.

Figure 7. Reconstruction of a peanut-shaped boundary for two incident fields (left) and an apple-shaped boundary for four incident fields (right). Here, we use θ=π/10,ω=7 and data with 3% noise.

Figure 7. Reconstruction of a peanut-shaped boundary for two incident fields (left) and an apple-shaped boundary for four incident fields (right). Here, we use θ=π/10,ω=7 and data with 3% noise.

To conclude, our examples have shown the feasibility of the proposed iterative scheme and the stability against noisy data. However, this method can only be applied to objects with smooth boundaries. In addition, the proposed method performs poorly for only one incident field which is the case also in the acoustic regime. The main reason is that we miss information since we linearize only the far-field equation. Thus, we had to consider few incident illuminations which improve considerably the reconstructions. The initial guess plays also an important role in this scheme.

Notes

No potential conflict of interest was reported by the authors.

References

  • Colton D , Kress R . Integral equation methods in scattering theory, Classics in applied mathematics. Philadelphia (PA): SIAM; 2013.
  • Colton D , Kress R . Inverse acoustic and electromagnetic scattering theory. 3rd ed. Vol. 93. Applied mathematical sciences. New York (NY): Springer-Verlag; 2013.
  • Caorsi S , Donelli M , Franceschini D , et al . A new methodology based on an iterative multiscaling for microwave imaging. IEEE Trans Microwave Theory Tech. 2003;51(4):1162–1173.
  • Caorsi S , Massa A , Pastorino M , et al . Detection of buried inhomogeneous elliptic cylinders by a memetic algorithm. IEEE Trans Antennas Propag. 2003;51(10):2878–2884.
  • Massa A , Pastorino M , Randazzo A . Reconstruction of two-dimensional buried objects by a differential evolution method. Inverse Prob. 2004;20(6):S135–S150.
  • Meng Q , Xu K , Shen F , et al . Microwave imaging under oblique illumination. Sensors. 2016;16(7):1046.
  • Oliveri G , Lizzi L , Pastorino M , et al . A nested multi-scaling inexact-Newton iterative approach for microwave imaging. IEEE Trans Antennas Propag. 2012;60(2):971–983.
  • Gintides D , Mindrinos L . The direct scattering problem of obliquely incident electromagnetic waves by a penetrable homogeneous cylinder. J. Integral Equ Appl. 2016;28(1):91–122.
  • Wang H , Nakamura G . The integral equation method for electromagnetic scattering problem at oblique incidence. Appl Numer Math. 2012;62:860–873.
  • Nakamura G , Wang H . The direct electromagnetic scattering problem from an imperfectly conducting cylinder at oblique incidence. J Math Anal Appl. 2013;397:142–155.
  • Cangellaris AC , Lee R . Finite element analysis of electromagnetic scattering from inhomogeneous cylinders at oblique incidence. IEEE Trans Antennas Propag. 1991;39:645–650.
  • Yan J , Gordon RK , Kishk AA . Electromagnetic scattering from impedance elliptic cylinders using finite difference method (oblique incidence). Electromagnetics. 1995;15(2):157–173.
  • Lucido M , Panariello G , Schettiho F . Scattering by polygonal cross-section dielectric cylinders at oblique incidence. IEEE Trans Antennas Propag. 2010;58:540–551.
  • Tsalamengas JL . Exponentially converging Nyström methods applied to the integral-integrodifferential equations of oblique scattering / hybrid wave propagation in presence of composite dielectric cylinders of arbitrary cross section. IEEE Trans Antennas Propag. 2007;55(11):3239–3250.
  • Martin OJF , Piller NB . Electromagnetic scattering in polarizable backgrounds. Phys Rev E. 1998;58(3):3909–3915.
  • Tsitsas NL , Alivizatos EG , Anastassiu HT , et al . Optimization of the method of auxiliary sources (MAS) for scattering by an infinite cylinder under oblique incidence. Electromagnetics. 2005;25(1):39–54.
  • Tsitsas NL , Alivizatos EG , Anastassiu HT , et al . Optimization of the method of auxiliary sources (MAS) for oblique incidence scattering by an infinite dielectric cylinder. Electr Eng. 2007;89:353–361.
  • Li R , Han X , Ren KF . Generalized Debye series expansion of electromagnetic plane wave scattering by an infinite multilayered cylinder at oblique incidence. Phys Rev E. 2009;79(3):036602.
  • Mao SC , Wu ZS , Li HY . Three-dimensional scattering by an infinite homogeneous anisotropic elliptic cylinder in terms of Mathieu functions. J Opt Soc Am A. 2009;26:2282–2291.
  • Zouros GP . Electromagnetic plane wave scattering by arbitrarily oriented elliptical dielectric cylinders. J Opt Soc Am A. 2011;28(11):2376–2384.
  • Zouros GP . Oblique electromagnetic scattering from lossless or lossy composite elliptical dielectric cylinders. J Opt Soc Am. 2013;30(2):196–205.
  • Nakamura G , Sleeman BD , Wang H . On uniqueness of an inverse problem in electromagnetic obstacle scattering for an impedance cylinder. Inverse Prob. 2012;28(5):055012.
  • Kress R , Rundell W . Nonlinear integral equations and the iterative solution for an inverse boundary value problem. Inverse Prob. 2005;21:1207–1223.
  • Ivanyshyn O , Johansson BT . Nonlinear integral equation methods for the reconstruction of an acoustically sound-soft obstacle. J Integral Equ Appl. 2007;19(3):289–308.
  • Ivanyshyn O , Johansson BT . Boundary integral equations for acoustical inverse sound-soft scattering. J Inverse Ill-posed Prob. 2008;16(1):65–78.
  • Chapko R , Gintides D , Mindrinos L . The inverse scattering problem by an elastic inclusion. Adv Comput Math. 2017.
  • Gintides D , Midrinos L . Inverse scattering problem for a rigid scatterer or a cavity in elastodynamics. ZAMM Z Angew Math Mech. 2011;91(4):276–287.
  • Eckel H , Kress R . Nonlinear integral equations for the inverse electrical impedance problem. Inverse Prob. 2007;23(2):475.
  • Johansson BT , Sleeman BD . Reconstruction of an acoustically sound-soft obstacle from one incident field and the far-field pattern. IMA J Appl Math. 2007;72:96–112.
  • Altundag A , Kress R . On a two-dimensional inverse scattering problem for a dielectric. Appl Anal. 2012;91(4):757–771.
  • Agarwal K , Pan L , Chen X . Subspace-based optimization method for reconstruction of 2-D complex anisotropic dielectric objects. IEEE Trans Microwave Theory Tech. 2010;58(4):1065–1074.
  • Cakoni F , Kress R . Integral equations for inverse problems in corrosion detection from partial cauchy data. Inverse Prob Imaging. 2007;1(2):229–245.
  • Cakoni F , Kress R , Schuft C . Integral equations for shape and impedance reconstruction in corrosion detection. Inverse Prob. 2010;26:095012.
  • Kress R . Linear integral equations. 3rd ed. New York (NY): Springer; 2014.
  • Lee KM . Inverse scattering problem from an impedance crack via a composite method. Wave Motion. 2015;56:43–51.
  • Hohage T , Schormann C . A Newton-type method for a transmission problem in inverse scattering. Inverse Prob. 1998;14(5):1207.
  • Kress R . On the numerical solution of a hypersingular integral equation in scattering theory. J Comput Appl Math. 1995;61(3):345–360.
  • Kress R . A collocation method for a hypersingular boundary integral equation via trigonometric differentiation. J Integral Equ Appl. 2014;26(2):197–213.