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

A practical method for recovering Sturm-Liouville problems from the Weyl function

Vladislav V. Kravchenko1, Sergii M. Torba1
1 Departamento de Matemáticas, Cinvestav, Unidad Querétaro,
Libramiento Norponiente #2000, Fracc. Real de Juriquilla, Querétaro, Qro., 76230 MEXICO.
e-mail: vkravchenko@math.cinvestav.edu.mx, storba@math.cinvestav.edu.mx
Abstract

In the paper we propose a direct method for recovering the Sturm-Liouville potential from the Weyl-Titchmarsh mm-function given on a countable set of points. We show that using the Fourier-Legendre series expansion of the transmutation operator integral kernel the problem reduces to an infinite linear system of equations, which is uniquely solvable if so is the original problem. The solution of this linear system allows one to reconstruct the characteristic determinant and hence to obtain the eigenvalues as its zeros and to compute the corresponding norming constants. As a result, the original inverse problem is transformed to an inverse problem with a given spectral density function, for which the direct method of solution from [28] is applied.

The proposed method leads to an efficient numerical algorithm for solving a variety of inverse problems. In particular, the problems in which two spectra or some parts of three or more spectra are given, the problems in which the eigenvalues depend on a variable boundary parameter (including spectral parameter dependent boundary conditions), problems with a partially known potential and partial inverse problems on quantum graphs.

1 Introduction

We consider the inverse problem of recovering the potential of a Sturm-Liouville equation on a finite interval together with the boundary conditions from values of the Weyl-Titchmarsh mm-function on a given set of points.

A fundamental result of Marchenko [30] states that the Weyl-Titchmarsh function m(z)m(z) determines uniquely the potential qL2(0,π)q\in L_{2}(0,\pi) as well as the boundary conditions. So some particular inverse problem is solved, at least in theory, if one finds a way to convert it to a corresponding problem for the mm-function. However, several difficulties arise. For some inverse spectral problems the given spectral data allow one to obtain the values of the mm-function at any point zz\in\mathbb{C}. This is the case of the classical inverse problem with a given spectral density function, formula (2.13) below provides the mm-function. While for many problems only the values of the mm-function on a countable set of points {zk}k=0\{z_{k}\}_{k=0}^{\infty} can be obtained. See Subsection 2.2 and references therein for some examples. Thus one faces two principle questions. Is it possible to recover uniquely the mm-function from the given values m(zk)m(z_{k}), k=0,1,k=0,1,\ldots? If yes, how can that be done?

Since the mm-function is meromorphic, the unique recovery is always possible if the sequence {zk}k=0\{z_{k}\}_{k=0}^{\infty} contains a subsequence znkz_{n_{k}} converging to a finite limit and such that the values m(znk)m(z_{n_{k}}) are bounded. This is usually not the case when we start with some inverse problem, the points zkz_{k} are real and go to infinity when kk\rightarrow\infty. Suppose all the points zkz_{k} are real. A simple sufficient condition for the unique recoverability of the mm-function from the given values m(zk)m(z_{k}) was given in [9], and a complete answer was given not so long ago in [20]. Simply speaking, the points zkz_{k} should be distributed sufficiently densely, like the points obtained as a union of two spectra.

The corresponding results are obtained by using infinite products and analytic continuation. And as it is stated in the following quote [7, p.106],

It is one of the guiding principles of computational analysis that such proofs often lead to elegant mathematics but are useless as far as constructive methods are concerned.

The only constructive method for finding the mm-function we are aware of is that from [4], [41] (based on the ideas of [34] and [7, Chapter 3]). This algorithm recovers a certain combination of the integral kernel of the transmutation operator with its derivative by using non-harmonic Fourier series. However it works only for sequences znz_{n} such that the sequence of functions {e±iznt}n\{e^{\pm i\sqrt{z_{n}}t}\}_{n\in\mathbb{N}} is a Riesz basis in L2(2π,2π)L_{2}(-2\pi,2\pi), requires additionally a beforehand knowledge of the parameter ω\omega (depending on the average of the potential qq and boundary parameter HH) and numerical realization of this method is expected to converge slowly (no numerical illustration is provided in [4] or [41]).

We use Fourier-Legendre series to represent the transmutation integral kernel and its derivative. This idea was originally proposed in [26] and resulted in a Neumann series of Bessel functions (NSBF) representation for the solutions and their derivatives of Sturm-Liouville equations. As a result, a constructive algorithm for recovering the mm-function from its values on a countable set of points is proposed. The problem is reduced to an infinite system of linear equations and the unique solvability of this system is proved. We would like to emphasize that no additional information is necessary beforehand.

Moreover, the proposed method leads to an efficient numerical algorithm. Since the NSBF representation possesses such a remarkable property as a uniform error bound for all real λ>0\lambda>0, one can obtain any finite set of approximate eigenvalues and corresponding norming constants with a non-deteriorating accuracy. That is, an efficient conversion of the original problem to an inverse problem by a given spectral density function is proposed. The method of solution introduced in [24], [25] and refined in [28] is adapted to recover the potential and the boundary conditions from the spectral density function.

For classical inverse spectral problems there are many results describing to what extent the potential can be recovered from a finite set of spectral data, see, e.g., [17], [32], [37], [35]. For the Weyl function, on the contrary, we are not aware of any single result. It can be seen that the values of zkz_{k}, 0kK0\leq k\leq K can be chosen neither from a too small interval nor too sparse. Indeed, taking all the values zkz_{k} from one spectrum is insufficient to recover the potential. On the other hand, as was mentioned in [7, Subsections 3.10.5 and 3.14], when all the values zkz_{k} are equal to the first eigenvalue λ0\lambda_{0} (for problems with different boundary conditions), the corresponding numerical problem is extremely ill-conditioned. So the proposed algorithm struggles as well when the points zkz_{k} either belong to a small interval or are too sparse. Nevertheless, it delivers excellent results for smooth potentials and sufficiently dense points zkz_{k}, does not require any additional information on the potential or boundary conditions and can be applied to a variety of inverse problems for which not so many (if any) specialized algorithms are known. So we believe the presented method will be a starting point for the future research of the question “to which extent a potential can be recovered from a given finite set of values of the Weyl-Titchmarsh mm-function”, as well as of a new class of numerical methods based on the Weyl-Titchmarsh mm-function.

The paper is organized as follows. In Section 2 we provide necessary information on the Weyl function, transmutation operators and Neumann series of Bessel functions representations. In Section 3 we show how the inverse problem of recovering the Sturm-Liouville problem from the Weyl function is reduced to an infinite system of linear algebraic equations, prove its unique solvability and show the reduction of the original problem to the classical problem of the recovery of a Sturm-Liouville problem from a given spectral density function; an algorithm is provided. Finally in Section 4 we present several numerical examples illustrating the efficiency and some limitations of the proposed method.

2 Preliminaries

2.1 The Weyl function and spectral data

Let qL2(0,π)q\in L_{2}(0,\pi) be real valued. Consider the Sturm-Liouville equation

y′′+q(x)y=ρ2y,x(0,π),-y^{\prime\prime}+q(x)y=\rho^{2}y,\quad x\in(0,\pi), (2.1)

where ρ\rho\in\mathbb{C}, with the boundary conditions

y(0)hy(0)=0,y(π)+Hy(π)=0,y^{\prime}(0)-hy(0)=0,\quad y^{\prime}(\pi)+Hy(\pi)=0, (2.2)

where hh and HH are arbitrary real constants, and the boundary value problem

y(0)hy(0)=1,y(π)+Hy(π)=0.y^{\prime}(0)-hy(0)=1,\quad y^{\prime}(\pi)+Hy(\pi)=0. (2.3)

If λ=ρ2\lambda=\rho^{2} is not an eigenvalue of the spectral problem (2.1), (2.2), then the boundary value problem (2.1), (2.3) possesses a unique solution, which we denote by Φ(ρ,x)\Phi(\rho,x). Let

M(λ):=Φ(ρ,0).M(\lambda):=\Phi(\rho,0).

The functions Φ(ρ,x)\Phi(\rho,x) and M(λ)M(\lambda) are called the Weyl solution and the Weyl function, respectively. See [42, Section 1.2.4] for additional details.

By φ(ρ,x)\varphi(\rho,x) and S(ρ,x)S(\rho,x) we denote the solutions of (2.1) satisfying the initial conditions

φ(ρ,0)=1andφ(ρ,0)=h\varphi(\rho,0)=1\quad\text{and}\quad\varphi^{\prime}(\rho,0)=h (2.4)

and

S(ρ,0)=0andS(ρ,0)=1,S(\rho,0)=0\quad\text{and}\quad S^{\prime}(\rho,0)=1, (2.5)

respectively. Denote additionally for λ=ρ2\lambda=\rho^{2}

Δ(λ):=φ(ρ,π)+Hφ(ρ,π)andΔ0(λ):=S(ρ,π)+HS(ρ,π).\Delta(\lambda):=\varphi^{\prime}(\rho,\pi)+H\varphi(\rho,\pi)\quad\text{and}\quad\Delta^{0}(\lambda):=S^{\prime}(\rho,\pi)+HS(\rho,\pi). (2.6)

Obviously, for all ρ\rho\in\mathbb{C} the function φ(ρ,x)\varphi(\rho,x) fulfills the first boundary condition, φ(ρ,0)hφ(ρ,0)=0\varphi^{\prime}(\rho,0)-h\varphi(\rho,0)=0, and thus, the spectrum of problem (2.1), (2.2) is the sequence of numbers {λn=ρn2}n=0\left\{\lambda_{n}=\rho_{n}^{2}\right\}_{n=0}^{\infty} such that

Δ(λn)=0.\Delta(\lambda_{n})=0.

It is easy to see that

Φ(ρ,x)=S(ρ,x)+M(λ)φ(ρ,x)\Phi(\rho,x)=S(\rho,x)+M(\lambda)\varphi(\rho,x) (2.7)

and

M(λ)=Δ0(λ)Δ(λ).M(\lambda)=-\frac{\Delta^{0}(\lambda)}{\Delta(\lambda)}. (2.8)

Moreover, the Weyl function is meromorphic with simple poles at λ=λn\lambda=\lambda_{n}, n0n\geq 0.

Remark 2.1.

There is another common definition of the Weyl-Titchmarsh mm-function [39], [10], [20]. Let v(ρ,x)v(\rho,x) be the solution of (2.1) satisfying

v(ρ,π)=1andv(ρ,π)=H.v(\rho,\pi)=1\quad\text{and}\quad v^{\prime}(\rho,\pi)=-H. (2.9)

Then the Weyl-Titchmarsh mm-function is defined as

m(ρ2)=v(ρ,0)v(ρ,0).m(\rho^{2})=\frac{v^{\prime}(\rho,0)}{v(\rho,0)}. (2.10)

One can verify that

m(ρ2)=hΔ(ρ2)Δ0(ρ2)=h+1M(ρ2),m(\rho^{2})=h-\frac{\Delta(\rho^{2})}{\Delta^{0}(\rho^{2})}=h+\frac{1}{M(\rho^{2})}, (2.11)

that is, the functions mm and MM can be easily transformed one into another.

Denote additionally

αn:=0πφ2(ρn,x)𝑑x.\alpha_{n}:=\int_{0}^{\pi}\varphi^{2}(\rho_{n},x)\,dx. (2.12)

The set {αn}n=0\left\{\alpha_{n}\right\}_{n=0}^{\infty} is referred to as the sequence of norming constants of problem (2.1), (2.2).

Theorem 2.2 (see [42, Theorem 1.2.6]).

The following representation holds

M(λ)=n=01αn(λλn).M(\lambda)=\sum_{n=0}^{\infty}\frac{1}{\alpha_{n}(\lambda-\lambda_{n})}. (2.13)

Let us formulate the inverse Sturm-Liouville problem.

Problem 2.3 (Recovery of a Sturm-Liouville problem from its Weyl function).

Given the Weyl function M(λ)M(\lambda), find a real valued qL2(0,π)q\in L_{2}(0,\pi), and the constants hh, HH\in\mathbb{R}, such that M(λ)M(\lambda) be the Weyl function of problem (2.1), (2.2).

We refer the reader to [42, Theorem 1.2.7] for the result on the unique solvability of Problem 2.3.

Having the Weyl function known completely, one can recover the sequence {λn}n=0\{\lambda_{n}\}_{n=0}^{\infty} as the sequence of its poles, and the sequence {αn}n=0\{\alpha_{n}\}_{n=0}^{\infty} can be immediately obtained from the corresponding residuals, the formula (1.2.14) from [42] states that

αn=1Resλ=λnM(λ)=1Δ0(λn)ddλΔ(λ)|λ=λn.\alpha_{n}=\frac{1}{\operatorname{Res}_{\lambda=\lambda_{n}}M(\lambda)}=-\frac{1}{\Delta^{0}(\lambda_{n})}\cdot\left.\frac{d}{d\lambda}\Delta(\lambda)\right|_{\lambda=\lambda_{n}}. (2.14)

Hence reducing the inverse problem to recovering the Sturm-Liouville problem by its spectral density function. Or, by taking the sequences of poles and zeros of the Weyl function one can reduce Problem 2.3 to recovering the Sturm-Liouville problem by two spectra, see [42, Remark 1.2.3].

However the situation is more complicated if one looks for a practical method of recovering qq, hh and HH from the Weyl function given only on a countable (even worse, finite) set of points from some (small) interval. The direct reduction to the inverse problems considered, in particular, in [28], is not possible. Let us formulate the corresponding problem and state some sufficient results for its unique solvability.

Problem 2.4 (Recovery of a Sturm-Liouville problem from its Weyl function given on a countable set of points).

Given the values Mn=M(zn)M_{n}=M(z_{n}), Mn{}M_{n}\in\mathbb{C}\cup\{\infty\}, of the Weyl function M(λ)M(\lambda) on a set of points {zn}n=0\{z_{n}\}_{n=0}^{\infty}, find a real valued qL2(0,π)q\in L_{2}(0,\pi), and the constants hh, HH\in\mathbb{R}, such that the Weyl function of problem (2.1), (2.2) takes the values MnM_{n} at the points znz_{n}, n0n\geq 0.

It is easy to see that Problem 2.4 is not always uniquely solvable. The following result shows that choosing the set of points znz_{n} “sufficiently dense” can guarantee the unique solvability.

Theorem 2.5 ([9, Theorem 2.1]).

Let a+=max(a,0)a_{+}=\max(a,0) for aa\in\mathbb{R}. Let {zn}n=0\{z_{n}\}_{n=0}^{\infty} be a sequence of distinct positive real numbers satisfying

n=1(zn14n2)+n2<.\sum_{n=1}^{\infty}\frac{\left(z_{n}-\frac{1}{4}n^{2}\right)_{+}}{n^{2}}<\infty. (2.15)

Let m1m_{1} and m2m_{2} be Weyl-Titchmarsh mm-functions for two Sturm-Liouville problems having the potentials q1q_{1} and q2q_{2} and the constants H1H_{1} and H2H_{2}, correspondingly. Suppose that m1(zn)=m2(zn)m_{1}(z_{n})=m_{2}(z_{n}) for all n0n\in\mathbb{N}_{0}. Then m1=m2m_{1}=m_{2} (and hence q1=q2q_{1}=q_{2} a.e. on [0,π][0,\pi] and H1=H2H_{1}=H_{2}).

See also [20, Theorem 1.5] for the necessary and sufficient result.

2.2 Some inverse problems which can be reduced to Problem 2.4

Below we briefly present some inverse problems, omitting any additional details like the conditions on a spectra ensuring the uniqueness of the solution. They can be consulted in the provided references.

2.2.1 Two spectra

Consider the second set of boundary conditions

y(0)=0,y(π)+Hy(π)=0.y(0)=0,\quad y^{\prime}(\pi)+Hy(\pi)=0. (2.16)

Let us denote the spectrum of problem (2.1), (2.16) as {νn=μn2}n=0\left\{\nu_{n}=\mu_{n}^{2}\right\}_{n=0}^{\infty}.

The inverse problem consists in recovering qq, hh and HH from given two spectra, {λn}n=0\{\lambda_{n}\}_{n=0}^{\infty} and {νn}n=0\{\nu_{n}\}_{n=0}^{\infty}. Note that this problem coincides with the one considered in [28] if one changes xx by πx\pi-x and hHh\leftrightarrow H. Now the equalities

M(λn)=andM(νn)=0,M(\lambda_{n})=\infty\quad\text{and}\quad M(\nu_{n})=0,

can be easily checked and give the reduction of the two spectra inverse problem to Problem 2.4.

2.2.2 Eigenvalues depending on a variable boundary parameter

There is a class of inverse problems where it is asked to recover the potential from given parts of three spectra [9], four spectra [33], NN spectra [19]. All these problems can be regarded as special cases of the following inverse problem. Suppose the sequences of numbers {λ~n}n=0\{\tilde{\lambda}_{n}\}_{n=0}^{\infty} and {hn}n=0\{h_{n}\}_{n=0}^{\infty} are given and such that λ~n\tilde{\lambda}_{n} is an eigenvalue (does not refer to nn-th indexed eigenvalue) of the problem (2.1) with the boundary conditions

y(0)hny(0)=0,y(π)+Hy(π)=0.y^{\prime}(0)-h_{n}y(0)=0,\quad y^{\prime}(\pi)+Hy(\pi)=0. (2.17)

The problem is to recover the potential qq. We refer the reader to [7, §3.10.5 and §3.15] for some discussion and numerical results.

The reduction to Problem 2.4 follows immediately when one observes that (2.17) leads to

m(λ~n)=hn.m(\tilde{\lambda}_{n})=h_{n}.

2.2.3 Partially known potential

The Hochstadt-Lieberman inverse problem from [18] reads as follows: suppose we are given the spectrum of the problem (2.1), (2.2) and the potential qq is known on the segment [0,π/2][0,\pi/2]. Recover qq on the whole segment [0,π][0,\pi]. See also [11], [9] for the case when the potential is known on a different portion of the segment. If the potential is known on less than half of the segment, more than one spectrum is necessary to recover the potential uniquely. For that reason the general formulation of the problem is as follows. Suppose a(0,π)a\in(0,\pi) and the potential qq is known on [0,a][0,a]. Moreover, two sequences of numbers are given as in Subsubsection 2.2.2. Recover qq on the whole [0,π][0,\pi].

Let m(x,ρ2)m(x,\rho^{2}) define the mm-function on [x,π][x,\pi] with the same boundary condition (2.9), i.e.,

m(x,ρ2)=v(ρ,x)v(ρ,x).m(x,\rho^{2})=\frac{v^{\prime}(\rho,x)}{v(\rho,x)}.

Consider an eigenvalue λ~n=ρn2\tilde{\lambda}_{n}=\rho_{n}^{2} and the corresponding boundary constant hnh_{n}. Then

m(0,λ~n)=m(λ~n)=hn,m(0,\tilde{\lambda}_{n})=m(\tilde{\lambda}_{n})=h_{n},

hence we may consider the following Cauchy problem for equation (2.1)

v(ρn,0)=1,v(ρn,0)=hn.v(\rho_{n},0)=1,\quad v^{\prime}(\rho_{n},0)=h_{n}.

Solving it on [0,a][0,a] (in practice it can be done efficiently for a large set of λ~n\tilde{\lambda}_{n} using the method developed in [26]), we obtain values v(ρn,a)v(\rho_{n},a) and v(ρn,a)v^{\prime}(\rho_{n},a), hence, m(a,λ~n)m(a,\tilde{\lambda}_{n}). And by rescaling we reduce the problem of a partially known potential to Problem 2.4.

2.2.4 Problems with analytic functions in the boundary condition

The following spectral parameter dependent boundary condition was considered in [4], [5], [41]

f1(λ)y(0)+f2(λ)y(0)=0,f_{1}(\lambda)y^{\prime}(0)+f_{2}(\lambda)y(0)=0, (2.18)

where f1f_{1} and f2f_{2} are entire functions, not vanishing simultaneously. The second boundary condition is the same as in (2.2). The functions f1f_{1} and f2f_{2} are supposed to be known, and an inverse problem consists in recovering the potential qq and the constant HH by the given parameter ω2\omega_{2} (see (2.39)), spectra λn\lambda_{n} and corresponding multiplicities mnm_{n}. See also [38] for a general spectral theory of such problems.

Supposing for simplicity that all the eigenvalues are simple, i.e., mn=1m_{n}=1 for all nn, the problem immediately reduces to Problem 2.4, since

m(λn)=f2(λn)f1(λn).m(\lambda_{n})=-\frac{f_{2}(\lambda_{n})}{f_{1}(\lambda_{n})}.

Many inverse problems can be reduced to a problem having boundary condition (2.18). In particular, the Hochstadt-Lieberman problem, the inverse transmission eigenvalue problem, partial inverse problems for quantum graphs. See [41] and references therein for more details.

2.3 Some practical applications

Inverse Sturm-Liouville problems arise in numerous applied field. For some examples in vibration models we refer to the book [14], for some biomedical engineering applications to the review [15], while here we add two more examples from different areas of physics.

In studying quantum physical properties of quantum dot nanostructures the problem of recovering symmetric potentials from a finite number of experimentally obtained eigenvalues is of considerable importance (see, e.g., [23]). Mathematically this means that the potential qq in (2.1) is symmetric: q(x)=q(πx)q(x)=q(\pi-x), and a finite part of one spectrum is given. For simplicity let us suppose that it corresponds to the Dirichlet conditions. Then, as it is well known (see [7, p. 103]), this problem reduces to the two spectra problem on the interval (0,π/2)(0,\pi/2). Indeed, for the eigenfunctions we have yn(x)=yn(πx)y_{n}(x)=y_{n}(\pi-x) if nn is odd and yn(x)=yn(πx)y_{n}(x)=-y_{n}(\pi-x) if nn is even. Thus, yn(0)=0y_{n}(0)=0, yn(π/2)=0y_{n}(\pi/2)=0 if nn is even and yn(0)=0y_{n}(0)=0, yn(π/2)=0y_{n}^{\prime}(\pi/2)=0 if nn is odd. This means that the original spectrum is split into two parts – the even-numbered eigenvalues forming the Dirichlet-Dirichlet eigenvalues on (0,π/2)(0,\pi/2) and the odd-numbered ones, the Dirichlet-Neumann eigenvalues on (0,π/2)(0,\pi/2). Thus the problem is reduced to the two spectra inverse problem, and symmetry of qq allows one to compute it over the entire interval (0,π)(0,\pi).

In order to proceed with the second example, following [13], we notice that besides the standard form (2.1) of the Sturm-Liouville equation, which is often the most convenient for analysis, two other forms

(a2(x)u(x))+λa2(x)u(x)=0\left(a^{2}(x)u^{\prime}(x)\right)^{\prime}+\lambda a^{2}(x)u(x)=0 (2.19)

and

u′′(x)+λr(x)u(x)=0u^{\prime\prime}(x)+\lambda r(x)u(x)=0 (2.20)

frequently arise in applications. When the functions aa, rr, qq are sufficiently smooth, each equation may be transformed into any of the other two. Equation (2.19) governs the modes of vibration of a thin straight rod in longitudinal or torsional vibration. It is often called the Sturm-Liouville equation in impedance form. Equation (2.20) models the transverse vibration of a taut string with mass density r(x)r(x).

Let us consider the problem of recovering the cross-sectional form F(x)F(x) of a rod with constant density rr and Young’s modulus EE (see [40, p.72]). The differential equation and the boundary conditions have the form

(EF(x)u(x))+rF(x)ω2u(x)=0,\displaystyle\left(EF(x)u^{\prime}(x)\right)^{\prime}+rF(x)\omega^{2}u(x)=0, (2.21)
u(0)=pEF(0),u(π)=0,\displaystyle u^{\prime}(0)=-\frac{p}{EF(0)},\qquad u^{\prime}(\pi)=0, (2.22)

where pp is a given constant. The values of FF at the end points F(0)F(0) and F(π)F(\pi) are supposed to be given, and the function F(x)F(x), x(0,π)x\in(0,\pi) needs to be recovered from the additional information on the solution:

u(ω,0)=f~(ω),ω[ω1,ω2].u(\omega,0)=\widetilde{f}(\omega),\qquad\omega\in\left[\omega_{1},\omega_{2}\right]. (2.23)

Note that for simplicity we substituted the Dirichlet condition at π\pi (in [40, p.72]) by the Neumann condition. Clearly, the problem with the Dirichlet condition can be considered similarly.

First of all, equation (2.21) can be written in the form (2.19) with a2=Fa^{2}=F and λ=rω2/E\lambda=r\omega^{2}/E. Next, for the function y(x)=u(x)/a(x)y(x)=u(x)/a(x) the problem (2.21), (2.22), (2.23) takes the form

y′′+q(x)y=ρ2y,x(0,π),\displaystyle-y^{\prime\prime}+q(x)y=\rho^{2}y,\qquad x\in(0,\pi),
y(ρ,0)hy(ρ,0)=c,y(ρ,π)+Hy(ρ,π)=0,\displaystyle y^{\prime}(\rho,0)-hy(\rho,0)=c,\qquad y^{\prime}(\rho,\pi)+Hy(\rho,\pi)=0, (2.24)
y(ρ,0)=f(ρ),ρ[ρ1,ρ2],\displaystyle y(\rho,0)=f(\rho),\qquad\rho\in\left[\rho_{1},\rho_{2}\right],

where h=a(0)/a(0)h=-a^{\prime}(0)/a(0), c=p/(EF3/2(0))c=-p/(EF^{3/2}(0)), H=a(π)/a(π)H=a^{\prime}(\pi)/a(\pi) and f(ρ)=f~(ω)/a(0)f(\rho)=\widetilde{f}(\omega)/a(0). The boundary conditions (2.24) indicate that y(ρ,x)=cΦ(ρ,x)y(\rho,x)=c\Phi(\rho,x). Thus, the original problem reduces to the problem of recovering the potential qq and the constants hh, HH from the Weyl function

M(λ)=f(ρ)cM(\lambda)=\frac{f(\rho)}{c}

given on the segment [ρ1,ρ2]\left[\rho_{1},\rho_{2}\right].

2.4 Integral representation of solutions via the transmutation operator

The solutions φ(ρ,x)\varphi(\rho,x) and S(ρ,x)S(\rho,x) admit the integral representations (see, e.g., [8], [29], [30], [31, Chapter 1], [36], [6])

φ(ρ,x)\displaystyle\varphi(\rho,x) =cosρx+xxK(x,t)cosρtdt,\displaystyle=\cos\rho x+\int_{-x}^{x}K(x,t)\cos\rho t\,dt, (2.25)
S(ρ,x)\displaystyle S(\rho,x) =sinρxρ+xxK(x,t)sinρtρ𝑑t,\displaystyle=\frac{\sin\rho x}{\rho}+\int_{-x}^{x}K(x,t)\frac{\sin\rho t}{\rho}\,dt, (2.26)

where the integral kernel KK is a continuous function of both arguments in the domain 0|t|xπ0\leq|t|\leq x\leq\pi and satisfies the equalities

K(x,x)=h2+120xq(t)𝑑tandK(x,x)=h2.K(x,x)=\frac{h}{2}+\frac{1}{2}\int_{0}^{x}q(t)\,dt\quad\text{and}\quad K(x,-x)=\frac{h}{2}. (2.27)

It is of crucial importance that K(x,t)K(x,t) is independent of ρ\rho.

The following result from [26] will be used.

Theorem 2.6 ([26]).

The integral transmutation kernel K(x,t)K(x,t) and its derivative K1(x,t):=xK(x,t)K_{1}(x,t):=\frac{\partial}{\partial x}K(x,t) admit the following Fourier-Legendre series representations

K(x,t)\displaystyle K(x,t) =n=0βn(x)xPn(tx),\displaystyle=\sum_{n=0}^{\infty}\frac{\beta_{n}(x)}{x}P_{n}\left(\frac{t}{x}\right), (2.28)
K1(x,t)\displaystyle K_{1}(x,t) =n=0γn(x)xPn(tx),0<|t|xπ,\displaystyle=\sum_{n=0}^{\infty}\frac{\gamma_{n}(x)}{x}P_{n}\left(\frac{t}{x}\right),\quad 0<|t|\leq x\leq\pi, (2.29)

where PkP_{k} stands for the Legendre polynomial of order kk.

For every x(0,π]x\in(0,\pi] the series converge in the norm of L2(x,x)L_{2}(-x,x). The first coefficients β0(x)\beta_{0}(x) and γ0(x)\gamma_{0}(x) have the form

β0(x)=φ(0,x)12,γ0(x)=β0(x)h2140xq(s)𝑑s,\beta_{0}(x)=\frac{\varphi(0,x)-1}{2},\quad\gamma_{0}(x)=\beta_{0}^{\prime}(x)-\frac{h}{2}-\frac{1}{4}\int_{0}^{x}q(s)\,ds, (2.30)

and the rest of the coefficients can be calculated following a simple recurrent integration procedure.

Corollary 2.7.

Equality (2.30) shows that the potential q(x)q(x) can be recovered from β0(x)\beta_{0}(x). Indeed,

q(x)=φ′′(0,x)φ(0,x)=2β0′′(x)2β0(x)+1.q(x)=\frac{\varphi^{\prime\prime}(0,x)}{\varphi(0,x)}=\frac{2\beta_{0}^{\prime\prime}(x)}{2\beta_{0}(x)+1}. (2.31)

Moreover, the constant hh is also recovered directly from β0(x)\beta_{0}(x), since

h=2β0(0).h=2\beta_{0}^{\prime}(0).

2.5 Neumann series of Bessel functions representations for solutions and their derivatives

The following series representations for the solutions φ(ρ,x)\varphi(\rho,x) and S(ρ,x)S(\rho,x) and for their derivatives were obtained in [26].

Theorem 2.8 ([26]).

The solutions φ(ρ,x)\varphi(\rho,x) and S(ρ,x)S(\rho,x) and their derivatives with respect to xx admit the following series representations

φ(ρ,x)\displaystyle\varphi(\rho,x) =cosρx+2n=0(1)nβ2n(x)j2n(ρx),\displaystyle=\cos\rho x+2\sum_{n=0}^{\infty}(-1)^{n}\beta_{2n}(x)j_{2n}(\rho x), (2.32)
S(ρ,x)\displaystyle S(\rho,x) =sinρxρ+2ρn=0(1)nβ2n+1(x)j2n+1(ρx),\displaystyle=\frac{\sin\rho x}{\rho}+\frac{2}{\rho}\sum_{n=0}^{\infty}(-1)^{n}\beta_{2n+1}(x)j_{2n+1}(\rho x), (2.33)
φ(ρ,x)\displaystyle\varphi^{\prime}(\rho,x) =ρsinρx+(h+120xq(s)𝑑s)cosρx+2n=0(1)nγ2n(x)j2n(ρx),\displaystyle=-\rho\sin\rho x+\left(h+\frac{1}{2}\int_{0}^{x}q(s)\,ds\right)\cos\rho x+2\sum_{n=0}^{\infty}(-1)^{n}\gamma_{2n}(x)j_{2n}(\rho x), (2.34)
S(ρ,x)\displaystyle S^{\prime}(\rho,x) =cosρx+12ρ(0xq(s)𝑑s)sinρx+2ρn=0(1)nγ2n+1(x)j2n+1(ρx),\displaystyle=\cos\rho x+\frac{1}{2\rho}\left(\int_{0}^{x}q(s)\,ds\right)\sin\rho x+\frac{2}{\rho}\sum_{n=0}^{\infty}(-1)^{n}\gamma_{2n+1}(x)j_{2n+1}(\rho x), (2.35)

where jk(z)j_{k}(z) stands for the spherical Bessel function of order kk (see, e.g., [1]). The coefficients βn(x)\beta_{n}(x) and γn(x)\gamma_{n}(x) are those from Theorem 2.6. For every ρ\rho\in\mathbb{C} all the series converge pointwise. For every x[0,π]x\in\left[0,\pi\right] the series converge uniformly on any compact set of the complex plane of the variable ρ\rho, and the remainders of their partial sums admit estimates independent of Reρ\operatorname{Re}\rho.

Moreover, the representations can be differentiated with respect to ρ\rho resulting in

φρ(ρ,x)\displaystyle\varphi^{\prime}_{\rho}(\rho,x) =xsinρx+2n=0(1)nβ2n(x)(2nρj2n(ρx)xj2n+1(ρx)),\displaystyle=-x\sin\rho x+2\sum_{n=0}^{\infty}(-1)^{n}\beta_{2n}(x)\left(\frac{2n}{\rho}j_{2n}(\rho x)-xj_{2n+1}(\rho x)\right), (2.36)
φx,ρ′′(ρ,x)=(1+hx+x20xq(s)𝑑s)sinρxρxcosρx+2n=0(1)nγ2n(x)(2nρj2n(ρx)xj2n+1(ρx)).\displaystyle\begin{split}\varphi^{\prime\prime}_{x,\rho}(\rho,x)&=-\left(1+hx+\frac{x}{2}\int_{0}^{x}q(s)\,ds\right)\sin\rho x-\rho x\cos\rho x\\ &\quad+2\sum_{n=0}^{\infty}(-1)^{n}\gamma_{2n}(x)\left(\frac{2n}{\rho}j_{2n}(\rho x)-xj_{2n+1}(\rho x)\right).\end{split} (2.37)

We refer the reader to [26] for the proof and additional details related to this result. The coefficients βn\beta_{n} and γn\gamma_{n} decay when nn\rightarrow\infty, and the decay rate is faster for smoother potentials. Namely, the following result is valid.

Proposition 2.9 ([27]).

Let qWp[0,π]q\in W_{\infty}^{p}[0,\pi] for some p0p\in\mathbb{N}_{0}, i.e., the potential qq is pp times differentiable with the last derivative being bounded on [0,π][0,\pi]. Then there exist constants cc and dd, independent of NN, such that

|βN(x)|cxp+2(N1)p+1/2and|γN(x)|dxp+1(N1)p1/2,Np+1.|\beta_{N}(x)|\leq\frac{cx^{p+2}}{(N-1)^{p+1/2}}\qquad\text{and}\qquad|\gamma_{N}(x)|\leq\frac{dx^{p+1}}{(N-1)^{p-1/2}},\qquad N\geq p+1.

Let us denote

hn:=γn(π)+Hβn(π),n0h_{n}:=\gamma_{n}(\pi)+H\beta_{n}(\pi),\qquad n\geq 0 (2.38)

and

ω:=h+H+120πq(t)𝑑tandω2:=ωh=H+120πq(t)𝑑t.\omega:=h+H+\frac{1}{2}\int_{0}^{\pi}q(t)\,dt\quad\text{and}\quad\omega_{2}:=\omega-h=H+\frac{1}{2}\int_{0}^{\pi}q(t)\,dt. (2.39)
Corollary 2.10.

The following representations hold for the characteristic functions Δ\Delta and Δ0\Delta^{0}

Δ(ρ2)\displaystyle\Delta(\rho^{2}) =ρsinρπ+ωcosρπ+2n=0(1)nh2nj2n(ρπ),\displaystyle=-\rho\sin\rho\pi+\omega\cos\rho\pi+2\sum_{n=0}^{\infty}(-1)^{n}h_{2n}j_{2n}(\rho\pi), (2.40)
Δ0(ρ2)\displaystyle\Delta^{0}(\rho^{2}) =cosρπ+ω2sinρπρ+2ρn=0(1)nh2n+1j2n+1(ρπ),\displaystyle=\cos\rho\pi+\omega_{2}\frac{\sin\rho\pi}{\rho}+\frac{2}{\rho}\sum_{n=0}^{\infty}(-1)^{n}h_{2n+1}j_{2n+1}(\rho\pi), (2.41)

and for the derivative with respect to ρ\rho

ddρΔ(ρ2)=(1+πω)sinρππρcosρπ+2n=0(1)nh2n(2nρj2n(ρπ)πj2n+1(ρπ)).\frac{d}{d\rho}\Delta(\rho^{2})=-(1+\pi\omega)\sin\rho\pi-\pi\rho\cos\rho\pi+2\sum_{n=0}^{\infty}(-1)^{n}h_{2n}\left(\frac{2n}{\rho}j_{2n}(\rho\pi)-\pi j_{2n+1}(\rho\pi)\right). (2.42)

2.6 The Gelfand-Levitan equation

Let

G(x,t):=K(x,t)+K(x,t)=n=02β2n(x)xP2n(tx)G(x,t):=K(x,t)+K(x,-t)=\sum_{n=0}^{\infty}\frac{2\beta_{2n}(x)}{x}P_{2n}\left(\frac{t}{x}\right)

and let

F(x,t)=n=0(cosρnxcosρntαncosnxcosntαn0),0t,x<πF(x,t)=\sum_{n=0}^{\infty}\left(\frac{\cos\rho_{n}x\cos\rho_{n}t}{\alpha_{n}}-\frac{\cos nx\cos nt}{\alpha_{n}^{0}}\right),\quad 0\leq t,\,x<\pi (2.43)

where

αn0={π/2,n>0,π,n=0.\alpha_{n}^{0}=\begin{cases}\pi/2,&n>0,\\ \pi,&n=0.\end{cases}

Then the function GG is the unique solution of the Gelfand-Levitan equation, see, e.g., [42, Theorem 1.3.1]

G(x,t)+F(x,t)+0xF(t,s)G(x,s)𝑑s=0,0<t<x<π.G(x,t)+F(x,t)+\int_{0}^{x}F(t,s)G(x,s)\,ds=0,\quad 0<t<x<\pi. (2.44)

The series (2.43) converges slowly and possesses a jump discontinuity (for ω0\omega\neq 0) at x=t=πx=t=\pi. To overcome these difficulties, in [22] another representation for the function FF was derived. Namely,

F(x,t)=n=1(cosρnxcosρntαncosnxcosntαn0+2ωπ2n(xsinnxcosnt+tsinntcosnx))+cosρ0xcosρ0tα01πωπ2(πmax{x,t}x2t2).\begin{split}F(x,t)&=\sum_{n=1}^{\infty}\left(\frac{\cos\rho_{n}x\,\cos\rho_{n}t}{\alpha_{n}}-\frac{\cos nx\,\cos nt}{\alpha_{n}^{0}}+\frac{2\omega}{\pi^{2}n}\Bigl{(}x\sin nx\,\cos nt+t\sin nt\,\cos nx\Bigr{)}\right)\\ &\quad+\frac{\cos\rho_{0}x\,\cos\rho_{0}t}{\alpha_{0}}-\frac{1}{\pi}-\frac{\omega}{\pi^{2}}\bigl{(}\pi\max\{x,t\}-x^{2}-t^{2}\bigr{)}.\end{split} (2.45)

3 Method of solution of Problem 2.4

3.1 Recovery of the parameters ω\omega and ω2\omega_{2} and of the characteristic functions Δ\Delta and Δ0\Delta^{0}

Let us rewrite (2.8) as

M(ρ2)Δ(ρ2)+Δ0(ρ2)=0M(\rho^{2})\Delta(\rho^{2})+\Delta^{0}(\rho^{2})=0

and substitute (2.40) and (2.41) into the last expression. We obtain

M(ρ2)(ρsinρπ+ωcosρπ+2n=0(1)nh2nj2n(ρπ))+cosρπ+ω2sinρπρ+2ρn=0(1)nh2n+1j2n+1(ρπ)=0,\begin{split}M(\rho^{2})&\left(-\rho\sin\rho\pi+\omega\cos\rho\pi+2\sum_{n=0}^{\infty}(-1)^{n}h_{2n}j_{2n}(\rho\pi)\right)\\ &+\cos\rho\pi+\omega_{2}\frac{\sin\rho\pi}{\rho}+\frac{2}{\rho}\sum_{n=0}^{\infty}(-1)^{n}h_{2n+1}j_{2n+1}(\rho\pi)=0,\end{split}

or

ωM(ρ2)cosρπ+ω2sinρπρ+2n=0h2n(1)nM(ρ2)j2n(ρπ)+2n=0h2n+1(1)nj2n+1(ρπ)ρ=M(ρ2)ρsinρπcosρπ.\begin{split}\omega\cdot M(\rho^{2})\cos\rho\pi&+\omega_{2}\cdot\frac{\sin\rho\pi}{\rho}+2\sum_{n=0}^{\infty}h_{2n}\cdot(-1)^{n}M(\rho^{2})j_{2n}(\rho\pi)+2\sum_{n=0}^{\infty}h_{2n+1}\cdot\frac{(-1)^{n}j_{2n+1}(\rho\pi)}{\rho}\\ &=M(\rho^{2})\rho\sin\rho\pi-\cos\rho\pi.\end{split} (3.1)

For the case when M(ρ2)=M(\rho^{2})=\infty, the last equality reduces to

ωcosρπ+2n=0h2n(1)nj2n(ρπ)=ρsinρπ.\omega\cos\rho\pi+2\sum_{n=0}^{\infty}h_{2n}\cdot(-1)^{n}j_{2n}(\rho\pi)=\rho\sin\rho\pi. (3.2)

Suppose that the Weyl function is known on a countable set of points {zk}k=0={ρ~k2}k=0\{z_{k}\}_{k=0}^{\infty}=\{\tilde{\rho}_{k}^{2}\}_{k=0}^{\infty}, that is, let the values Mk:=M(zk)M_{k}:=M(z_{k}), Mk{}M_{k}\in\mathbb{R}\cup\{\infty\} be given. Then considering the equality (3.1) for all ρ=ρ~k\rho=\tilde{\rho}_{k} we obtain an infinite system of linear algebraic equations for the unknown ω\omega, ω2\omega_{2} and {hn}n=0\{h_{n}\}_{n=0}^{\infty}.

One can easily see from (2.28) and (2.29) that the functions 2πhn2n+1\frac{\sqrt{2\pi}h_{n}}{\sqrt{2n+1}} are the Fourier coefficients of the square integrable function K1(π,t)+HK(π,t)K_{1}(\pi,t)+HK(\pi,t) in the space L2(π,π)L_{2}(-\pi,\pi) with respect to the orthonormal basis

{pn(t)}n=0:={2n+1Pn(t/π)2π}n=0.\left\{p_{n}(t)\right\}_{n=0}^{\infty}:=\left\{\frac{\sqrt{2n+1}P_{n}(t/\pi)}{\sqrt{2\pi}}\right\}_{n=0}^{\infty}. (3.3)

Let us introduce new unknowns

ξn=2πhn2n+1,n0.\xi_{n}=\frac{\sqrt{2\pi}h_{n}}{\sqrt{2n+1}},\qquad n\geq 0. (3.4)

Then {ξn}n=02\{\xi_{n}\}_{n=0}^{\infty}\in\ell_{2} (as a sequence of the Fourier coefficients). Let us rewrite the infinite linear system of equations as

n=0ξ2n(1)n4n+2Mkj2n(ρ~kπ)π+n=0ξ2n+1(1)n4n+2j2n+1(ρ~kπ)πρ~k+ωMkcosρ~kπ+ω2sinρ~kπρ~k=Mkρ~nsinρ~kπcosρ~kπ,k0\begin{split}&\sum_{n=0}^{\infty}\xi_{2n}\cdot\frac{(-1)^{n}\sqrt{4n+2}M_{k}j_{2n}(\tilde{\rho}_{k}\pi)}{\sqrt{\pi}}+\sum_{n=0}^{\infty}\xi_{2n+1}\cdot\frac{(-1)^{n}\sqrt{4n+2}j_{2n+1}(\tilde{\rho}_{k}\pi)}{\sqrt{\pi}\tilde{\rho}_{k}}\\ &+\omega\cdot M_{k}\cos\tilde{\rho}_{k}\pi+\omega_{2}\cdot\frac{\sin\tilde{\rho}_{k}\pi}{\tilde{\rho}_{k}}=M_{k}\tilde{\rho}_{n}\sin\tilde{\rho}_{k}\pi-\cos\tilde{\rho}_{k}\pi,\qquad k\in\mathbb{N}_{0}\end{split} (3.5)

(equation

n=0ξ2n(1)n4n+2j2n(ρ~kπ)π+ωcosρ~kπ=ρ~ksinρ~kπ\sum_{n=0}^{\infty}\xi_{2n}\cdot\frac{(-1)^{n}\sqrt{4n+2}j_{2n}(\tilde{\rho}_{k}\pi)}{\sqrt{\pi}}+\omega\cdot\cos\tilde{\rho}_{k}\pi=\tilde{\rho}_{k}\sin\tilde{\rho}_{k}\pi (3.6)

should be used if for some kk we have Mk=M_{k}=\infty).

Theorem 3.1.

Suppose that the original Problem 2.4 is solvable and the numbers {zk}k=0\{z_{k}\}_{k=0}^{\infty} satisfy the condition (2.15). Then the infinite system (3.5) possesses a unique 2\ell_{2} solution {ω,ω2,{ξn}n=0}\left\{\omega,\omega_{2},\{\xi_{n}\}_{n=0}^{\infty}\right\}.

Proof.

Since the original Problem 2.4 is solvable, we can recover the Weyl funcion MM and by [42, Theorem 1.2.7] recover the potential qq and the constants hh and HH. From (2.39) we obtain ω\omega and ω2\omega_{2}, and by Theorem 2.6 we obtain the square integrable with respect to the second variable integral kernel KK and its derivative K1K_{1}, as well as the series representations (2.28), (2.29). The coefficients βn\beta_{n} and γn\gamma_{n} give us the sequence {ξn}n=02\{\xi_{n}\}_{n=0}^{\infty}\in\ell_{2}. One can verify using (2.40) and (2.41) that the obtained numbers ω\omega, ω2\omega_{2} and ξn\xi_{n}, n0n\geq 0 satisfy the system (3.5). Now suppose that there is another 2\ell_{2} solution {ω~,ω~2,{ξ~n}n=0}\left\{\tilde{\omega},\tilde{\omega}_{2},\{\tilde{\xi}_{n}\}_{n=0}^{\infty}\right\} of the system (3.5). Let h~n=2n+12πξ~n\tilde{h}_{n}=\frac{\sqrt{2n+1}}{\sqrt{2\pi}}\tilde{\xi}_{n}, n0n\geq 0. Consider the following functions

G~(t)\displaystyle\tilde{G}(t) =2n=0h~2nπP2n(tπ),\displaystyle=2\sum_{n=0}^{\infty}\frac{\tilde{h}_{2n}}{\pi}P_{2n}\left(\frac{t}{\pi}\right), (3.7)
S~(t)\displaystyle\tilde{S}(t) =2n=0h~2n+1πP2n+1(tπ),t[0,π].\displaystyle=2\sum_{n=0}^{\infty}\frac{\tilde{h}_{2n+1}}{\pi}P_{2n+1}\left(\frac{t}{\pi}\right),\qquad t\in[0,\pi]. (3.8)

Due to condition {ξ~n}n=02\{\tilde{\xi}_{n}\}_{n=0}^{\infty}\in\ell_{2} one has G~L2(0,π)\tilde{G}\in L_{2}(0,\pi) and S~L2(0,π)\tilde{S}\in L_{2}(0,\pi), hence the following functions

Δ~(λ):=Δ~(ρ2)\displaystyle\tilde{\Delta}(\lambda):=\tilde{\Delta}(\rho^{2}) =ρsinρπ+ω~cosρπ+0πG~(t)cosρtdt,λ,\displaystyle=-\rho\sin\rho\pi+\tilde{\omega}\cos\rho\pi+\int_{0}^{\pi}\tilde{G}(t)\cos\rho t\,dt,\qquad\lambda\in\mathbb{C}, (3.9)
Δ~0(λ):=Δ~0(ρ2)\displaystyle\tilde{\Delta}^{0}(\lambda):=\tilde{\Delta}^{0}(\rho^{2}) =cosρπ+ω~2sinρπρ+1ρ0πS~(t)sinρtdt,λ{0}\displaystyle=\cos\rho\pi+\tilde{\omega}_{2}\frac{\sin\rho\pi}{\rho}+\frac{1}{\rho}\int_{0}^{\pi}\tilde{S}(t)\sin\rho t\,dt,\qquad\lambda\in\mathbb{C}\setminus\{0\} (3.10)

are well defined (to specify ρ\rho we use the square root branch with Imλ0\operatorname{Im}\sqrt{\lambda}\geq 0). The function Δ~0\tilde{\Delta}^{0} can be continuously extended to λ=0\lambda=0 as well. The resulting functions Δ~\tilde{\Delta} and Δ~0\tilde{\Delta}^{0} are entire functions of order 1/21/2. Moreover, representations (2.40) and (2.41) hold for the functions Δ~\tilde{\Delta} and Δ~0\tilde{\Delta}^{0} if one changes ω\omega, ω2\omega_{2} and hnh_{n} by ω~\tilde{\omega}, ω~2\tilde{\omega}_{2} and h~n\tilde{h}_{n} respectively. Let us define

M~(λ)=Δ~0(λ)Δ~(λ),λ.\tilde{M}(\lambda)=-\frac{\tilde{\Delta}^{0}(\lambda)}{\tilde{\Delta}(\lambda)},\qquad\lambda\in\mathbb{C}.

Then one can verify that

M~(zk)=Mk=M(zk),k0.\tilde{M}(z_{k})=M_{k}=M(z_{k}),\quad k\geq 0.

We would like to emphasize here that we do not know if the function M~\tilde{M} is the Weyl function corresponding to some potential qq, so Theorem 2.5 can not be applied directly. Nevertheless, the proof of this theorem in [9] requires only that M~\tilde{M} is a quotient of two entire functions satisfying some basic asymptotic conditions, which can be easily verified for the functions Δ~\tilde{\Delta} and Δ~0\tilde{\Delta}^{0}. Hence following the proof of Theorem 2.1 from [9] we obtain that M~M\tilde{M}\equiv M, a contradiction. ∎

Now suppose that only a finite set of pairs {zk,Mk}k=0K:={zk,M(zk)}k=0K\{z_{k},\,M_{k}\}_{k=0}^{K}:=\{z_{k},\,M(z_{k})\}_{k=0}^{K} is given. We assume that the numbers zkz_{k} are real (the system of equations (3.5) can be formulated for complex numbers zkz_{k} as well) and ordered: zk<zk+1z_{k}<z_{k+1}, 0kK10\leq k\leq K-1. Clearly any finite sequence of numbers zkz_{k}, kKk\leq K, can be extended to an infinite one satisfying the condition (2.15), which is not of a great use since the truncated system (3.5) for some particular choices of zkz_{k}, kKk\leq K may not be solved uniquely. For example, taking zk=k2z_{k}=k^{2} for all kKk\leq K makes it impossible to find the coefficient ω2\omega_{2}. As a rule of thumb we can ask that the terms in (2.15) remain small and bounded, even better if they decay as kk increases. So we can formulate the following empirical requirement for the numbers zk=ρ~k2z_{k}=\tilde{\rho}_{k}^{2}:

ρ~k<k2+1or equivalentlyzk<k24+k+1,\tilde{\rho}_{k}<\frac{k}{2}+1\qquad\text{or equivalently}\qquad z_{k}<\frac{k^{2}}{4}+k+1,

at least starting from some k=K<Kk=K^{\prime}<K. Such requirement is sufficient to deal, for example, with the two spectra inverse problem, for which {zk}k=02K={λk}k=0K{νk}k=0K\{z_{k}\}_{k=0}^{2K}=\{\lambda_{k}\}_{k=0}^{K}\cup\{\nu_{k}\}_{k=0}^{K}.

Let us consider a truncated system (3.5),

n=0[(K2)/2]ξ2n(1)n4n+2Mkj2n(ρ~kπ)π+n=0[(K3)/2]ξ2n+1(1)n4n+2j2n+1(ρ~kπ)πρ~k+ωMkcosρ~kπ+ω2sinρ~kπρ~k=Mkρ~ksinρ~kπcosρ~kπ,0kK.\begin{split}&\sum_{n=0}^{[(K-2)/2]}\xi_{2n}\cdot\frac{(-1)^{n}\sqrt{4n+2}M_{k}j_{2n}(\tilde{\rho}_{k}\pi)}{\sqrt{\pi}}+\sum_{n=0}^{[(K-3)/2]}\xi_{2n+1}\cdot\frac{(-1)^{n}\sqrt{4n+2}j_{2n+1}(\tilde{\rho}_{k}\pi)}{\sqrt{\pi}\tilde{\rho}_{k}}\\ &+\omega\cdot M_{k}\cos\tilde{\rho}_{k}\pi+\omega_{2}\cdot\frac{\sin\tilde{\rho}_{k}\pi}{\tilde{\rho}_{k}}=M_{k}\tilde{\rho}_{k}\sin\tilde{\rho}_{k}\pi-\cos\tilde{\rho}_{k}\pi,\qquad 0\leq k\leq K.\end{split} (3.11)

Solving this system one obtains approximate values of the parameters ω\omega, ω2\omega_{2} and the coefficients h0,,hK2h_{0},\ldots,h_{K-2}.

The coefficient matrix of the truncated system (3.11) can be badly conditioned. Since we know that the coefficients hnh_{n} decay, see Proposition 2.9, there is no need to look for the same number of the unknowns as the number of points λ~k\tilde{\lambda}_{k}. One may consider less unknowns and treat the system (3.11) as an overdetermined one. Asking for the condition number of the resulting coefficient matrix to be relatively small one estimates an optimum number MM, MK2M\leq K-2, of the coefficients h0,,hMh_{0},\ldots,h_{M}. We refer the reader to [28, Subsection 3.2 and Section 5] for additional details.

3.2 Recovery of eigenvalues λn\lambda_{n} and norming constants αn\alpha_{n}

Suppose we have the coefficients ω\omega, ω2\omega_{2} and {hn}n=0\{h_{n}\}_{n=0}^{\infty} recovered as described in Subsection 3.1. Then we can use the representation (2.40) to calculate the characteristic function Δ(ρ2)\Delta(\rho^{2}) for any value of ρ\rho. Recalling that the eigenvalues of the problem (2.1), (2.2) are real and coincide with zeros of Δ(λ)\Delta(\lambda), the recovery of the eigenvalues reduces to finding zeros of the entire function Δ(λ)\Delta(\lambda).

Having the spectrum λk=ρk2\lambda_{k}=\rho_{k}^{2}, the corresponding norming constants can be found using (2.14), (2.41) and (2.42). Indeed,

αk=1Δ0(λk)ddλΔ(λ)|λ=λk=12ρkΔ0(ρk2)ddρΔ(ρ2)|ρ=ρk.\alpha_{k}=-\frac{1}{\Delta^{0}(\lambda_{k})}\cdot\left.\frac{d}{d\lambda}\Delta(\lambda)\right|_{\lambda=\lambda_{k}}=-\frac{1}{2\rho_{k}\Delta^{0}(\rho_{k}^{2})}\left.\frac{d}{d\rho}\Delta(\rho^{2})\right|_{\rho=\rho_{k}}.

Suppose we have only a finite number of coefficients h0,,hMh_{0},\ldots,h_{M}. Then approximate eigenvalues are sought as zeros of the function

ΔM(ρ2)=ρsinρπ+ωcosρπ+2n=0[M/2](1)nh2nj2n(ρπ),\Delta_{M}(\rho^{2})=-\rho\sin\rho\pi+\omega\cos\rho\pi+2\sum_{n=0}^{[M/2]}(-1)^{n}h_{2n}j_{2n}(\rho\pi), (3.12)

and afterwards for each approximate eigenvalue λk=ρk2\lambda_{k}=\rho_{k}^{2} the corresponding norming constant can be obtained from

αk(1+πω)sinρkππρkcosρkπ+2n=0[M/2](1)nh2n(2nρkj2n(ρkπ)πj2n+1(ρkπ))2ρkcosρkπ+2ω2sinρkπ+4n=0[(M1)/2](1)nh2n+1j2n+1(ρkπ).\alpha_{k}\approx\frac{(1+\pi\omega)\sin\rho_{k}\pi-\pi\rho_{k}\cos\rho_{k}\pi+2\sum_{n=0}^{[M/2]}(-1)^{n}h_{2n}\left(\frac{2n}{\rho_{k}}j_{2n}(\rho_{k}\pi)-\pi j_{2n+1}(\rho_{k}\pi)\right)}{2\rho_{k}\cos\rho_{k}\pi+2\omega_{2}\sin\rho_{k}\pi+4\sum_{n=0}^{[(M-1)/2]}(-1)^{n}h_{2n+1}j_{2n+1}(\rho_{k}\pi)}. (3.13)

3.3 Main system of equations

Known the eigenvalues {λn}n=0\{\lambda_{n}\}_{n=0}^{\infty} and the corresponding norming constants {αn}n=0\{\alpha_{n}\}_{n=0}^{\infty}, the potential qq and the parameters hh and HH can be recovered by solving the Gelfand-Levitan equation (2.44). In [24] (see also [25]) using representation (2.28) the Gelfand-Levitan equation was reduced to an infinite system of linear algebraic equations for the coefficients β2n(x)\beta_{2n}(x). In [28] a more accurate modification of the method was proposed. We present the final result from [28] below.

For this subsection and without loss of generality, we suppose that ρ0=0\rho_{0}=0. This always can be achieved by a simple shift of the potential. If originally ρ00\rho_{0}\neq 0, then we can consider the new potential q~(x):=q(x)ρ02\widetilde{q}(x):=q(x)-\rho_{0}^{2}. Obviously, the eigenvalues {λn}n=0\left\{\lambda_{n}\right\}_{n=0}^{\infty} shift by the same amount, while the numbers hh, HH and {αn}n=0\left\{\alpha_{n}\right\}_{n=0}^{\infty} do not change. After recovering q~(x)\widetilde{q}(x) from the shifted eigenvalues, one gets the original potential q(x)q(x) by adding ρ02\rho_{0}^{2} back.

Let us denote

c~km(x)\displaystyle\widetilde{c}_{km}(x) =ωx8π(δm,k1(2k3/2)32δm,k(2k1/2)3+δm,k+1(2k+1/2)3)\displaystyle=-\frac{\omega x}{8\pi}\left(\frac{\delta_{m,k-1}}{(2k-3/2)_{3}}-\frac{2\delta_{m,k}}{(2k-1/2)_{3}}+\frac{\delta_{m,k+1}}{(2k+1/2)_{3}}\right)
+(1)k+mn=1[j2k(ρnx)j2m(ρnx)αn2j2k(nx)j2m(nx)π\displaystyle\quad+(-1)^{k+m}\sum_{n=1}^{\infty}\left[\frac{j_{2k}(\rho_{n}x)j_{2m}(\rho_{n}x)}{\alpha_{n}}-\frac{2j_{2k}(nx)j_{2m}(nx)}{\pi}\right. (3.14)
+2ωπ2n(xj2k(nx)j2m+1(nx)+xj2k+1(nx)j2m(nx)2(k+m)j2k(nx)j2m(nx)n)],\displaystyle\quad+\left.\frac{2\omega}{\pi^{2}n}\left(xj_{2k}(nx)j_{2m+1}(nx)+xj_{2k+1}(nx)j_{2m}(nx)-\frac{2(k+m)j_{2k}(nx)j_{2m}(nx)}{n}\right)\right],
C~0m(x)\displaystyle\widetilde{C}_{0m}(x) =(1α01π+2ωx23π2)δ0m+2ωx215π2δ1m+c~0m(x),\displaystyle=\left(\frac{1}{\alpha_{0}}-\frac{1}{\pi}+\frac{2\omega x^{2}}{3\pi^{2}}\right)\delta_{0m}+\frac{2\omega x^{2}}{15\pi^{2}}\delta_{1m}+\widetilde{c}_{0m}(x), (3.15)
C~1m(x)\displaystyle\widetilde{C}_{1m}(x) =2ωx215π2δ0m+c~1m(x),\displaystyle=\frac{2\omega x^{2}}{15\pi^{2}}\delta_{0m}+\widetilde{c}_{1m}(x), (3.16)
C~km(x)\displaystyle\widetilde{C}_{km}(x) =c~km(x),k=2,3,,m0,\displaystyle=\widetilde{c}_{km}(x),\qquad k=2,3,\ldots,\ m\in\mathbb{N}_{0}, (3.17)

and

d~k(x)=(1α01π+4ωx23π2ωxπ)δk02ωx215π2δk1(1)kn=1[cosρnxαnj2k(ρnx)2cosnxπj2k(nx)+2ωπ2n(xsinnxj2k(nx)+xcosnxj2k+1(nx)2kncosnxj2k(nx))],\begin{split}\widetilde{d}_{k}(x)&=-\left(\frac{1}{\alpha_{0}}-\frac{1}{\pi}+\frac{4\omega x^{2}}{3\pi^{2}}-\frac{\omega x}{\pi}\right)\delta_{k0}-\frac{2\omega x^{2}}{15\pi^{2}}\delta_{k1}\\ &\quad-(-1)^{k}\sum_{n=1}^{\infty}\left[\frac{\cos\rho_{n}x}{\alpha_{n}}j_{2k}(\rho_{n}x)-\frac{2\cos nx}{\pi}j_{2k}(nx)\right.\\ &\quad+\left.\frac{2\omega}{\pi^{2}n}\left(x\sin nxj_{2k}(nx)+x\cos nxj_{2k+1}(nx)-\frac{2k}{n}\cos nxj_{2k}(nx)\right)\right],\end{split} (3.18)

where δk,m\delta_{k,m} stands for the Kronecker delta and (k)m(k)_{m} is the Pochhammer symbol.

Then the following result is valid.

Theorem 3.2 ([28]).

The coefficients β2m(x)\beta_{2m}(x) satisfy the system of linear algebraic equations

β2k(x)(4k+1)x+m=0β2m(x)C~km(x)=d~k(x)2,k=0,1,.\frac{\beta_{2k}(x)}{(4k+1)x}+\sum_{m=0}^{\infty}\beta_{2m}(x)\widetilde{C}_{km}(x)=\frac{\widetilde{d}_{k}(x)}{2},\qquad k=0,1,\ldots. (3.19)

For all x(0,π]x\in\left(0,\pi\right] and k=0,1,k=0,1,\ldots the series in (3.19) converges.

It is of crucial importance the fact that for recovering the potential qq as well as the constants hh and HH it is not necessary to compute many coefficients β2m(x)\beta_{2m}(x) (that would be equivalent to approximate the solution of the Gelfand-Levitan equation) but instead the sole β0(x)\beta_{0}(x) is sufficient for this purpose, see Corollary 2.7.

For the numerical solution of the system (3.19) it is natural to truncate the infinite system, i.e., to consider m,kNm,k\leq N. As was shown in [28], the truncation process possesses some very nice properties. Namely, the truncated system is uniquely solvable for all sufficiently large NN, the approximate solutions converge to the exact one, the condition numbers of the coefficient matrices are uniformly bounded with respect to NN, and the solution is stable with respect to small errors in the coefficients. Moreover, a reduced number of equations in a truncated system is sufficient for recovering the coefficient β0\beta_{0} with a high accuracy. However it should be noted that a lot of (approximate) eigenvalues and norming constants are necessary to obtain values of the series (3.14)–(3.18) accurately enough. As was shown in Subsection 3.2, it is not a problem, thousands of approximate eigendata can be computed.

3.4 Algorithm of solution of Problem 2.4

Given a finite set of points {zn}n=0N={ρ~n2}n=0N\left\{z_{n}\right\}_{n=0}^{N}=\left\{\tilde{\rho}_{n}^{2}\right\}_{n=0}^{N} and the corresponding values Mn=M(zn)M_{n}=M(z_{n}), 0nN0\leq n\leq N of the Weyl function at these points, the following direct method for recovering the potential qq and the numbers hh and HH is proposed.

  1. 1.

    Solving (overdetermined) system (3.11) find the coefficients ω\omega, ω2\omega_{2} and h0,,hMh_{0},\ldots,h_{M}.

  2. 2.

    Compute h=ωω2h=\omega-\omega_{2}.

  3. 3.

    Find approximate eigenvalues λk\lambda_{k}, kKk\leq K as zeros of (3.12). Compute the corresponding norming constants αk\alpha_{k}, kKk\leq K using (3.13).

  4. 4.

    If ρ00\rho_{0}\neq 0, perform the shift of the eigenvalues

    ρ~k=ρk2ρ02,k0,\tilde{\rho}_{k}=\sqrt{\rho_{k}^{2}-\rho_{0}^{2}},\qquad k\geq 0,

    so that 0 becomes the first eigenvalue of the spectral problem. The parameters ω\omega and ω2\omega_{2} have to be shifted as well,

    ω~=ωπ2ρ02,ω~2=ω2π2ρ02.\tilde{\omega}=\omega-\frac{\pi}{2}\rho_{0}^{2},\qquad\tilde{\omega}_{2}=\omega_{2}-\frac{\pi}{2}\rho_{0}^{2}.

    Let us denote the square roots of the shifted eigenvalues by the same expression ρk\rho_{k}, and similarly for the shifted parameters ω\omega and ω2\omega_{2}.

  5. 5.

    For a set of points {xl}\left\{x_{l}\right\} from (0,π](0,\pi], compute the approximate values of the coefficients C~km(x)\widetilde{C}_{km}(x) and d~k(x)\widetilde{d}_{k}(x) for k,m=0,,Nk,m=0,\ldots,N with the aid of the formulas (3.14)–(3.18) and solve the truncated system (3.19) obtaining thus β0(x)\beta_{0}(x).

  6. 6.

    Compute qq from (2.31). Take into account that φ(0,x)\varphi(0,x) is an eigenfunction associated with the first eigenvalue λ0\lambda_{0} and hence does not have zeros on [0,π][0,\pi] (see, e. g., [2, Theorem 8.4.5]). This justifies the division over φ(0,x)\varphi(0,x).

  7. 7.

    Compute HH using (2.39). For this compute the mean of the potential 0πq(t)𝑑t\int_{0}^{\pi}q(t)\,dt, and thus,

    H=ω2120πq(t)𝑑t.H=\omega_{2}-\frac{1}{2}\int_{0}^{\pi}q(t)\,dt.
  8. 8.

    Recall that one has to add the original eigenvalue ρ02\rho_{0}^{2} back to the recovered potential to return to the original potential qq.

Remark 3.3.

The idea of interval flipping proposed in [28, Subsection 3.6] to improve the recovery of the potential near x=πx=\pi can be adapted for the proposed method as well, the formulas (3.39) and (3.40) from [28] are directly applicable with the data obtained on Steps 14. We leave the details to the reader.

Remark 3.4.

The proposed method with few modifications can be adapted to the case of Dirichlet boundary condition at x=πx=\pi, corresponding to the value H=H=\infty. The Weyl function in such case is given by

M(λ)=S(ρ,π)φ(ρ,π).M_{\infty}(\lambda)=-\frac{S(\rho,\pi)}{\varphi(\rho,\pi)}.

Following the described steps one can similarly obtain an infinite system of equations for the coefficients βn(π)\beta_{n}(\pi). The coefficients ω\omega and ω2\omega_{2} are no longer necessary. Squares of the zeros of the function φ(ρ,π)\varphi(\rho,\pi) are the eigenvalues, and the corresponding norming constants can be obtained similarly. Now one has to solve an inverse problem by the given spectral density function in the case when one boundary condition is of the Dirichlet type. It can be done with the use of a Gelfand-Levitan equation with a different kernel function FF, see [29, Chapter 2, §9], [16] and references therein. We leave the details for a separate paper.

Below, in Section 4 we illustrate the performance of this algorithm with several numerical examples.

4 Numerical examples

The proposed algorithm can be implemented directly, similarly to the algorithm from [28], only several observations should be made. On Step 1, to determine the number of unknowns to look for, we used the following criteria: we try to recover at least 10 (if less than 30 points znz_{n} were given) or at least 20 coefficients hnh_{n} (for more than 30 points znz_{n} given); if the condition number of the coefficient matrix of the system (3.11) is less than 1000, we increase the number of recovered coefficients hnh_{n} until the condition number surpasses 1000. A significant time saving can be achieved by computing the spherical Bessel functions jm(ρ~kπ)j_{m}(\tilde{\rho}_{k}\pi) using a backwards recursion, see [28], [3], [12] for details.

The number KK of the eigenvalues to compute on Step 3 can be estimated from the decay of the coefficients hnh_{n}. Faster decay means smoother potential, larger value of KK. Say, K=104K=10^{4} for smooth potentials (coefficients hnh_{n} decay very fast) and up to K=103K=10^{3} for non-smooth potentials (coefficients hnh_{n} decay slowly). For several examples originating from the inverse problems considered in Subsection 2.2, one can take an arbitrary value of the parameter hh when transforming the Weyl-Titchmarsh mm-function into the function MM using (2.11). In all these examples we took h=0h=0. Since the algorithm recovers two parameters ω\omega and ω2\omega_{2} which are equal due to the choice h=0h=0, we used the difference |ωω2||\omega-\omega_{2}| as some additional indicator of the accuracy of the recovered coefficients. Smaller value |ωω2||\omega-\omega_{2}| suggests to take more approximate eigenvalues (larger value of KK used, as was aforementioned).

Similarly to [28], the number of points xlx_{l} taken on Step 4 should not be large, about one hundred uniformly spaced points works best. We used 8 equations in the truncated system (3.11) in all the examples. All the computations were performed in Matlab 2017 on an Intel i7-7600U equipped laptop computer. We used spline approximation and differentiated the obtained spline on Step 6. In all examples where eigenvalues or solution of a Cauchy problem were necessary, we applied the method from [26].

First we consider several inverse spectral problems which can be reduced to Problem 2.4. In the last subsection we illustrate that the proposed algorithm can be applied even for complex points znz_{n}, but may fail for some particular choices of the points.

4.1 Two spectra

Consider an inverse problem of recovering the potential and boundary parameters by two spectra, see Subsubsection 2.2.1. The same problem was considered in [28] with the only difference that the shared boundary condition was at the point 0 (which is not essential since one can apply the change of variable xπxx\leftrightarrow\pi-x). Since for this problem the given values of the Weyl function are either 0 or \infty, the system (3.1) splits into two independent systems. One (given by (3.2)) coincides with the system (3.8) in [28], but the second system is different from (4.6) from [28]. Moreover, in [28] we recovered the parameters ω\omega and ω2\omega_{2} and obtained larger index eigenvalues from the eigenvalue asymptotics, and used the solutions of the systems only to compute the norming constants. While in this example we applied the proposed algorithm as is, without separating the system (3.1) or previously obtaining the values of ω\omega and ω2\omega_{2}, and computing all the eigenvalues from the obtained approximate characteristic function. Of course, one can not expect obtaining the same accuracy as those of [28].

We considered three potentials: smooth q1(x)=16π2x2exp(28xπ)q_{1}(x)=\frac{16}{\pi^{2}}x^{2}\exp\left(2-\frac{8x}{\pi}\right) (potential from [7], adapted to the interval [0,π][0,\pi]), non-smooth continuous q2(x)=|3|x23||q_{2}(x)=\bigl{|}3-|x^{2}-3|\bigr{|} (potential from [21]) and discontinuous

q3(x)={0,x[0,π8][3π8,3π5),12xπ+32,x(π8,π4],12xπ92,x(π4,3π8),4,x[3π5,4π5),2,x[4π5,π].q_{3}(x)=\begin{cases}0,&x\in[0,\frac{\pi}{8}]\cup[\frac{3\pi}{8},\frac{3\pi}{5}),\\ -\frac{12x}{\pi}+\frac{3}{2},&x\in(\frac{\pi}{8},\frac{\pi}{4}],\\ \frac{12x}{\pi}-\frac{9}{2},&x\in(\frac{\pi}{4},\frac{3\pi}{8}),\\ 4,&x\in[\frac{3\pi}{5},\frac{4\pi}{5}),\\ 2,&x\in[\frac{4\pi}{5},\pi].\end{cases}

(potential matching the one from [7]). For all potentials we took h=1h=1 and H=2H=2.

On Figure 1 we present the recovered potential q1q_{1} and its absolute error. 6, 10 or 16 eigenvalue pairs were used for recovery. As one can see, the error almost stabilizes. For 16 eigenvalue pairs the parameter qq was obtained with L1(0,π)L_{1}(0,\pi) error of 8.61078.6\cdot 10^{-7}, the constants hh and HH were obtained with the errors of 4.610104.6\cdot 10^{-10} and 2.31072.3\cdot 10^{-7}.

Refer to caption
Refer to caption
Figure 1: On the left: exact (blue line), recovered from 6 eigenvalue pairs (red dots) and recovered from 10 eigenvalue pairs (black dots) potential q1q_{1} from Subsection 4.1. On the right: absolute error of the recovered potential, 6, 10 or 16 eigenvalue pairs were used.

On Figure 2 we present the potential q2q_{2} recovered from 40 and 201 eigenvalue pairs. For 40 eigenvalue pairs the potential qq was obtained with L1(0,π)L_{1}(0,\pi) error of 4.41024.4\cdot 10^{-2}, the constants hh and HH were obtained with the errors of 0.0170.017 and 0.0240.024. For 201 eigenvalue pairs, with 1.21021.2\cdot 10^{-2}, 0.0110.011 and 0.0140.014, respectively.

Refer to caption
Refer to caption
Figure 2: On the left: exact (blue line) and recovered from 40 eigenvalue pairs (black dots); on the right: exact (blue line) and recovered from 201 eigenvalue pairs (black dots) potential q2q_{2} from Subsection 4.1.

On Figure 3 we present the potential q3q_{3} recovered from 30 and 201 eigenvalue pairs. While for 30 eigenvalue pairs the algorithm failed near the interval endpoints, inside the interval (0.5,π0.5)(0.5,\pi-0.5) the result is acceptable. For 201 eigenvalue pairs the algorithm was able to recover both potential and boundary parameters hh and HH, L1L_{1} error of the potential was 0.240.24, absolute errors of the boundary parameters were 0.00760.0076 and 0.0290.029, respectively.

Refer to caption
Refer to caption
Figure 3: On the left: exact (blue line) and recovered from 30 eigenvalue pairs (black dots); on the right: exact (blue line) and recovered from 201 eigenvalue pairs (black dots) potential q3q_{3} from Subsection 4.1.

4.2 Inverse problems with known parts of several spectra

As it is stated in [9], “Two-thirds of the spectra of three spectral problems uniquely determine qq.” In this subsection we numerically illustrate the solution of such inverse problems.

To illustrate the performance of the algorithm we considered two potentials, the same smooth potential q1q_{1} and the following potential, possessing a discontinuous second derivative q4(x)=0xq2(s)𝑑sq_{4}(x)=\int_{0}^{x}q_{2}(s)\,ds, where q1q_{1} and q2q_{2} were introduced in Subsection 4.1. We took H=2H=2 and h{1,2,3}h\in\{1,2,3\}, and eigenvalues indexed 0,1,3,4,6,7,0,1,3,4,6,7,\ldots from the first spectrum, 0,2,3,5,6,8,9,0,2,3,5,6,8,9,\ldots from the second spectrum and 1,2,4,5,7,8,1,2,4,5,7,8,\ldots from the third spectrum.

Refer to caption
Refer to caption
Figure 4: On the left: exact (blue line), recovered from 12 eigenvalues (red dots) and recovered from 18 eigenvalues (black dots) potential q1q_{1} from Subsection 4.2. On the right: absolute error of the recovered potential, 12, 18 or 30 eigenvalues were used.

On Figure 4 we present the recovered potential q1q_{1} and its absolute error. The absolute error rapidly decreases with more eigenvalues used and stabilizes at 24 eigenvalues (8 from each spectra). L1L_{1} error of the recovered q1q_{1} was 3.91073.9\cdot 10^{-7}, the parameter HH had an absolute error of 3.11073.1\cdot 10^{-7}.

Refer to caption
Refer to caption
Figure 5: On the left: exact (blue line), recovered from 60 eigenvalues (red dots) and recovered from 120 eigenvalues (black dots) potential q4q_{4} from Subsection 4.2. On the right: absolute error of the recovered potential, 60 or 120 eigenvalues were used.
Refer to caption
Refer to caption
Figure 6: Exact potential q1q_{1} from Subsection 4.2 (blue line) and recovered one (black dots). On the left: recovered from 10 eigenvalues λ0(h)\lambda_{0}(h), h{0,1,2,3,4,5,10,20,50,100}h\in\{0,1,2,3,4,5,10,20,50,100\}, on the right: recovered from 30 eigenvalues λ0(h),λ1(h),λ2(h)\lambda_{0}(h),\lambda_{1}(h),\lambda_{2}(h), h{0,1,2,3,4,5,10,20,50,100}h\in\{0,1,2,3,4,5,10,20,50,100\}.

On Figure 5 we present the recovered potential q4q_{4} and its absolute error. In this case the potential is of finite smoothness, so the algorithm converges slowly. With 120 eigenvalues (40 from each spectrum) we obtained the following errors: L1L_{1} error of the potential was 3.71033.7\cdot 10^{-3} and the error of the parameter HH was 1.51031.5\cdot 10^{-3}.

In [7, Section 3.14] the following inverse problem was considered. Suppose that the first eigenvalue is known for spectral problems having the same potential but different boundary conditions at x=0x=0. From countably many values the potential can be recovered uniquely. However, if only finitely many values are given, the situation changes. The corresponding numerical problem is extremely ill-posed, so the method from [34] was not able to solve it, and the modification from [7] was able to recover only the general shape of the potential.

We tried the proposed algorithm on this problem. For a numerical example we took the potential q1q_{1} from Subsection 4.1, H=0H=0 and h{0,1,2,3,4,5,10,20,50,100}h\in\{0,1,2,3,4,5,10,20,50,100\}. On Figure 6, left plot, we present the recovered potential. The algorithm failed, it is a case of very small interval for zkz_{k}. And the result does not improve if we take more values of the parameter hh. However the situation greatly improves if instead of 1 eigenvalue for each spectral problem we take several. On Figure 6, right plot, we present the recovered potential for the inverse problem when the first 3 eigenvalues were taken from each spectrum.

4.3 Partially known potential

In this subsection we illustrate the performance of the proposed method applied to solution of inverse problems with a partially known potential. For comparison we have taken two potentials from [21], a smooth potential q5(x)=exx212q_{5}(x)=\frac{e^{x}-x^{2}}{12} and a non-smooth continuous potential q2(x)=|3|x23||q_{2}(x)=|3-|x^{2}-3||. The potentials are known on [0,π/2][0,\pi/2], hence one spectrum is necessary to recover the potential. While in [21] the Dirichlet boundary conditions are considered, we decided to take mixed boundary conditions with h=1h=1 and H=2H=2 for both potentials.

To solve the inverse problem we followed the idea presented in Subsubsection 2.2.3: solve the Cauchy problems with initial data v(ρn,0)=1v(\rho_{n},0)=1, v(ρn,0)=hv^{\prime}(\rho_{n},0)=h on the segment [0,π/2][0,\pi/2], compute the values of the Weyl-Titchmarsh mm-function

m(π/2,ρn2)=v(ρn,π/2)v(ρn,π/2),m(\pi/2,\rho_{n}^{2})=\frac{v^{\prime}(\rho_{n},\pi/2)}{v(\rho_{n},\pi/2)},

transform them to the Weyl function MM using (2.11). After a simple rescaling we get Problem 2.4.

For the numerical example 40 eigenvalues with 1 missing eigenvalue were considered in [21]. For the potential q5q_{5} we similarly considered 40 eigenvalues, but with 2 eigenvalues missing: λ10\lambda_{10} and λ35\lambda_{35}. For the potential q2q_{2} one eigenvalue λ35\lambda_{35} was missing. We would like to emphasize that we used the proposed method “as is”, without recovering missing eigenvalues (which is impossible, since the algorithm has no information if the points znz_{n} correspond to any eigenvalues).

On Figures 7 and 8 we present the recovered potentials and their absolute errors. One can appreciate an especially excellent accuracy in the case of a smooth potential.

Refer to caption
Refer to caption
Figure 7: On the left: exact (blue line) and recovered (black dots) potential q5q_{5} from Subsection 4.3. On the right: absolute error of the recovered potential. Eigenvalues {λn}n=039{λ10,λ35}\{\lambda_{n}\}_{n=0}^{39}\setminus\{\lambda_{10},\lambda_{35}\} were used. The error of the recovered parameter HH is 5.010105.0\cdot 10^{-10}, and the L1L_{1} error of the recovered potential is 3.01093.0\cdot 10^{-9}.
Refer to caption
Refer to caption
Figure 8: On the left: exact (blue line) and recovered (black dots) potential q2q_{2} from Subsection 4.3. On the right: absolute error of the recovered potential. Eigenvalues {λn}n=039{λ35}\{\lambda_{n}\}_{n=0}^{39}\setminus\{\lambda_{35}\} were used. The error of the recovered parameter HH is 2.51022.5\cdot 10^{-2}, and the L1L_{1} error of the recovered potential is 0.110.11.

4.4 Recovery of the potential from values of the Weyl function: complex points znz_{n} and some problematic choices

The proposed algorithm works even when the numbers znz_{n} are complex (however for numerical stability the imaginary parts should be relatively small comparing to the real parts). For illustration we considered the potential q1q_{1} from Subsection 4.1, h=1h=1, H=2H=2 and the following sets of points

zn(k)=(15+n2+ki)2,0n40,k{0,1,2}.z_{n}^{(k)}=\left(\frac{1}{5}+\frac{n}{2}+k\cdot i\right)^{2},\qquad 0\leq n\leq 40,\ k\in\{0,1,2\}. (4.1)

The obtained results are presented on Figure 9, left plot. The algorithm was able to recover both the potential and the boundary conditions, however the accuracy decreases when the points are taken farther from the real line.

Refer to caption
Refer to caption
Figure 9: Absolute errors of the recovered potential q1q_{1} from Subsection 4.4. On the left: using the points (4.1) with different values of kk. On the right: using the points shown on the legend, 0n400\leq n\leq 40.

It should be mentioned that the proposed algorithm may fail for some particular choices of the points znz_{n} even for nice potentials. We considered the same problem and the following set of points

yn(1)=(14+n2)2,0n40.y_{n}^{(1)}=\left(\frac{1}{4}+\frac{n}{2}\right)^{2},\qquad 0\leq n\leq 40.

The truncated system (3.11) was not ill conditioned, but its solution was nowhere close to the exact one and the situation does not improve when taking more or less unknowns. For example, for 9 unknowns ξ0,,ξ8\xi_{0},\ldots,\xi_{8} the parameter h=ωω2h=\omega-\omega_{2} resulted to be 0.79990.7999 instead of 11. The best value was achieved for 10 unknowns and that was 0.86770.8677. On Figure 9, right plot we present the absolute error of the recovered potential in comparison with those recovered using a very close set of points

yn(2)=(0.249+n2)2,0n40.y_{n}^{(2)}=\left(0.249+\frac{n}{2}\right)^{2},\qquad 0\leq n\leq 40.

For the latter choice of points, the recovered parameter hh resulted to be 0.9999970.999997 and the potential was recovered to 5 extra decimal digits. It should be noted that the sets {yn(1)}n=0\{y_{n}^{(1)}\}_{n=0}^{\infty} and {yn(2)}n=0\{y_{n}^{(2)}\}_{n=0}^{\infty} do not satisfy neither the condition (2.15) nor the condition from [20], so any behavior can be expected for finite subsets. And the set {yn(1)}n=0\{y_{n}^{(1)}\}_{n=0}^{\infty} is, in some sense, the worst possible since the points yn(1)y_{n}^{(1)} are the most distant from any eigenvalue asymptotics (numbers close either to integers or integers plus one half). The situation changes completely if we consider the set {0}{yn(1)}n=0\{0\}\cup\{y_{n}^{(1)}\}_{n=0}^{\infty}. This set satisfies the condition (2.15), and the proposed method works for it without any problem, see Figure 9, right plot. The recovered parameter hh was 0.99999999910.9999999991.

5 Conclusions

A direct method for recovering the Sturm-Liouville problem from the Weyl-Titchmarsh function given on a countable set of points is developed, which leads to an efficient numerical algorithm for solving a variety of inverse Sturm-Liouville problems. The main role in the proposed approach is played by the coefficients of the Fourier-Legendre series expansion of the transmutation operator integral kernel. We show how the given spectral data lead to an infinite system of linear algebraic equations for the coefficients, and the crucial observation is that for recovering the potential it is sufficient to find the first coefficient only. In practical terms this observation translates into the fact that very few equations in the truncated system are enough for solving the inverse Sturm-Liouville problem.

The method is simple, direct and accurate. Its limitations are all related to a possible insufficiency of the input data. A too reduced number of eigendata or a special choice of points at which the Weyl function is given may lead to inaccurate results, as we illustrate by numerical examples.

Acknowledgements

Research was supported by CONACYT, Mexico via the project 284470. Research of Vladislav Kravchenko was supported by the Regional mathematical center of the Southern Federal University, Russia.

References

  • [1] M. Abramovitz and I. A. Stegun, Handbook of mathematical functions, New York: Dover, 1972.
  • [2] F. V. Atkinson, Discrete and continuous boundary problems, New York: Academic Press, 1964.
  • [3] A. R. Barnett, The calculation of spherical Bessel and Coulomb functions, in Computational Atomic Physics, Electron and Positron Collisions with Atoms and Ions, Berlin: Springer-Verlag, 1996, 181–202.
  • [4] N. P. Bondarenko, Inverse Sturm-Liouville problem with analytical functions in the boundary condition, Open Mathematics 18 (2020), 512–528.
  • [5] N. P. Bondarenko, Solvability and stability of the inverse Sturm–Liouville problem with analytical functions in the boundary condition, Math. Meth. Appl. Sci. 43 (2020), 7009–7021.
  • [6] H. Campos, V. V. Kravchenko and S. M. Torba, Transmutations, L-bases and complete families of solutions of the stationary Schrödinger equation in the plane, J. Math. Anal. Appl. 389 (2012), no. 2, 1222–1238.
  • [7] Kh. Chadan, D. Colton, L. Päivärinta and W. Rundell, An introduction to inverse scattering and inverse spectral problems. SIAM, Philadelphia, 1997.
  • [8] K. Chadan and P. C. Sabatier, Inverse problems in quantum scattering theory. Second edition, Springer-Verlag, New York, 1989.
  • [9] F. Gesztesy, R. del Rio and B. Simon, Inverse spectral analysis with partial information on the potential, III. Updating boundary conditions, Internat. Math. Research Notices 15 (1997), 751–758.
  • [10] F. Gesztesy and B. Simon, A new approach to inverse spectral theory II. General real potentials and the connection to the spectral measure, Ann. of Math. 152 (2000), 593–643.
  • [11] F. Gesztesy and B. Simon, Inverse spectral analysis with partial information on the potential, II. The case of discrete spectrum, Trans. Amer. Math. Soc. 352 (2000), 2765–2787.
  • [12] E. Gillman and H. R. Fiebig, Accurate recursive generation of spherical Bessel and Neumann functions for a large range of indices, Comput. Phys. 2 (1988), 62–72.
  • [13] G. M. L. Gladwell, Inverse problems in vibration – II, Appl. Mech. Rev. 49(10S) (1996), S25–S34.
  • [14] G. M. L. Gladwell, Inverse problems in vibration. Second edition, Kluwer Academic Publishers, Dordrecht, 2005.
  • [15] K. Gou and Z. Chen, Inverse Sturm-Liouville problems and their biomedical engineering applications, JSM Math. Stat. 2 (2015), 1008.
  • [16] T. N. Harutyunyan and A. M. Tonoyan, Some properties of the kernel of a transmutation operator, 2021, to appear.
  • [17] H. Hochstadt, The inverse Sturm–Liouville problem, Comm. Pure Appl. Math., 26 (1973), 715–729.
  • [18] H. Hochstadt and B. Lieberman, An inverse Sturm-Liouville problem with mixed given data, SIAM J. Appl. Math. 34 (1978), 676–680.
  • [19] M. Horváth, On the inverse spectral theory of Schrödinger and Dirac operators, Trans. Amer. Math. Soc. 353 (2001), 4155–4171.
  • [20] M. Horváth, Inverse spectral problems and closed exponential systems, Ann. of Math. 162 (2005), 885–918.
  • [21] A. Kammanee and C. Böckmann, Determination of partially known Sturm–Liouville potentials, Appl. Math. Comput. 204 (2008) 928–937.
  • [22] K. V. Khmelnytskaya, V. V. Kravchenko and S. M. Torba, A representation of the transmutation kernels for the Schrödinger operator in terms of eigenfunctions and applications, Appl. Math. Comput. 353 (2019), 274–281.
  • [23] K. V. Khmelnytskaya and T. V. Torchynska, Reconstruction of potentials in quantum dots and other small symmetric structures, Math. Methods Appl. Sci. 33 (2010), 469–472.
  • [24] V. V. Kravchenko, On a method for solving the inverse Sturm–Liouville problem, J. Inverse Ill-posed Probl. 27 (2019), 401–407.
  • [25] V. V. Kravchenko, Direct and inverse Sturm-Liouville problems: A method of solution, Birkhäuser, Cham, 2020.
  • [26] V. V. Kravchenko, L. J. Navarro and S. M. Torba, Representation of solutions to the one-dimensional Schrödinger equation in terms of Neumann series of Bessel functions, Appl. Math. Comput. 314 (2017), 173–192.
  • [27] V. V. Kravchenko and S. M. Torba, A Neumann series of Bessel functions representation for solutions of Sturm-Liouville equations, Calcolo 55 (2018), article 11, 23pp.
  • [28] V. V. Kravchenko and S. M. Torba, A direct method for solving inverse Sturm-Liouville problems, Inverse Probl. 37 (2021), 015015 (32pp).
  • [29] B. M. Levitan, Inverse Sturm-Liouville problems, VSP, Zeist, 1987.
  • [30] V. A. Marchenko, Some questions on one-dimensional linear second order differential operators, Transactions of Moscow Math. Soc., 1 (1952), 327–420.
  • [31] V. A. Marchenko, Sturm-Liouville operators and applications: revised edition, AMS Chelsea Publishing, 2011.
  • [32] M. Neher, Enclosing solutions of an inverse Sturm-Liouville problem with finite data, Computing 53 (1994), 379–395.
  • [33] V. Pivovarchik, An inverse problem by eigenvalues of four spectra, J. Math. Anal. Appl. 396 (2012) 715–723.
  • [34] W. Rundell and P. E. Sacks, Reconstruction techniques for classical inverse Sturm–Liouville problems, Math. Comput. 58 (1992), 161–183.
  • [35] A. M. Savchuk, Reconstruction of the potential of the Sturm–Liouville operator from a finite set of eigenvalues and normalizing constants, Math. Notes 99 (2016), 715–728.
  • [36] E. L. Shishkina and S. M. Sitnik, Transmutations, singular and fractional differential equations with applications to mathematical physics, Elsevier, Amsterdam, 2020.
  • [37] A. M. Savchuk and A. A. Shkalikov, Recovering of a potential of the Sturm-Liouville problem from finite sets of spectral data, in Spectral Theory and Differential Equations, Amer. Math. Soc. Transl. Ser. 2, 233 (2014), 211–224.
  • [38] A. A. Shkalikov, Boundary problems for ordinary differential equations with parameter in the boundary conditions, J. Sov. Math. 33 (1986), 1311–1342.
  • [39] B. Simon, A new approach to inverse spectral theory I. Fundamental formalism, Ann. of Math. 150 (1999), 1029–1057.
  • [40] A. O. Vatulyan, Inverse problems of solid mechanics, Fizmatlit, Moscow, 2007 (in Russian).
  • [41] C. F. Yang, N. P. Bondarenko and X. C. Xu, An inverse problem for the Sturm-Liouville pencil with arbitrary entire functions in the boundary condition, Inverse Probl. Imag. 14 (2020), 153–169.
  • [42] V. A. Yurko, Introduction to the theory of inverse spectral problems, Fizmatlit, Moscow, 2007 (in Russian).