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

On the series solutions of
integral equations in scattering

Mirza Karamehmedović and Faouzi Triki
Abstract.

We study the validity of the Neumann or Born series approach in solving the Helmholtz equation and coefficient identification in related inverse scattering problems. Precisely, we derive a sufficient and necessary condition under which the series is strongly convergent. We also investigate the rate of convergence of the series. The obtained condition is optimal and it can be much weaker than the traditional requirement for the convergence of the series. Our approach makes use of reduction space techniques proposed by Suzuki [20]. Furthermore we propose an interpolation method that allows the use of the Neumann series in all cases. Finally, we provide several numerical tests with different medium functions and frequency values to validate our theoretical results.

Key words and phrases:
Helmholtz equation; Born series; Scattering
1991 Mathematics Subject Classification:
35R30; 34L25; 78A46

1. Introduction and main results

Let d=1,2,d=1,2,\dots, fix a positive k0k_{0} and k^Sd1\widehat{k}\in S^{d-1}, let qL(𝑹d)q\in L^{\infty}(\bm{R}^{d}) be compactly supported with q(x)>1q(x)>-1, and consider the following Helmholtz equation in 𝑹d\bm{R}^{d} with the Sommerfeld radiation condition:

(1) {(Δ+k02(1+q(x)))u=k02q(x)eik0k^.xin𝑹d,lim|x||x|(d1)/2(|x|ik)u=0uniformly inx/|x|Sd1.\left\{\begin{array}[]{rcl}(\Delta+k_{0}^{2}(1+q(x)))u&=&-k_{0}^{2}q(x)e^{ik_{0}\widehat{k}.x}\quad\text{in}\,\,\,\bm{R}^{d},\\ \lim_{|x|\rightarrow\infty}|x|^{(d-1)/2}\left(\partial_{|x|}-ik\right)u&=&0\quad\text{uniformly in}\,\,\,x/|x|\in S^{d-1}.\end{array}\right.

Convolving the PDE in (1) with the outgoing fundamental solution111Here H0(1)H_{0}^{(1)} the Hankel function of the first kind and order zero. of the Helmholtz operator Δ+k02\Delta+k_{0}^{2} in 𝑹d\bm{R}^{d},

Φd(x)={(2π|x|)(1d)/2(2ik0)1|x|(d1)/2eik0|x|,x𝑹d{0},dodd,(2π|x|)(2d)/2(4i)1|x|(d2)/2H0(1)(k0|x|),x𝑹d{0},deven,\Phi_{d}(x)=\begin{cases}(-2\pi|x|)^{(1-d)/2}(2ik_{0})^{-1}\partial_{|x|}^{(d-1)/2}e^{ik_{0}|x|},\quad&x\in\bm{R}^{d}\setminus\{0\},\,\,\,d\,\,\,\text{odd},\\ (-2\pi|x|)^{(2-d)/2}(4i)^{-1}\partial_{|x|}^{(d-2)/2}H_{0}^{(1)}(k_{0}|x|),\quad&x\in\bm{R}^{d}\setminus\{0\},\,\,\,d\,\,\,\,\text{even},\end{cases}

and integrating by parts, we get the Lippmann-Schwinger equation

(2) (IVq(k0))u=Vq(k0)eik0k^()in𝑹d,(I-V_{q}(k_{0}))u=V_{q}(k_{0})e^{ik_{0}\widehat{k}(\cdot)}\quad\text{in}\,\,\,\bm{R}^{d},

where

Vq(k0)u(x)=k02ysuppqΦd(xy)q(y)u(y)𝑑yV_{q}(k_{0})u(x)=k_{0}^{2}\int_{y\in{\rm supp}\,q}\Phi_{d}(x-y)q(y)u(y)dy

exists as an improper integral for each x𝑹dx\in\bm{R}^{d}. It is well-known that the integral equation (2) is equivalent with (1), and that it suffices to solve (2) in, say, a bounded open ball B𝑹dB\subset\bm{R}^{d} that includes suppq{\rm supp}\,q, followed by the continuous extension u(x)=Vq(k0)[u()+exp(ik0k^())](x)u(x)=V_{q}(k_{0})[u(\cdot)+\exp(ik_{0}\widehat{k}(\cdot))](x) for x𝑹dBx\in\bm{R}^{d}\setminus B. The mapping Vq(k0):L2(B)L2(B)V_{q}(k_{0}):L^{2}(B)\rightarrow L^{2}(B) is compact, and we shall in the following consider only the restriction of the Lippmann-Schwinger equation in (2) to BB. The objective of the paper is to study the successive approximations for solving the integral equation (2):

(3) u0=Vq(k0)eik0k^;un+1=u0+Vq(k0)un,n𝑵.u_{0}=V_{q}(k_{0})e^{ik_{0}\widehat{k}};\;u_{n+1}=u_{0}+V_{q}(k_{0})u_{n},\;\;n\in\bm{N}.

The computational advantage of this iterative method is that it does need to solve the partial differential equation (1) in the whole space and deal with the radiation conditions. Instead, one can obtain a good approximation unu_{n} of the solution uu by applying successively the integral operator Vq(k0)V_{q}(k_{0}) if the sequence converges.

On the other hand the strong convergence of the sequence (un)n𝑵(u_{n})_{n\in\bm{N}} to the solution uu of the integral equation (2) is equivalent to the convergence of the Neumann series:

(4) limnun=j=0Vqj+1(k0)eik0k^=(IVq(k0))1Vq(k0)eik0k^.\lim_{n\to\infty}u_{n}=\sum_{j=0}^{\infty}V_{q}^{j+1}(k_{0})e^{ik_{0}\widehat{k}\cdot}=(I-V_{q}(k_{0}))^{-1}V_{q}(k_{0})e^{ik_{0}\widehat{k}\cdot}.

In inverse scattering problems the Neumann series approach known more under the name of Born approximation was initially employed to study the quantum mechanical inverse backscattering problem in one dimension (see for instance [25] and references therein). The principal advantage of using this technique in inverse medium problem is that it requires solving a linear equation instead of an oscillatory nonlinear one [10, 11]. It has also been applied to various other inverse problems, including optical and electrical impedance tomographies, acoustic and electromagnetic parameters identification [2, 30, 1, 23, 24, 26, 16, 13, 14, 4, 3]. However, it is important to note, that the strategies considered in these works are based on purely formal analysis or assume strong conditions on the targeted physical parameters.

It is well known that a sufficient condition for the convergence of the Neumann series (4) is that the spectral radius Spr(Vq(k0))\textrm{Spr}(V_{q}(k_{0})) of the compact operator Vq(k0)V_{q}(k_{0}) is strictly less than one, that is Spr(Vq(k0))<1\textrm{Spr}(V_{q}(k_{0}))<1. But this latter condition while it is optimal for the expansion of the operator (IVq(k0))1(I-V_{q}(k_{0}))^{-1} in L2(B)L^{2}(B), it is obviously too restrictive for the convergence of (4). Then is it possible to derive a necessary and sufficient condition for the convergence of only (4)? On the other hand the strong convergence

(5) Vq(k0)jeik0k^0,j,V_{q}(k_{0})^{j}e^{ik_{0}\widehat{k}\cdot}\rightarrow 0,\;\;j\to\infty,

in L2(B)L^{2}(B), is evidently a necessary condition for the convergence of the series (4). Suzuki in his seminal work [20] wondered if this condition is also sufficient. Surprisingly, it turns out that this condition also guarantees the convergence of the series. The main idea of the proof is to derive a minimal invariant space where the expansion of the restriction of (IVq(k0))1(I-V_{q}(k_{0}))^{-1} to that space is equivalent to the convergence of the series (4).
Let

(6) Lk0,k^2(B)=Span(Vq(k0)eik0k^,Vq(k0)2eik0k^,,Vq(k0)j+1eik0k^,).L^{2}_{k_{0},\hat{k}}(B)=\textrm{Span}\left(V_{q}(k_{0})e^{ik_{0}\widehat{k}\cdot},V_{q}(k_{0})^{2}e^{ik_{0}\widehat{k}\cdot},\cdots,V_{q}(k_{0})^{j+1}e^{ik_{0}\widehat{k}\cdot},\cdots\right).

By construction Lk0,k^2(B)L^{2}_{k_{0},\hat{k}}(B) is invariant by Vq(k0)V_{q}(k_{0}). Denote V~q(k0)\widetilde{V}_{q}(k_{0}) the restriction of Vq(k0)V_{q}(k_{0}) to Lk0,k^2(B)L^{2}_{k_{0},\hat{k}}(B). Suzuki showed that condition (5) indeed implies Spr(V~q(k0))<1\textrm{Spr}(\widetilde{V}_{q}(k_{0}))<1, and hence ensures the convergence of the Neumann series to the unique solution.

Proposition 1.

The convergence of the Neumann series (4) is equivalent to the condition (5).

Remark 1.

Since Vq(k0)V_{q}(k_{0}) is a compact operator the strong convergence (5) can be replaced by a weak convergence of a subsequence. Notice that Lk0,k^2(B)L^{2}_{k_{0},\hat{k}}(B) can also be generated by finite sums of the sequence

Lk0,k^2(B)=Span(j=0JVq(k0)j+1eik0k^;J𝑵).L^{2}_{k_{0},\hat{k}}(B)=\textrm{Span}\left(\sum_{j=0}^{J}V_{q}(k_{0})^{j+1}e^{ik_{0}\widehat{k}\cdot};\;\;J\in\bm{N}\right).

Recall that the traditional condition to ensure the convergence of the Neumann series is [11]

(7) Vq(k0)Ck0,q=(BB|k02Φd(xy)q(x)|2𝑑x𝑑y)1/2<1.\|V_{q}(k_{0})\|\leq C_{k_{0},q}=\left(\int_{B}\int_{B}|k_{0}^{2}\Phi_{d}(x-y)q(x)|^{2}dxdy\right)^{1/2}<1.

This condition occurs in the situation for weak scattering, and is not valid for high wave number k0k_{0}, or large magnitude of the medium function qq. But since eik0k^e^{ik_{0}\widehat{k}\cdot} is sparse we expect that Lk0,k^2(B)L^{2}_{k_{0},\hat{k}}(B) has a lower dimensionality than the whole space L2(B)L^{2}(B), and consequently the convergence of the Neumann series (4) may happen beyond the conventional limitation (7). In other words Spr(V~q(k0))<1\textrm{Spr}(\widetilde{V}_{q}(k_{0}))<1 can be satisfied by a larger class of wave numbers and medium functions not necessary within the weak scattering regime. This was also observed in many numerical experiments in the past, has fueled many discussions and was the origin of several investigations [23, 24, 26, 16, 13, 14, 3]. This pattern is clearly confirmed by many numerical examples in section 4.

Theorem 1.

Assume that the condition (5) is satisfied, that is

limnVq(k0)neik0k^=0.\lim_{n\to\infty}\|V_{q}(k_{0})^{n}e^{ik_{0}\widehat{k}\cdot}\|=0.

Then there exists a constant C>0C>0 independent of nn such that the following error estimate

(8) uunCVq(k0)neik0k^,\|u-u_{n}\|\leq C\|V_{q}(k_{0})^{n}e^{ik_{0}\widehat{k}\cdot}\|,

holds for all n𝐍.n\in\bm{N}.

The rest of the paper is organized as follows. In section 2, we provide the proofs for Proposition 1, and Theorem 1. Section 3 is devoted to the construction of a preconditioner for the integral equation (2). Precisely, we propose an interpolation method that allows the use of the Neumann series independently of the fact that the condition (5) is fulfilled or not. We present then several numerical experiments to show the effectiveness of the derived theoretical results in section 4.

2. Proof of the main results

In this section we shall prove the main results of the paper. To ease the notation we set

A=Vq(k0);ψ=Vq(k0)eik0k^;=L2(B);0=Lk0,k^2(B).A=V_{q}(k_{0});\;\;\psi=V_{q}(k_{0})e^{ik_{0}\widehat{k}\cdot};\;\;\mathfrak{H}=L^{2}(B);\;\;\mathfrak{H}_{0}=L^{2}_{k_{0},\hat{k}}(B).
Proof of Proposition 1.

If the series

(9) j=0AjΨ=(IA)1Ψ,\sum_{j=0}^{\infty}A^{j}\Psi=(I-A)^{-1}\Psi,

is convergent then obviously we will have AjΨ0A^{j}\Psi\rightarrow 0 strongly in \mathfrak{H}. Now assume that AjΨA^{j}\Psi converges strongly to zero in \mathfrak{H}, and focus on the nontrivial opposite direction.

The main observation of Suzuki is that the convergence of the series (9) depends more on the specific local behavior of the operator AA relative to the given vector ψ\psi rather than its global properties on the whole space \mathfrak{H} which requires that Spr(A)<1\textrm{Spr}(A)<1. Indeed consider the Hilbert subspace 0\mathfrak{H}_{0}\subset\mathfrak{H} space generated by the vectors AjΨ,j𝑵A^{j}\Psi,\,j\in\bm{N}, that is

(10) 0=Span(Ψ,AΨ,,AjΨ,).\mathfrak{H}_{0}=\textrm{Span}\left(\Psi,A\Psi,\cdots,A^{j}\Psi,\cdots\right).

Clearly 0\mathfrak{H}_{0} is invariant by AA, and since Ψ\Psi lies in 0\mathfrak{H}_{0} to prove that the series (9) strongly in 0\mathfrak{H}_{0} it is sufficient to show that A0A_{0} the restriction of AA to 0\mathfrak{H}_{0} verifies Spr(A0)<1\textrm{Spr}(A_{0})<1. Remark that since 0\mathfrak{H}_{0}\subset\mathfrak{H} we have Spr(A0)Spr(A)\textrm{Spr}(A_{0})\leq\textrm{Spr}(A).

Let 𝔐\mathfrak{M} be the linear manifold formed by the vectors v0v\in\mathfrak{H}_{0} satisfying AjvA^{j}v tends strongly to zero as jj\to\infty. We first remark that the fact jAjΨj\to A^{j}\Psi tends strongly to zero, 𝔐\mathfrak{M} contains all the vectors AjΨ,j𝑵A^{j}\Psi,\,j\in\bm{N}, and consequently is dense in 0\mathfrak{H}_{0}.

Let σ(A0)\sigma(A_{0}) denotes the spectrum of A0A_{0}, and set σ(A0)={λΣ(A0);|λ|<1}\sigma_{-}(A_{0})=\{\lambda\in\Sigma(A_{0});\;|\lambda|<1\} and σ+(A0)={λΣ(A0);|λ|1}\sigma_{+}(A_{0})=\{\lambda\in\Sigma(A_{0});\;|\lambda|\geq 1\}. Since A0A_{0} is compact σ+(A0)\sigma_{+}(A_{0}) is finite, in addition there exists a rectifiable Jordan curve 𝒞+\mathcal{C}_{+} in the resolvent set surrounding σ+(A0)\sigma_{+}(A_{0}) and does not contain other eigenvalues of σ(A0)\sigma(A_{0}). Similarly there exists a rectifiable Jordan curve 𝒞\mathcal{C}_{-} in the resolvent set surrounding only σ(A0)\sigma_{-}(A_{0}). Then following [18], the spectral projections

(11) P±=12iπ𝒞±(λIA0)1𝑑λ,\displaystyle P_{\pm}=\frac{1}{2i\pi}\int_{\mathcal{C}_{\pm}}(\lambda I-A_{0})^{-1}d\lambda,

verify the following identities

(12) P+P+=I;PP+=P+P=0;P±A=AP±.P_{-}+P_{+}=I;\;P_{-}P_{+}=P_{+}P_{-}=0;\;P_{\pm}A=AP_{\pm}.

Recalling that Spr(A0)=supλσ(A0)|λ|.\textrm{Spr}(A_{0})=\sup_{\lambda\in\sigma(A_{0})}{|\lambda|}. Since σ(A0)\sigma(A_{0}) is a sequence of complex values that may converge to zero, proving that Spr(A0)<1\textrm{Spr}(A_{0})<1 is equivalent to show that σ+(A0)\sigma_{+}(A_{0}) is an empty set. Let now vP+0v\in P_{+}\mathfrak{H}_{0}. By the density of the set 𝔐\mathfrak{M} in 0\mathfrak{H}_{0}, there exists a sequence (vn)n𝑵0𝔐(v_{n})_{n\in\bm{N}_{0}}\in\mathfrak{M} that converges strongly to vv. 𝑵0\bm{N}_{0} here is the set 𝑵{0}\bm{N}\setminus\{0\}. Denote vn,±=P±vnv_{n,\pm}=P_{\pm}v_{n}. Therefore vn=vn,++vn,v_{n}=v_{n,+}+v_{n,-}. Remarking that Avn,=APvn=PAvnAv_{n,-}=AP_{-}v_{n}=P_{-}Av_{n} converges strongly to Pv=0P_{-}v=0 leads to vn,𝔐v_{n,-}\in\mathfrak{M}. Hence vn,+=vnvn,v_{n,+}=v_{n}-v_{n,-} lies in fact in 𝔐P+0\mathfrak{M}\cap P_{+}\mathfrak{H}_{0}. This shows that 𝔐P+0\mathfrak{M}\cap P_{+}\mathfrak{H}_{0} is indeed dense in P+0P_{+}\mathfrak{H}_{0}. But since σ+(A0)\sigma_{+}(A_{0}) is finite P+0P_{+}\mathfrak{H}_{0} is finite dimensional space and consequently 𝔐P+0=P+0\mathfrak{M}\cap P_{+}\mathfrak{H}_{0}=P_{+}\mathfrak{H}_{0}. This is clear not correct if P+0P_{+}\mathfrak{H}_{0} is not trivial (take any eigenvector of A0A_{0} associated to λσ+(A0)\lambda\in\sigma_{+}(A_{0}), it obviously does not belong to 𝔐\mathfrak{M}). Then σ+(A0)\sigma_{+}(A_{0}) is an empty set, and finally Spr(A0)<1\textrm{Spr}(A_{0})<1, which achieves the proof.

Proof of Theorem 1.

Since ψ𝔐\psi\in\mathfrak{M} we deduce from Proposition 1 that the Neumann series (9) is convergent. On the other hand we deduce from (12) Ajψ=AjPψ=A0jψA^{j}\psi=A^{j}P_{-}\psi=A_{0}^{j}\psi. Therefore

(13) j=0AjΨ=j=0A0jΨ=(IA0)1Ψ(IA0)1Ψ.\|\sum_{j=0}^{\infty}A^{j}\Psi\|=\|\sum_{j=0}^{\infty}A_{0}^{j}\Psi\|=\|(I-A_{0})^{-1}\Psi\|\leq\|(I-A_{0})^{-1}\|\|\Psi\|.

Let n𝑵n\in\bm{N} be fixed. The fact that ψ𝔐\psi\in\mathfrak{M} implies that An+1ψ𝔐A^{n+1}\psi\in\mathfrak{M}. Applying inequality (13) to the vector An+1ψA^{n+1}\psi leads to

(14) j=n+1AjΨ=j=0AjAn+1Ψ(IA0)1An+1ΨCAnΨ,\|\sum_{j=n+1}^{\infty}A^{j}\Psi\|=\|\sum_{j=0}^{\infty}A^{j}A^{n+1}\Psi\|\leq\|(I-A_{0})^{-1}\|\|A^{n+1}\Psi\|\leq C\|A^{n}\Psi\|,

with C=(IA0)1A0C=\|(I-A_{0})^{-1}\|\|A_{0}\|, which finishes the proof of the Theorem.

Remark 2.

The proofs stay valid for any general compact operator AA and even if \mathfrak{H} is a Banach space. In the particular case where AA is in addition normal, that is AA=AAAA^{*}=A^{*}A, the obtained results are straightforward. Indeed if σ(A)={λk;k𝐍0}\sigma(A)=\{\lambda_{k};k\in\bm{N}_{0}\} the eigenvalues of AA, and PkP_{k} is the orthogonal spectral projection associated to λk\lambda_{k}, we have

A=k=1λkPk,A=\sum_{k=1}^{\infty}\lambda_{k}P_{k},

and it is clear that the condition ψ𝔐\psi\in\mathfrak{M} is equivalent to

ψ=|λk|<1Pkψ.\psi=\sum_{|\lambda_{k}|<1}P_{k}\psi.

Therefore

j=n+1AjΨ2=|λk|<1λkn+11λkPkΨ2=|λk|<1λk2(n+1)(1λk)2PkΨ2\displaystyle\left\|\sum_{j=n+1}^{\infty}A^{j}\Psi\right\|^{2}=\left\|\sum_{|\lambda_{k}|<1}\frac{\lambda_{k}^{n+1}}{1-\lambda_{k}}P_{k}\Psi\right\|^{2}=\sum_{|\lambda_{k}|<1}\frac{\lambda_{k}^{2(n+1)}}{(1-\lambda_{k})^{2}}\|P_{k}\Psi\|^{2}
r02(1r0)2AnΨ2,\displaystyle\leq\frac{r_{0}^{2}}{(1-r_{0})^{2}}\|A^{n}\Psi\|^{2},

where r0=max|λk|<1|λk|=A0=Spr(A0).r_{0}=\max_{|\lambda_{k}|<1}|\lambda_{k}|=\|A_{0}\|=\textrm{Spr}(A_{0}). One can verify that C=(IA0)1A0=r01r0.C=\|(I-A_{0})^{-1}\|\|A_{0}\|=\frac{r_{0}}{1-r_{0}}. Finally it is easy to find examples of AA such that the inequalities

Spr(A0)1Spr(A)=A,\textrm{Spr}(A_{0})\ll 1\ll\textrm{Spr}(A)=\|A\|,

hold, and where the benefit of considering the reduced space 0\mathfrak{H}_{0} is indeed remarkable.

3. Preconditioning

By ’preconditioning’ we here mean the transformation of the original Lippmann-Schwinger equation (IVq(k0))u=ψ(I-V_{q}(k_{0}))u=\psi to an integral equation solvable by a convergent Neumann series regardless of the value of Vq(k0)L2(B)L2(B)\|V_{q}(k_{0})\|_{L^{2}(B)\rightarrow L^{2}(B)} and of whether or not the sequence (Vq(k0)jψL2)j𝑵(\|V_{q}(k_{0})^{j}\psi\|_{L^{2}})_{j\in\bm{N}} converges to zero. See [5, 28, 29, 30, 21] for related approaches. Throughout this section we assume the problem dimension d{1,2,3}d\in\{1,2,3\}.

Lemma 1.

If q(x)0q(x)\geq 0, q0q\not\equiv 0, then there is a complex constant γ\gamma such that the solution of the equation (IVq(k0))u=ψ(I-V_{q}(k_{0}))u=\psi in L2(B)L^{2}(B) is expressible in terms of the convergent Neumann series

u=j=0Mjγψ,u=\sum_{j=0}^{\infty}M^{j}\gamma\psi,

where M=(1γ)I+γVq(k0)M=(1-\gamma)I+\gamma V_{q}(k_{0}).

Proof.

Let Vq(k0)φ=λφV_{q}(k_{0})\varphi=\lambda\varphi in 𝑹d\bm{R}^{d} with nonzero λ\lambda and with φ\varphi not identically zero outside any bounded ball in 𝑹d\bm{R}^{d}. Then

(15) {φ′′+k02(1+q(x)/λ)φ=0,x]R,R[,φ(R)=ik0φ(R),φ(R)=ik0φ(R)\left\{\begin{array}[]{rcl}\varphi^{\prime\prime}+k_{0}^{2}(1+q(x)/\lambda)\varphi&=&0,\qquad\quad x\in]-R,R[,\\ -\varphi^{\prime}(-R)&=&ik_{0}\varphi(-R),\\ \varphi^{\prime}(R)&=&ik_{0}\varphi(R)\end{array}\right.

for d=1d=1, and

{(Δ+k02(1+q(x)/λ))φ=0in𝑹d,lim|x||x|(d1)/2(|x|ik)φ=0uniformly inx/|x|Sd1,\left\{\begin{array}[]{rcl}(\Delta+k_{0}^{2}(1+q(x)/\lambda))\varphi&=&0\quad\text{in}\,\,\,\bm{R}^{d},\\ \lim_{|x|\rightarrow\infty}|x|^{(d-1)/2}\left(\partial_{|x|}-ik\right)\varphi&=&0\quad\text{uniformly in}\,\,\,x/|x|\in S^{d-1},\end{array}\right.

for d{2,3}d\in\{2,3\}. Thus, for sufficiently large R>0R>0 we have

0\displaystyle 0 =|x|<R(φ¯Δφ+k02|φ|2+k02λ1q(x)|φ|2)\displaystyle=\int_{|x|<R}(\overline{\varphi}\Delta\varphi+k_{0}^{2}|\varphi|^{2}+k_{0}^{2}\lambda^{-1}q(x)|\varphi|^{2})
(16) =|x|<R(k02|φ|2|φ|2)+|x|=Rφ¯rφ+k02λ1xsuppqq(x)|φ|2,\displaystyle=\int_{|x|<R}(k_{0}^{2}|\varphi|^{2}-|\nabla\varphi|^{2})+\int_{|x|=R}\overline{\varphi}\partial_{r}\varphi+k_{0}^{2}\lambda^{-1}\int_{x\in{\rm supp}\,q}q(x)|\varphi|^{2},

as well as (Δ+k02)φ=0(\Delta+k_{0}^{2})\varphi=0 in {|x|>R}\{|x|>R\}. In the case d=1d=1 we readily find that

|x|=Rφ¯rφ=k0(|φ(R)|2+|φ(R)|2)>0,\Im\int_{|x|=R}\overline{\varphi}\partial_{r}\varphi=k_{0}(|\varphi(-R)|^{2}+|\varphi(R)|^{2})>0,

while in the case d{2,3}d\in\{2,3\} we can follow the argument in, e.g., [32, Theorem 2.13, p. 38] to find

|x|=Rφ¯rφ>0.\Im\int_{|x|=R}\overline{\varphi}\partial_{r}\varphi>0.

Hence

xsuppqq(x)|φ|2𝑑x>0,\int_{x\in{\rm supp}\,q}q(x)|\varphi|^{2}dx>0,

and this in conjunction with (3) gives

(λ1)=|x|=Rφ¯rφk02xsuppqq(x)|φ|2<0,\Im(\lambda^{-1})=-\frac{\Im\int_{|x|=R}\overline{\varphi}\partial_{r}\varphi}{k_{0}^{2}\int_{x\in{\rm supp}\,q}q(x)|\varphi|^{2}}<0,

so λ>0\Im\lambda>0 and finally (eiα(1λ))>0\Re(e^{i\alpha}(1-\lambda))>0 if α]π/2,π/2[\alpha\in]-\pi/2,\pi/2[ satisfies

(17) tanα>maxλσ(Vq(k0))λ1λ,\tan\alpha>\max_{\lambda^{\prime}\in\sigma(V_{q}(k_{0}))}\frac{\Re\lambda^{\prime}-1}{\Im\lambda^{\prime}},

where σ(Vq(k0))\sigma(V_{q}(k_{0})) is the spectrum of Vq(k0)V_{q}(k_{0}). The existence of the maximum in (17) follows from the fact that the eigenvalues of the compact operator Vq(k0):L2(B)L2(B)V_{q}(k_{0}):L^{2}(B)\rightarrow L^{2}(B) reside in the closed ball {λ𝑪,|λ|Vq(k0)L2(B)L2(B)}\{\lambda^{\prime}\in\bm{C},\,\,\,|\lambda^{\prime}|\leq\|V_{q}(k_{0})\|_{L^{2}(B)\rightarrow L^{2}(B)}\} and can accumulate only at zero. These facts also imply that there exists ε>0\varepsilon>0 such that |γ(1λ)1|<1|\gamma(1-\lambda^{\prime})-1|<1 for all λσ(Vq(k0))\lambda^{\prime}\in\sigma(V_{q}(k_{0})), where γ=εeiα\gamma=\varepsilon e^{i\alpha}. It remains to notice that each eigenvalue μ\mu of MM is of the form μ=1γ+γλ\mu=1-\gamma+\gamma\lambda^{\prime}, where λ\lambda^{\prime} is some eigenvalue of Vq(k0)V_{q}(k_{0}), and finally that the equation (IM)u=γψ(I-M)u=\gamma\psi is equivalent with the equation (IVq(k0))u=ψ(I-V_{q}(k_{0}))u=\psi. ∎

We can in fact be more specific in a special case in dimension one. Let LL be a positive constant, set B=]0,L[B=]0,L[, and let q(x)q0=const.>0q(x)\equiv q_{0}={\rm const.}>0 for xB¯x\in\overline{B}.

Lemma 2.

If ε>k0Lq0/2\varepsilon^{\prime}>k_{0}Lq_{0}/2,

α=arctan1+k0Lq02εεk0Lq02,\alpha=\arctan\frac{1+\frac{k_{0}Lq_{0}}{2}\varepsilon^{\prime}}{\varepsilon^{\prime}-\frac{k_{0}Lq_{0}}{2}},
0<ε<11+k0Lq0/2|tan(2max{α,arctanε})|tan(max{α,arctanε}),0<\varepsilon<\frac{1}{{1+k_{0}Lq_{0}/2}}\frac{|\tan(2\max\{\alpha,\arctan\varepsilon^{\prime}\})|}{\tan(\max\{\alpha,\arctan\varepsilon^{\prime}\})},

and γ=εeiα\gamma=\varepsilon e^{i\alpha}, then each eigenvalue μ\mu of M=(1γ)I+γVq(k0)M=(1-\gamma)I+\gamma V_{q}(k_{0}) satisfies |μ|<1|\mu|<1.

Proof.

Assume λ𝑪{0}\lambda\in\bm{C}\setminus\{0\} and φL2(]0,L[)\varphi\in L^{2}(]0,L[), φ0\varphi\not\equiv 0, satisfy

(18) Vq(k0)φ(x)=λφ(x),x(0,L).V_{q}(k_{0})\varphi(x)=\lambda\varphi(x),\quad x\in(0,L).

Integration by parts readily shows the equivalence of the Lippmann-Schwinger equation (18) with the Helmholtz system

(19) {φ′′(x)+k02s2φ(x)=0,x]0,L[,φ(0)=ik0φ(0),φ(L)=ik0φ(L),\left\{\begin{array}[]{rcl}\varphi^{\prime\prime}(x)+k_{0}^{2}s^{2}\varphi(x)&=&0,\qquad\quad x\in]0,L[,\\ -\varphi^{\prime}(0)&=&ik_{0}\varphi(0),\\ \varphi^{\prime}(L)&=&ik_{0}\varphi(L),\end{array}\right.

where s=(1+q0/λ)1/2s=\left(1+q_{0}/\lambda\right)^{1/2}. The eigenvectors of the Laplacian on ]0,L[]0,L[ are generally of the form

(20) φ(x)=Aexp(ik0sx)+Bexp(ik0sx),\varphi(x)=A\exp(ik_{0}sx)+B\exp(-ik_{0}sx),

and we then readily find that (19) is equivalent with (20) together with

(21) s1,B=A(s+1)/(s1),e2ik0Ls=(s+1)2/(s1)2.s\neq 1,\quad B=A(s+1)/(s-1),\quad e^{2ik_{0}Ls}=(s+1)^{2}/(s-1)^{2}.

Next define the constant T(k0L)>0T(k_{0}L)>0 by T(k0L)sinh(k0LT(k0L))=1T(k_{0}L)\sinh(k_{0}LT(k_{0}L))=1. Using the last condition in (21), we find that, necessarily,

sS±(k0L)={coshk0Lt±1t2sinh2k0Ltsinhk0Lt+it,t(0,T(k0L)]},s\in S_{\pm}(k_{0}L)=\left\{\frac{-\cosh k_{0}Lt\pm\sqrt{1-t^{2}\sinh^{2}k_{0}Lt}}{\sinh k_{0}Lt}+it,\,\,t\in(0,T(k_{0}L)]\right\},

which in turn implies

λΛ(q0,k0L)={q0s21,sS(k0L)S+(k0L)}.\lambda\in\Lambda(q_{0},k_{0}L)=\left\{\frac{q_{0}}{s^{2}-1},\,\,s\in S_{-}(k_{0}L)\cup S_{+}(k_{0}L)\right\}.

As an example, Figure 1 shows the set Λ(q0,k0L)\Lambda(q_{0},k_{0}L), as well as the numerically computed spectrum, for two sets of parameter values for k0k_{0}, LL, and q0q_{0}.

Refer to caption
Refer to caption
Figure 1. The theoretically predicted set Λ(q0,k0L)\Lambda(q_{0},k_{0}L) that includes the eigenvalues of Vq(k0)V_{q}(k_{0}), plotted against the numerically computed eigenvalues. The parameter values are: top, k0=1k_{0}=1, L=1L=1, and q0=5q_{0}=5; bottom, k0=50k_{0}=50, L=1L=1, and q0=1q_{0}=1. The reference circle is centered at the origin and has radius one.

Since λ>0\Im\lambda>0 and since the eigenvalues of Vq(k0)V_{q}(k_{0}) accumulate precisely at zero, we have

limt0ω(t)1ω(t)=\lim_{t\rightarrow 0}\frac{\Re\omega(t)-1}{\Im\omega(t)}=-\infty

for

ω(t)=q0s(t)21,t(0,T(k0L)];\omega(t)=\frac{q_{0}}{s(t)^{2}-1},\quad t\in(0,T(k_{0}L)];

here s(t)s(t) is given by the above definition of S±(k0L)S_{\pm}(k_{0}L). Furthermore, if ω(t)<1\Re\omega(t)<1 then (ω(t)1)/ω(t)<0(\Re\omega(t)-1)/\Im\omega(t)<0, while ω(t)1\Re\omega(t)\geq 1 implies

1t2sinh2k0Lt+cosh(k0Lt)1t2sinh2k0Ltsinh2k0Ltq02;\frac{1-t^{2}\sinh^{2}k_{0}Lt+\cosh(k_{0}Lt)\sqrt{1-t^{2}\sinh^{2}k_{0}Lt}}{\sinh^{2}k_{0}Lt}\leq\frac{q_{0}}{2};

the latter can be seen by rewriting ω10\Re\omega-1\geq 0 as

((s)2(s)21)((s)2(s)21q0)+4(s)2(s)20,((\Re s)^{2}-(\Im s)^{2}-1)((\Re s)^{2}-(\Im s)^{2}-1-q_{0})+4(\Re s)^{2}(\Im s)^{2}\leq 0,

and noting that (s)2(s)21>0(\Re s)^{2}-(\Im s)^{2}-1>0 and (s)2(s)2>0(\Re s)^{2}(\Im s)^{2}>0. Now since also

1t2sinh(k0Lt)2cosh(k0Lt)1t2sinh(k0Lt)2tsinh(k0Lt)(cosh(k0Lt)1t2sinh(k0Lt)2)\displaystyle\frac{1-t^{2}\sinh(k_{0}Lt)^{2}-\cosh(k_{0}Lt)\sqrt{1-t^{2}\sinh(k_{0}Lt)^{2}}}{t\sinh(k_{0}Lt)(\cosh(k_{0}Lt)-\sqrt{1-t^{2}\sinh(k_{0}Lt)^{2}})}
1t2sinh(k0Lt)2+cosh(k0Lt)1t2sinh(k0Lt)2tsinh(k0Lt)(cosh(k0Lt)+1t2sinh(k0Lt)2),\displaystyle\leq\frac{1-t^{2}\sinh(k_{0}Lt)^{2}+\cosh(k_{0}Lt)\sqrt{1-t^{2}\sinh(k_{0}Lt)^{2}}}{t\sinh(k_{0}Lt)(\cosh(k_{0}Lt)+\sqrt{1-t^{2}\sinh(k_{0}Lt)^{2}})},

we have

ω(t)1ω(t)<ω(t)ω(t)<1t2sinh2k0Lt+cosh(k0Lt)1t2sinh2k0Lttsinh(k0Lt)(coshk0Lt+1t2sinh2k0Lt),\frac{\Re\omega(t)-1}{\Im\omega(t)}<\frac{\Re\omega(t)}{\Im\omega(t)}<\frac{1-t^{2}\sinh^{2}k_{0}Lt+\cosh(k_{0}Lt)\sqrt{1-t^{2}\sinh^{2}k_{0}Lt}}{t\sinh(k_{0}Lt)(\cosh k_{0}Lt+\sqrt{1-t^{2}\sinh^{2}k_{0}Lt})},

so ω(t)1\Re\omega(t)\geq 1 implies

ω(t)1ω(t)\displaystyle\frac{\Re\omega(t)-1}{\Im\omega(t)} <q02sinhk0Ltt(coshk0Lt+1t2sinh2k0Lt)q02tanhk0Ltt\displaystyle<\frac{q_{0}}{2}\frac{\sinh k_{0}Lt}{t(\cosh k_{0}Lt+\sqrt{1-t^{2}\sinh^{2}k_{0}Lt})}\leq\frac{q_{0}}{2}\frac{\tanh k_{0}Lt}{t}
q02limτ0tanhk0Lττ=q0k0L2,\displaystyle\leq\frac{q_{0}}{2}\lim_{\tau\searrow 0}\frac{\tanh k_{0}L\tau}{\tau}=\frac{q_{0}k_{0}L}{2},

that is, we get a similar estimate on (ω(t)1)/ω(t)(\Re\omega(t)-1)/\Im\omega(t) as we do on Vq(k0)\|V_{q}(k_{0})\|. Next, define

(22) ξ+=arctansupωΛ(q0,k0L)(eiα(1ω))(eiα(1ω))\xi_{+}=\arctan\sup_{\omega\in\Lambda(q_{0},k_{0}L)}\frac{\Im\left(e^{i\alpha}(1-\omega)\right)}{\Re\left(e^{i\alpha}(1-\omega)\right)}

and

(23) ξ=arctaninfωΛ(q0,k0L)(eiα(1ω))(eiα(1ω)).\xi_{-}=\arctan\inf_{\omega\in\Lambda(q_{0},k_{0}L)}\frac{\Im\left(e^{i\alpha}(1-\omega)\right)}{\Re\left(e^{i\alpha}(1-\omega)\right)}.

With α(0,π/2)\alpha\in(0,\pi/2), ε>0\varepsilon>0, and γ=εeiα\gamma=\varepsilon e^{i{\alpha}}, we have arg(γ(1ω))[ξ,ξ+]\arg(\gamma(1-\omega))\in[\xi_{-},\xi_{+}] for all ωΛ(q0,k0L)\omega\in\Lambda(q_{0},k_{0}L), and if

(24) ε<11+k0Lq0/2|tan(2max{|ξ+|,|ξ|})|tan(max{|ξ+|,|ξ|})\varepsilon<\frac{1}{{1+k_{0}Lq_{0}/2}}\frac{|\tan(2\max\{|\xi_{+}|,|\xi_{-}|\})|}{\tan(\max\{|\xi_{+}|,|\xi_{-}|\})}

then |γ(1ω)1|<1|\gamma(1-\omega)-1|<1 for all ωΛ(q0,k0L)\omega\in\Lambda(q_{0},k_{0}L), and specifically |γ(1λ)1|<1|\gamma(1-\lambda^{\prime})-1|<1 for all eigenvalues λ\lambda^{\prime} of Vq(k0)V_{q}(k_{0}); the condition (24) can be deduced by requiring (γ(1ω))<|tan(π2arg(γ(1ω)))|\Im(\gamma(1-\omega))<|\tan(\pi-2\arg(\gamma(1-\omega)))| and using |1ω|1+Vq(k0)|1-\omega|\leq 1+\|V_{q}(k_{0})\|. Note that we must choose α<π/2\alpha<\pi/2 rather than α=π/2\alpha=\pi/2 since

(eiα(1ω(t)))(eiα(1ω(t)))t0tanα,\frac{\Im(e^{i\alpha}(1-\omega(t)))}{\Re(e^{i\alpha}(1-\omega(t)))}\underset{t\rightarrow 0}{\longrightarrow}\tan\alpha,

which for α=π/2\alpha=\pi/2 forces ξ+=π/2\xi_{+}=\pi/2 and thus ε<0\varepsilon<0. Note furthermore that we can bound (22)–(23) analytically for

α=arctan1+k0Lq02εεk0Lq02,\alpha=\arctan\frac{1+\frac{k_{0}Lq_{0}}{2}\varepsilon^{\prime}}{\varepsilon^{\prime}-\frac{k_{0}Lq_{0}}{2}},

with ε>k0Lq0/2\varepsilon^{\prime}>k_{0}Lq_{0}/2, since

cotα(eiα(1ω))(eiα(1ω))tanα,ωΛ(q0,k0L),ω1,-\cot\alpha\leq\frac{\Im\left(e^{i\alpha}(1-\omega)\right)}{\Re\left(e^{i\alpha}(1-\omega)\right)}\leq\tan\alpha,\quad\omega\in\Lambda(q_{0},k_{0}L),\quad\Re\omega\leq 1,

since furthermore (recall that (ω1)/ω<k0Lq0/2<tanα(\Re\omega-1)/\Im\omega<k_{0}Lq_{0}/2<\tan\alpha)

k0Lq02tanα+1tanαk0Lq02(eiα(1ω))(eiα(1ω))<0,ωΛ(q0,k0L),ω>1,-\frac{\frac{k_{0}Lq_{0}}{2}\tan\alpha+1}{\tan\alpha-\frac{k_{0}Lq_{0}}{2}}\leq\frac{\Im\left(e^{i\alpha}(1-\omega)\right)}{\Re\left(e^{i\alpha}(1-\omega)\right)}<0,\quad\omega\in\Lambda(q_{0},k_{0}L),\quad\Re\omega>1,

and finally since

cotα>k0Lq02tanα+1tanαk0Lq02=ε.-\cot\alpha>-\frac{\frac{k_{0}Lq_{0}}{2}\tan\alpha+1}{\tan\alpha-\frac{k_{0}Lq_{0}}{2}}=-\varepsilon^{\prime}.

Remark 3.

A drawback of preconditioning is that, with increasing k0Lq0k_{0}Lq_{0}, the acceptable values of α\alpha and of arctanε\arctan\varepsilon^{\prime} tend to π/2\pi/2, and ε\varepsilon therefore tends to zero. Thus, while the Neumann series remains convergent, the equation

(IM)u=γψ(I-M)u=\gamma\psi

may be said, especially in a numerical context, to lose information about the operator Vq(k0)V_{q}(k_{0}) and about the original inhomogeneity ψ\psi, as both are multiplied with γ=εeıα\gamma=\varepsilon e^{\i\alpha} there.

Remark 4.

Instead of using the bound on ε\varepsilon stated in Lemma 2, we can estimate ξ+\xi_{+} and ξ\xi_{-} from (22)–(23) numerically and arrive at a larger sufficiently small ε\varepsilon using (24).

We show numerical examples of the use of preconditioning in Section 4.

4. Numerical examples

We here present several numerical examples in dimension one. Fix a positive wavenumber k0k_{0}, obstacle size L>0L>0, medium function qL2(]0,L[)q\in L^{2}(]0,L[), q(x)>1q(x)>-1, and consider the following system for the scattered wave u(x)u(x) corresponding to the left excitation exp(ik0x)\exp(ik_{0}x) in dimension one:

(25) {ψ′′(x)+k02(1+q(x))ψ(x)=k02q(x)exp(ik0x),x]0,L[,ψ(0)=ik0ψ(0),ψ(L)=ik0ψ(L).\left\{\begin{array}[]{rcl}\psi^{\prime\prime}(x)+k_{0}^{2}(1+q(x))\psi(x)&=&-k_{0}^{2}q(x)\exp(ik_{0}x),\quad x\in]0,L[,\\ -\psi^{\prime}(0)&=&ik_{0}\psi(0),\\ \psi^{\prime}(L)&=&ik_{0}\psi(L).\end{array}\right.

The function G(x,y)=(i/2k0)exp(ik0|xy|)G(x,y)=(i/2k_{0})\exp(ik_{0}|x-y|), x,y[0,L]x,y\in[0,L], is the free-space Green’s function associated with the boundary problem (25), since

{(y2+k02)G(x,y)=δ(xy),x,y]0,L[,yG(x,0)=ik0G(x,0),x]0,L[,yG(x,L)=ik0G(x,L),x[0,L].\left\{\begin{array}[]{rcl}(\partial_{y}^{2}+k_{0}^{2})G(x,y)&=&-\delta(x-y),\quad x,y\in]0,L[,\\ -\partial_{y}G(x,0)&=&ik_{0}G(x,0),\quad x\in]0,L[,\\ \partial_{y}G(x,L)&=&ik_{0}G(x,L),\quad x\in[0,L].\end{array}\right.

Multiplying the differential equation in (25) with G(x,y)G(x,y) and integrating by parts, we get the Lippmann-Schwinger equation

(26) (IVq(k0))ψ(x)=Vq(k0)exp(ik0)(x),x]0,L[,(I-V_{q}(k_{0}))\psi(x)=V_{q}(k_{0})\exp(ik_{0}\cdot)(x),\quad x\in]0,L[,

where

Vq(k0)u(x)=ik02y=0Leik0|xy|q(y)u(y)dy,x]0,L[.V_{q}(k_{0})u(x)=\frac{ik_{0}}{2}\int_{y=0}^{L}e^{ik_{0}|x-y|}q(y)u(y)dy,\quad x\in]0,L[.

The operator Vq(k0):L2(]0,L[)L2(]0,L[)V_{q}(k_{0}):L^{2}(]0,L[)\rightarrow L^{2}(]0,L[) is compact, with norm satisfying

Vq(k0)2x=0Ly=0L|ik02eik0|xy|q(y)|2𝑑y𝑑x=k02Lq224.\|V_{q}(k_{0})\|^{2}\leq\int_{x=0}^{L}\int_{y=0}^{L}\left|\frac{ik_{0}}{2}e^{ik_{0}|x-y|}q(y)\right|^{2}dydx=\frac{k_{0}^{2}L\|q\|^{2}_{2}}{4}.

Now let q0]1,+[q_{0}\in]-1,+\infty[, and assume that q|]0,L[q0q|_{]0,L[}\equiv q_{0}. One can easily prove

0=Span(Vq(k0)jeik0;j)=Span(V1(k0)jeik0;j).\mathfrak{H}_{0}=\textrm{Span}\left(V_{q}(k_{0})^{j}e^{ik_{0}\cdot};\;j\in\mathbb{N}\right)=\textrm{Span}\left(V_{1}(k_{0})^{j}e^{ik_{0}\cdot};\;j\in\mathbb{N}\right).

On the other hand, since Vq(k0)=q0V1(k0)V_{q}(k_{0})=q_{0}V_{1}(k_{0}), we have

Spr((Vq)0(k0))=|q0|Spr((V1)0(k0)),\textrm{Spr}\left((V_{q})_{0}(k_{0})\right)=|q_{0}|\textrm{Spr}\left((V_{1})_{0}(k_{0})\right),

where (Vq)0(k0)(V_{q})_{0}(k_{0}) and (V1)0(k0)(V_{1})_{0}(k_{0}) are the restrictions of Vq(k0)V_{q}(k_{0}) and V1(k0)V_{1}(k_{0}), respectively, to the space 0\mathfrak{H}_{0}. Following the proof of Proposition 1, the Born series

j=1(Vq(k0))jeik0\sum_{j=1}^{\infty}\left(V_{q}(k_{0})\right)^{j}e^{ik_{0}\cdot}

is convergent if and only if Spr((Vq)0(k0))<1\textrm{Spr}\left((V_{q})_{0}(k_{0})\right)<1, that is, if and only if

|q0|<1Spr((V1)0(k0)).|q_{0}|<\frac{1}{\textrm{Spr}\left((V_{1})_{0}(k_{0})\right)}.

To illustrate this, we show in Figure 2 two cases of repeated application of Vq(k0)V_{q}(k_{0}) on the original right-hand side Vq(k0)eik0V_{q}(k_{0})e^{ik_{0}\cdot}.

Refer to caption
Refer to caption
Figure 2. Top: k0=1k_{0}=1, L=1L=1, q0=1q_{0}=1. Bottom: k0=50k_{0}=50, L=1L=1, q0=1q_{0}=1.

In the case where Spr((Vq)0(k0))<1((V_{q})_{0}(k_{0}))<1, the series (Vq(k0)Jeik0)(V_{q}(k_{0})^{J}e^{ik_{0}\cdot}) converges strongly to zero, while (IVq(k0))j=1JVq(k0)jeik0(I-V_{q}(k_{0}))\sum_{j=1}^{J}V_{q}(k_{0})^{j}e^{ik_{0}\cdot} converges strongly to Vq(k0)eik0V_{q}(k_{0})e^{ik_{0}\cdot}, The second of the two convergence processes plateaus for large values of JJ due to the effect of numerical errors. In contrast, neither sequence converges in the case where Spr((Vq)0(k0))1((V_{q})_{0}(k_{0}))\geq 1.

Finally, we illustrate Lemma 2 numerically in Figures 3 and 4. We indeed get a convergent Neumann series solution for the case k0=50k_{0}=50, L=1L=1, q0=1q_{0}=1, albeit the convergence is rather slow.

Refer to caption
Refer to caption
Refer to caption

a) k0k_{0}=1, L=1L=1, q0=5q_{0}=5, Vq(k0)=2.4183\|V_{q}(k_{0})\|=2.4183, ε=5.2\varepsilon^{\prime}=5.2, α=1.3803\alpha=1.3803, ε=0.02\varepsilon=0.02, T=0.9320T=0.9320.

Refer to caption
Refer to caption
Refer to caption

b) k0=50k_{0}=50, L=1L=1, q0=1q_{0}=1, Vq(k0)=7.1791\|V_{q}(k_{0})\|=7.1791, ε=50\varepsilon^{\prime}=50, α=1.5508\alpha=1.5508, ε=3105\varepsilon=3\cdot 10^{-5}, T=0.0677T=0.0677.

Figure 3. The original theoretically predicted and numerically computed spectra of Vq(k0)V_{q}(k_{0}); the transformed curve {z=1γ(1ω),ωΛ(q0,k0L)}𝑪\{z=1-\gamma(1-\omega),\,\,\,\omega\in\Lambda(q_{0},k_{0}L)\}\subset\bm{C} is included in the unit open disk centered at 1.
Refer to caption
Figure 4. Preconditioning of the equation (IVq(k0))u=Vq(k0)eik0(I-V_{q}(k_{0}))u=V_{q}(k_{0})e^{ik_{0}\cdot} results in a convergent Neumann series solution. The parameters here are as in Figure 2 (bottom) and Figure 3 b).

References

  • [1] H. Ammari, J. Garnier, H. Kang, M. Lim, and K. Solna, Multistatic imaging of extended targets, SIAM J. Imaging Science, 5 (2012), 564–600.
  • [2] G. Bao, Y. Chen, and F. Ma, Regularity and stability for the scattering map of a linearized inverse medium problem, J. Math. Anal. Appl., 247 (2000), 255–271.
  • [3] S. Arridge, S. Moskow and J. C. Schotland, Inverse Born series for the Calderon problem, Inverse Probl. 28, 035003 (2012).
  • [4] P. Bardsley and F. Guevara Vasquez, Restarted inverse Born series for the Schrödinger problem with discrete internal measurements, Inverse Probl. 30, 045014 (2014).
  • [5] H. W. Engl and M. Z. Nashed. Generalized Inverses of Random Linear Operators in Banach Spaces. Journal of Mathematical Analysis and Applications, 83, 582–610 (1981)
  • [6] B. Osting and M. I. Weinstein. Long-lived Scattering Resonances and Bragg Structures. SIAM Journal on Applied Mathematics, 73(2) (2013)
  • [7] A. McIntosh and M. Mitrea. Clifford Algebras and Maxwell’s Equations in Lipschitz Domains. Mathematical Methods in the Applied Sciences, 22, 1599–1620 (1999)
  • [8] C.-C. Chou, J. Yao, and D. J. Kouri. Volterra inverse scattering series method for one-dimensional quantum barrier scattering. International Journal of Quantum Chemistry, 117:e25403 (2017)
  • [9] R. G. Newton. Scattering Theory of Waves and Particles, Second Edition, Springer, 1982.
  • [10] G. Bao and F. Triki. Stability for the multifrequency inverse medium problem. Journal of Differential Equations, 269(9), 7106-7128 (2020)
  • [11] G. Bao and F. Triki. Error estimates for the recursive linearization of inverse medium problems. J. Comput. Math. 28, No. 6, 725–744
  • [12] M. Karamehmedović and K. Linder-Steinlein. Spectral properties of radiation for the Helmholtz equation with a random coefficient. In review.
  • [13] K. Kilgore, S. Moskow and J. C. Schotland, Convergence of the Born and inverse Born series for electromagnetic scattering, Applicable Analysis 96, 1737 (2017).
  • [14] K. Kilgore, S. Moskow and J. C. Schotland, Inverse Born series for scalar waves, J. Computational Math. 30, 601 (2012).
  • [15] A. Louis, Approximate inverse for linear and some nonlinear problems, Inverse Probl. 12, 175 (1996).
  • [16] M. Machida and J. C. Schotland, Inverse Born series for the radiative transport equation, Inverse Probl. 31, 095009 (2015).
  • [17] V. Markel, J. A. O’Sullivan and J. C. Schotland, Inverse problem in optical diffusion tomography.IV nonlinear inversion formulas, J. Opt. Soc. Am A, 20, 903 (2003).
  • [18] T. Kato. Perturbation theory for linear operators. Vol. 132. Springer Science & Business Media, 2013.
  • [19] A. Lechleiter, K. S. Kazimierski and M. Karamehmedović. Tikhonov regularization in LpL^{p} applied to inverse medium scattering. Inverse Problems 29, 075003 (2013)
  • [20] N. Suzuki. On the Convergence of Neumann Series in Banach Space. Mathematische Annalen 220, 143–146 (1976)
  • [21] G. Osnabrugge, S. Leedumrongwatthanakun and I. M. Vellekoop. A convergent Born series for solving the inhomogeneous Helmholtz equation in arbitrarily large media. Journal of Computational Physics 322, 113–124 (2016)
  • [22] H. Lopez-Menchon, J. M. Rius, A. Heldring and E. Ubeda. Acceleration of Born Series by Change of Variables. IEEE Transactions on Antennas and Propagation 69(9), 5750–5760 (2021)
  • [23] S. Moskow, and J. C. Schotland, Numerical studies of the inverse born series for diffuse waves, Inverse Problems, 25 (2009) 095007-25.
  • [24] S. Moskow, and J. C. Schotland, Convergence and stability of the inverse scattering series for diffuse waves, Inverse Problems, 24 (2008) 065005-2
  • [25] S. Moskow, and J. C. Schotland, Inverse Born Series, in The Radon Transform: The First 100 Years and Beyond, edited by R. Ramlau and O. Scherzer (De Gruyter, 2019)
  • [26] G. Panasyuk, V. A. Markel, P. S. Carney and J. C. Schotland, Nonlinear inverse scattering and three dimensional near-field optical imaging, App. Phys. Lett. 89, 221116 (2006)
  • [27] C. Brezinski. Convergence acceleration during the 20th century. Journal of Computational and Applied Mathematics 122, 1–21 (2000)
  • [28] R. E. Kleinman and G. F. Roach. Iterative solutions of boundary integral equations in acoustics. Proc. R. Soc. Lond. A 417, 45–57 (1988)
  • [29] R. E. Kleinman, G. F. Roach, L. S. Schuetz, and J. Shirron. An iterative solution to acoustic scattering by rigid objects. J. Acoust. Soc. Am. 84(1), 385 (1988)
  • [30] R. E. Kleinman, G. F. Roach and P. M. van den Berg. Convergent Born series for large refractive indices. Journal of The Optical Society of America A 7(5), 890–897 (1990)
  • [31] M. E. Taylor. Partial Differential Equations I: Basic Theory, Second Edition, Springer, 2011
  • [32] D. Colton and R. Kress. Inverse Acoustic and Electromagnetic Scattering Theory, Fourth Edition, Springer, 2019