Brought to you by:
Paper

The inside–outside duality for scattering problems by inhomogeneous media

and

Published 17 September 2013 © 2013 IOP Publishing Ltd
, , Citation Andreas Kirsch and Armin Lechleiter 2013 Inverse Problems 29 104011DOI 10.1088/0266-5611/29/10/104011

0266-5611/29/10/104011

Abstract

This paper investigates the relationship between interior transmission eigenvalues k0 > 0 and the accumulation point 1 of the eigenvalues of the scattering operator when k approaches k0. As is well known, the spectrum of is discrete, the eigenvalues μn(k) lie on the unit circle in and converge to 1 from one side depending on the sign of the contrast. Under certain (implicit) conditions on the contrast it is shown that interior transmission eigenvalues k0 can be characterized by the fact that one eigenvalue of converges to 1 from the opposite side if k tends to k0 from below. The proof uses the Cayley transform, Courant's maximum–minimum principle, and the factorization of the far field operator. For constant contrasts that are positive and large enough or negative and small enough, we show that the conditions necessary to prove this characterization are satisfied at least for the smallest transmission eigenvalue.

Export citation and abstractBibTeXRIS

1. Introduction

Interior transmission eigenvalue problems occur in the study of scattering problems of plane time-harmonic fields by an inhomogeneous medium filling a bounded domain D. In scattering theory, they play about the same role as the eigenvalue problem for −Δ in D with respect to, e.g., Dirichlet boundary conditions for the scattering problem by a sound-soft obstacle. We refer to [13] for their first appearance and to [4, 5, 9, 15, 18] for an incomplete list of references. Moreover, interior transmission eigenvalue problems are interesting objects of research in themselves because they fail to be self-adjoint. There are a few results for the corresponding inverse spectral problem, see [1, 2, 17], where the task is to recover information about the index of refraction from knowledge of the interior transmission eigenvalues. From the point of view of inverse scattering problems one would like to recover this information from the far field patterns of scattered waves or from the far field operator F, the linear integral operator whose kernel is formed by the far field patterns. It has been known for a long time that the injectivity of the far field operator F corresponding to the scattering by an inhomogeneous medium D is assured if the wavenumber k is not an eigenvalue of the corresponding interior transmission eigenvalue problem in D. (This is actually easy to prove, see theorem 2.3 below.) In other words, the scattering of a superposition of plane waves of the form

cannot result in a vanishing scattered field us if the wave number k is not an eigenvalue of the corresponding interior transmission eigenvalue problem in D. Therefore, there exists a relationship between the injectivity of the far field operator F—a property of the exterior of D—and the eigenvalues of an eigenvalue problem—a property of the interior of D. This relationship is sometimes called the inside–outside duality in scattering theory (see, e.g., [11, 12]) and has been used to determine the interior transmission eigenvalues from the far field patterns. The computation of these eigenvalues as those wavenumbers for which the far field operator fails to be injective suffers—at least theoretically—from the fact that injectivity is only a necessary but not a sufficient condition for k being no eigenvalue. It has been shown in [19] that it is very well possible that k is an eigenvalue but at the same time F is injective. Indeed, in theorem 2.3 below we recall a result which states that non-injectivity of the far field operator F for some k is equivalent to the fact that k is an interior transmission eigenvalue plus an additional condition on the corresponding eigenfunctions. In [19] it has been shown for D being a rectangular box that the far field operator is always one-to-one. In this case it is hence—at least theoretically—not possible to determine interior transmission eigenvalues merely by checking the injectivity of far field operators in an interval of frequencies. It is the goal of this paper to weaken the condition of non-injectivity of the far field operator F for wavenumber k into a condition which is exactly equivalent to k being an interior transmission eigenvalue.

In our proof we follow the paper [12] in which the scattering by an impenetrable, sound-soft obstacle has been studied. We (AK and AL) were not aware of this paper until recently and want to point out that this impressive work already contains the factorization of the far field operator in exactly the same form as in [14]3. The notation and partly also the analytical techniques this paper uses differ from standard notation and tools used in the community interested in mathematical inverse scattering theory. This might be a possible explanation why the paper [12] and its results apparently remained largely unknown in this community.

The approach presented below strongly relies on the fact that the far field operator is normal and does not straightforwardly extend to, e.g., absorbing media. Analogous results for different scattering problems leading to normal far field operators will be presented in forthcoming papers.

The paper is organized as follows. In section 2 we introduce the scattering problem by an inhomogeneous medium and the corresponding far field operator F and scattering operator . We recall properties of these operators, prove the mentioned relationship between non-injectivity of F to the interior transmission eigenvalue problem and recall the factorization of F as in [16]. In section 3 we characterize the interior transmission eigenvalues by the operator T which appears in the factorization of F. Section 4 studies the phases δn of the eigenvalues μn of the scattering operator . The eigenvalues lie on the unit circle in because of the unitarity of . We recall from [16] that μn tends to 1 from 'above' or 'below' as n, depending on the sign of the contrast. We translate this into a condition on the phases δn and prove a characterization of the 'first' eigenvalue4 using the Cayley transform and Courant's maximum–minimum principle. Finally, in sections 5 and 6 we consider this 'first' eigenvalue as a function of the wavenumber k and show that it tends to 1, too, as k approaches an interior eigenvalue k0. However, roughly speaking, this 'first' eigenvalue approaches 1 from 'below' or 'above' if μn approaches 1 from 'above' or 'below' (see our main theorem 6.3 for a precise formulation). We also give the corresponding result for the far field operator F instead of the scattering operator (see corollary 6.4) and finish with an illustrative numerical example for the case of scattering from a penetrable ball.

2. Transmission eigenvalues and the far field operator

We will model time-harmonic scattering from an inhomogeneous medium by the Helmholtz equation with a refractive index n that is equal to one outside the scattering object D (compare equation (2.2) below). This motivates to introduce the contrast function q := n2 − 1. Our general assumptions on the contrast and on the scatterer D in the entire paper are as follows.

Assumption 2.1. Let be open and bounded such that the complement is connected. Furthermore, we assume that the boundary ∂D of D is smooth enough such that the imbedding of H1(D) into L2(D) is compact. Let qL(D) be real-valued and satisfy

  • (1)  
    There exists c0 > 0 with 1 + q(x) ⩾ c0 for almost all xD.
  • (2)  
    Either q(x) > 0 for almost all xD or q(x) < 0 for almost all xD.
  • (3)  
    |q| is locally bounded below, i.e. for every compact subset MD there exists c > 0 (depending on M) such that

We extend q by zero outside of D.

We note that part (3) of this assumption is satisfied for continuous contrasts q that vanish at most on the boundary of D.

By k > 0 we denote the wavenumber. For any incident plane wave of direction , the (direct) scattering problem is to determine the scattered field such that the total field u := us + ui satisfies the Helmholtz equation

and such that us satisfies the Sommerfeld radiation condition

From now on, we call any solution v of the Helmholtz equation Δv + k2v = 0 outside some ball containing D that satisfies the radiation condition a radiating solution. If appropriate, we indicate the dependence of all fields on the incident direction by writing , and .

It is well known that this direct scattering problem is uniquely solvable (see, e.g., [7] or lemma 2.4 below). Furthermore, any radiating solution v of the Helmholtz equation has the asymptotic behavior

uniformly with respect to . The complex valued function v is called the far field pattern. In the special case where v is the scattered field us, the far field pattern depends on the direction of observation and the direction of the incident plane wave. We indicate this dependence by writing . Furthermore we define the far field operator F from L2(S2) into itself as the integral operator whose kernel is this far field pattern; that is,

The far field operator F is closely related to the scattering operator (or scattering matrix) , namely

We note that the factor 8π2 in the denominator—instead of the more common factor 2π—stems from our definition (2.4) of the far field pattern.

It is well known (see, e.g., [16, theorem 4.4] or [6]) that is unitary (that is, ) and that F is normal (that is, F and its adjoint F* commute).

Due to assumption 2.1, |q| is positive within D. This makes it convenient to introduce the weighted L2-space L2(D, |q| dx) which is the completion of with respect to the inner product

Obviously, L2(D, |q| dx) contains L2(D) as a dense subspace and coincides with L2(D) if |q| is bounded below in D by a (global) positive constant.

The starting point of this paper is the following connection between the injectivity of F and the following interior transmission eigenvalue problem.

Definition 2.2. k > 0 is an interior transmission eigenvalue if there exists a non-trivial pair (u, w) ∈ L2(D, |q| dx) × L2(D, |q| dx) such that uwH2(D) and

The differential equations are understood in the ultra-weak sense; that is,

and analogously for u. The boundary conditions can be reformulated as .

Theorem 2.3. F fails to be one-to-one if, and only if, k > 0 is an interior transmission eigenvalue such that the corresponding solution w of Δw + k2w = 0 can be extended to all of as a Herglotz wave function,

for some gL2(S2), .

Proof. We sketch the proof for the convenience of the reader. Let gL2(S2) be in the null space of F; that is

By linearity, Fg is the far field pattern of the scattered field which corresponds to the incident field

Rellich's lemma (see [7]) and unique continuation implies that the scattered field vanishes outside D. (Here we use the fact that the exterior of D is connected.) Therefore, the total field ug coincides with the incident field outside D. Therefore, the Cauchy data of ug and coincide on ∂D, and solves (2.7), (2.8).

If, on the other hand, (u, w) solves (2.7), (2.8) and w has the form (2.9) for some gL2(S2) then, by the same arguments, g is in the null space of F. The proof is finished by noting that is equivalent to (see, e.g., [7]). □

It is the aim of this paper to characterize the interior transmission eigenvalues by the far field operator F—or, equivalently, the scattering operator . We note that injectivity of F is equivalent to the fact that 1 is not an eigenvalue of . Since is unitary its eigenvalues lie on the unit circle in the complex plane. This translates into the fact that the eigenvalues λj of F lie on the circle of radius 8π2/k centered at (8π2/k)i on the imaginary axes. They tend to zero because F is compact. We will show that they tend to zero from one side only, depending on the sign of q. To prove this we will need properties of the factorization method. We collect them in the following theorem 2.5, after introducing some notation. We denote by H: L2(S2) → L2(D, |q| dx) the linear and compact Herglotz operator, defined by

Moreover, we introduce the constant

Finally, T: L2(D, |q| dx) → L2(D, |q| dx) is defined by Tf = f + k2v|D, where is the radiating weak solution to

that is,

for all with compact support and, additionally, v satisfies (2.3). Existence and uniqueness is formulated in the following lemma.

Lemma 2.4. For all k > 0 and all fL2(D, |q| dx) there exists a unique radiating solution of (2.11). For every R > 0 such that and every compact set K with the solution operators fv|B and f↦(v|K, ∇v|K) and fv are bounded from L2(D, |q| dx) into H1(B), into C(K) × C(K)3, and into L2(S2), respectively, the latter two even compact. Furthermore, these operators depend continuously on k.

There are several ways to prove this lemma, e.g., by reformulating the problem into the Lippmann–Schwinger integral equation. We omit the proof but refer to, e.g., [16].

Theorem 2.5. 

  • (a)  
    Let F: L2(S2) → L2(S2) be defined by (2.5). Then
  • (b)  
    The mapping fv|D is compact from L2(D, |q| dx) into itself. Therefore, T is a compact perturbation of the identity.
  • (c)  
    for all fL2(D, |q| dx).

For the proof we follow the presentation of [16], theorems 4.5 and 4.8.

  • (a)  
    Note that the scattered field us satisfies the differential equation
    Define the operator G from L2(D, |q| dx) into L2(S2) by Gf = v where v solves (2.11). Then F = k2GH by the superposition principle. The adjoint of H is given by
    which is the far field pattern w of the volume potential
    It satisfies Δw + k2w = −σq ψ in and is radiating. From
    and the definition of T we conclude that Gf = v = σH*Tf. Substituting this into F = k2GH yields the factorization (2.13).
  • (b)  
    This is easily seen by a regularity argument.
  • (c)  
    For fL2(D, |q| dx) and the corresponding field v we have with Tf = f + k2v that
    By the radiation condition we conclude that
    as R tends to infinity. Taking the imaginary part and the limit as R yields

3. A characterization of transmission eigenvalues

Now we want to characterize the interior transmission eigenvalues by the operator T. To this end, we denote the closure of the range of the Herglotz operator H in L2(D, |q| dx) by X; that is,

The last set equality is due to the density of Herglotz wave functions in the set of all solutions to the Helmholtz equation (see [8]). Then we have:

Theorem 3.1. 

  • (a)  
    Let k > 0 be an interior transmission eigenvalue with corresponding non-trivial pair (u, w). Then wX∖{0}, and w satisfies .
  • (b)  
    Let k > 0 and let wX∖{0} satisfy . Then there exists uL2(D) such that k is an interior transmission eigenvalue with corresponding pair (u, w). The function uw belongs to and does not vanish. In particular, for all wX∖{0} if k > 0 is not an interior transmission eigenvalue.

Proof. 

  • (a)  
    Let k > 0 be an interior transmission eigenvalue with corresponding u, wL2(D, |q| dx). We define in D and note that . We extend v by zero outside D and observe that v is the radiating solution of
    Therefore, with compact support and wX satisfies (2.11). Let such that Δwn + k2wn = 0 in D and wnw in L2(D, |q| dx). Then Tw = w + k2v|D and thus
    by Green's second theorem. The left-hand side converges to which proves that . Finally, because otherwise also v vanishes as a radiating solution of (3.2), too.
  • (b)  
    Let now wX such that . We denote by the radiating solution of (2.11) for f = w. From (2.16) we conclude that v = 0. Rellich's lemma yields that v vanishes outside D; that is, . We set u = w + k2v. For we conclude that
    Therefore, (u, w) satisfies the conditions of the interior transmission eigenvalue problem.

 □

Remark 3.2. From part (b) of the previous proof we observe that for wX implies that the radiating solution to (3.2) vanishes outside D and thus .

Corollary 3.3. Assume that for some wX. Define again to be the radiating solution of (3.2). Then

Proof. As we noted in the preceding remark 3.2, v vanishes outside D; that is, . Hence, (2.15) yields

The claim follows from passing to the limit as R. □

4. Phases of the spectrum of the scattering operator

In this section, we collect well-known properties of the eigenvalues of the (normal) far field operator and the related scattering operator. These properties will be useful later on to prove an inside–outside duality between the interior transmission eigenvalues and the eigenvalues of the far field operator.

We recall that the eigenvalues λn of F lie on the circle of radius 8π2/k centered at (8π2/k)i on the imaginary axes and converge to zero.

Lemma 4.1. Let k be no interior transmission eigenvalue. Then λn converges from the right to zero if σ = +1 and from the left if σ = −1; that is, σReλn > 0 for sufficiently large n.

Proof. Let ψnL2(S2) be the eigenfunctions of F corresponding to the eigenvalues λn such that forms an orthonormal basis in L2(S2). Define by

From the factorization we conclude that

with sn = λn/|λn| and δn, m = 0 for and δn, m = 1 for n = m. We note that |sn| = 1 and Imsn > 0. Since λn tends to zero, the only accumulation points of sn can be 1 or −1. From T = I + C for some compact operator C we conclude that

First we show that the sequence ϕn is bounded. Otherwise there exists a subsequence such that . The sequence satisfies

Since it is bounded there exists a weakly convergent subsequence . The compactness of C implies and thus . Taking the imaginary part yields and thus also

theorem 3.1 shows that which contradicts . Therefore, the sequence ϕn is bounded and contains a weakly convergent subsequence ϕn⇀ϕ. Again, converges to . Also, taking the imaginary part of (4.1), yields and thus again and . We conclude again ϕ = 0 and thus . Equation (4.1) yields and enforces σ to be the accumulation point of (sn) rather than −σ. □

We translate this into a condition on the eigenvalues of the scattering matrix .

Corollary 4.2. Let k be no interior transmission eigenvalue and let be the eigenvalues of the scattering operator . Then |μn| = 1 for all n, they converge to 1 and σImμn > 0 for sufficiently large n.

Let k be not an interior transmission eigenvalue. Then 1 is not an eigenvalue of and the far field operator F is one-to-one with dense range. We define the Cayley transform by

It is easily seen that is self-adjoint, its spectrum is discrete, and μn = exp ( − 2iδn) is an eigenvalue of for some δn ∈ (0, π) if, and only if, is an eigenvalue of . Since (μn) tends to one with σImμn > 0 we conclude that

In particular, the numbers

are well defined and belong to (0, π) if k is not an interior transmission eigenvalue. We call them the 'first' eigenvalue.

Lemma 4.3. Let k be no interior transmission eigenvalue.

  • (a)  
    Let first σ = −1; that is, q < 0 in D. Then
  • (b)  
    Second, let σ = +1; that is, q > 0 in D. Then

Note that in both cases the denominator is strictly positive because of the assumption on k and theorem 3.1(b).

Proof. (a) Courant's max–min principle (see [10]) yields

Now we make use of the form and the factorization (2.13). In the following we write (Th, h) instead of . Then

which proves the assertion. Part (b) follows the same lines by just replacing the infimum by the supremum and noting that

 □

5. Spectral behavior of the scattering operator at a transmission eigenvalue

In the following we study the dependence of the eigenvalues of the scattering operator on the wavenumber k and write X(k), T(k), δ±(k), and so on to indicate this dependence. The eigenvalue characterization from lemma 4.3 uses k-dependent spaces over which the infimum or supremum is taken. In a first step, we will transform this k-dependence into the quadratic forms by introducing the orthogonal projection operator P(k) from L2(D, |q| dx) into X(k). This step will hence eliminate the k-dependence of the considered function spaces.

For the sake of notational simplicity, we write kk0 and kk0 to indicate that tends to from below and above, respectively. More precisely, if kk0 (kk0) the inequality k < k0 (k > k0) is always satisfied in the limiting process.

Recall that the numbers δ±(k) were defined in (4.2). They allow us to state the following simple result.

Lemma 5.1. Let k0 > 0 and w0X(k0) such that and and assume that

  • (a)  
    Let σ = −1; that is, q < 0 in D. Then
  • (b)  
    Let σ = +1; that is, q > 0 in D. Then

Note that we will show in lemmas 5.2 and 5.3 that is indeed a differentiable function in k0 for every fixed w0X(k0).

Proof. (a) We note from theorem 3.1 that implies that k0 is an interior transmission eigenvalue. Let I = (k0 − ε, k0 + ε) be an interval containing no other transmission eigenvalue. From the previous lemma we have for kI∖{k0}

Furthermore, from Taylor's theorem we have that

with r(k) = o(|kk0|) and Imr(k) < 0. Here we used that

Therefore, we have

which converges to − provided kk0 such that α(kk0) < 0.

Part (b) is proven in the same way. □

Now we want to investigate the derivative of at an eigenvalue k0. We begin with the following result.

Lemma 5.2. Let k0 > 0 be an interior transmission eigenvalue. Due to theorem 3.1 there exists w0X(k0) such that and . Then is differentiable in k0 and

where v0 is the radiating solution of (2.11) for k = k0 and f = w0; that is,

Proof. First we note that v0 vanishes outside D by remark 3.2. In particular, . Let v' be the unique radiating solution of

and vk the solution of (5.2) corresponding to k instead of k0. Then it is easily seen that

Now we eliminate w0 from this equation by using (5.2)

where we used the definition of v'. This term vanishes only for constant functions v0; that is, for v0 = 0 because of . □

Now we study the dependence of the orthogonal projection P(k): L2(D, |q| dx) → X(k) on k. To give an explicit representation of P(k), let us denote by W the completion of with respect to the semi-norm . Then P(k) is explicitly given by

where solves the fourth order system

in the variational sense; that is, solves the W-coercive variational problem

Then we have the following extension of the previous lemma:

Lemma 5.3. Let k0 > 0 be an interior transmission eigenvalue and w0X(k0) such that and . Then is differentiable in k0 and

where v0 is the solution of (5.2).

Proof. First we note that kP(k)w0 is Frechét-differentiable and, by differentiating the characterization of P(k)w0,

where solves (5.4) for g = w0 and solves

By the chain rule we have

The first term on the right-hand side has been computed at k = k0 in the previous lemma. The adjoint T* of T is given by where vg is the radiating solution of

Indeed, if fL2(D, |q| dx) with corresponding vf satisfying

then

From one concludes as in the proof of theorem 2.5 that v which corresponds to T*(k0)w0 vanishes outside D and, therefore, coincides with . This shows that T*(k0)w0 = T(k0)w0 and thus

because P(k0)w0 = w0. Therefore also and

and solves

We compute, noting that P(k0)w0 = w0,

This proves the assertion. □

6. Inside–outside duality

We would like to find conditions on the contrast q ensuring that the derivative in (5.5) is either strictly positive or strictly negative, that is,

where

for all such that . Due to lemma 5.1, such a sign property allows, generally speaking, to characterize when k0 > 0 is an interior transmission eigenvalue, merely knowing the far field operators F(k) in some interval around k0. Unfortunately, we are able to prove this sign property under strong assumptions on the contrast q only, roughly speaking for constant contrast that is either positive and large enough or negative and small enough. Weakening these assumptions is an issue of ongoing research.

Theorem 6.1. Let k0 be the smallest interior transmission eigenvalue and q(x) = q0 > 0 for xD being constant such that

Here, and ρ1 denote the smallest Dirichlet eigenvalues of Δ2 and −Δ, respectively, in D.5 Then

for all such that .

Proof. Multiplication of (5.2) with , integration and Green's first identity yields

that is, taking the real part,

We write

Now we use that and arrive at

It has been shown in [15] that for q0 satisfying (6.3) the smallest transmission eigenvalue k0 satisfies

Therefore, for this eigenvalue we have that A > 0. □

Theorem 6.2. Let k0 be the smallest interior transmission eigenvalue and q(x) = q0 for q0 ∈ ( − 1, 0). Then there exists such that

for all such that .

Proof. We exploit again (6.5) together with the Poincaré inequality involving the first Dirichlet eigenvalue ρ1 of −Δ in D to estimate that

If R+ denotes the radius of the smallest ball B = B(0, R+) containing then

where ρ1, B denotes the smallest Dirichlet eigenvalue of the ball B = B(0, R+). Substituting this estimate into the previous estimate for A we arrive at

The last term on the right is negative if, and only if,

From the proof of theorem 3.2 and corollary 3.1 in [3] we know that the smallest transmission eigenvalue k0 of the obstacle D for the contrast q0 satisfies

where is the smallest transmission eigenvalue of the unit ball for the contrast q0 and R is the radius of the largest ball contained in . Hence, whenever

we can conclude that A is negative at least for the smallest transmission eigenvalue.

For constant q0, the smallest transmission eigenvalue of the unit ball can be estimated from above by the smallest positive zero of

It is well-known that positive roots of W are transmission eigenvalues. Setting , we find that

We use that j0(λ) = sin (λ)/λ and observe that j0(0) = 1 and . Therefore,

with ϕ(λ) = λcos λ − sin λ. Elementary arguments (consider ϕ'(λ) = −λsin λ) show that the first positive zero of W( ·, 0) is in the interval (π, 2π). Furthermore,

and thus

The implicit function theorem assures existence of and an interval I around such that

is uniquely solvable in and as n0 → 0. Since the limit n0 → 0 corresponds to q0 → −1 we obtain in particular that the smallest transmission eigenvalue remains bounded as q0 → −1, since

In consequence, the left-hand side of (6.7) remains bounded as q0 → −1 while the right-hand side obviously tends to + as q0 → −1. This shows that there exists such that (6.7) is indeed satisfied. □

Now we are ready to prove the main result of this paper.

Theorem 6.3 (Inside–outside duality for ). Let k0 > 0 and I = (k0 − ε, k0 + ε)∖{k0} such that no kI is an interior transmission eigenvalue. For kI let be the eigenvalues of with phases δn(k) ∈ (0, π).

  • (a)  
    Let k0 be the smallest interior transmission eigenvalue and let q = q0 satisfy (6.3). Then
    where δ(k) = min nδn(k).
  • (b)  
    Let (6.8) hold and q > 0 in D. Then k0 is an interior transmission eigenvalue.
  • (c)  
    Let k0 be the smallest interior transmission eigenvalue and let q = q0 satisfy the assumption of (6.6) in theorem 6.2. Then
    where δ+(k) = max nδn(k).
  • (d)  
    Let (6.9) hold and q < 0 in D. Then k0 is an interior transmission eigenvalue.

Proof. (a) Let k0 be the smallest interior transmission eigenvalue and w0X(k0) with and . We apply part (b) of lemma 5.1 and note that α > 0 by theorem 6.1. This yields (6.8).

(b) Let now (6.8) hold and assume, on the contrary, that k0 is not an interior transmission eigenvalue. By lemma 4.3 the condition (6.8) is equivalent to

Therefore, there exist sequences kjI, kjk0, and wjX(kj) with such that

for sufficiently large j. Let be the corresponding radiating solutions of (5.2). Then wjw0 weakly in L2(D, |q| dx) for some subsequence and some w0L2(D, |q| dx). It is easy to see that w0X(k0) and vjv0 weakly in H1(B(0, R)) for every ball B(0, R) where is the radiating weak solution to in (see lemma 2.4 for an existence proof). From (2.16) for f = wj we conclude that

The left-hand side converges to zero, the right-hand side to by lemma 2.4 again. Therefore, and thus v0 vanishes outside D by Rellich's lemma. Because k0 is not an interior transmission eigenvalue we conclude that w0 and v0 vanish everywhere; that is, wj and vj converge weakly to zero. Now we compute, similarly to (2.14),

and thus, taking the real part,

which converges to zero by lemma 2.4 and the compact imbedding of H1(B(0, R)) into L2(B(0, R)). Therefore, vj converges to zero in H1(B(0, R)) for every B(0, R) which implies wj → 0 in L2(D, |q| dx), a contradiction to . This ends the proof of part (b).

(c) As for the proof of part (a), we apply lemma 5.1(a), noting that α = 2σk0A > 0 by theorem 6.1 and the assumption q < 0 (i.e., σ = −1). This yields (6.9). The proof of part (d) follows in the same way as the proof of part (b), with obvious adaptions due to the different sign of q. □

The last theorem has been formulated for the scattering operator . Of course, there is an analogous result for the far field operator F. Let λn(k) be the eigenvalues of F(k). From lemma 4.1 we recall that the projections sn(k) = λn(k)/|λn(k)| onto the unit circle in satisfy Imsn(k) > 0 for all n and limnsn(k) = ±1 provided q≷0. In particular, if q > 0 there exists a unique with |s(k)| = 1 and Ims(k) > 0 and

while for q < 0 there exists a unique with |s+(k)| = 1 and Ims+(k) > 0 and

Corollary 6.4 (Inside–outside duality for F). Let k0 > 0 and I = (k0 − ε, k0 + ε)∖{k0} such that no kI is an interior transmission eigenvalue. Let λn(k) be the eigenvalues of F(k). Define s(k) and s+(k) as above by (6.10) and (6.11), respectively. (That is, as the projection of λn onto the unit circle which is furthest to the left if q > 0 and furthest to the right if q < 0, respectively.)

  • Let k0 be the smallest interior transmission eigenvalue and let q = q0 satisfy (6.3). Then
  • Let (6.12) hold and q > 0 in D. Then k0 is an interior transmission eigenvalue.
  • Let k0 be the smallest interior transmission eigenvalue and let q = q0 satisfy the assumption of (6.6) in theorem 6.2. Then
  • Let (6.13) hold and q < 0 in D. Then k0 is an interior transmission eigenvalue.

Proof. We only consider the case q > 0 of (a) and (b) since the case q < 0 of (c) and (d) can be shown analogously. By we observe that μn is an eigenvalue of if, and only if, the number

is an eigenvalue of F. Furthermore, we note that

and this is minimal for minimal δn (provided δn ∈ (0, π/2)). Furthermore, is equivalent to . Now the claim follows directly from theorem 6.3. □

We want to illustrate the statements of theorem 6.3 and of lemma 4.1 with a simple and explicit numerical example. Consider the scattering from a penetrable ball of radius R > 0, centered at the origin, and assume that the refractive index inside BR equals n0 = (1 + q0)1/2 for a constant q0 > −1. For this setting it is well known that one can compute F(k): L2(S2) → L2(S2) semi-analytically. Indeed, explicit computations using spherical Bessel and Hankel functions jl and hl of order , respectively, and spherical harmonics , one computes that the far field operator can in this special case be represented as

with Fourier coefficients of gL2(S2) given by

see, e.g., [7] for similar computations in the case of a sound-soft ball. Hence, the eigenvalues of F(k) are

Interior transmission eigenvalues are precisely those k0 > 0 for which the numerator in the last expression vanishes. Note that the order of the eigenvalues comes from the series expression of the far field operator in (6.14); in particular, the order is not chosen in accordance with (6.10) or (6.11).

According to lemma 4.1, as l the eigenvalues should approach zero from the right (left) if the contrast is positive (negative). We confirm this statement for the above-described spherical setting by plotting the eigenvalues of F(k) for k = 20 and for the contrast q0 = ±0.5. Figure 1 shows that for q0 = 0.5 the eigenvalues converge from the right to zero, whereas for q0 = −0.5, they converge from the left to zero. In this and in all the following examples, the radius R of the ball is equal to one.

Figure 1. Refer to the following caption and surrounding text.

Figure 1. The eigenvalues from (6.15) of F(20) for the cases (a) q0 = 0.5 and (b) q0 = −0.5 in the complex plane. The size of the markers increases in l = 1, ..., 50 while the transparency of the markers decreases.

Standard image High-resolution image

Figure 2 confirms the statement of theorem 6.3. For q0 = 0.9 > 0, the function k↦δ(k) plotted in (a) tends to 0 as k approaches the smallest transmission eigenvalue from below (and all further ones occurring in the considered interval). For q0 = −0.9 < 0, the function k↦δ+(k) plotted in (b) tends to π as k approaches the smallest transmission eigenvalue from below (and all further ones occurring in the considered interval). For each k, the minimum and the maximum used to define and , respectively, is computed using the first 300 values of .

Figure 2. Refer to the following caption and surrounding text.

Figure 2. (a) The function k↦δ(k) plotted for k ∈ (4, 25) for contrast q0 = 0.9. (b) The function k↦δ+(k) plotted for k ∈ (3, 25) for contrast q0 = −0.9. Both for (a) and (b) the domain D is the unit ball.

Standard image High-resolution image

Finally, we plot that the eigenvalue curves of F(k) (see (6.15)) for the special spherical setting indicated above. To this end, we use 100 equidistributed wavenumbers between kmin  and kmax . Again, we consider positive and negative constant contrasts q0 = 1.5 and q0 = −0.9, respectively. Figure 3(a) shows that for q0 = 1.5 the eigenvalue approaches 0 twice from the left as k increases. For q0 = −0.9, figure 3(b) shows that approaches 0 twice from the right as k increases. Both observations fit the statement of corollary 4.1, since s+(k) → −1 and s(k) → 1 as kk0 means that the corresponding eigenvalue of F(k) approaches 0 from the left and right, respectively. In the special case of a spherical scatterer, zero is an eigenvalue of F(k0) for any transmission eigenvalue k0 and, even more, the curves depend smoothly on k > 0. This is not valid in general. We already mentioned in the introduction that it may happen that the far field operator at a transmission eigenvalue is injective.

Figure 3. Refer to the following caption and surrounding text.

Figure 3. The eigenvalue from (6.15) for the contrasts (a) q0 = 1.5 and (b) q0 = −0.9 plotted in the complex plane for 100 equidistributed wavenumbers k between kmin  and kmax . In (a), kmin  = 1 and kmax  = 13.2 while in (b) kmin  = 1 and kmax  = 12.2. The size of the markers increases in k. The three dashed lines indicate circles containing the eigenvalues of F(k) for k = 1 and the two interior transmission eigenvalues visible in the plots where vanishes.

Standard image High-resolution image

Acknowledgments

We would like to thank the anonymous referees for carefully reading the manuscript, in particular for finding a serious mistake in a former version of the text. The research of AL was supported through an exploratory project granted by the University of Bremen in the framework of its institutional strategy, funded by the excellence initiative of the federal and state governments of Germany.

Footnotes

  • On the other hand, the authors of [12] were apparently not aware of the existing literature on boundary integral operators such as, e.g., the first edition of the monograph [7].

  • See (4.2) for the precise definition of what we mean by first eigenvalue.

  • Note that , see [18].

10.1088/0266-5611/29/10/104011
undefined