1. Introduction
In the recent few years, there has been an increased interest in the inverse spectral problems for the so-called Sturm–Liouville operators with frozen arguments that are given by the differential expression
and subject to some boundary conditions. Here
is a fixed (or “frozen”) value of the argument
x and
q is a function in
. The corresponding differential operator is non-self-adjoint, unbounded, and has compact resolvent in
. A natural question arises, what spectra such operators may have and whether their eigenvalues completely characterize potentials
q, i.e., whether
q can be reconstructed from the known freezing point
a and the corresponding eigenvalues.
A thorough study of these questions was made in several recent articles. Buterin and Vasiliev [
1] studied the problem of reconstructing the potential
q from the spectrum of the Sturm–Liouville-type operator (
1) in case when
is a complex-valued function and
is a rational number. They used the so-called transformation operators to get an integral representation of the solutions of the equation
, derived an integral representation of the characteristic function, and then obtained asymptotics of the eigenvalues and eigenfunctions. That allowed the authors to study the inverse spectral problem and identify the iso-spectral sets of potentials
q sharing the same spectrum and thus causing non-uniqueness of reconstruction; uniqueness was established for the class of potentials possessing some symmetry. Bondarenko et al. [
2] extended the above research to the case of various boundary conditions. A more difficult case when
is irrational was recently discussed in [
3]; uniqueness of
q was proved and the reconstruction algorithm suggested. In [
4], the authors considered the periodic boundary conditions for the differential expression (
1); then the unperturbed operator with
has eigenvalues of multiplicity 2, and the analysis of both direct and inverse spectral problems becomes technically more involved. Trace formulas and the inverse nodal problems were discussed in [
5].
We also mention the paper of Nizhnik et al. [
6], in which Sturm–Liouville eigenvalue problems (
1) on
were considered with
and non-local boundary conditions
. Here,
q is an arbitrary complex-valued function in the Hilbert space
and
denotes the scalar product therein. The corresponding operator is self-adjoint in
, and the authors studied the inverse spectral problem of reconstructing the potential
q from its eigenvalues. In [
7,
8], Nizhnik extended the results of [
6] to some other differential expressions and boundary conditions. The inverse problem for a class of self-adjoint perturbations that are integral operators with degenerate kernels was recently studied by Zolotarev [
9].
The purpose of this note is to suggest a different approach to the spectral study of the Sturm–Liouville-type operators with frozen arguments. Namely, the operator generated by (
1) can be viewed as a rank-one perturbation of the reference operator
A corresponding to
; indeed, the term
can be represented as
with
being the Dirac delta-function,
, and
denoting the pairing in the Hilbert space scale generated by
A; see details in the next section. We observe that this perturbation is unbounded as
does not belong to
; however, it is bounded (and even compact) relative to
A, and that allows a generalization of the preliminary results obtained for the bounded case with
.
Generic bounded rank one perturbations of self-adjoint operators are studied quite well, see e.g., the reference lists of [
10,
11,
12,
13]. In our resent work [
12,
13], we gave a complete characterization of possible spectra
of bounded non-self-adjoint rank-one perturbations
of self-adjoint operators
A with simple discrete spectrum. In particular, we proved therein that geometric multiplicities are at most 2 while the algebraic ones can be arbitrary; the only essential restriction on the eigenvalues of
B comes from their asymptotics.
In this paper, we shall first show that even when
or
are allowed to be singular, the statements on the multiplicity of eigenvalues of
B remain valid while their asymptotic distribution should accordingly be modified. The effective tool for proving the main results of the paper is the characteristic function
F of the operator
B; namely, the eigenvalues of
B that are not in the spectrum of
A are zeros of
F of required multiplicity. By studying the latter, we completely characterize eigenvalue asymptotics as stated in Theorems 1 and 2. We stress that this asymptotics differs from the one derived in [
13] for the bounded case
, and its derivation requires essential changes in the proofs. Next, zeros of
F allow for a unique reconstruction of
F, thus specifying
up to an iso-spectral set, see Corollary 5. Since
F is given by explicit formulae, this approach suggests a constructive algorithm of determining
from the spectrum of the perturbation
B. After establishing these abstract results, we specialize them to a wide variety of differential operators with frozen arguments including those of Sturm–Liouville type (
1) thus providing a unified treatment of the questions discussed in the papers [
1,
2,
3,
4].
The paper is organized as follows. In the next section, we introduce necessary definitions and specify the setting of the paper, recall auxiliary results, derive the characteristic function and explain why the results of [
12,
13] are applicable here. In
Section 3, we establish the asymptotics of the eigenvalues of
B and in
Section 4, we solve the inverse problem of reconstructing the vector
given
. Several examples are given in
Section 5.
2. Preliminaries
In this section, we collect some properties of rank-one perturbations of self-adjoint operators
A acting in a fixed separable (infinite-dimensional) Hilbert space
established in [
12,
13] that will be used to prove the main results of this work. The reader can find further references and examples of applications in the monographs [
10,
11].
Throughout the paper, we shall assume that
- (A1)
the operator A is self-adjoint and has simple discrete spectrum.
Without loss of generality, we assume that A is either bounded below or else is unbounded from both below and above. In both cases, we list the eigenvalues in increasing order as , , where in the former case and in the latter case. Since the motivation for this work stems from differential operators, we make an additional assumption that
- (A2)
the eigenvalues of
A are
d-separated, i.e.,
To treat rank-one perturbations (
2) with
not a regular function as in (
1), we introduce the scale of Hilbert spaces
,
[
14]. Without loss of generality, we assume that
A is invertible, adding to it
otherwise. Then for
,
coincides with
and is equipped with the scalar product
. For negative
,
is the completion of
in the norm generated by the above scalar product. The standard scalar product
extends by continuity to
and
via
and is called the
pairing between
and
.
In what follows, we assume that
but
with some
. Then the rank-one operator
is compact relative to
A ([
15], Ch. IV.1): indeed, in view of the relation
we find that
so that
is bounded and
is compact. It follows that the operator
is well defined and closed on the domain
of
A and has compact resolvent ([
15], Ch. IV.1).
Next, for
, we introduce the
characteristic function
this function (in a slightly different form) appears in the Krein resolvent formula for
B [
11,
12], and its zeros characterise the spectrum of
B. The standard form of
F as discussed in the previous work was
this formula also makes sense in the current setting if we interpret the scalar product as the pairing between
and
as explained above.
To see that both interpretations of
F coincide, we suggest yet another representation of
F using the spectral theorem for the operator
A. Namely, let
be a normalized eigenvector of
A corresponding to the eigenvalue
(so that the set
is an orthonormal basis of
), and let
and
denote the corresponding Fourier coefficients of the vectors
and
, so that
We point out that the Fourier coefficients
of
are well defined since the formula
makes sense as a pairing between
and
. Set also
then the characteristic function
F of (
4) can be written as
and thus can be analytically extended to
; we keep the notation
F for this extension. (Note that in (
6) and in what follows, the summations and products over the index sets that are not bounded from below and above are understood in the principal value sense.) It is known [
12] that
is the common part of the spectra of
A and
B, while the spectrum of
B in
coincides with the set of zeros of
F. For convenience, we set
.
It turns out that the function
F also characterizes eigenvalue multiplicities of the operator
B. We recall that the
geometric multiplicity of an eigenvalue
of
B is the dimension of the null-space of the operator
, while its
algebraic multiplicity is the dimension of the corresponding root subspace, i.e., of the set of all
such that
for some
. As proved in [
12], the geometric multiplicity of every eigenvalue
of
B is at most 2; multiplicity 2 is only possible when
, i.e.,
for some
and, in addition,
. It should be pointed out that the equality
implies that the subspace
is invariant under both
B and
and thus is reducing for
B. Denoting by
the closed linear span of all such subspaces, we conclude that
and
are reducing for
B and the operators
A and
B coincide on
. As a result, only the part of
B in
is of interest, and we may assume that
without loss of generality.
Under such an assumption, every eigenvalue
of
B is geometrically simple and the main results of [
12] can be summarised as follows:
- (a)
the algebraic multiplicity m of an eigenvalue coincides with the multiplicity l of as a zero of F;
- (b)
if , then the above multiplicities m and l satisfy the relation ;
- (c)
for any n-tuple of pairwise distinct complex numbers and any n-tuple of natural numbers, there exists a rank-one perturbation B of A such that every is an eigenvalue of B of algebraic multiplicity ;
- (d)
the eigenvalues of B can be enumerated as , , in such a way that as ; in particular, B has at most finitely many non-simple eigenvalues.
Property (c) means that locally the spectrum of
B can be arbitrary, while (d) describes the asymptotic behavior of the eigenvalues of
B at infinity. One of the main aims of this note is to provide a complete characterization of the possible spectra of
A under rank-one perturbations by refining the asymptotics of
, cf. Theorems 1 and 2. In view of (a) and (b), this task amounts to the study of zero distribution of the characteristic function
F of (
4), which will be done in
Section 3 and
Section 4.
3. Eigenvalue Distribution of the Operator
As we mentioned in the previous section, the eigenvalues of the operator B are determined by the characteristic function F which, in turn, is completely determined by the Fourier coefficients and of the vectors and through their products , . For typical applications we have in mind (with A a differential operator and the Dirac delta-function), the eigenfunctions are uniformly bounded and so are the Fourier coefficients . Therefore, we assume throughout the rest of the paper that
- (A3)
the Fourier coefficients of are uniformly bounded.
Under (A3), the sequence belongs to ; however, we need a more precise characterization of .
Definition 1. For φ as above, we denote by the set of sequences of the form with and the Fourier coefficients of φ in the basis .
Under assumption (A3), we have ; moreover, the inclusion is strict if some of are zero or if . The main result of this section establishes the asymptotic distribution of the eigenvalues of B in the following form.
Theorem 1. Under the above assumptions, the eigenvalues of the operator B can be labelled as , , in such a way that the sequence belongs to ; in particular,and all but finitely many eigenvalues of B are simple. We should point out the effect on the asymptotic distribution of eigenvalues that singularity of
makes: for regular
, the sequence of
was absolutely summable ([
13], Theorem 3.1), while here it is only square summable.
As explained in the previous section, the spectrum of
B is the union of two parts,
and
;
is the common part of the spectra of
A and
B, while
is the set of zeros of the characteristic function
in the domain
; moreover, the algebraic multiplicity of an eigenvalue
is determined by its multiplicity as a zero of the characteristic function
F.
First, we shall show that large enough elements of
are located near
, which will enable their proper enumeration. To begin with, for
we define the functions
and
by the formulae (here and hereafter, the symbol
will denote summation over the index set
)
and introduce the sets
where we replace
with
if
. Due to the assumption (A2), the sets
are pairwise disjoint and also
if
.
Lemma 1. The seriesconverges locally uniformly in and its sum is uniformly bounded therein. Proof. If
, then we choose
such that
if
, then we set
for convenience and define
as above provided
; otherwise, set
. Observing that
, we conclude that
if
; as a result, we derive the uniform bound
The proof is complete. □
Remark 1. We observe that the same statements on uniform convergence and uniform boundedness hold when the radius of the circles is replaced by an arbitrary ε or when z is taken within the union of but is omitted in the sum; these modifications will be used below.
Lemma 2. For every , there exist integers and such that the following holds:
- (a)
for every k with and every - (b)
for every
Proof. Since the sequence
belongs to
, for every
there exists a
K such that
In view of the Cauchy–Bunyakowsky–Schwarz inequality and Lemma 1, we find that
Part (a) follows by choosing
so that the above value is less than
and denoting by
the corresponding integer
K.
For part (b), note that
if
and
with
; therefore, by choosing
large enough, we arrive at (
10). □
Corollary 1. Take an arbitrary and fix as in the above lemma; thenIndeed, it suffices to note that if z does not belong to the above set, thenwhich together with part (b) of Lemma 2 shows thatso that such z cannot be an eigenvalue of B. Lemma 3. There exists a such that for all with the following holds:
- (a)
the function F has exactly one zero in ;
- (b)
the functions and F have the same number of zeros in .
Proof. Fix an ; we shall show that (a) and (b) hold for of Lemma 2.
If
k satisfies
, then by Lemma 2 for every
we get
On the other hand,
if
satisfies
and
, and then
for all
. By the choice of
we conclude that then
for all such
z. As the functions
and
F both have the same number of poles in
(namely, a simple pole at
), by estimate (
11) and Rouché’s theorem [
16] they have the same number of zeros in the set
. Since
for large enough
, the unique zero
of the function
belongs to the circle
for all
with
, and thus the function
F has exactly one zero in
for such
k as well. This completes the proof of part (a).
Next, by the definition of the set
, we see that
belongs to the set
; repeating the arguments used in the proof of part (a) of Lemma 2, we conclude that
and
if
and
. Also, by part (b) of Lemma 2 we have
as soon as
and
. Combining estimates (
12) and (
13), we conclude that
for all
k with
and all
. It follows that for
k with
and for all
Since the functions
and
F have the same poles in
(namely, simple poles
for
with
), we conclude by Rouché’s theorem that they have the same number of zeros in
for all
. The proof is complete. □
Remark 2. Take k larger than K of the above lemma and denote by the cardinality of the set . The function is a ratio of two polynomials of degree and due to (14) all its zeros are in . Therefore, the function F has precisely zeros in counting with multiplicities. Corollary 2. The zeros of F in can be labelled (counting with multiplicities) as with in such a way that for all with .
Recalling the results of the previous section on the relation between the eigenvalues of B and zeros of the function F in , we arrive at the following conclusion.
Corollary 3. Eigenvalues of the operator B can be labelled (counting with multiplicities) as with in such a way that when , K being the constant of Lemma 3.
Combining the above corollary with Lemma 4.3 of [
12], we conclude that
as
goes to infinity, cf. Theorem 4.7(ii) of [
12]. However, the estimates established above will enable us to prove a stronger statement of Theorem 1 on the asymptotics of
.
Proof of Theorem 1 We fix an enumeration of as in Corollary 3. Then for all with large enough , whence it suffices to prove that the differences for of sufficiently large absolute value are square summable.
We take
and
as in Lemma 2; then, according to Corollary 3, for every
with
the eigenvalue
is a zero of
F, so that
and
By virtue of Lemma 2 we conclude that
so that
for all
with
. Since the sequence
belongs to
, the proof is complete. □
4. Inverse Spectral Problem
The purpose of this section is two-fold. Firstly, we show that Theorem 1 gives not only necessary but also sufficient conditions on the eigenvalues of B. Secondly, we study the inverse spectral problem of reconstructing the operator B from its spectrum assuming that the operator A and the vector are known. As a by-product, we come up with the constructive algorithm of determining the vector in the rank-one perturbation from the given data—the operator A, the vector , and the spectrum of the rank-one perturbation B.
Thus we fix an operator
A satisfying the standing assumptions (A1) and (A2), i.e., is self-adjoint and has a simple discrete spectrum
that is
d-separated as in (
3). Assume further that
is a vector in the Hilbert space
for some
such that (A3) holds. Then the following statement holds true.
Theorem 2. Assume that a sequence of complex numbers can be enumerated as , , in such a way that the differences form an element of . Then there exists a vector such that the spectrum of the operator B coincides with counting with multiplicities.
Observe that the assumptions of the theorem imply that the series
converges and that
for every
such that
. More generally, denote by
the set of indices
for which
appears in
and set
; then every
for which
belongs to
. By virtue of (
16), for every
there exists a
such that
for all
with
. Therefore, if
and
, then
, and without loss of generality we may assume that
for all
.
We also set
,
, and introduce the function
We first show that
is well defined; to this end, we take an arbitrary
, introduce the sets
and prove the following result.
Lemma 4. For each , the product in (17) converges locally uniformly in to a function that is uniformly bounded in . Proof. It is enough to show that the series
converges locally uniformly on the set
. By the standard bound on
and the Cauchy–Bunyakovski–Schwarz inequality, we get
Convergence of the series (
16) along with Lemma 1 results in the locally uniform convergence of the product for
on the set
as well as in the uniform boundedness of
therein. □
Similar arguments based on the Lebesgue dominated convergence theorem justify passage to the limit in
as a result, we get
Corollary 4. The function tends to 1 along the imaginary line, i.e., We next develop
in the sum of simple fractions. The function
is meromorphic in
, and its residue at the point
is
the minus sign is introduced here for convenience. We also set
for
.
Lemma 5. Under the assumptions of Theorem 2, the sequence belongs to .
Proof. In view of (
19) and the assumption of Theorem 2, it suffices to prove that the sequence
is uniformly bounded in
.
Applying the same reasoning as in the proof of Lemma 1 (see also Remark 1), we conclude that the sum of the series
has an
n-independent bound, which implies that the sequence (
20) is uniformly bounded. The proof is complete.
Given the above lemma and the uniform bound established in the proof of Lemma 1, the series
converges locally uniformly in
for every
. It follows that the function
is well defined and analytic in the set
and has simple poles at the points
.
Lemma 6. For every , the function F is uniformly bounded in the set and, moreover, Proof. To prove the first statement, it suffices to apply the Cauchy–Bunyakowski–Schwarz inequality to the sum in (
21) and use Lemmas 1 and 5. Since for real
u we have
and since the series
converges, the Lebesgue dominated convergence theorem justifies term-wise passage to the limit in the series of (
21) and thus produces the required limit. □
Now we are ready to show that the functions F and coincide.
Lemma 7. The function is equal to zero identically in .
Proof. The function
is entire: indeed, it is meromorphic in
with possible single poles at the points of
, and since the residua of
F and
at the point
are equal to
, each such a singularity is removable. Being uniformly bounded over
by virtue of Lemmas 4 and 5,
G is constant by the Liouville theorem. Since
according to Corollary 4 and Lemma 6, this constant is zero. The proof is complete. □
Proof of Theorem 2 Given any sequence
of complex numbers satisfying the assumption of the theorem, we construct the meromorphic function
via (
17). Next, we calculate the residua
of
at the points
via (
19) and define the sequence
via
and
Since the sequence
belongs to
, it follows that the sequence
defined this way belongs to
. Therefore, there exists a vector
in the Hilbert space
whose Fourier coefficients in the basis
are equal to
.
We now consider the operator
B of the form (
2) with the given vector
and the vector
constructed above and conclude by virtue of Lemma 7 that the corresponding characteristic function
F of (
6) coincides with
. Therefore, zeros of
F are precisely the elements of the subsequence
, both counting multiplicity; namely, if a number
occurs
k times in
, it is a zero of
F of multiplicity
k. The analysis of the paper [
12] summarised in
Section 2 shows that each element
of
is an eigenvalue of
B and its multiplicity is equal to the number of times
is repeated in the sequence
. The proof is complete. □
The above proof also suggests an algorithm of constructing a particular operator
B whose spectrum corresponds to a sequence
of complex numbers satisfying (
16). Namely, given such a sequence
, we
- 1.
construct the product
of (
17);
- 2.
then calculate the residua of at the points ;
- 3.
construct the sequence
via (
22) and (
23);
- 4.
determine the function
from its Fourier coefficients
via (
5).
Corollary 5. The above analysis allows to completely describe the isospectral set
of all vectors for which the corresponding rank-one perturbations B of (2) have spectrum counting with multiplicity. Namely, all such ψ have fixed Fourier coefficients for , which are given by (22), while must be zero for those for which . Therefore, the “degree of freedom” in coincides with the number of zero Fourier coefficients of the vector φ; in particular, ψ is uniquely determined by if and only if all Fourier coefficients of φ are nonzero.