401
Views
2
CrossRef citations to date
0
Altmetric
Original Articles

Born approximation for the magnetic Schrödinger operator

&
Pages 422-438 | Received 10 Apr 2017, Accepted 14 Apr 2018, Published online: 08 May 2018

ABSTRACT

We prove the existence of scattering solutions for multidimensional magnetic Schrödinger equation such that the scattered field belongs to the weighted Lebesgue space L-δ2(Rn) (n2) with some δ>12. As a consequence of this we provide the mathematical foundation of the direct Born approximation for the magnetic Schrödinger operator. Connection to the inverse Born approximation is discussed with numerical examples illustrating the applicability of the method.

AMS SUBJECT CLASSIFICATIONS:

1. Introduction

We develop a time harmonic scattering theory for the magnetic Schrödinger operator, analogous to the well-known theory for the Schrödinger operator. For this purpose we consider a Lippmann–Schwinger equation and prove that it is a Fredholm equation on the weighted Lebesgue space Lδ2. We then establish the asymptotic behaviour of the scattering operator and its Born, or linear, approximation and an explicit formula for the first nonlinear term in the Born series.

The direct Born approximation is known as the most applicable approximate method in the numerous practical problems. It is also known that the inverse scattering Born approximation is well-defined and perfectly works (as a mathematical tool) in the case of linear and nonlinear Schrödinger operators and for all types of scattering data: full scattering, backscattering, fixed angle scattering and fixed energy scattering. For some scattering data, it is possible to get the uniqueness and reconstruction procedure while for some data we are able to reconstruct singularities and jumps of unknowns even when there is no uniqueness. We mention here the results of Päivärinta and Somersalo [Citation1], Nachman [Citation2,Citation3], Sun and Uhlmann [Citation4], Isakov and Sylvester [Citation5], Päivärinta, Serov and Somersalo [Citation6], Päivärinta and Serov [Citation7Citation9], Ola, Päivärinta and Serov [Citation10], Ruiz [Citation11], Ruiz and Vargas [Citation12], Päivärinta and Serov [Citation13], Reyes [Citation14], Serov [Citation15], Serov and Harju [Citation16,Citation17], Serov and Sandhu [Citation18], Lechleiter [Citation19], Reyes and Ruiz [Citation20]. The main point of all these results is the precise calculation of the first (quadratic) nonlinear term in the Born series.

For the magnetic Schrödinger operator, the direct scattering problem (i.e. existence of the scattering solutions) as well as the inverse scattering Born approximation are much less familiar. We are only aware of [Citation21Citation23]. Indeed, in [Citation21] the authors assumed that magnetic and electric potentials with derivatives up to orders 4 and 5, respectively, decay exponentially. They were able to prove that the scattering amplitude at fixed energy determine uniquely the electric potential. In [Citation22] the authors relaxed these assumptions to one derivative of magnetic potential and just electric potential decaying exponentially while proving the same main result. The inverse backscattering problem for magnetic Schrödinger was considered in [Citation23] in the two-dimensional case under the assumptions that the magnetic potential, its gradient and electric potential have some polynomial decay at infinity.

In this work we continue to study the inverse backscattering problem in the multidimensional case. The big interest to this problem is connected to the fact that the knowledge of the scattering amplitude with backscattering data allows us to obtain essential information about the unknowns. The main goal of the present work is to justify the direct and inverse Born approximation for the magnetic Schrödinger operator. What is more, we will give the first (to the best of our knowledge) numerical examples illustrating this in two dimensions using backscattering data.

We consider the magnetic Schrödinger operatorH=-(+iW(x))2+V(x),xRn,

in dimensions n2 where the coefficients W(x) and V(x) are assumed to be real-valued. We assume generally that(1) WWp,σ1(Rn),VLσp(Rn),(1)

where(2) σ>max1;n1-2p,p>n.(2)

Here Lσp(Rn) denotes usual weighted Lebesgue space defined by finiteness of the normfLσp(Rn):=Rn(1+|x|)pσ|f(x)|pdx1/p.

The Sobolev space Wp,σ1(Rn) is understood so that f belongs to Wp,σ1(Rn) if and only if f and f belong to Lσp(Rn). It is the same (up to equivalent norms) that (1+|x|2)σ/2fWp1(Rn). For the case p=2 instead of the symbol W2,σ1(Rn) we use the symbol Hσ1(Rn).

It is known, see [Citation24], that under these conditions for the coefficients of the magnetic Schrödinger operator the following Gårding’s inequality holds:(Hu,u)L2(Rn)νuL2(Rn)2-CuL2(Rn)2,

where 0<ν<1 and C>0. This inequality allows us to define symmetric operator H by the method of quadratic forms. Then H has a self-adjoint Friedrichs extension with the domain (in general)D(H)={fW21(Rn):HfL2(Rn)}.

In our particular case it is possible to prove that actually(3) D(H)=W22(Rn),(3)

see Appendix 1.

In the scattering theory the main role are played by the special solutions (scattering solutions) of the equationHu(x)=k2u(x)

which are of the formu(x)=u0(x)+usc(x),

where u0(x)=eik(x,θ) is incident plane wave with direction θSn-1 and the scattered wave usc(x) satisfies the Sommerfeld radiation condition at the infinity, i.e.limr+rn-12usc(x)r-ikusc(x)=0,r=|x|.

In this case the total field u satisfies the so-called Lippmann–Schwinger equation(4) u(x)=u0(x)+RnGk+(|x-y|)2i(W(y)u(y))-q(y)u(y)dy,(4)

where q:=iW+|W|2+V and Gk+ is the kernel of integral operator (-Δ-k2-i0)-1. Since WWp1(Rn)L(Rn),p>n thenqLσp(Rn)

with the same p and σ as in (Equation2)

In Section 2, we show that, for every incident wave u0(x), there exists a unique scattering solution to the equation Hu=k2u such that the scattered field usc belongs to the weighted Lebesgue space L-δ2(Rn) with some δ>12. We do this by showing that the Lippmann–Schwinger integral Equation (Equation4) is Fredholm on this Hilbert space.

In Section 3, (in order to define the scattering amplitude) we prove the asymptoticuscceikrrn-12A(k,θ,θ)

as r=|x|, where θ=x|x|. The function A(k,θ,θ) here is the scattering amplitude, the kernel of the relative scattering operator (also called the far field operator), which summarizes all the data that can be obtained from scattering experiments.

The dependence of the scattering operator on the potentials V and W is not linear. The Born, or single scattering approximation, provides insight into the full non-linear problem, and is often a good enough approximation to provide meaningful results in many applied inverse problems. In Section 3, we compute the Born approximation of the scattering operator for the magnetic Schrödinger operator. Even more, we go one step further and compute the second term in the Born series. In Section 4, we discuss connection to inverse Born approximation and give numerical examples of it in Section 5.

2. Existence of the scattering solutions

Using the representation u=u0+usc we rewrite the integral Equation (Equation4) only for scattered field usc as(5) usc(x)=u~0(x)+RnGk+(|x-y|)2i(W(y)usc(y))-q(y)usc(y)dy,(5)

where u~0 is equal tou~0(x)=RnGk+(|x-y|)2i(W(y)u0(y))-q(y)u0(y)dy.

We use the following results of Agmon [Citation25, Remark 2, Appendix A]: for any gH-δ2(Rn) it is true that1|k|gH-δ2(Rn)+gH-δ1(Rn)+|k|gL-δ2(Rn)c(Δ+k2)gLδ2(Rn),|k|1,

where δ>12 and H-δs(Rn),s=0,1,2, denotes the weighted Sobolev spaces. As a consequence we have uniformly in |k|1 the crucial estimates(6) (-Δ-k2-i0)-1fL-δ2(Rn)β|k|fLδ2(Rn),(-Δ-k2-i0)-1fH-δ1(Rn)βfLδ2(Rn).(6)

The constant β will be used below precisely.

The operator (-Δ-k2-i0)-1 is the integral operator of convolution type. Then using duality we can conclude that uniformly in |k|1 it maps continuously Hδ-1(Rn) to L-δ2(Rn) with the same norm estimate β as above, i.e.(7) (-Δ-k2-i0)-1fL-δ2(Rn)βfHδ-1(Rn),(7)

where Hδ-1(Rn) denotes the dual of the Sobolev space H-δ1(Rn).

We rewrite (Equation5) as the operator equationusc=u~0+Lk(usc),u~0=Lk(u0),

where the integral operator Lk is defined as(8) Lkf(x):=RnGk+(|x-y|)2i(W(y)f(y))-q(y)f(y)dy.(8)

The mapping properties and asymptotic expansion of the operator Lk are studied in the following lemmas which have also independent interest.

Lemma 2.1:

Let us assume that conditions (Equation1) and (Equation2) are fulfilled. Then u~0L-σ/22(Rn) and the integral operator Lk maps L-σ/22(Rn) into itself, where σ is the same as in (Equation2). Moreover, uniformly in |k|1 the following estimates hold(9) u~0L-σ/22(Rn)β2WLσ/22(Rn)+qLσ/22(Rn)(9)

and(10) LkfL-σ/22(Rn)β2WLσ(Rn)+CpqLσp(Rn)fL-σ/22(Rn),(10)

where the constant Cp is equal toCp=1(2π)nΓ(p-n2)Γ(p2)1p.

Proof Conditions (Equation2) for p and σ imply that σ2>12 andLσp(Rn)Lσ/22(Rn).

It is therefore true that under the conditions (Equation1) and (Equation2) the functions V, W, W and |W|2 belong to Lσ/22(Rn). Since u0 is bounded then using Agmon’s result (Equation6) we obtainu~0L-σ/22(Rn)β|k|2|k|WLσ/22(Rn)+qLσ/22(Rn).

Hence, the estimate (Equation9) is proved. Next, applying (Equation7) we obtain for any fL-σ/22(Rn) thatLkfL-σ/22(Rn)β2(Wf)Hσ/2-1(Rn)+qfHσ/2-1(Rn)β2WfLσ/22(Rn)+qfHσ/2-1(Rn)β2WLσ(Rn)fL-σ/22(Rn)+qfHσ/2-1(Rn)

since the conditions (Equation1) and (Equation2) guarantee that W belongs to Lσ(Rn). In order to estimate the second term qfHσ/2-1(Rn) we proceed as follows. First, we rewrite and estimate this norm (using Hölder’s and Hausdorff-Young inequalities) asqfHσ/2-1(Rn)=q~f~H-1(Rn)=F(q~f~)L-12(Rn)C0F(q~f~)L2pp-2(Rn)C0(2π)-n/pq~f~L2pp+2(Rn)C0(2π)-n/pq~Lp(Rn)f~L2(Rn),p>n,

where F is the Fourier transform, q~(x)=(1+|x|2)σ/2q(x), f~(x)=(1+|x|2)-σ/4f(x) and the constant C0 is equal toC0=Rn1(1+|x|2)p/2dx1/p=πnΓ(n2)0rn-22(1+r)-p/2dr1/p.

Combining this constant C0 with the latter inequality we obtain the value for the constant Cp from lemma and the inequality (Equation10). Thus, lemma is proved.

Corollary 2.1:

The integral operator Lk for fixed k>0 maps L(Rn) into itself with the norm estimate depending on the norms WLσp(Rn) and qLσp(Rn).

Lemma 2.2:

Under the same assumptions as in Lemma 2.1, for any fixed k>0 and for any fC0(Rn) the following asymptotical representation holds:(11) Lkf(x)=Ceik|x|kn-32|x|n-12Rne-ik(θ,y)2kθW(y)f(y)+q(y)f(y)dy+O1|x|n+12,|x|,(11)

where θ=x|x| and constant C>0 depends only on n.

Proof Since fC0(Rn) then integration by parts in (Equation8) leads toLkf(x)=-2iRnyGk+(|x-y|)W(y)f(y)dy-RnGk+(|x-y|)q(y)f(y)dy.

In view of this version of (Equation8) one must study the behaviour for |x| of the functions Gk+ and Gk+, i.e.i4k2π|x-y|n-22Hn-22(1)(k|x-y|),kx-y|x-y|i4k2π|x-y|n-22Hn2(1)(k|x-y|),

respectively, where Hν(1) denotes the Hankel function of first kind and of order ν. Since k>0 is fixed and y is bounded then k|x-y| as |x|. Thus, we may use the behaviour of the Hankel functions Hν(1) for large argument (see, e.g. [Citation26]), i.e.Hn-22(1)(z)=cneizz+O1z3/2,Hn2(1)(z)=-icneizz+O1z3/2,

where z+ and the constant cn (which is the same for both asymptotic) is equal to (see, e.g. [Citation26])cn=2πe-iπ4(n-1),n=2,3,

Hence, we obtainGk+(|x-y|)=icn4(2π)n-22eik|x-y|kn-32|x-y|n-12+O1|x|n+12,|x|+

and (using x-y|x-y|=x|x|+O(1|x|)) we have alsoyGk+(|x-y|)=cn4(2π)n-22θkeik|x-y|kn-32|x-y|n-12+O1|x|n+12,|x|+

with θ=x|x|. In addition to these asymptotics, for bounded y and |x|+ we have|x-y|-n-12=|x|-n-12+O(|x|-n+12),eik|x-y|=eik|x|e-i(θ,y)+O(|x|-1),

and thereforeLkf(x)=c~neik|x|kn-32|x|n-12Rne-ik(θ,y)2kθW(y)f(y)+q(y)f(y)dy+O1|x|n+12,

where c~n=-icn4(2π)n-22. Thus, lemma is proved.

These lemmas allow us to achieve the main goal of this section in the form of the first main result of this work. Let us denote by α and γ the following norms:(12) α:=2WLσ(Rn)+CpqLσp(Rn),γ:=2WLσ/22(Rn)+qLσ/22(Rn).(12)

Theorem 2.1:

Assume that the conditions (Equation1) and (Equation2) for the coefficients of H are satisfied and assume that αβ<1. Then for any |k|1 the integral Equation (Equation5) has a unique solution usc from the space L-σ/22(Rn). Moreover, uniformly in |k|1 the following estimate holds(13) uscL-σ/22(Rn)βγ1-βα.(13)

Proof Lemma 2.1 and the conditions of this theorem say that the integral operator Lk maps in L-σ/22(Rn) with the norm estimate(14) LkL-σ/22(Rn)L-σ/22(Rn)βα<1.(14)

In particular,LkH-σ/21/2(Rn)L-σ/22(Rn)βα.

Since u~0 also belongs to L-σ/22(Rn) (see Lemma 2.1) with the norm estimate βγ then the integral equation has a unique solution usc that can be obtained by the iterations asusc=(I-Lk)-1(u~0)=j=0Lkj(u~0).

The norm estimate follows now from Lemma 2.1 and the latter representation. It proves the theorem.

Corollary 2.2:

Under the conditions of Theorem 2.1, usc(x,k,θ) belongs to L(Rn) in x for any fixed k>0 and uniformly in θSn-1.

3. Scattering amplitude and direct backscattering Born approximation

In this section, we will consider the direct backscattering Born approximation for the magnetic Schrödinger operator H with conditions (Equation1)–(Equation2) but with σ>n-n/p there. The motivation for this problem is connected to the fact that the knowledge of the scattering amplitude with the backscattering data gives essential information about the unknown functions V and W.

Theorem 2.1, Lemma 2.2 (see representation (Equation11)) and Corollary 2.1 yield the following asymptotical representation (we may integrate by parts in (Equation5) with these new conditions (Equation1)–(Equation2) for the coefficients of H since u is bounded for fixed k>0) for the scattering solutions u(x,k,θ) with fixed k>0 as |x|+:(15) u(x,k,θ)=eik(x,θ)+Ceik|x|kn-32|x|n-12A(k,θ,θ)+o1|x|n-12,(15)

where the function A is called the scattering amplitude and defined by(16) A(k,θ,θ)=Rne-ik(θ,y)2kθW(y)u(y)+q(y)u(y)dy.(16)

Here (as above) q=iW+|W|2+V, the constant C>0 depends only on n and the latter equality is understood in the sense of tempered distributions.

Substituting u=u0+usc into (Equation16) gives that(17) A(k,θ,θ)=Rne-ik(θ,y)2kθW(y)u0(y)+q(y)u0(y)dy+Rne-ik(θ,y)2kθW(y)usc(y)+q(y)usc(y)dy=:AB(k,θ,θ)+R(k,θ,θ).(17)

The function AB is called the direct Born approximation. It can be easily checked that AB is actually equal to(18) AB(k,θ,θ)=2kθF(W)(k(θ-θ))+F(q)(k(θ-θ))=k(θ+θ)F(W)(k(θ-θ))+F(|W|2+V)(k(θ-θ)),(18)

where F denotes the n-dimensional Fourier transformF(f)(ξ)=Rnf(x)ei(x,ξ)dx.

Next we establish a connection between the direct Born approximation and the coefficients of the magnetic Schrödinger operator. In the future, this might give insight into the inverse scattering problem with full scattering data.

Proposition 3.1:

Let ξ0 be an arbitrary vector from Rn and let ξ^ be the unit vector that is orthogonal to ξ. Let also k>0 be such that ξ24k2. Let us choose θ and θ as follows:θ=ξ2k+ξ^2k4k2-ξ2,θ=-ξ2k+ξ^2k4k2-ξ2.

Then θ,θSn-1, ξ=k(θ-θ), and(19) F(|W|2+V)(ξ)=12AB(k,θ,θ)+AB(k,-θ,-θ)4k2-ξ2(ξ^,F(W)(ξ))Rn=12AB(k,θ,θ)-AB(k,-θ,-θ).(19)

Proof Follows straightforwardly from (Equation18).

Our next main interest (with respect to inverse problems) concerns to the particular case θ=-θ. This case leads to the so-called direct backscattering Born approximation, i.e.(20) ABb(k,-θ,θ)=F(|W|2+V)(2kθ).(20)

Formulae (Equation17)–(Equation20) show that in the frame of the backscattering Born approximationA(k,-θ,θ)F(|W|2+V)(2kθ).

But we want to write precisely more terms in the Born series. This is presented in the following theorems which constitute the second main result of this work.

Theorem 3.1:

The backscattering amplitude A(k,-θ,θ) admits the following representation(21) A(k,-θ,θ)=F(|W|2+V)(2kθ)+2kθh1(kθ)+h0(kθ),(21)

where the functions h1(η),h0(η) both belong to L(Rn) with the norm estimates(22) h1L(Rn),h0L(Rn)βγ21-βα.(22)

Proof The formula (Equation17) shows thatR(k,-θ,θ)=-2kθRneik(θ,y)W(y)usc(y,k,θ)dy+Rneik(θ,y)q(y)usc(y,k,θ)dy.

Sinceusc(x,k,θ)=j=1Lkj(u0),u0(x,k,θ)=eik(θ,x)

then the representation (Equation21) follows from the latter formulas. The estimates (Equation22) can be easily obtained from (Equation10) and (Equation13) by using the Hölder’s inequality since W,q belong to Lσ/22(Rn) and usc belongs to L-σ/22(Rn). Thus, theorem is proved.

Concerning the latter terms in the Born series (Equation21) we have the following result.

Theorem 3.2:

The sum of the functions 2kθh1 and h0 admits the following representation:(23) 2kθh1(kθ)+h0(kθ)=-1(2π)nRnF(q¯)(kθ+η)F(q)(kθ-η)η2-k2-i0dη+4k(2π)nRnθF(W)(kθ+η)ηF(W)(kθ-η)η2-k2-i0dη+hrest(kθ),(23)

where q¯ denotes the complex conjugate of q, and where the function hrest has the following estimate for |k|1 (24) hrestL(Rn)3β2αγ21-βα.(24)

Proof The formulas (Equation17) and (Equation21) imply that the R(k,-θ,θ)=2kθh1+h0=R1+R2, whereR1=-2kθRneik(θ,y)W(y)Lku0(y,k,θ)dy+Rneik(θ,y)q(y)Lku0(y,k,θ)dy,R2=-2kθRneik(θ,y)W(y)j=1Lkju0~(y,k,θ)dy+Rneik(θ,y)q(y)j=1Lkju0~(y,k,θ)dy

with Lku0 equal to (see (Equation8))Lku0(y,k,θ)=-2kθRneik(θ,z)W(z)Gk+(|y-z|)dz-Rneik(θ,z)q¯(z)Gk+(|y-z|)dz.

Using integration by parts, R2 can be written asR2=iRneik(θ,y)·W(y)j=1Lkju0~(y,k,θ)dy+2iRneik(θ,y)W(y)·j=1Lkju0~(y,k,θ)dy-Rneik(θ,y)(|W|2+V)(y)j=1Lkju0~(y,k,θ)dy.

Then we can easily obtain the representation(25) R1=4k2RnRneik(θ,y+z)Gk+(|y-z|)θW(y)θW(z)dydz+2kRnRneik(θ,y+z)Gk+(|y-z|)θW(y)(q¯(z)-q(z))dydz-RnRneik(θ,y+z)Gk+(|y-z|)q(y)q¯(z)dydz=:I1+I2+I3(25)

in the sense of tempered distributions. Let us considerI1=4k2RnθW(z)eik(θ,2z)Rneik(θ,u)Gk+(|u|)θW(u+z)dudz.

The integral with respect to u can be written as the Fourier transform of a product. Using F(ϕ·ψ)=(2π)-nF(ϕ)F(ψ) this Fourier transform becomes the convolution(2π)-nF(Gk+(|u|))F(θW(u+z))(kθ).

Using F(Gk+)(s)=1s2-k2-i0 the latter expression can be written as(2π)-nRnF(θW(u+z))(η)(kθ-η)2-k2-i0dη=(2π)-nRne-i(z,η)θF(W)(η)η2-2k(θ,η)-i0dη

after elementary simplifications. This allows us to compute I1 asI1=4k2(2π)nRnθF(W)(η)η2-2k(θ,η)-i0RnθW(z)ei(2kθ-η,z)dzdη=4k2(2π)nRnθF(W)(η)θF(W)(2kθ-η)η2-2k(θ,η)-i0dη=4k2(2π)nRnθF(W)(kθ+η)θF(W)(kθ-η)η2-k2-i0dη.

Proceeding similarly for I2 and I3 yields(26) I2=-4ik(2π)nRnθF(W)(kθ+η)F(W)(kθ-η)η2-k2-i0dη,I3=-1(2π)nRnF(q¯)(kθ+η)F(q)(kθ-η)η2-k2-i0dη.(26)

Since I2 is actually equal toI2=-4ik(2π)nRnθF(W)(kθ+η)(-ikθ+iη)F(W)(kθ-η)η2-k2-i0dη=-I1+4k(2π)nRnθF(W)(kθ+η)ηF(W)(kθ-η)η2-k2-i0dη

then the main part in the representation (Equation23) is proved. It remains now to estimate hrest (or R2). Indeed (see, (Equation12)–(Equation14)),|R2|Rn|·W(y)|j=1Lkju0~(y,k,θ)dy+2Rn|W(y)|j=1Lkju0~(y,k,θ)dy+Rn(|W|2+|V|)(y)j=1Lkju0~(y,k,θ)dyqLσ/22(Rn)j=1Lkju0~L-σ/22(Rn)+2WHσ/21(Rn)j=1Lkju0~H-σ/2-1(Rn)qLσ/22(Rn)+2WHσ/21(Rn)β2αγ1-βα3β2αγ21-βα.

The latter inequality together with (Equation24)–(Equation26) show that Theorem 3.2 is completely proved.

Summarizing our considerations (see (Equation20)–(Equation26)) we may conclude that neglecting hrest the following direct backscattering Born approximation for the magnetic Schrödinger operator holds(27) A(k,-θ,θ)A~(k,-θ,θ):=F(|W|2+V)(2kθ)-1(2π)nRnF(q¯)(kθ+η)F(q)(kθ-η)η2-k2-i0dη+4k(2π)nRnθF(W)(kθ+η)ηF(W)(kθ-η)η2-k2-i0dη,(27) (28) ub(x,k,θ)=eik(x,θ)+Ceik|x|kn-32|x|n-12A~(k,-θ,θ).(28)

These formulas give very good approximation for the backscattering amplitude A and for the scattering solutions u uniformly in θSn-1. It is important that for this approximation we need to have only the magnetic potential W and electric potential V, but we do not need to have the exact solution u(x,k,θ) of the differential equation Hu(x)=k2u(x).

4. Inverse Born approximation

The direct approximation (Equation28) can be used for the inverse backscattering Born approximation as well. Theorems 3.1 and 3.2 give us the key to define the inverse backscattering Born approximation as(29) qBb(x):=1(2π)n0kn-1dkSn-1e-ik(θ,x)Ak2,-θ,θdθ.(29)

Due to this definition and formula (Equation20) (see also (Equation28)) we conclude thatqBb(x)=(|W|2+V)(x)+qquad(x)+qrest(x),

where the quadratic form qquad can easily be calculated precisely from (Equation27) as(30) qquad(x)=-1(2π)nFξx-1RnF(q¯)(ξ-η)F(q)(η)η2-(ξ,η)-i0dη+1(2π)nFξx-1RnξF(W)(η)(2η-ξ)F(W)(ξ-η)η2-(ξ,η)-i0dη.(30)

This equality must be understood in the sense of tempered distributions. The precise form (Equation30) of the quadratic term qquad allows us to estimate its smoothness in the two dimensional case. Indeed, we have the following result from [Citation23, Theorem 2.2].

Proposition 4.1:

[n=2] Let us assume that conditions (Equation1) and (Equation2) are fulfilled. Let us also assume that F(q),F(W)Ls(R2) for some 1<s<2. Then qquad belongs to the space

(1)

Ht(R2),t<4/s-2 for 4/3<s<2;

(2)

H1(R2) for s=4/3;

(3)

C(R2)L(R2) for 1<s<4/3.

The latter proposition means that the term qquad is smoother than the original potential V (note that W is continuous and bounded). It means that we can recover the main singularities of V (such as jumps over smooth curves) using the inverse Born backscattering approximation.

Remark 4.1:

If we assume that VHα(Rn) and WHα+1(Rn) with compact supports and withn-12<α<n2,n=2,3

then qquadC(Rn)L(Rn). This further simplifies the recovery of singularities of V. More precisely, using the inverse Born backscattering approximation we can reconstruct all jumps and singularities of V.

5. Numerical examples

In this section we study numerically in two dimensions how well qBb approximates |W|2+V. To this end, we employ the following scheme proposed in [Citation27,Citation28]. For given V,W we compute an approximation to the scattered field usc asusc(x)j=1Juj,uj=Lkuj-1,(see the proof of Theorem 3.1) by performing numerical integration over the supports of W,V, see [Citation27,Citation28] for details.

Then (synthetic) backscattering data is obtained from the approximationA(k,-θ,θ)4πkR-1-ie-ikRusc(-Rθ,k,θ)

after putting x=Rθ=-Rθ in (Equation15). We use J=2,R=105 in our examples and we add 1% of Gaussian noise to scattering data.

In order to compute qBb we invert (Equation29) to read(31) R2eik(x,θ)qBb(x)dx=A(k/2,-θ,θ).(31)

We represent qBb in the discrete formqBb(x)=j=1Nfjχrj(x),

where rj is a subdivision of unit cube, our area of interest, and χΩ is the characteristic function of ΩR2. If we substitute this representation into (Equation31) and evaluate it at several points k,θ we obtain a linear system Ef=g for the solution of unknown coefficients f=(fj)j=1N.

We use N=802, 12 values for k and 40 values for θ uniformly from the unit circle. Hence our linear system is of size 480×6400. The regularized solution of our underdetermined and ill-conditioned linear system is obtained by truncated singular value decomposition (TSVD) as follows. If E=USV with S=diag(s1,,sn) is the singular value decomposition of E then we computef=VLUg,

where L=diag(1/s1,,1/sr,0,,0) and sr is the last singular value exceeding a prescribed tolerance stol=0.02.

Figure 1. Potential combination |W|2+V (left) and TSVD reconstruction (right), Example 1.

Figure 1. Potential combination |W→|2+V (left) and TSVD reconstruction (right), Example 1.

Figure 2. Potential combination |W|2+V (left) and TSVD reconstruction (right), Example 2.

Figure 2. Potential combination |W→|2+V (left) and TSVD reconstruction (right), Example 2.

Next we describe our sample scatterers. Since W is actually bounded and continuous we concentrate our attention to recovering jumps of V. We put V(x)=0.5χΩ(x), where the domain Ω is an ellipse in Example 1 and a rectangle in Example 2, see Figures and . For W=(W1,W2) we use the (infinitely smooth) bump functionWj(x)=wjexp(1)χ|x-cj|2<rj2(x)exp(rj2/(|x-cj|2-rj2)),j=1,2

supported in the ball of radius rj centered at cj with scaled height wj. The following table summarizes the parameters of magnetic potentials for both examples.

In Figures and , the left panel shows the unknown combination |W|2+V. The right panel shows the reconstruction of qBb. In both figures white line indicates the true geometry of the support of the scatterers. We see that the inverse Born approximation is able to locate quite accurately the shape and location of these supports from noisy data, even in the case of non-smooth support of V.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

This work was supported by the Academy of Finland; Luonnontieteiden ja Tekniikan Tutkimuksen Toimikunta [application number 250215, Finnish Programme for Centres of Excellence in Research 2012–2017].

References

  • Päivärinta L , Somersalo E . Inversion of discontinuities for the Schrödinger equation in three dimensions. SIAM J Math Anal. 1991;22:480–499.
  • Nachman AI . Inverse scattering at fixed energy. In: Proceedings of 10th International Congress on Mathematical Physics, Leipzig, August 1991. Berlin: Springer; 1992.
  • Nachman AI . Global uniqueness for a two-dimensional boundary value problem. Ann Math. 1996;143:71–96.
  • Sun Z , Uhlmann G . Recovery of singularities from formally determined inverse problems. Commun Math Phys. 1993;153:431–445.
  • Isakov V , Sylvester J . Global uniqueness for a semi-linear elliptic inverse problem. Commun Pure Appl Math. 1994;47:1403–1410.
  • Päivärinta L , Serov V , Somersalo E . Reconstruction of singularities of the potential for the Schrödinger operator in two dimensions. Adv Appl Math. 1994;15:97–113.
  • Päivärinta L , Serov V . Recovery of singularities of a multi-dimensional scattering potential. SIAM J Math Anal. 1998;29:697–711.
  • Päivärinta L , Serov V . Inverse scattering problems for two-dimensional Schrödinger operator. J Inverse Ill-posed Probl. 2006;14:295–305.
  • Päivärinta L , Serov V . Recovery of jumps and singularities in the multi-dimensional Schrödinger operator from limited data. Inverse Probl Imaging. 2007;1:525–535.
  • Ola P , Päivärinta L , Serov V . Recovering singularities from backscattering in two dimensions. Commun PDE. 2001;26:697–715.
  • Ruiz A . Recovery of the singularities of a potential from fixed angle scattering data. Commun PDE. 2001;26:1721–1738.
  • Ruiz A , Vargas A . Partial recovery of a potential from backscattering data. Commun PDE. 2005;30:67–96.
  • Päivärinta L , Serov V . New estimates of the Green-Faddeev function and recovering of singularities in the two-dimensional Schrödinger operator with fixed energy. Inverse Probl. 2005;21:1291–1301.
  • Reyes JM . Inverse backscattering for the Schrödinger equation in 2D. Inverse Probl. 2007;23:625–643.
  • Serov V . Inverse Born approximation for the nonlinear two-dimensional Schrödinger operator. Inverse Probl. 2007;23:1259–1270.
  • Serov V , Harju M . Partial recovery of the potentials in generalized nonlinear Schrödinger equation on the line. J Math Phys. 2007;48:18 pp.
  • Serov V , Harju M . A uniqueness theorem and reconstruction of singularities for a two-dimensional nonlinear Schrödinger equation. Nonlinearity. 2008;21:1323–1337.
  • Serov V , Sandhu J . Inverse backscattering problem for the generalized nonlinear Schrödinger operator in two-dimensions. J Phys A Math Theor. 2010;43:325206.
  • Lechleiter A . Explicit characterization of the support of non-linear inclusions. Inverse Probl Imaging. 2011;5:675–694.
  • Reyes JM , Ruiz A . Reconstruction of the singularities of a potential from backscattering data in 2D and 3D. Inverse Probl Imaging. 2012;6:321–355.
  • Eskin G , Ralston J . Inverse scattering problem for the Schrödinger equation with magnetic potential at a fixed energy. Commun Math Phys. 1995;173:199–224.
  • Päivärinta L , Salo M , Uhlmann G . Inverse scattering for the magnetic Schrödinger operator. J Func Anal. 2010;259:1771–1798.
  • Serov V . Inverse backscattering Born approximation for the a two-dimensional magnetic Schrödinger operator. Inverse Probl. 2013;29:075015.
  • Simon B . Schrödinger operators with singular magnetic vector potentials. Math Z. 1973;131:361–370.
  • Agmon S . Spectral properties of Schrödinger operators and scattering theory. Ann Scuola Norm Sup Pisa. 1975;2:151–218.
  • Lebedev NN . Special functions and their applications. New York (NY): Dover; 1972.
  • Harju M . Numerical computation of the inverse Born approximation for the nonlinear Schrödinger equation in two dimensions. Comput Methods Appl Math. 2016;16(1):133–143.
  • Fotopoulos G , Harju M . Inverse scattering with fixed observation angle data in 2D. Inverse Probl Sci Eng. 2017;25(10):1492–1507.

Appendix 1

Proof of (Equation3)

We rewrite first the magnetic Schrödinger operator in the formH=-Δ-2iW·+q¯,

where q=iW+|W|2+V. If now fW22(Rn) then for inclusion W22(Rn)D(H) it is enough to show that W·f and q¯f belong to L2(Rn), since obviously ΔfL2(Rn). We have WWp,σ1(Rn)Wp1(Rn)L(Rn) for σ>0 and p>n. This fact implies that W·fL2(Rn) and q¯Lp(Rn). Further, since W22(Rn)L(Rn) for n=2,3 thenRn|q¯|2|f|2dx=|q¯|<1|q¯|2|f|2dx+|q¯|>1|q¯|2|f|2dxfL2(Rn)2+fL(Rn)2q¯Lp(Rn)p<

for p>n. If n5 then using the imbedding W22(Rn)L2nn-4(Rn) we obtain analogously thatRn|q¯|2|f|2dxfL2(Rn)2+fL2n/(n-4)(Rn)2q¯Lp(Rn)4p/n<,p>n.

The same result is valid for n=4 due to the imbedding W22(R4)Ls(R4) for any s<. ThusW22(Rn)D(H),n2.

For the opposite inclusion we need to show that if g=Hf belongs to L2(Rn) with fW21(Rn) then actually fW22(Rn). To this end, we use the identityf=(-Δ+I)-1(g+2iW·f-q¯f+f).

Using now the imbedding W21(Rn)L2nn-2(Rn) for n3 and W21(R2)Ls(R2) for any s< we may easily obtain thatg+2iW·f-q¯f+fL2(Rn),n2.

Thus, fW22(Rn) since(-Δ+I)-1:L2(Rn)W22(Rn),n2.

This proves the needed opposite imbedding.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.