\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2016 (2016), No. 11, pp. 1--12.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2016 Texas State University.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2016/11\hfil Inverse spectral problems] {Inverse spectral problems for energy-dependent Sturm-Liouville equations with finitely many point $\delta $-interactions} \author[M. Dzh. Manafov \hfil EJDE-2016/11\hfilneg] {Manaf Dzh. Manafov} \dedicatory{In memory of M. G. Gasymov (1939-2008)} \address{Manaf Dzh. Manafov \newline Ad\i{y}aman University, Faculty of Science and Arts, Department of Mathematics, \newline Ad\i{y}aman, Turkey} \email{mmanafov@adiyaman.edu.tr} \thanks{Submitted August 20, 2015. Published January 6, 2016.} \subjclass[2010]{34A55, 34B24, 34L05} \keywords{Energy-dependent Sturm-Liouville equation; \hfill\break\indent inverse spectral problems; point delta-interactions} \begin{abstract} In this study, inverse spectral problems for a energy-dependent Sturm-Liouville equations with finitely many point $\delta $-interactions. The uniqueness theorems for the inverse problems of reconstruction of the boundary value problem from the Weyl function, from the spectral data and from two spectra are proved and a constructive procedure for finding its solution are obtained. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{algorithm}[theorem]{Algorithm} \allowdisplaybreaks \section{Introduction} We consider inverse problems for the boundary value problem L (BVP L) generated by the differential equation \begin{equation} -y''+q(x)y=\lambda ^2y,\quad x\in \cup _{s=0}^{m}( a_s,a_{s+1})\quad (0=a_0,\;a_{m+1}=\pi ) \label{1.1} \end{equation} with the Robin boundary conditions \begin{equation} \begin{gathered} U(y):=y'(0)-hy(0)=0,\\ V(y):=y'(\pi )+Hy(\pi )=0, \end{gathered} \label{1.2} \end{equation} and the conditions at the points $x=a_s$: \begin{equation} I(y):=\{ \begin{gathered} y(a_s+0)=y(a_s-0)=y(a_s) \\ y'(a_s+0)-y'(a_s-0)=2\alpha _s\lambda y(a_s), \end{gathered} \label{1.3} \end{equation} $q(x)$ is a complex-valued function in $L_2(0,\pi )$; $h,H$ and $\alpha _s$ are complex numbers, $s= \overline{1,m}:= 1,2,\dots m$; and $\lambda $ is a spectral parameter. Note that, we can understand problem \eqref{1.1} and \eqref{1.3} as studying the equation \begin{equation} y''+(\lambda ^2-2\lambda p(x)-q(x)) y=0,\quad x\in (0,\pi ) , \label{1.5} \end{equation} when $p(x)=\sum_{s=1}^{m}\alpha _s\delta (x-a_s) $, where $\delta (x) $ is the Dirac function (see \cite{2}). Sturm-Liouville spectral problems with potentials depending on the spectral parameter arise in various models of quantum and classical mechanics. For instance, the evolution equations that are used to model interactions between colliding relativistic spineless particles can be reduced to the form \eqref{1.1}. Then $\lambda ^2$ is associated with the energy of the system (see \cite{15,17}). Spectral problems of differential operators are studied in two main branches, namely, direct and inverse problems. Direct problems of spectral analysis consist in investigating the spectral properties of an operator. On the other hand, inverse problems aim at recovering operators from their spectral characteristics. One takes for the main spectral data, for instance, one, two, or more spectra, the spectral function, the spectrum, and the normalized constants, the Weyl function. Direct and inverse problems for the classical Sturm-Liouville operators have been extensively studied (see \cite{8,16,18} and the references therein). Some classes of direct and inverse problems for discontinuous $BVPs$ (special case $p(x)\equiv 0$) in various statements have been considered in \cite{3,11,13,21,23} and for further discussion see the references therein. Notice that, inverse spectral problems for non-selfadjoint Sturm-Liouville operators on a finite interval with possibly multiple spectra have been investigated in \cite{B1, B2}. Non-linear dependence of \eqref{1.5} on the spectral parameter $\lambda $ should be regarded as a spectral problem for a quadratic operator pensil. The problem with $p(x)\in W_2^{1}(0,1)$ and $q(x)\in L_2(0,1)$ and with Robin boundary conditions was discussed in \cite{10}. Such problems for separated and non-separated boundary conditions were considered (see \cite{9,12,24} and the references therein). In this aspect, various inverse spectral problems for the equation \eqref{1.5} have been investigated in \cite{19,20} for the case $m=1$. In this article we establish various uniqueness results for inverse spectral problems of energy-dependent Sturm-Liouville equations with finitely many point $\delta $-interactions, and obtain a process for finding its solution. \section{Properties of the spectral characteristics of BVP L} In this section, we provide the spectral characteristics of the BVP L and present the relationship among these spectral characteristics. Moreover, we formulate the inverse problem of the reconstruction of BVP L: from the Weyl function, from the spectral data, and from two spectra. The technique employed is similar to those used in \cite{8}. Let $y(x)$ and $z(x)$ be continuously differentiable functions on the intervals $(0,a_1) $, $(a_1,a_2) ,\dots,(a_{m-1},a_m) $ and $(a_m,\pi ) $. Denote $\langle y,z\rangle :=yz'-y'z$. If $y(x)$ and $z(x)$ satisfy the conditions \eqref{1.3}, then \begin{equation} \langle y,z\rangle _{x=a_s-0}=\langle y,z\rangle _{x=a_s+0},\quad s=\overline{1,m}, \label{2.1} \end{equation} i.e. the function $\langle y,z\rangle $ is continuous on $(0,\pi ) $. Let $\varphi (x,\lambda ) $, $\psi (x,\lambda ) $ be solutions of \eqref{1.1} under the conditions \begin{equation} \begin{gathered} \varphi (0,\lambda ) =\psi (\pi ,\lambda ) =1 \\ \varphi '(0,\lambda ) =h,\quad \psi '(\pi ,\lambda ) =-H, \end{gathered} \label{2.2} \end{equation} and under condition \eqref{1.3}. Denote $\Delta (\lambda ) :=\langle \varphi (x,\lambda ) ,\psi (x,\lambda ) \rangle $. By \eqref{2.1} and the Ostrogradskii-Liouville theorem (see \cite[p.83]{5}), $\Delta (\lambda ) $ does not depend on $x$. The function $\Delta (\lambda ) $ is called the characteristic function of $L$. Clearly, \begin{equation} \Delta (\lambda ) =-V(\varphi ) =U(\psi) . \label{2.3} \end{equation} Obviously, the function $\Delta (\lambda ) $ is entire in $\lambda $ and it has at most a countable set of zeros $\{ \lambda_n\} $. We confine ourselves, for simplicity, to the case when all zeros of the characteristic function $\Delta (\lambda ) $ are simple. \begin{lemma} \label{lem2.1} The eigenvalues $\{ \lambda _n^2\} _{n\geq0} $ of the BVP L coincide with the zeros of the characteristic function. The functions $\varphi (x,\lambda _n) $ and $\psi (x,\lambda _n) $ are eigenfunctions, and there exists a sequence $\{ \beta _n\} $ such that \begin{equation*} \psi (x,\lambda _n) =\beta _n\varphi (x,\lambda_n), \quad \beta _n\neq 0. \end{equation*} \end{lemma} \begin{proof} Let $\Delta (\lambda _0) =0$. Then by $\langle \varphi (x,\lambda _0) ,\psi (x,\lambda _0)\rangle =0$, we have $\psi (x,\lambda _0) =C\varphi(x,\lambda _0) $ for some constant $C$. Hence $\lambda _0$ is an eigenvalue and $\varphi (x,\lambda _0) $, $\psi (x,\lambda _0) $ are eigenfunctions related to $\lambda _0$. Conversely, let $\lambda _0$ be an eigenvalue of $L$. We shall show that $\Delta (\lambda _0) =0$. Assuming the converse, suppose that $\Delta (\lambda _0) \neq 0$. In this case the functions $ \varphi (x,\lambda _0) $ and $\psi (x,\lambda_0) $ are linearly independent. Then $y(x,\lambda _0)=c_1\varphi (x,\lambda _0) +c_2\psi (x,\lambda _0) $ is a general solution of the problem $L$. If $c_1\neq 0$, we can write \begin{equation*} \varphi (x,\lambda _0) =\frac{1}{c_1}y(x,\lambda _0)-\frac{c_2}{c_1}\psi (x,\lambda _0) . \end{equation*} Then we have \begin{equation*} \langle \varphi (x,\lambda _0) ,\psi (x,\lambda_0) \rangle =-\frac{1}{c_1}[ y'(\pi,\lambda _0) +Hy(\pi ,\lambda _0) ] =0, \end{equation*} which is a contradiction. Note that for each eigenvalue there exists only one eigenfunction. Therefore there exists sequence $\beta _n$ such that $\psi ( x,\lambda _n) =\beta _n\varphi (x,\lambda _n) $. \end{proof} Denote \begin{equation*} \gamma _n=\int_0^{\pi }\varphi ^2(x,\lambda _n) dx -\frac{1}{\lambda _n}\sum_{s=1}^{m}\alpha _s\varphi ^2(a_s,\lambda_n) . \end{equation*} The set $\{ \lambda _n,\gamma _n\} _{n=0,\pm 1,\pm 2,\dots}$ is called the \textit{spectral data} of $L$. \begin{lemma}\label{lem2.2} The equality \begin{equation} \dot{\Delta }(\lambda _n) =-2\lambda _n\beta _n\gamma _n \label{2.4} \end{equation} holds. Here $\dot{\Delta }(\lambda ) =\frac{d}{d\lambda }\Delta (\lambda ) $. \end{lemma} \begin{proof} Since \begin{equation*} -\varphi ^{''}(x,\lambda _n) +q(x)\varphi (x,\lambda _n) =\lambda _n^2\varphi (x,\lambda _n) ,\quad -\psi ^{''}(x,\lambda )+q(x)\psi (x,\lambda ) =\lambda^2 \psi (x,\lambda ) , \end{equation*} we obtain \begin{equation*} \frac{d}{dx}\langle \varphi (x,\lambda _n) ,\psi ( x,\lambda ) \rangle =(\lambda ^2-\lambda _n^2) \varphi (x,\lambda _n) \psi (x,\lambda ) . \end{equation*} Integrating from $0$ to $\pi $ and using the conditions \eqref{1.2}, \eqref{1.3}, we obtain \begin{equation*} \int_0^{\pi }\varphi (x,\lambda _n) \psi (x,\lambda) dx -\frac{2}{\lambda _n+\lambda }\sum_{s=1}^{m}\alpha _s\varphi (a_s,\lambda _n) \psi (a_s,\lambda ) =\frac{\Delta (\lambda _n) -\Delta (\lambda ) }{\lambda ^2-\lambda _n^2}. \end{equation*} Because $\lambda \to \lambda _n$, by Lemma \ref{lem2.1}, this yields \begin{equation*} \dot{\Delta }(\lambda _n) =-2\lambda _n\beta_n\gamma _n. \end{equation*} \end{proof} \begin{corollary} For BVP L, $\lambda _n\neq 0$, $\gamma _n\neq 0$ and $\lambda _n\neq \lambda _{k}$ $(n\neq k)$ hold. \end{corollary} Now, consider the solutions $\varphi (x,\lambda ) $ and $\psi (x,\lambda ) $. Let $\tau :=\operatorname{Im}\lambda $. The following asymptotic estimates hold uniformly with respect to $x\in ( a_s,a_{s+1}) $, $0\leq s\leq m$, as $| \lambda | \to \infty$: \begin{equation} \label{2.7} \begin{aligned} &\varphi (x,\lambda ) \\ &=\alpha _1'\alpha _2'\dots\alpha _s'\cos (\lambda x-\theta _1-\theta _2-\dots-\theta _s) \\ &\quad +\sum_{1\leq i\leq s}\alpha _1'\dots\alpha _{i}\dots\alpha _s'\sin [ \lambda (x-2a_{i}) +\theta _1+\dots+\theta _{i-1}-\theta _{i+1}-\dots-\theta _s] \\ &\quad +\sum_{1\leq i0$ such that \begin{equation} | \Delta (\lambda ) | \geq C_{\delta }| \lambda | \exp (| \tau | \pi ) \quad \text{for all }\lambda \in G_{\delta }\quad (\delta >0). \label{2.14} \end{equation} \item For sufficiently large values of $n$, one has \begin{gather*} | \Delta (\lambda ) -\Delta _0(\lambda ) | <\frac{C_{\delta }}{2}\exp (| \tau |\pi ) ,\\ \lambda \in \Gamma _n:=\{ \lambda :| \lambda | =| \lambda _n^{0}| +\frac{1}{2} \inf_{n\neq m}| \lambda _n^{0}-\lambda _m^{0}|\} . \end{gather*} Then \begin{equation} \lambda _n=\lambda _n^{0}+o(1),\text{ }n\to \infty .\label{2.15} \end{equation} Substituting \eqref{2.7}, \eqref{2.8} and \eqref{2.14} into \eqref{2.3} we obtain \begin{equation} \gamma _n=\gamma _n^{0}+o(1),\text{ }n\to \infty , \label{2.16} \end{equation} where \begin{equation*} \gamma _n^{0}=\int_0^{\pi }\varphi ^2(x,\lambda _n^{0}) dx-\frac{1}{\lambda _n^{0}}\sum_{s=1}^{m}\alpha _s\varphi ^2( a_s,\lambda _n^{0}) . \end{equation*} \end{enumerate} \section{Formulation of the inverse problem. Uniqueness theorems} In this section, we study three inverse problems of recovering L from its spectral characteristics, namely \begin{itemize} \item[i] from the Weyl function, \item[ii] from the so-called spectral data, \item[iii] from two spectra. \end{itemize} For each class of inverse problems we prove the corresponding uniqueness theorems and show the connection between the different spectral characteristics. Let $\Phi (x,\lambda ) $ be the solution of \eqref{1.5} under the conditions $U(\Phi )=1$ and $V(\Phi )=0$. We set $M(\lambda ):=\Phi (\lambda )$. The functions $\Phi (x,\lambda ) $ and $M(\lambda )$ are called the Weyl solution and the Weyl function for BVP L, respectively. Clearly \begin{gather} \Phi (x,\lambda ) =\frac{\psi (x,\lambda ) }{\Delta (\lambda ) }=S(x,\lambda ) +M(\lambda )\varphi (x,\lambda ), \label{4.1} \\ M(\lambda )=\frac{\psi (0,\lambda ) }{\Delta (\lambda ) }, \label{4.2} \end{gather} where $\psi (0,\lambda ) $ is the characteristic function of BVP L$_1$ which is equation \eqref{1.5} with the boundary conditions $U(y)=0$, $y(\pi )=0$, and $S(x,\lambda ) $ is defined from the equality \begin{equation*} \psi (x,\lambda ) =\psi (0,\lambda ) \varphi ( x,\lambda ) +\Delta (\lambda ) S(x,\lambda ) . \end{equation*} Note that, by equalities $\langle \varphi (x,\lambda ) ,S(x,\lambda ) \rangle \equiv 1$ and \eqref{4.1}, one has \begin{equation} \langle \Phi (x,\lambda ) ,\varphi (x,\lambda ) \rangle \equiv 1. \label{4.3} \end{equation} Let $\{ \mu _n^2\} _{n\geq 0\text{ }}$ be zeros of $\psi (0,\lambda ) $, i.e. the eigenvalues of $L_1$. First, let us prove the uniqueness theorems for the solutions of the problems (i)-(iii). For this purpose we agree that together with L we consider a BVP $\widetilde{L}$ of the same form but with different coefficients $\widetilde{q}(x)$, $\widetilde{h}$, $\widetilde{H}$, $\widetilde{a_s}$ and $\widetilde{\alpha _s}$, $s=\overline{1,m}$. Everywhere below if a certain symbol $e$ denotes an object related to $L$, then the corresponding symbol $\widetilde{e}$ with tilde denotes the analogous object related to $\widetilde{L}$. \begin{theorem}\label{theorem 4.1} If $M(\lambda )=\widetilde{M}(\lambda )$, then $L=\widetilde{L}$. Thus, the specification of the Weyl function $M(\lambda )$ uniquely determines $L$. \end{theorem} \begin{proof} Let us define the matrix $P(x,\lambda )=[ P_{jk}(x,\lambda )]_{j,k=1,2}$ by the formula \begin{equation} P(x,\lambda )\begin{bmatrix} \widetilde{\varphi }(x,\lambda ) & \widetilde{\Phi }( x,\lambda ) \\ \widetilde{\varphi }'(x,\lambda ) & \widetilde{\Phi } '(x,\lambda ) \end{bmatrix} = \begin{bmatrix} \varphi (x,\lambda ) & \Phi (x,\lambda ) \\ \varphi '(x,\lambda ) & \Phi '(x,\lambda ) \end{bmatrix}. \label{4.4} \end{equation} By \eqref{4.1}, we calculate \begin{equation} \begin{gathered} P_{j1}(x,\lambda )=\varphi ^{(j-1) }(x,\lambda ) \widetilde{\Phi }'(x,\lambda ) -\Phi ^{( j-1) }(x,\lambda ) \widetilde{\varphi }'(x,\lambda ) , \\ P_{j2}(x,\lambda )=\Phi ^{(j-1) }(x,\lambda ) \widetilde{\varphi }(x,\lambda ) -\varphi ^{(j-1) }(x,\lambda ) \widetilde{\Phi }(x,\lambda ) , \end{gathered} \label{4.5} \end{equation} and \begin{equation} \begin{gathered} \varphi (x,\lambda ) =P_{11}(x,\lambda )\widetilde{\varphi } (x,\lambda ) +P_{12}(x,\lambda )\widetilde{\varphi }'(x,\lambda ) , \\ \Phi (x,\lambda ) =P_{11}(x,\lambda )\widetilde{\Phi }( x,\lambda ) +P_{12}(x,\lambda )\widetilde{\Phi }'( x,\lambda ) . \end{gathered} \label{4.6} \end{equation} Taking \eqref{2.7}, \eqref{2.8} and \eqref{2.14} into account, we infer that \begin{equation} | P_{11}(x,\lambda )-1| \leq C_{\delta }| \lambda | ^{-1},\quad | P_{12}(x,\lambda )|\leq C_{\delta }| \lambda | ^{-1},\quad \lambda \in G_{\delta }, \label{4.7} \end{equation} where $G_{\delta }$ is defined in \eqref{2.14} and $C_{\delta }$ is a constant. On the other hand according to \eqref{4.2} and \eqref{4.5}, \begin{gather*} P_{11}(x,\lambda )=\varphi (x,\lambda ) \widetilde{S}'(x,\lambda ) -S(x,\lambda )\widetilde{\varphi }'(x,\lambda ) +(\widetilde{M}(\lambda )-M(\lambda ))\varphi (x,\lambda ) \widetilde{\varphi }'(x,\lambda ) , \\ P_{12}(x,\lambda )=S(x,\lambda )\widetilde{\varphi }(x,\lambda ) -\varphi (x,\lambda ) \widetilde{S}(x,\lambda )+(M(\lambda )- \widetilde{M}(\lambda ))\varphi (x,\lambda ) \widetilde{\varphi } (x,\lambda ) . \end{gather*} Since $M(\lambda )=\widetilde{M}(\lambda )$, it follows that for each fixed $x$ the functions $P_{1k}(x,\lambda )$, $k=1,2$ are entire in $\lambda $. With the help of \eqref{4.7} and well-known Liouville's theorem, this yields $P_{11}(x,\lambda )\equiv 1$, $P_{12}(x,\lambda )\equiv 0$. Substituting into \eqref{4.6}, we obtain $\varphi (x,\lambda ) \equiv \widetilde{ \varphi }(x,\lambda ) $, $\Phi (x,\lambda ) = \widetilde{\Phi }(x,\lambda ) $ for all $x\in \cup _{s=0}^{m}(x_s,x_{s+1}) $ and $\lambda $. Taking this into account, from \eqref{1.1} we obtain $q(x)=\widetilde{q}(x)$ a.e. on $( 0,\pi ) $, from \eqref{2.2} we obtain $h=\widetilde{h}$, $H=\widetilde{ H}$, and from \eqref{1.3} we conclude that $a_s=\widetilde{a_s}$, $ \alpha _s=\widetilde{\alpha _s}$ $(s=\overline{1,m}) $. Consequently, $L=\widetilde{L}$. \end{proof} \begin{theorem}\label{thm4.2} If $\lambda _n=\widetilde{\lambda _n}$, $\gamma _n= \widetilde{\gamma _n}$, $n=0,\pm 1,\pm 2,\dots$, then $L=\widetilde{L}$. Thus, the specification of the spectral data $\{ \lambda _n,\gamma _n\} _{n=0,\pm 1,\pm 2,\dots}$ uniquely determines the operator. \end{theorem} \begin{proof} It follows from \eqref{4.2} that the Weyl function $M(\lambda )$ is meromorphic with simple poles at points $\lambda _n^2$. Using \eqref{4.2}, Lemma \ref{lem2.1} and equality $\dot{\Delta }(\lambda_n)=-2\lambda _n\beta _n\gamma _n$, we have \begin{equation} \operatorname{Re}s_{\lambda-\lambda _n} M(\lambda ) =\frac{\psi (0,\lambda ) }{\dot{\Delta }(\lambda _n)} =\frac{\beta _n}{\dot{ \Delta }(\lambda _n)} =\frac{1}{-2\lambda _n\gamma _n}. \label{4.8} \end{equation} Since the Weyl function $M(\lambda )$ is regular for $\lambda \in \Gamma _n$, applying the Rouche theorem \cite[p.112]{6}, we conclude that \begin{equation*} M(\lambda )=\frac{1}{2\pi i}\int_{\Gamma _n}\frac{M(\mu )}{\lambda -\mu } d\mu ,\text{ }\lambda \in int\Gamma _n, \end{equation*} where the contour $\Gamma _n$ is assumed to have the counterclockwise circuit. Taking \eqref{2.14} and \eqref{4.2} into account, we arrive at $|M(\lambda )| \leq C_{\delta }| \lambda | ^{-1}$, $\lambda \in G_{\delta }$. Hence, by the residue theorem, we have \begin{equation} M(\lambda )=\sum_{n=-\infty }^{\infty }\frac{1}{2\lambda _n\gamma _n(\lambda -\lambda _n) }. \label{4.9} \end{equation} Under the hypothesis of the theorem we obtain, in view of \eqref{4.9}, that $M(\lambda )=\widetilde{M}(\lambda )$ and consequently by Theorem \ref{theorem 4.1}, $L=\widetilde{L}$. \end{proof} \begin{theorem} \label{theorem 4.3} If $\lambda _n=\widetilde{\lambda _n}$ and $\mu _n= \widetilde{\mu _n}$, $n=0,\pm 1,\pm 2,\dots$, then $L=\widetilde{L}$. Thus the specification of two spectra $\{ \lambda _n,\mu _n\}_{n=0,\pm 1,\pm 2,\dots}$ uniquely determines $L$. \end{theorem} \begin{proof} It is obvious that characteristic functions $\Delta (\lambda ) $ and $\psi (0,\lambda ) $ are uniquely determined by the sequences $\{ \lambda _n^2\} $ and $\{ \mu_n^2\} $ $(n=0,\pm 1,\pm 2,\dots) $, respectively. If $\lambda _n=\widetilde{\lambda _n}$, $\mu _n=\widetilde{\mu _n}$, $n=0,\pm 1,\pm 2,\dots$, then $\Delta (\lambda ) \equiv \widetilde{\Delta }(\lambda ) $, $\psi (0,\lambda ) =\widetilde{\psi }(0,\lambda ) $. Together with \eqref{4.2} this yields $M(\lambda )=\widetilde{M}(\lambda )$. By Theorem \ref{theorem 4.1} we obtain $L=\widetilde{L}$. \end{proof} \begin{remark} \label{remark 4.4} \rm By \eqref{4.2}, the specification of two spectra $\{ \lambda _n,\mu _n\} _{n=0,\pm 1,\pm 2,\dots}$ is equivalent to the specification of the Weyl function $M(\lambda )$. On the other hand, it follows from \eqref{4.9} that the specification of the Weyl function $M(\lambda )$ is equivalent to the specification of the spectral data $\{ \lambda _n,\gamma _n\} _{n=0,\pm 1,\pm 2,\dots}$. Consequently, three statements of the inverse problem of reconstruction of the problem $L$ from the Weyl function, from the spectral data and from two spectra are equivalent. \end{remark} \section{Solution of the inverse problem} In this section, we will solve the inverse problems of recovering the BVP L by the method of spectral mappings with the help of Cauchy's integral formula and Residue theorem. Finally, we provide the algorithms for the solution of the inverse problems by using the solution of the main equation. For definiteness we consider the inverse problem of recovering L from the spectral data $\{ \lambda _n,\gamma _n\} _{n=0,\pm 1,\pm 2,\dots}$. Let BVPs L and $\widetilde{L}$ be such that \begin{equation} a_s=\widetilde{a_s},\quad s=\overline{1,m},\quad \sum_{n=-\infty}^{\infty }\zeta _n| \lambda _n| <\infty , \label{5.1} \end{equation} where $\zeta _n:=| \lambda _n-\widetilde{\lambda } _n| +| \gamma _n-\widetilde{\gamma }_n| $. Denote \begin{gather*} \lambda _{n0}=\lambda _n,\text{ }\lambda _{n1}=\widetilde{\lambda }_n, \quad \gamma _{n_0}=\gamma _n,\gamma _{n1}=\widetilde{\gamma }_n, \\ \varphi _{ni}(x)=\varphi (x,\lambda _{ni}) ,\quad \widetilde{ \varphi }_{ni}(x)=\widetilde{\varphi }(x,\lambda _{ni}) , \\ Q_{kj}(x,\lambda ) =\frac{\langle \varphi (x,\lambda ) ,\varphi _{kj}(x)\rangle }{2\lambda _{kj}\gamma _{kj}( \lambda -\lambda _{kj}) }=\frac{1}{2\lambda _{kj}\gamma _{kj}} \int_0^{x}\varphi (t,\lambda ) \varphi _{kj}(t)dt, \\ Q_{ni,kj}(x) =Q_{kj}(x,\lambda _{ni}) \end{gather*} for $i,j=0,1$ and $n,k=0,\pm 1,\pm 2,\dots$, where $\widetilde{\varphi }(x,\lambda ) $ is the solution of \eqref{1.5} with the potential $\widetilde{q}$ under the initial conditions $\widetilde{\varphi }(0,\lambda ) =1$, $\widetilde{\varphi }'(0,\lambda ) =\widetilde{h}$. Analogously, we can define $\widetilde{Q} _{kj}(x,\lambda ) $ by replacing $\varphi $ with $\widetilde{\varphi }$ in the above definition. Using Schwarz's lemma \cite[p.130]{6} and \eqref{2.7}-\eqref{2.10}, \eqref{2.15} we obtain the following asymptotic estimates: \begin{gather} | \varphi _{nj}^{(\nu )}(x) | \leq C(| \lambda _n^{0}| +1) ^{\nu }, \label{5.2} \\ | Q_{ni,kj}(x) | \leq \frac{C}{| \lambda _n^{0}-\lambda _{k}^{0}| +1},\quad |Q_{ni,kj}^{(\nu +1)}(x) | \leq C(|\lambda _n^{0}| +| \lambda _{k}^{0}| +1) ^{(\nu )}, \label{5.3} \end{gather} where $n,k=0,\pm 1,\pm 2,\dots$, $i,j,\nu =0,1$ and $C$ is a positive constant. Analogous estimates are also valid for $\widetilde{\varphi }_{ni}(x)$, $\widetilde{Q}_{ni,kj}(x)$. \begin{lemma}\label{lemma 5.1} Let $\varphi _{nj}(x) $ and $Q_{ni,kj}(x) $ be defined as above. Then the following representations hold for $i,j=0,1$ and $n,k=0,\pm 1,\pm 2,\dots$: \begin{equation} \widetilde{\varphi }_{ni}(x)=\varphi _{n_{i}}(x)+\sum_{l=-\infty }^{\infty }\Big(\widetilde{Q}_{ni,l0}(x)\varphi _{k0}(x)-\widetilde{Q} _{ni,l1}(x)\varphi _{k1}(x)\Big) . \label{5.4} \end{equation} Series \eqref{5.4} converge absolutely and uniformly with respect to $x\in [ 0,\pi] \backslash \{ a_s\} _{s=1}^{m}$. \end{lemma} \begin{proof} By \eqref{5.1} we have $a_s=\widetilde{a_s}$, $\alpha _s=\widetilde{\alpha _s}$ for $s=\overline{1,m}$. Then it follows from \eqref{2.7}-\eqref{2.8} that \begin{equation} | \varphi ^{(\nu )}(x,\lambda ) -\widetilde{\varphi }^{(\nu )}(x,\lambda ) | \leq C| \lambda | ^{\nu -1}\exp (| \tau | x) . \label{5.5} \end{equation} Similarly, \begin{equation} | \psi ^{(\nu )}(x,\lambda ) -\widetilde{\psi} ^{(\nu )}(x,\lambda ) | \leq C| \lambda | ^{\nu -1}\exp (| \tau | (\pi -x) ) . \label{5.6} \end{equation} Denote $G_{\delta }^{0}=G_{\delta }\cap \widetilde{G}_{\delta }$. In view of \eqref{2.12}, \eqref{2.14}, \eqref{4.1} and \eqref{5.6} we obtain \begin{equation} | \Phi ^{(\nu )}(x,\lambda ) -\widetilde{\Phi} ^{(\nu )}( x,\lambda ) | \leq C_{\delta }| \lambda | ^{\nu -2}\exp (-| \tau | x) ,\text{ }\lambda \in G_{\delta }^{0}. \label{5.7} \end{equation} Let $P(x,\lambda )$ be the matrix defined in Section 3. Since for each fixed $x$, the functions $P_{1k}(x,\lambda )$ are meromorphic in $\lambda $ with simple poles $\lambda _n$ and $\widetilde{\lambda }_n$, we obtain by Cauchy's theorem \cite[p.84]{6} \begin{equation} P_{1k}(x,\lambda )-\delta _{1k}=\frac{1}{2\pi i}\int_{\Gamma _n}\frac{ P_{1k}(x,\zeta )-\delta _{1k}}{\lambda -\zeta },\text{ }k=1,2, \label{5.9} \end{equation} where $\lambda \in int\Gamma _n$, and $\delta _{jk}$ is the Kronecker delta. Further, \eqref{4.3} and \eqref{4.5} imply \begin{equation} P_{11}(x,\lambda )=1+(\varphi (x,\lambda ) -\widetilde{ \varphi }(x,\lambda ) ) \widetilde{\Phi }'( x,\lambda ) -(\Phi (x,\lambda ) -\widetilde{\Phi } (x,\lambda ) ) \widetilde{\varphi }'(x,\lambda ) . \label{5.10} \end{equation} Using \eqref{2.11}, \eqref{2.12}, \eqref{4.5}, \eqref{5.5}, \eqref{5.7} and \eqref{5.10} we infer \begin{equation} | P_{1k}(x,\lambda )-\delta _{1k}| \leq C_{\delta }| \lambda | ^{-1},\quad \lambda \in G_{\delta }^{0}. \label{5.11} \end{equation} By \eqref{5.9} and \eqref{5.11} \begin{equation*} P_{1k}(x,\lambda )-\delta _{1k}=\lim_{n\to \infty }\frac{1}{2\pi i} \int_{\Gamma _{n1}}\frac{P_{1k}(x,\zeta )}{\zeta -\lambda }d\zeta , \end{equation*} where $\Gamma _{n1}=\{ \lambda :| \lambda | =| \lambda _n^{0}| ,\text{ }n=0,\pm 1,\pm 2,\dots\}$. Substituting this into \eqref{4.6}, we obtain \begin{equation*} \varphi (x,\lambda ) =\widetilde{\varphi }(x,\lambda ) +\lim_{n\to \infty }\frac{1}{2\pi i}\int_{\Gamma _{n1}}\frac{ \widetilde{\varphi }(x,\lambda ) P_{11}(x,\zeta )+\widetilde{ \varphi }'(x,\lambda ) P_{12}(x,\zeta )}{\lambda -\zeta }d\zeta . \end{equation*} Taking \eqref{4.1} and \eqref{4.5} into account we calculate \begin{equation} \widetilde{\varphi }(x,\lambda ) =\varphi (x,\lambda ) +\lim_{n\to \infty }\frac{1}{2\pi i} \int_{\Gamma _{n1}}\frac{\langle \widetilde{\varphi }(x,\lambda ) , \widetilde{ \varphi }(x,\zeta ) \rangle }{\lambda -\zeta }\widetilde{M} (\zeta ) \varphi (x,\zeta ) d\zeta . \label{5.12} \end{equation} It follows from \eqref{4.8} that \begin{equation*} \operatorname{Re}s_{\zeta =\lambda _{kj}}\frac{\langle \widetilde{\varphi } (x,\lambda ) ,\widetilde{\varphi }(x,\zeta ) \rangle }{\lambda -\zeta }\widetilde{M}(\zeta ) \varphi (x,\zeta ) =\widetilde{Q}_{kj}(x,\lambda ) \varphi _{kj}(x). \end{equation*} Consequently, calculating the integral in \eqref{5.12} by the residue theorem \cite[p.112]{6} and then taking $\lambda =\lambda _{ni}$, we arrive at \eqref{5.4} as $n\to \infty $. Furthermore, according to asymptotic formulaes \eqref{5.2} and \eqref{5.3}, we derive for $x\in (a_s,a_{s+1}) $, $0\leq s\leq m$ that the series converges absolutely and uniformly on $x\in [ 0,\pi] \backslash \{ a_s\}_{s=1}^{m}$. \end{proof} From the above arguments, it is seen that, for each fixed $x\in [ 0,\pi ] \backslash \{ a_s\} _{s=1}^{m}$, the relation \eqref{5.4} can be considered as a system of linear equations with respect to $\varphi _{ni}(x)$ for $n=0,\pm 1,\pm 2,\dots$ and $i=0,1$. But the series in \eqref{5.4} converges only ``with brackets''. Therefore, it is not convenient to use \eqref{5.4} as a main equation of the inverse problem. Let $V$ be a set of indices $u=(n,i) $, $n=0,\pm 1,\pm 2,\dots$ and $i=0,1$. For each fixed $x\in [ 0,\pi ] \backslash \{a_s\} _{s=1}^{m}$, we define the vectors \begin{equation*} \phi (x)=[ \phi _{u}(x)] _{u\in V}=[ \begin{bmatrix} \phi _{n0}(x) \\ \phi _{n1}(x) \end{bmatrix},\quad \widetilde{\phi }(x)=[ \widetilde{\phi }_{u}(x)] _{u\in V} =\begin{bmatrix} \widetilde{\phi }_{n0}(x) \\ \widetilde{\phi }_{n1}(x) \end{bmatrix} _{n=0,\pm 1,\pm 2,\dots} \end{equation*} by the formulas \begin{equation} \begin{bmatrix} \phi _{n0}(x) \\ \phi _{n1}(x) \end{bmatrix} =\begin{bmatrix} \zeta _n^{-1} & -\zeta _n^{-1} \\ 0 & 1 \end{bmatrix} \begin{bmatrix} \varphi _{n0}(x) \\ \varphi _{n1}(x) \end{bmatrix},\quad \begin{bmatrix} \widetilde{\phi }_{n0}(x) \\ \widetilde{\phi }_{n1}(x) \end{bmatrix}= \begin{bmatrix} \zeta _n^{-1} & -\zeta _n^{-1} \\ 0 & 1 \end{bmatrix} \begin{bmatrix} \widetilde{\varphi }_{n0}(x) \\ \widetilde{\varphi }_{n1}(x) \end{bmatrix}. \label{5.13} \end{equation} If $\zeta _n=0$ for a certain $n$, then we put $\phi _{n0}(x)=\widetilde{\phi }_{n0}(x)=0$. Further, we define the block matrix \begin{equation*} H(x)=[ H_{u,v}(x)] _{u,v\in V}= \begin{bmatrix} H_{n0,k0}(x) & H_{n0,k1}(x) \\ H_{n1,k0}(x) & H_{n1,k1}(x) \end{bmatrix}_{n,k=0,\pm 1,\pm 2,\dots}, \end{equation*} where $u=(n,i) $, $v=(k,j) $ and \begin{equation*} \begin{bmatrix} H_{n0,k0}(x) & H_{n0,k1}(x) \\ H_{n1,k0}(x) & H_{n1,k1}(x) \end{bmatrix}= \begin{bmatrix} \zeta _n^{-1} & -\zeta _n^{-1} \\ 0 & 1 \end{bmatrix} \begin{bmatrix} Q_{n0,k0}(x) & Q_{n0,k1}(x) \\ Q_{n1,k0}(x) & Q_{n1,k1}(x) \end{bmatrix} \begin{bmatrix} \zeta _{k} & \zeta _{k} \\ 0 & -1 \end{bmatrix}. \end{equation*} Analogously we define $\widetilde{\phi }(x)$, $\widetilde{H}(x)$ by replacing in the previous definitions $\varphi _{ni}(x)$, $Q_{ni,kj}(x)$ by $\widetilde{\varphi }_{ni}(x)$, $\widetilde{Q}_{ni,kj}(x)$, respectively. It follows from \eqref{2.7}-\eqref{2.10}, \eqref{2.15}, \eqref{2.16}, \eqref{5.2}, \eqref{5.3} and Schwarz's lemma that \begin{gather} | \phi _{nj}^{(\nu )}(x) | ,\text{ } | \widetilde{\phi }_{nj}^{(\nu )}(x) | \leq C(| \lambda _n^{0}| +1) ^{\nu },\quad \nu =0,1, \nonumber \\ | H_{ni,kj}(x)| ,\text{ }| \widetilde{H} _{ni,kj}(x)| \leq \frac{C\zeta _{k}}{| \lambda _n^{0}-\lambda _{k}^{0}| +1}, \label{5.14} \\ | H_{ni,kj}^{(\nu +1) }(x)| ,\; | \widetilde{H}_{ni,kj}^{(\nu +1) }(x)| \leq C(| \lambda _n^{0}| +| \lambda _{k}^{0}| +1) ^{\nu }\zeta _{k},\quad \nu =0,1. \label{5.15} \end{gather} Let us consider the Banach space $B$ of bounded sequences $\alpha =[\alpha _{u}] _{u\in V}$ with the norm $\| \alpha \|_{B}=\sup_{u\in V}| \alpha _{u}| $. It follows from \eqref{5.14}, \eqref{5.15} that for each fixed $x\in (a_s,a_{s+1}) $, $0\leq s\leq m$, the operator $E+\widetilde{H}(x)$ and $E-H(x)$ (here $E$ is the identity operator), acting from $B$ to $B$, are linear bounded operator, and \begin{equation*} \| H(x)\| ,\| \widetilde{H}(x)\| \leq C\sup_n\sum_{k=-\infty }^{\infty }\frac{\zeta _{k}}{| \lambda _n^{0}-\lambda _{k}^{0}| +1}<\infty . \end{equation*} Here, we give the main result of the section. \begin{theorem}\label{theorem 5.1} For each fixed $x\in [ 0,\pi ] \backslash \{ a_s\} _{s=1}^{m}$, the vector $\phi (x)\in B$ satisfies the equation \begin{equation} \widetilde{\phi }(x)=(E+\widetilde{H}(x)) \phi (x) \label{5.16} \end{equation} in the Banach space $B$. Moreover, the operator $E+\widetilde{H}(x)$ has a bounded inverse operator, i.e. the equation \eqref{5.16} is uniquely solvable. \end{theorem} \begin{proof} Using the notation $\widetilde{\phi }(x)$ we rewrite \eqref{5.4} as \begin{equation*} \widetilde{\phi }_{ni}(x)=\phi _{ni}(x) +\sum_{k,j}\widetilde{H} _{ni,kj}(x)\phi _{kj}(x),\quad (n,i) \in V,\quad (k,j) \in V, \end{equation*} which is equivalent to \eqref{5.4}. Similarly, using the notation $H(x)$: \begin{equation*} H_{ni,lj}(x)-\widetilde{H}_{ni,lj}(x)+\sum_{k,t}\widetilde{H} _{ni,kt}(x)H_{kt,lj}(x)=0 \end{equation*} for $(n,i) $, $(l,j) $, $(k,t) \in V$ or in the other form \begin{equation*} (E+\widetilde{H}(x)) (E-H(x)) =E. \end{equation*} Interchanging places for $L$ and $\widetilde{L}$, we obtain analogously \begin{equation*} \phi (x)=(E-H(x)) \widetilde{\varphi }(x),\quad (E-H(x)) (E+\widetilde{H}(x)) =E. \end{equation*} Hence the operator $(E+\widetilde{H}(x)) ^{-1}$ exists, and it is a linear bounded operator. \end{proof} Equation \eqref{5.16} is called the \textit{basic }equation of the inverse problem. Solving \eqref{5.16} we find the vector $\phi (x)$, and consequently, the functions $\varphi _{ni}(x)$. Thus, we obtain the following algorithms for the solution of our inverse problems. \begin{algorithm} \label{a1} \rm Given the spectral data $\{ \lambda _n,\gamma _n\}_{n=0,\pm 1,\pm 2,\dots}$, construct $q(x)$ and $h,H;a_s,\alpha _s$, $s=\overline{1,m}$. \begin{enumerate} \item Choose $\widetilde{L}$ and calculate $\widetilde{\phi }(x)$ and $ \widetilde{H}(x)$; \item Find $\phi (x)$ by solving the equation \eqref{5.16} and calculate $ \varphi _{no}(x)$ via \eqref{5.13}; \item Choose some $n$ (e.g., $n=0$) and construct $q(x)$ and $ h,H;a_s,\alpha _s$, $s=\overline{1,m}$ by formulae \begin{gather*} q(x)=\frac{\varphi _{n0}''(x)}{\varphi _{n0}(x)}+\lambda _n, \quad h=\varphi _{n0}'(0),\quad H=-\frac{\varphi _{n0}'(\pi )}{\varphi _{n0}(\pi )}; \\ \varphi _{n0}(a_s+0)=\varphi _{n0}(a_s-0); \quad \alpha _s=\frac{\varphi _{n0}'(a_s+0) -\varphi _{n0}'(a_s-0)}{2\lambda _n\varphi _{n0}(a_s-0)};\quad s=\overline{1,m}. \end{gather*} \end{enumerate} \end{algorithm} \begin{algorithm} \label{a2} \rm Given $M(\lambda )$, we construct $q(x)$ and $h,H;a_s,\alpha _s$, $s=\overline{1,m}$. \begin{enumerate} \item According to \eqref{4.9} construct the spectral data $\{ \lambda _n,\gamma _n\} _{n=0,\pm 1,\pm 2,\dots}$. \item By Algorithm \ref{a1}, construct $q(x)$ and $h,H;a_s,\alpha _s$, $s=\overline{1,m}$. \end{enumerate} \end{algorithm} \begin{algorithm} \label{a3} \rm Given two spectra $\{ \lambda _n,\mu _n\} _{n=0,\pm 1,\pm 2,\dots}$, we construct $q(x)$ and $h,H;a_s,\alpha _s$, $s= \overline{1,m}$. \begin{enumerate} \item By \eqref{4.2} calculate $M(\lambda )$; \item By considering Algorithm \ref{a2}, we construct $q(x)$ and $h,H;a_s,\alpha _s$, $s=\overline{1,m}$. \end{enumerate} \end{algorithm} \begin{thebibliography}{99} \bibitem{2} Albeverio, S.; Gesztesy, F.; H$\phi $egh-Krohn, R.; Holden, H.; (with an appendix by Exner, P.) \textit{Solvable Models in Quantum Mechanics} \textit{(second edition).} AMS Chelsea Publ., (2005). \bibitem{3} Amirov, R. Kh.; On Sturm-Liouville operators with discontinuty conditions inside an interval,\textit{J. Math. Anal. Appl.}, 317 (2006), 163-176. \bibitem{4} Bellmann, R.; Cooke, K.; \textit{Differential-Difference Equations,} Academic Press, New York, 1963. \bibitem{B1} Buterin, S. A.; On inverse problem for non-selfadjoint Sturm-Liouville operator on a finite interval, \textit{J. Math. Anal. Appl.,} 335 (2007), 739-749. \bibitem{B2} Buterin, S. A.; Shieh, C.-T.; Yurko, V. A.; Inverse spectral problems for non-selfadjoint second-order differential operators with Dirichlet boundary conditions, \textit{Boundary Value Problems} 2013, 2013:180, 24p. \bibitem{5} Coddington, E. A.; Levinson, N.; \textit{Theory of Ordinary Differential Equations}, McGraw-Hill, New York, (1955). \bibitem{6} Conway, J. B.; \textit{Functions of One Complex Variable,} vol. 1, Springer, New York, USA 2nd edition, 1995. \bibitem{8} Frelling, G.; Yurko, V.; \textit{Inverse Sturm-Liouville Problems and Their Applications.} Nova Science Publ., Inc: Huntington, NY, 2001. \bibitem{9} Frelling, G.; Yurko, V.; Recovering nonselfadjoint differential pensils with nonseperated boundary conditions. \textit{Applicable Analysis: An Intern. J.,} DOI: 10.1080/00036811. 2014.940918. \bibitem{10} Gasymov, M. G.; Guseinov, G. Sh.; Determination of a diffusion operator from spectral data. \textit{Dokl. Akad. Nauk Azerb. SSR}, 37:2 (1980), 19-23. \bibitem{11} Guo, Y., Wei, G.; On the Reconstuction of the Sturm-Liouville Problems with Spectral Parameter in the Discontinuity Conditions, \textit{Results. Math.}, 65 (2014), 385-398. \bibitem{12} Guseinov, I. M.; Mammadova, L. I.; Reconstruction of the diffusion equation with singular coefficients for two spectra. \textit{Dokl. Akad. Nauk, }471:1 (2014), 13-16; English transl.: \textit{Doklady Mathematics, }90:1 (2014), 401-404. \bibitem{13} Hald, O. H.; Discontinuous inverse eigenvalue problems. \textit{Comm. on Pure and Appl. Math}., 37:5 (1986), 53-72. \bibitem{15} Jonas, P.; On the spectral theory of operators associated with perturbed Klein-Gordon and wave type equations. \textit{J. Operator Theory}, 29 (1993), 207-224. \bibitem{16} Levitan, B. M.; \textit{Inverse Sturm-Liouville Problems.} VSP: Zeist, 1987. \bibitem{17} Markus, A. S.; \textit{Introduction to the Spectral Theory of Polynomial Operator Pensils}. Shtinitsa, Kishinev, 1986; English transl., AMS, Providense, 1988. \bibitem{18} Marchenko, V. A.; \textit{Sturm-Liouville Operators and Their Applications}. Operator Theory: Advanced and Application, Birkhauser: Basel, 22 (1986). \bibitem{19} Manafov, M. Dzh.; Half-inverse spectral problem for differential pensils with interaction-point and eigenvalue-dependent boundary conditions. \textit{Hacettepe J. of. Math. and Stats}., 42:4 (2013), 339-345. \bibitem{20} Manafov, M. Dzh.; Kablan, A.; Inverse spectral and inverse nodal problems for energy-dependent Sturm-Liouville equations with $\delta $-interaction. \textit{Electronic J. of Diff. Equa.}, vol. 2015 (2015), no. 26, 1-10. \bibitem{21} Shepelsky, D. G.; The inverse problem of reconstruction of the mediums conductivity in a class of discontinuous and increasing functions. \textit{Advances in Soviet Mathematics}, 19 (1994), 209-231. \bibitem{23} Yurko, V. A.; Boundary Value Problems with Discontinuity Conditions in an Interior Point of the Interval. Differential'nye Uravneniya, 36 (2000), No. 8, 1139-1140; translation in Differential Equations, 36 (2000), No. 8, 1266-1269. \bibitem{24} Yurko, V. A.; An inverse problem for pensils of differentials operators. \textit{Math. Sb.}, 191:10 (2000), 137-160; English transl.: \textit{Sb. Math}., 191:10 (2000), 1561-1586. \end{thebibliography} \end{document}