Sbornik: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
License agreement
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Mat. Sb.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Sbornik: Mathematics, 2023, Volume 214, Issue 1, Pages 39–57
DOI: https://doi.org/10.4213/sm9709e
(Mi sm9709)
 

Structure of the spectrum of a nonselfadjoint Dirac operator

A. S. Makin

MIREA — Russian Technological University, Moscow, Russia
References:
Abstract: For the Dirac operator with two-point boundary conditions and an arbitrary complex-valued $L_2$-integrable potential $V(x)$ the spectral problem is considered. Necessary and sufficient conditions on an entire function to be the characteristic function of such a boundary value problem are obtained. Necessary and sufficient conditions on the spectrum of the above operator are established in the case when the boundary conditions are regular.
Bibliography: 16 titles.
Keywords: Dirac operator, characteristic function, spectrum.
Received: 15.12.2021 and 31.08.2022
Russian version:
Matematicheskii Sbornik, 2023, Volume 214, Number 1, Pages 43–60
DOI: https://doi.org/10.4213/sm9709
Bibliographic databases:
Document Type: Article
MSC: Primary 34L40; Secondary 34A55
Language: English
Original paper language: Russian

§ 1. Introduction

An important class of inverse spectral problems is the problem of recovering a system of differential equations from its spectral data. Problems of this kind for the Dirac and Dirac-type operators are best studied. In particular, these problems for the canonical Dirac system on a finite interval

$$ \begin{equation} B\mathbf{y}'+V\mathbf{y} =\lambda\mathbf{y}, \end{equation} \tag{1.1} $$
where $\mathbf{y}=\operatorname{col}(y_1(x),y_2(x))$,
$$ \begin{equation*} B=\begin{pmatrix} 0&1 \\ -1&0 \end{pmatrix}\quad\text{and} \quad V(x)=\begin{pmatrix} p(x)&q(x) \\ q(x)&-p(x) \end{pmatrix}, \end{equation*} \notag $$
are well studied when the operator is selfadjoint. In the case of Dirichlet or Neumann boundary conditions, the continuous potential was recovered from two spectra in [1]. Similar results for the Dirac operator with an integrable potential were deduced in [2]. The above problems for nonseparable boundary conditions, including periodic, antiperiodic and quasi-periodic ones, were solved in [3] and [4]. In the nonselfadjoint case the problem of the recovery of the potential $V(x)$ from the spectral data is much more complicated, since many methods used successfully for selfadjoint operators are inapplicable. For example, the spectra of periodic (antiperiodic) boundary-value problems for the operator (1.1) were characterized in [3] in terms of special conformal mappings, which do not exist for complex-valued potentials, while the alternation property of eigenvalues of the corresponding Dirichlet and Neumann problems, which is often used to prove the solvability of the basic equation, has no sense in the complex case. Inverse problems for nonselfadjoint differential systems of the first order were considered in [5] and [6]. Among the recent publications, we note [7].

This paper studies the Dirac system (1.1) for complex-valued functions $p$ and $q$ in the space $L_2(0,\pi)$ $(V\in L_2(0,\pi))$, with the two-point boundary conditions

$$ \begin{equation} \begin{gathered} \, U_1(\mathbf{y})= a_{11}y_1(0)+a_{12}y_2(0)+a_{13}y_1(\pi)+ a_{14}y_2(\pi)=0, \\ U_2(\mathbf{y})= a_{21}y_1(0)+a_{22}y_2(0)+a_{23}y_1(\pi)+ a_{24}y_2(\pi)=0, \end{gathered} \end{equation} \tag{1.2} $$
where the coefficients $a_{ij}$ are arbitrary complex numbers and the matrix
$$ \begin{equation*} A=\begin{pmatrix} a_{11}&a_{12}&a_{13} &a_{14} \\ a_{21}&a_{22}&a_{23} &a_{24} \end{pmatrix} \end{equation*} \notag $$
has linearly independent rows. Our main goal is to investigate the structure of the spectrum of the eigenvalue problem for (1.1) with boundary conditions like (1.2) and nonsmooth complex-valued potential $V(x)$.

We let $\|\mathbf{f}\|=(|f_1|^2+|f_2|^2)^{1/2}$ denote the norm of an arbitrary vector $\mathbf{f}=\operatorname{col}(f_1,f_2)\in\mathbb{C}^2$ and set $\langle \mathbf{f},\mathbf{g}\rangle=f_1\overline g_1+f_2\overline g_2$; we let $\|W\|= \sup_{\|\mathbf{f}\|=1}\|W\mathbf{f}\|$ denote the norm of an arbitrary $2\times2$ matrix $W$. We let $L_{2,2}(a,b)$ denote the space of two-dimensional vector functions $\mathbf{f}(t)=\operatorname{col}(f_1(t),f_2(t))$ with norm $\displaystyle\|\mathbf{f}\|_{L_{2,2}(a,b)}=\biggl(\int_a^b\|\mathbf{f}(t)\|\,dt\biggr)^{1/2}$ and $L_{2,2}^{2,2}(a,b)$ denote the space of matrix functions $W(t)$ of size $2\,{\times}\,2$ with norm $\displaystyle\|W\|_{L_{2,2}^{2,2}(a,b)}=\biggl(\int_a^b\|W(t)\|\,dt\biggr)^{1/2}$. The operator $\mathbb{L}\mathbf{y}={B\mathbf{y}'+V\mathbf{y}}$ is regarded as a linear operator in the space $L_{2,2}(0,\pi)$ with domain $D(\mathbb{L})=\{\mathbf{y}\in W_1^1[0,\pi]\colon \mathbb{L}\mathbf{y}\in L_{2,2}(0,\pi),\, U_j(\mathbf{y})=0,\, j=1,2\}$.

We let

$$ \begin{equation} E(x,\lambda)=\begin{pmatrix} c_1(x,\lambda)&-s_2(x,\lambda) \\ s_1(x,\lambda)&c_2(x,\lambda) \end{pmatrix} \end{equation} \tag{1.3} $$
denote the matrix of the fundamental system of solutions of equation (1.1) with boundary conditions $E(0,\lambda)=I$, where $I$ is the identity matrix, and $E_0(x,\lambda)$ denotes the matrix of the fundamental system of solutions of the nonperturbed equation $B\mathbf{y}'=\lambda\mathbf{y}$ with boundary conditions $E_0(0,\lambda)=I$. It is obvious that
$$ \begin{equation*} E_0(x,\lambda)=\begin{pmatrix} \cos\lambda x&-\sin\lambda x \\ \sin\lambda x&\cos\lambda x \end{pmatrix}. \end{equation*} \notag $$
It is well known that the entries of $E(x,\lambda)$ satisfy
$$ \begin{equation} c_1(x,\lambda)c_2(x,\lambda)+s_1(x,\lambda)s_2(x,\lambda)=1 \end{equation} \tag{1.4} $$
for any $x$ and $\lambda$.

Eigenvalues of the problem (1.1), (1.2) are roots of the characteristic equation

$$ \begin{equation*} \Delta(\lambda)=0, \end{equation*} \notag $$
where
$$ \begin{equation*} \Delta(\lambda)= \biggl|\begin{matrix} U_1(E^{[1]}(\,\cdot\,,\lambda)) &U_1(E^{[2]}(\,\cdot\,,\lambda)) \\ U_2(E^{[1]}(\,\cdot\,,\lambda)) &U_2(E^{[2]}(\,\cdot\,,\lambda)) \end{matrix} \biggr| \end{equation*} \notag $$
and $E^{[k]}(x,\lambda)$ is the $k$th column of the matrix (1.3).

We let $J_{ij}$ denote the determinant formed of the $i$th and $j$th columns of the matrix $A$. We introduce the notation $J_0=J_{12}+J_{34}$, $J_1=J_{14}-J_{23}$ and $J_2=J_{13}+J_{24}$.

It is well known that the characteristic function $\Delta(\lambda)$ of the problem (1.1), (1.2) is reducible to the form

$$ \begin{equation} \begin{aligned} \, \notag \Delta(\lambda) &=J_{12}+J_{34}+J_{14}c_2(\pi,\lambda)-J_{23}c_1(\pi,\lambda)-J_{13}s_2(\pi,\lambda) -J_{24}s_1(\pi,\lambda) \\ &=\Delta_0(\lambda)+\int_0^\pi r_1(t)e^{-i\lambda t}\,dt+\int_0^\pi r_2(t)e^{i\lambda t}\,dt, \end{aligned} \end{equation} \tag{1.5} $$
where
$$ \begin{equation} \Delta_0(\lambda)=J_0+J_1\cos\pi\lambda-J_2\sin\pi\lambda=J_0+\frac{J_1+iJ_2}{2}\,e^{i\pi\lambda} +\frac{J_1-iJ_2}{2}\,e^{-i\pi\lambda} \end{equation} \tag{1.6} $$
is the characteristic function of the nonperturbed problem
$$ \begin{equation} B\mathbf{y}'=\lambda\mathbf{y}, \qquad U(\mathbf{y})=0, \end{equation} \tag{1.7} $$
and the functions $r_j$ are in $L_2(0,\pi)$, $j=1,2$.

The boundary conditions (1.2) can be divided into four basic types.

Definition 1. The boundary conditions (1.2) are called regular if

$$ \begin{equation} J_1^2+J_2^2=(J_{14}+J_{32})^2+(J_{13}+J_{24})^2\ne0 \end{equation} \tag{1.8} $$
and strongly regular if, in addition,
$$ \begin{equation} J_0^2\ne J_1^2-J_2^2. \end{equation} \tag{1.9} $$

Definition 2. The boundary conditions (1.2) are called regular but not strongly regular if (1.8) holds but (1.9) does not, that is,

$$ \begin{equation*} J_0^2=J_1^2-J_2^2. \end{equation*} \notag $$

Definition 3. The boundary conditions (1.2) are called irregular if

$$ \begin{equation*} J_0\ne0, \quad J_1+iJ_2\ne 0 \quad\text{and}\quad J_1-iJ_2=0; \end{equation*} \notag $$
or
$$ \begin{equation*} J_0\ne0, \quad J_1+iJ_2=0 \quad\text{and} \quad J_1-iJ_2\ne0. \end{equation*} \notag $$

Definition 4. Boundary conditions (1.2) are called degenerate if

$$ \begin{equation*} J_1=J_2=0; \quad\text{or}\quad J_0=0, \quad J_1+iJ_2\ne0 \quad\text{and} \quad J_1-iJ_2=0; \end{equation*} \notag $$
or
$$ \begin{equation*} J_0=0, \quad J_1+iJ_2=0 \quad\text{and} \quad J_1-iJ_2\ne0. \end{equation*} \notag $$
It is straightforward to see that the boundary conditions (1.2) are degenerate if and only if either the characteristic equation $\Delta_0(\lambda)=0$ has no roots or $\Delta_0(\lambda)\equiv0$.

We introduce the notation $c_j(\lambda)=c_j(\pi,\lambda)$ and $s_j(\lambda)=s_j(\pi,\lambda)$, $j=1,2$. We also let $\mathrm{PW}_\sigma$ denote the class of entire functions $f(z)$ of exponential type not exceeding $\sigma$ such that $\|f\|_{L_2(R)}<\infty$. It is known (see [8]) that the functions $c_j(\lambda)$ and $s_j(\lambda)$ admit the representations

$$ \begin{equation} c_j(\lambda)= \cos\pi\lambda+g_j(\lambda)\quad\text{and} \quad s_j(\lambda)=\sin\pi\lambda+h_j(\lambda), \end{equation} \tag{1.10} $$
where $g_j,h_j\in \mathrm{PW}_\pi$, $j=1,2$.

Lemma (see [3]). Two entire functions $u(\lambda)$ and $v(\lambda)$ admit the representations

$$ \begin{equation} u(\lambda)=\sin\pi \lambda+h(\lambda)\quad\textit{and} \quad v(\lambda)=\cos\pi\lambda+g(\lambda), \end{equation} \tag{1.11} $$
where $h,g\in \mathrm{PW}_\pi$, if and only if
$$ \begin{equation*} u(\lambda)=-\pi(\lambda_0 -\lambda)\prod_{n=-\infty,\,n\ne0}^\infty \frac{\lambda_n-\lambda}{n} \end{equation*} \notag $$
for $\lambda_n=n+\epsilon_n$, $\{\epsilon_n\}\in l_2$, and
$$ \begin{equation*} v(\lambda)=\prod_{n=-\infty}^\infty \frac{\lambda_n-\lambda}{n-1/2}, \end{equation*} \notag $$
for $\lambda_n=n-1/2+\kappa_n$, $\{\kappa_n\}\in l_2$.

The convergence of infinite products is understood in the sense of principal value.

§ 2. Characteristic function

In this paper we consider the problem (1.1), (1.2) under the assumptions

$$ \begin{equation} J_{13}\ne J_{24}, \quad J_{14}=0, \quad J_{13}\ne0\quad\text{and}\quad J_{24}\ne0. \end{equation} \tag{2.1} $$
Relations (2.1) are satisfied by a wide class of boundary conditions, for example, conditions specified by a matrix
$$ \begin{equation*} A=\begin{pmatrix} a_1&b_1&c_1 &d_1 \\ 0&b_2&c_2 &0 \end{pmatrix}, \end{equation*} \notag $$
such that $a_1d_1b_2c_2\ne0$ and $ b_2d_1\ne -a_1c_2$. These include:

$\bullet$ strongly regular conditions for

$$ \begin{equation*} A=\begin{pmatrix} 1&1&1 &1 \\ 0&1&2 &0 \end{pmatrix}; \end{equation*} \notag $$

$\bullet$ regular but not strongly regular conditions for

$$ \begin{equation*} A=\begin{pmatrix} 1&1&1+\sqrt{2} &1 \\ 0&1&2 &0 \end{pmatrix}; \end{equation*} \notag $$

$\bullet$ irregular conditions for

$$ \begin{equation*} A=\begin{pmatrix} 1&1&3-i &1 \\ 0&1&2 &0 \end{pmatrix}; \end{equation*} \notag $$

$\bullet$ degenerate conditions for

$$ \begin{equation*} A=\begin{pmatrix} 1&1&\dfrac{1+3i}{2}&2 \\ 0&2&1 &0 \end{pmatrix}. \end{equation*} \notag $$

We consider the Dirac system (1.1), (1.2), (2.1). It follows from (1.5), (1.6) that the characteristic function $\Delta(\lambda)$ of this problem reduces to

$$ \begin{equation} \Delta(\lambda)=J_0-J_{23} c_1(\lambda)-J_{13}s_2(\lambda)-J_{24}s_1(\lambda)=\Delta_0(\lambda)+f(\lambda), \end{equation} \tag{2.2} $$
where $\Delta_0(\lambda)=J_0-(J_{13}+J_{24})\sin\pi\lambda-J_{23} \cos\pi\lambda$ and $f\in \mathrm{PW}_\pi$. The converse assertion is also true.

Theorem 1. For any function $f\in\mathrm{PW}_\pi$ there exists $V\in L_2(0,\pi)$ such that the characteristic function $\Delta(\lambda)$ of problem (1.1), (1.2), (2.1) with potential $V(x)$ satisfies (2.2).

Proof. Let $f$ be an arbitrary function in the class $\mathrm{PW}_\pi$. The Paley-Wiener theorem and Lemma 1.3.1 in [9] imply that
$$ \begin{equation} \lim_{|\lambda|\to\infty}e^{-\pi|{\operatorname{Im}\lambda}|}f(\lambda)=0. \end{equation} \tag{2.3} $$
Let $\{\lambda_n\}$, $n\in\mathbb{Z}$, be a strictly monotonically increasing sequence of real numbers satisfying the conditions
$$ \begin{equation*} \lambda_n=n-\frac12\ \text{ for }n>N_0\quad\text{and}\quad \lambda_n=-\lambda_{-n+1}\ \text{ for each integer } n, \end{equation*} \notag $$
which we call conditions $(*)$.

We introduce the notation

$$ \begin{equation*} c(\lambda)=\prod_{n=-\infty}^\infty \frac{\lambda_n-\lambda}{n-1/2}. \end{equation*} \notag $$

The above lemma yields

$$ \begin{equation} c(\lambda)=\cos\pi \lambda+g(\lambda), \end{equation} \tag{2.4} $$
where $g\in \mathrm{PW}_\pi$. It follows from the Paley-Wiener theorem and Lemma 1.3.1 in [9] that
$$ \begin{equation*} \lim_{|\lambda|\to\infty}e^{-\pi |{\operatorname{Im}\lambda}|}g(\lambda)=0; \end{equation*} \notag $$
therefore,
$$ \begin{equation} |c(\lambda)|\geqslant c_0e^{\pi |{\operatorname{Im}\lambda}|} \end{equation} \tag{2.5} $$
for $|{\operatorname{Im}\lambda}|\geqslant M$, where $c_0>0$ and $M$ is some sufficiently large number.

Differentiating (2.4) we obtain

$$ \begin{equation} \dot c(\lambda)=-\pi\sin\pi \lambda+\dot g(\lambda). \end{equation} \tag{2.6} $$
Since the function $\dot g$ is in $\mathrm{PW}_\pi$, according to [3], we have
$$ \begin{equation*} \dot c(\lambda_n)=-\pi\sin\pi \lambda_n+\tau_n, \end{equation*} \notag $$
where
$$ \begin{equation*} \sum_{n=-\infty}^\infty |\tau_n|^2 <\infty. \end{equation*} \notag $$
The last equality and the definition of $\lambda_n$ imply that
$$ \begin{equation} \dot c(\lambda_n)=\pi(-1)^n+\rho_n, \end{equation} \tag{2.7} $$
where
$$ \begin{equation*} \sum_{n=-\infty}^\infty |\rho_n|^2 <\infty. \end{equation*} \notag $$
It follows that $\dot c(\lambda_n)>0$ for all even $n$ that are sufficiently large in absolute value. We can easily see that $\dot c(\lambda_n)\dot c(\lambda_{n+1})<0$ for all $n\in\mathbb{Z}$. This yields the inequality
$$ \begin{equation} (-1)^n\dot c(\lambda_n)>0 \end{equation} \tag{2.8} $$
for all $n$. Note that (2.7) implies that
$$ \begin{equation} \frac{1}{\dot c(\lambda_n)}=\frac{(-1)^n}{\pi}+\sigma_n, \end{equation} \tag{2.9} $$
where
$$ \begin{equation*} \sum_{n=-\infty}^\infty |\sigma_n|^2 <\infty. \end{equation*} \notag $$

We set

$$ \begin{equation*} \alpha=-J_{13}, \quad \beta=-J_{24}, \quad\gamma=-J_{23}\quad\text{and} \quad u_+(\lambda)=(\alpha +\beta)\sin\pi\lambda+\gamma \cos\pi\lambda+f(\lambda). \end{equation*} \notag $$
Note that
$$ \begin{equation} \alpha\ne0, \quad\beta\ne0 \quad\text{and} \quad\alpha\ne\beta. \end{equation} \tag{2.10} $$
We consider the equation
$$ \begin{equation} \alpha w^2-u_+(\lambda_n)w+\beta=0. \end{equation} \tag{2.11} $$
It has the roots
$$ \begin{equation*} s_n^\pm=\frac{u_+(\lambda_n)\pm\sqrt{u_+^2(\lambda_n)-4\alpha\beta}}{2\alpha}. \end{equation*} \notag $$
Substituting the expression for $u_+(\lambda_n)$ into the last equality we obtain
$$ \begin{equation*} \begin{aligned} \, s_n^\pm &=\frac1{2\alpha}\bigr((\alpha +\beta)\sin\pi\lambda_n+\gamma \cos\pi\lambda_n+f(\lambda_n)\bigl) \\ &\qquad \pm\frac1{2\alpha}\bigl(((\alpha +\beta)\sin\pi\lambda_n+\gamma\cos\pi\lambda_n+f(\lambda_n))^2-4\alpha\beta\bigr)^{1/2} \\ &=\frac1{2\alpha}\bigl((\alpha +\beta)\sin\pi\lambda_n+\gamma \cos\pi\lambda_n+f(\lambda_n)\bigr) \\ &\qquad \pm\frac1{2\alpha}\bigl((\alpha-\beta)^2+2(\alpha +\beta)\sin\pi\lambda_n(\gamma \cos\pi\lambda_n+f(\lambda_n)) \\ &\qquad\qquad+(\gamma \cos\pi\lambda_n+f(\lambda_n))^2-(\alpha +\beta)^2\cos^2\pi\lambda_n\bigr)^{1/2}. \end{aligned} \end{equation*} \notag $$

We introduce the following notation:

$$ \begin{equation*} \begin{aligned} \, F^\pm(\lambda) &=\frac1{2\alpha}\bigl((\alpha +\beta)\sin\pi\lambda+\gamma \cos\pi\lambda+f(\lambda)\bigr) \\ &\qquad\pm\frac1{2\alpha}\bigl(((\alpha +\beta)\sin\pi\lambda+\gamma\cos\pi\lambda+f(\lambda))^2-4\alpha\beta\bigr)^{1/2} \\ &=\frac1{2\alpha}\bigl((\alpha +\beta)\sin\pi\lambda+\gamma \cos\pi\lambda+f(\lambda)\bigr) \\ &\qquad\pm\frac1{2\alpha}\bigl((\alpha-\beta)^2+2(\alpha +\beta)\sin\pi\lambda(\gamma \cos\pi\lambda+f(\lambda)) \\ &\qquad\qquad+(\gamma \cos\pi\lambda+f(\lambda))^2-(\alpha +\beta)^2\cos^2\pi\lambda\bigr)^{1/2} \end{aligned} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{aligned} \, F_0^\pm(\lambda) &=\frac1{2\alpha}\bigl((\alpha +\beta)\sin\pi\lambda+\gamma \cos\pi\lambda\bigr) \\ &\qquad\pm\frac1{2\alpha}\bigl(((\alpha +\beta)\sin\pi\lambda+\gamma\cos\pi\lambda)^2-4\alpha\beta\bigr)^{1/2} \\ &=\frac1{2\alpha}\bigl((\alpha +\beta)\sin\pi\lambda+\gamma \cos\pi\lambda\bigr) \\ &\qquad\pm\frac1{2\alpha}\bigl((\alpha-\beta)^2+(\alpha +\beta)\gamma \sin\pi\lambda\cos\pi\lambda+(\gamma^2-(\alpha +\beta)^2)\cos^2\pi\lambda\bigr)^{1/2}. \end{aligned} \end{equation*} \notag $$

We also let $\Gamma(z,r)$ denote the circle with centre $z$ and radius $r$. Now we consider several cases.

1) $\operatorname{Im}({\beta}/{\alpha})\ne0$. Let $l_0$ be the straight line through the points $-1$ and ${\beta}/{\alpha}$ and $l$ be the line through the origin which is parallel to $l_0$. Clearly, there exists $\varepsilon_0$ such that the discs $\Gamma(-1,\varepsilon_0)$ and $\Gamma({\beta}/{\alpha},\varepsilon_0)$ lie strictly on the same side of $l$. It follows from (2.3) and (2.10) that there exists an odd positive integer $\widetilde N$ such that

$$ \begin{equation} |F^\pm(\lambda)-F_0^\pm(\lambda)|<\frac{\varepsilon_0}2 \end{equation} \tag{2.12} $$
for all $\lambda=k-1/2$, $k\in \mathbb Z$, satisfying $|\lambda|\geqslant\widetilde N-1$. Inequalities (2.10) imply that there exists $\delta_0$, $0<\delta_0<1/4$, such that
$$ \begin{equation} |F_0^\pm(\lambda)-F_0^\pm(\lambda+z)|<\frac{\varepsilon_0}2 \end{equation} \tag{2.13} $$
for $\lambda=\pm(\widetilde N-1/2)$ and $|z|<\delta_0$. We define a sequence $\{\lambda_n\}$, $n\in \mathbb Z$, as follows: $\lambda_n=n-1/2$ if $n\geqslant\widetilde N+1$, $\lambda_n=\widetilde N-1/2+\epsilon_n$, where $0<\epsilon_1<\dots<\epsilon_{\widetilde N}<\delta$, if $n=1,\dots, \widetilde N$, and $\lambda_n=-\lambda_{-n+1}$. It is obvious that $\{\lambda_n\}$ satisfies conditions $(*)$. It is easy to see that if $n\geqslant \widetilde N+1$ or $n\leqslant -\widetilde N$, then for odd $n$ the points $s_n^+$ are in the disc $\Gamma(1,\varepsilon_0)$, and for even $n$ the points $s_n^-$ are in the disc $\Gamma(-1,\varepsilon_0)$. Let $s_n=s_n^+$ for odd $n$ and $s_n=s_n^-$ for even $n$; thus, all points $(-1)^ns_n$ are in the disc $\Gamma(-1,\varepsilon_0)$.

Let $-\widetilde N+1\leqslant n\leqslant\widetilde N$. Then it follows from (2.12) and (2.13) that

$$ \begin{equation*} \biggl|F^\pm(\lambda_n)-F_0^\pm\biggl(\widetilde N-\frac12\biggr)\biggr|<\varepsilon_0. \end{equation*} \notag $$
We can easily see that, as $\widetilde N$ is odd, all points $s_n^+$ lie in the disc $\Gamma(1,\varepsilon_0)$, while all the $s_n^-$ lie in the disc $\Gamma({\beta}/{\alpha},\varepsilon_0)$. We set $s_n=s_n^+$ if $n$ is odd; then all points $(-1)^ns_n$ lie in $\Gamma(-1,\varepsilon_0)$. We set $s_n=s_n^-$ if $n$ is even; then the points $(-1)^ns_n$ lie in $\Gamma({\beta}/{\alpha},\varepsilon_0)$. Thus, all points $(-1)^ns_n$, $n\in \mathbb Z$, are strictly on one side of the line $l$.

2) $\operatorname{Im}(\beta/\alpha)=0$, $\operatorname{Re}(\beta/\alpha)<0$. There obviously exists a number $\varepsilon_0$ such that the discs $\Gamma(-1,\varepsilon_0)$ and $\Gamma(\beta/\alpha,\varepsilon_0)$ lie strictly to the left of the imaginary axis. Reasoning similarly to the previous case, we deduce that all points $(-1)^ns_n$ lie strictly to the left of the imaginary axis.

3) $\operatorname{Im}(\beta/\alpha)=0$ and $\operatorname{Re}(\beta/\alpha)>0$. We set $\widetilde \alpha=\alpha\overline\alpha$, $\widetilde \beta=\beta\overline\alpha$, $\widetilde \gamma=\gamma\overline\alpha$ and $\widetilde f(\lambda)=\overline\alpha f(\lambda)$; then $\widetilde \alpha>0$ and $\widetilde \beta>0$.

It is easy to see that

$$ \begin{equation*} \begin{aligned} \, s_n^\pm &=\frac1{2\widetilde\alpha}\bigl((\widetilde\alpha +\widetilde\beta)\sin\pi\lambda_n+\widetilde\gamma \cos\pi\lambda_n+\widetilde f(\lambda_n)\bigr) \\ &\qquad \pm\frac1{2\widetilde\alpha}\bigl(((\widetilde\alpha +\widetilde\beta)\sin\pi\lambda_n+\widetilde\gamma\cos\pi\lambda_n+\widetilde f(\lambda_n))^2-4\widetilde\alpha\widetilde\beta\bigr)^{1/2}. \end{aligned} \end{equation*} \notag $$

We introduce the notation

$$ \begin{equation*} R^\pm=\frac{\widetilde\gamma\pm\sqrt{\widetilde\gamma^2-4\widetilde\alpha\widetilde\beta}} {2\widetilde\alpha}. \end{equation*} \notag $$
It is clear that $R^+R^->0$; consequently, $\operatorname{Im}R^+\operatorname{Im}R^-<0$ or $\operatorname{Im}R^+=0=\operatorname{Im}R^-=0$.

3.1) $\operatorname{Im}R^+\operatorname{Im}R^-<0$. In this case

$$ \begin{equation} \widetilde\gamma^2-4\widetilde\alpha\widetilde\beta\ne0. \end{equation} \tag{2.14} $$
For definiteness let $\operatorname{Im}R^+>0$; then $\operatorname{Im}R^-<0$. Inequality (2.14) shows that there exists $\varepsilon_1>0$ such that $\Gamma(R^+,\varepsilon_1)$ and $\Gamma(R^-,\varepsilon_1)$ lie strictly on opposite sides of some straight line $l$ through the origin which is distinct from the real axis. Clearly, there exists $\varepsilon_2$, where $0<\varepsilon_2<\varepsilon_1$, such that the disc $\Gamma(-1,\varepsilon_2)$ does not intersect $l$; therefore, $\Gamma(-1,\varepsilon_2)$ lies strictly on one side of $l$. It follows from (2.3) and (2.10) that there exists an odd positive integer $\widehat N$ such that
$$ \begin{equation} |F^\pm(\lambda)-F_0^\pm(\lambda)|<\frac{\varepsilon_1}2 \end{equation} \tag{2.15} $$
for all $\lambda=k-1/2$, $k\in \mathbb Z$, satisfying $|\lambda|\geqslant\widehat N-1$. By virtue of (2.14) there exists $\delta_1$, $0<\delta_1<1/4$, such that
$$ \begin{equation} |F_0^\pm(\lambda-F_0^\pm(\lambda+z)|<\frac{\varepsilon_2}2 \end{equation} \tag{2.16} $$
if $\lambda=\pm(\widehat N-1)$ and $|z|<\delta_1$.

We define a sequence $\{\lambda_n\}$, $n\in \mathbb Z$, as follows: $\lambda_n=n-1/2$ if $n\geqslant\widehat N+1$, $\lambda_n=\widehat N-1+\epsilon_n$, where $0<\epsilon_1<\dots<\epsilon_{\widehat N}<\delta_1$, if $n=1,\dots, \widehat N$; also let $\lambda_n=-\lambda_{-n+1}$. It is obvious that $\{\lambda_n\}$ satisfies conditions $(*)$. It is straightforward to see that if $n\geqslant \widehat N+1$ or $n\leqslant -\widehat N$, then $s_n^+$ lies in the disc $\Gamma(1,\varepsilon_1)$ if $n$ is odd, while $s_n^-$ lies in the disc $\Gamma(-1,\varepsilon_1)$ if $n$ is even. We set $s_n=s_n^+$ if $n$ is odd and $s_n=s_n^-$ if $n$ is even; thus, all points $(-1)^ns_n$ lie in $\Gamma(-1,\varepsilon_1)$.

Let $-\widehat N+1\leqslant n\leqslant\widehat N$; then it follows from (2.15) and (2.16) that

$$ \begin{equation*} |F^\pm(\lambda_n)-F_0^\pm(\widehat N-1)|<\varepsilon_1. \end{equation*} \notag $$
Then the points $s_n^+$ belong to $\Gamma(R^+,\varepsilon_1)$, while the $s_n^-$ belong to $\Gamma(R^-,\varepsilon_1)$. Assume, for example, that the disc $\Gamma(R^-,\varepsilon_1)$ lies on the same side of $l$ as $\Gamma(-1,\varepsilon_0)$. We set $s_n=s_n^+$ if $n$ is odd and $s_n=s_n^-$ if $n$ is even; then all points $(-1)^ns_n$, $n\in \mathbb Z$, lie strictly on the same side of $l$.

3.2) $\operatorname{Im}R^+=\operatorname{Im}R^-=0$. Clearly, there exists $\varepsilon_3>0$ such that the disc $\Gamma(-1,\varepsilon_3)$ lies strictly below the line $l_1\colon y=-x$, whereas the disc $\Gamma\Bigl(i\sqrt{{\widetilde\beta}/{\widetilde\alpha}},\varepsilon_3\Bigr)$ lies strictly above and $\Gamma\Bigl(-i\sqrt{{\widetilde\beta}/{\widetilde\alpha}},\varepsilon_3\Bigr)$ strictly below $l_1$. It is obvious that ${\widetilde\gamma}/{\widetilde\alpha}$ is real; therefore, $\widetilde\gamma$ is real. We consider the equation

$$ \begin{equation*} (\widetilde\alpha +\widetilde\beta)\sin t+\widetilde\gamma \cos t=0. \end{equation*} \notag $$
Since $\widetilde\alpha +\widetilde\beta\ne0$, this equation has the roots $t_n=-\arctan({\widetilde\gamma}/(\widetilde\alpha +\widetilde\beta))+\pi n$, where $n\in \mathbb Z$. We introduce the notation
$$ \begin{equation*} h_0=\frac{-\arctan(\widetilde\gamma/(\widetilde\alpha +\widetilde\beta))}{\pi}. \end{equation*} \notag $$

Relations (2.3) and (2.10) imply that there exists a positive number $\widehat N$ such that

$$ \begin{equation} |F^\pm(\lambda)-F_0^\pm(\lambda)|<\frac{\varepsilon_3}2 \end{equation} \tag{2.17} $$
for all $\lambda=k-1/2$, $k\in \mathbb Z$, satisfying $|\lambda|\geqslant\widetilde N-1$ and also for $\lambda =\widehat\lambda=t_{\widehat N-1}/\pi$ and that there also exists $\delta_3>0$ such that
$$ \begin{equation} |F_0^\pm(\lambda)-F_0^\pm(\lambda+z)|<\frac{\varepsilon_3}2 \end{equation} \tag{2.18} $$
if $\lambda=\pm \widehat \lambda$ and $|z|<\delta_3$.

We define a sequence $\{\lambda_n\}$, $n\in \mathbb Z$, as follows: $\lambda_n=n - 1/2$ for $n \geqslant \widetilde N + 1$, $\lambda_n = \widehat\lambda + \epsilon_n$ for $1 \leqslant n \leqslant \widehat N$, where $0<\epsilon_1<\dots<\epsilon_{\widehat N}<\min(\delta,1/2-h_0)$, and $\lambda_n=-\lambda_{-n+1}$. It is clear that $\{\lambda_n\}$ satisfies conditions $(*)$. We can easily see that if $n\geqslant \widetilde N+1$ or $n\leqslant -\widetilde N$, then for odd $n$ all the $s_n^+$ lie in the disc $\Gamma(1,\varepsilon_0)$ while for even $n$ all the $s_n^-$ lie in $\Gamma(-1,\varepsilon_0)$. We set $s_n=s_n^+$ if $n$ is odd and $s_n=s_n^-$ if $n$ is even; thus, all points $(-1)^ns_n$ lie in $\Gamma(-1,\varepsilon_0)$.

Let $-\widetilde N+1\leqslant n\leqslant\widetilde N$; then $s_n^+\in\Gamma(i\sqrt{\beta/\alpha}, \varepsilon_2)$ and $s_n^-\in\Gamma(-i\sqrt{\beta/\alpha}, \varepsilon_2)$. Let $s_n=s_n^+$ if $n$ is odd and $s_n=s_n^-$ if $n$ is even. Then all points $(-1)^ns_n$ lie strictly below the line $l$.

Since $\{f(\lambda_n)\}\in l_2$ (see [10]), the definition of $s_n$ shows that

$$ \begin{equation} s_n=(-1)^{n+1}+\vartheta_n, \end{equation} \tag{2.19} $$
where $\{\vartheta_n\}\in l_2$. It follows from the definition of $s_n$ and (2.8) that in all cases all points $z_n=s_n/{\dot c(\lambda_n)}$ lie strictly on the same side of some straight line passing through the origin, while it follows from (2.9) and (2.19) that
$$ \begin{equation} z_n=-\frac{1}{\pi}+\rho_n, \end{equation} \tag{2.20} $$
where $\{\rho_n\}\in l_2$. Let $\beta_n=s_n-\sin\pi\lambda_n$; equality (2.19) implies that $\{\beta_n\}\in l_2$. We introduce the notation
$$ \begin{equation*} h(\lambda)=c(\lambda)\sum_{n=-\infty}^\infty \frac{\beta_n}{\dot c(\lambda_n)(\lambda-\lambda_n)}. \end{equation*} \notag $$
By Theorem 28 in [11] the function $h$ belongs to $\mathrm{PW}_\pi$ and $h(\lambda_n)=\beta_n$. We set $s(\lambda)=\sin\pi\lambda+h(\lambda)$. It is true that $s(\lambda_n)=s_n\ne0$; hence $s(\lambda)$ and $c(\lambda)$ have no common zeros.

Let

$$ \begin{equation*} Y_0(x,\lambda)= \begin{pmatrix} \cos\lambda x \\ \sin\lambda x \end{pmatrix}. \end{equation*} \notag $$
It was established in [7] that the system of vectors $Y_0(x,\lambda_n)$, $n\in\mathbb{Z}$, is complete in $L_{2,2}(0,\pi)$.

We set

$$ \begin{equation} F(x,t)=-\sum_{n=-\infty}^\infty \biggl(\frac{s_n}{\dot c(\lambda_n)}\bigl(Y_0(x,\lambda_n) Y_0^T(t,\lambda_n)\bigr)+\frac{1}{\pi}Y_0\biggl(x,n-\frac12\biggr) Y_0^T\biggl(t,n-\frac12\biggr)\biggr). \end{equation} \tag{2.21} $$
It follows from [12] and (2.20) that
$$ \begin{equation*} \|F(\cdot,x)\|_{L_{2,2}^{2,2}(0,\pi)}+\|F(x,\cdot)\|_{L_{2,2}^{2,2}(0,\pi)}<C, \end{equation*} \notag $$
where $C$ is independent of $x$.

We prove that for each $x\in[0,\pi]$ the homogeneous equation

$$ \begin{equation} \mathbf{f}^T(t)+\int_0^x\mathbf{f}^T(s)F(s,t)\,ds=0, \end{equation} \tag{2.22} $$
where $\mathbf{f}(t)=\operatorname{col}(f_1(t),f_2(t))$, $\mathbf{f}\in L_{2,2}(0,x)$ and $\mathbf{f}(t)=0$ for $x<t\leqslant\pi$, has only the trivial solution. Multiplying (2.22) by $\overline{\mathbf{f}^T(t)}$ and integrating the resulting equation over the interval $[0,x]$ we obtain
$$ \begin{equation} \|\mathbf{f}\|^2_{L_{2,2}(0,x)}+\int_0^x\biggl\langle\int_0^x\mathbf{f}^T(s)F(s,t)\,ds, \overline{\mathbf{f}^T(t)}\biggr\rangle\, dt=0. \end{equation} \tag{2.23} $$
Based on (2.21), we infer via simple calculations that
$$ \begin{equation} \begin{aligned} \, \notag &\mathbf{f}^T(s)F(s,t) \\ \notag &=-\biggl\{\sum_{n=-\infty}^\infty \biggl\{z_n\bigl[f_1(s)\cos\lambda_ns\cos\lambda_nt+f_2(s)\sin\lambda_ns\cos\lambda_nt, \\ \notag &\qquad\qquad f_1(s)\cos\lambda_ns\sin\lambda_nt +f_2(s)\sin\lambda_ns\sin\lambda_nt\bigr] \\ \notag &\qquad +\frac{1}{\pi}\biggl[f_1(s)\cos \biggl(n-\frac12\biggr)s\cos \biggl(n-\frac12\biggr)t+f_2(s)\sin \biggl(n-\frac12\biggr)s\cos \biggl(n-\frac12\biggr)t, \\ \notag &\qquad\qquad f_1(s)\cos \biggl(n-\frac12\biggr)s\sin \biggl(n-\frac12\biggr)t +f_2(s)\sin \biggl(n-\frac12\biggr)s\sin \biggl(n-\frac12\biggr)t\biggr]\biggr\}\biggr\} \\ \notag &=-\biggl\{\sum_{n=-\infty}^\infty \biggl\{z_n\bigl[f_1(s)\cos\lambda_ns\cos\lambda_nt+f_2(s)\sin\lambda_ns\cos\lambda_nt\bigr] \\ \notag &\qquad+\frac{1}{\pi}\biggl[f_1(s)\cos \biggl(n-\frac12\biggr)s\cos \biggl(n-\frac12\biggr)t+f_2(s)\sin \biggl(n-\frac12\biggr)s\cos \biggl(n-\frac12\biggr)t\biggr], \\ \notag &\qquad\qquad z_n\bigl[f_1(s)\cos\lambda_ns\sin\lambda_nt+f_2(s)\sin\lambda_ns\sin\lambda_nt\bigr] \\ &\qquad+\frac{1}{\pi} \biggl[f_1(s)\cos \biggl(n-\frac12\biggr) s\sin\biggl(n-\frac12\biggr)t +f_2(s)\sin \biggl(n-\frac12\biggr)s\sin \biggl(n-\frac12\biggr)t\biggr]\biggr\}\biggr\}. \end{aligned} \end{equation} \tag{2.24} $$
Substituting the right-hand side of (2.24) into the second term on the left-hand side of (2.23), transforming the iterated integrals into products of integrals, and using the fact that all numbers $\lambda_n$ are real we obtain
$$ \begin{equation*} \begin{aligned} \, &\int_0^x\biggl\langle \int_0^x \mathbf{f}^T(s)F(s,t)\,ds,\overline{\mathbf{f}^T(t)}\bigg\rangle \,dt \\ &=-\biggl\{\sum_{n=-\infty}^\infty\int_0^x\biggr(\int_0^x\biggl\{z_n\bigl[f_1(s) \cos\lambda_ns\cos\lambda_nt+f_2(s)\sin\lambda_ns\cos\lambda_nt\bigr] \\ &\quad\qquad +\frac{1}{\pi}\biggl[f_1(s)\cos \biggl(n-\frac12\biggr)s\cos \biggl(n-\frac12\biggr)t \\ &\quad\qquad +f_2(s)\sin \biggl(n-\frac12\biggr)s\cos \biggl(n-\frac12\biggr)t\biggr]\biggr\}\,ds\biggr)\overline{f_1(t)}\,dt \\ &\quad +\sum_{n=-\infty}^\infty\int_0^x\biggr(\int_0^x \biggl\{z_n\bigl[f_1(s)\cos\lambda_ns\sin\lambda_nt+f_2(s)\sin\lambda_ns\sin\lambda_nt\bigr] \\ &\quad\qquad +\frac{1}{\pi} \biggl[f_1(s)\cos \biggl(n-\frac12\biggr)s\sin \biggl(n-\frac12\biggr)t \\ &\quad\qquad+f_2(s)\sin \biggl(n-\frac12\biggr)s\sin \biggl(n-\frac12\biggr)t\biggr]\biggr\}ds\biggr)\overline{f_2(t)}\,dt\biggr\} \\ &=-\biggl\{\sum_{n=-\infty}^\infty\biggr(\int_0^x z_n\bigl[f_1(s)\cos\lambda_ns +f_2(s)\sin\lambda_ns\bigr]\,ds\int_0^x\cos\lambda_nt\overline{f_1(t)}\,dt \\ &\quad\qquad +\frac{1}{\pi}\int_0^x\biggl[f_1(s)\cos \biggl(n-\frac12\biggr)s +f_2(s)\sin \biggl(n-\frac12\biggr)s\biggr]\,ds\biggr) \\ &\quad\qquad\qquad\times\int_0^x\cos \biggl(n-\frac12\biggr)t\overline{f_1(t)}\,dt \\ &\quad\qquad +\sum_{n=-\infty}^\infty\biggr(\int_0^x z_n\bigl[f_1(s)\cos\lambda_ns+f_2(s)\sin\lambda_ns\bigr]\,ds \int_0^x\sin\lambda_nt\overline{f_2(t)}\,dt \\ &\quad\qquad\qquad +\frac{1}{\pi} \int_0^x\biggl[f_1(s)\cos \biggl(n-\frac12\biggr)s +f_2(s)\sin \biggl(n-\frac12\biggr)s\biggr]\,ds \\ &\quad\qquad\qquad\qquad\times\int_0^x\sin \biggl(n-\frac12\biggr)t\overline{f_2(t)}\,dt \biggr)\biggr\} \\ &=-\biggl\{\sum_{n=-\infty}^\infty\biggr(\int_0^x z_n[f_1(s)\cos\lambda_ns +f_2(s)\sin\lambda_ns]\,ds\int_0^x\cos\lambda_nt\overline{f_1(t)}\,dt \\ &\quad\qquad+\int_0^x z_n\bigl[f_1(s)\cos\lambda_ns+f_2(s)\sin\lambda_ns\bigr]\,ds\int_0^x \sin\lambda_nt\overline{f_2(t)}\,dt\biggr) \\ &\quad\qquad +\frac{1}{\pi}\sum_{n=-\infty}^\infty\biggr(\int_0^x\biggl[f_1(s)\cos \biggl(n-\frac12\biggr)s \\ &\quad\qquad +f_2(s)\sin \biggl(n-\frac12\biggr)s\biggr]\,ds\int_0^x \cos\biggl(n-\frac12\biggr)t\overline{f_1(t)}\,dt \\ &\quad\qquad+\int_0^x\biggl[f_1(s)\cos \biggl(n-\frac12\biggr)s +f_2(s)\sin \biggl(n-\frac12\biggr)s\biggr]\,ds \\ &\quad\qquad\qquad\times\int_0^x\sin \biggl(n-\frac12\biggr)t\overline{f_2(t)}\,dt \biggr)\biggr\} \\ &=-\biggl\{\sum_{n=-\infty}^\infty\biggr(\int_0^x z_n[f_1(t)\cos\lambda_nt +f_2(t)\sin\lambda_nt]\,dt\int_0^x\cos\lambda_nt\overline{f_1(t)}\,dt \\ &\quad\qquad+\int_0^x z_n\bigl[f_1(t)\cos\lambda_nt+f_2(t)\sin\lambda_nt\bigr]\,dt \int_0^x\sin\lambda_nt\overline{f_2(t)}\,dt\biggr) \\ &\quad\qquad+\frac{1}{\pi}\sum_{n=-\infty}^\infty\biggr(\int_0^x\biggl[f_1(t)\cos \biggl(n-\frac12\biggr)t \\ &\quad\qquad+f_2(t)\sin \biggl(n-\frac12\biggr)t\biggr]\,dt\int_0^x\cos \biggl(n-\frac12\biggr)t\overline{f_1(t)}\,dt \\ &\quad\qquad+\int_0^x\biggl[f_1(t)\cos nt+f_2(t)\sin \biggl(n-\frac12\biggr)t\biggr]\,dt\int_0^x\sin \biggl(n-\frac12\biggr)t\overline{f_2(t)}\,dt\biggr)\biggr\} \\ &=-\biggl\{\sum_{n=-\infty}^\infty z_n\int_0^x [f_1(t)\cos\lambda_nt+f_2(t)\sin\lambda_nt]\,dt \\ &\quad\qquad\qquad\times \int_0^x\bigl[\overline{f_1(t)}\cos\lambda_nt+\overline{f_2(t)}\sin\lambda_nt\bigr]\,dt \\ &\quad\qquad+\sum_{n=-\infty}^\infty \frac{1}{\pi}\int_0^x\biggl[f_1(t)\cos \biggl(n-\frac12\biggr)t+f_2(t)\sin \biggl(n-\frac12\biggr)t\biggr]\,dt \\ &\quad\qquad\qquad\times \int_0^x\biggl[\overline{f_1(t)}\cos \biggl(n-\frac12\biggr)t+\overline{f_2(t)}\sin \biggl(n-\frac12\biggr)t\biggr]\,dt\biggr\} \\ &=-\sum_{n=-\infty}^\infty z_n\biggl|\int_0^x\langle\mathbf{f}(t),Y_0(t,\lambda_n)\rangle \,dt\biggr|^2-\sum_{n=-\infty}^\infty \frac{1}{\pi}\biggl|\int_0^x\biggl\langle \mathbf{f}(t),Y_0\biggl(t,\biggl(n-\frac12\biggr)\biggr)\biggr\rangle \,dt\biggr|^2. \end{aligned} \end{equation*} \notag $$

Parseval’s identity implies that

$$ \begin{equation*} \|\mathbf{f}\|^2_{L_{2,2}(0,x)} =\sum_{n=-\infty}^\infty \frac{1}{\pi}\biggl|\int_0^x\biggl\langle \mathbf{f}(t),Y_0\biggl(t,\biggl(n-\frac12\biggr)\biggr)\biggr\rangle \,dt\biggr|^2; \end{equation*} \notag $$
hence
$$ \begin{equation} \sum_{n=-\infty}^\infty z_n\biggl|\int_0^x\langle \mathbf{f}(t),Y_0(t,\lambda_n)\rangle\, dt\biggr|^2=0. \end{equation} \tag{2.25} $$
Since all numbers $z_n$ are strictly on one side of some straight line through the origin, it follows from (2.25) that $\displaystyle\int_0^x\langle \mathbf{f}(t),Y_0(t,\lambda_n)\rangle \,dt=0$. This fact and the completeness in $L_{2,2}(0,\pi)$ of the vector system $\{Y_0(t,\lambda_n)\}$ yield $\mathbf{f}(t)\equiv0$.

We infer from this (see [8]) that the functions $c(\lambda)$ and $-s(\lambda)$ are the entries in the first row of the monodromy matrix

$$ \begin{equation*} \widetilde U(\pi,\lambda)= \begin{pmatrix} \widetilde c_1(\pi,\lambda) &-\widetilde s_2(\pi,\lambda) \\ \widetilde s_1(\pi,\lambda) &\widetilde c_2(\pi,\lambda) \end{pmatrix} \end{equation*} \notag $$
of problem (1.1), (1.2), (2.1) with some potential $\widetilde V\in L_2(0,\pi)$, that is,
$$ \begin{equation} c(\lambda)=\widetilde c_1(\pi,\lambda)\quad\text{and} \quad s(\lambda)=\widetilde s_2(\pi,\lambda). \end{equation} \tag{2.26} $$
From (2.2) we derive that the corresponding characteristic determinant is given by
$$ \begin{equation*} \widetilde\Delta(\lambda)=J_0+\gamma \widetilde c_1(\lambda)+\alpha\widetilde s_2(\lambda)+\beta \widetilde s_1(\lambda)=J_0+(\alpha +\beta)\sin\pi\lambda+\gamma \cos\pi\lambda+\widetilde f(\lambda), \end{equation*} \notag $$
where $\widetilde f\in \mathrm{PW}_\pi$. Relations (1.4), (2.11) and (2.26) imply that
$$ \begin{equation*} \begin{aligned} \, \widetilde\Delta(\lambda_n) &=J_0+\alpha\widetilde s_2(\pi,\lambda_n)+\beta \widetilde s_1(\pi,\lambda_n)=J_0+\frac{\beta}{\widetilde s_2(\pi,\lambda_n)}+\alpha\widetilde s_2(\pi,\lambda_n) \\ &=J_0+\frac{\beta}{s(\lambda_n)}+\alpha s(\lambda_n)=J_0+u_+(\lambda_n)=U(\lambda_n). \end{aligned} \end{equation*} \notag $$
It follows that
$$ \begin{equation*} \Phi(\lambda)=\frac{U(\lambda)-\widetilde\Delta(\lambda)}{c(\lambda)}=\frac{f(\lambda)-\widetilde f(\lambda)}{c(\lambda)} \end{equation*} \notag $$
is an entire function on the whole complex plane. Since
$$ \begin{equation} |f(\lambda)-\widetilde f(\lambda)|<c_1e^{\pi|{\operatorname{Im}\lambda}|}, \end{equation} \tag{2.27} $$
from (2.5) we obtain $|\Phi(\lambda)|\leqslant c_2$ for $|{\operatorname{Im}\lambda}|\geqslant M$. We let $H$ denote the union of the vertical line segments $\{z\colon |{\operatorname{Re} z}|=n,\ |{\operatorname{Im}\lambda}|\leqslant M\}$, where $|n|=N_0+1,N_0+2, \dots$ . Since $c(\lambda)$ is a sine-type function (see [13]), we have $|c(\lambda)|>\delta>0$ if $\lambda\in H$. We deduce from the last inequality, (2.27) and the maximum modulus principle for analytic functions that $|\Phi(\lambda)|<c_3$ in the strip $|{\operatorname{Im}\lambda}|\leqslant M$; therefore, the function $\Phi(\lambda)$ is bounded on the whole complex plane and, by Liouville’s theorem, is a constant. Let $|{\operatorname{Im}\lambda}|=M$. Then it follows from (2.3) that $\lim_{|\lambda|\to\infty}(f(\lambda)-\widetilde f(\lambda))=0$; hence $\Phi(\lambda)\equiv0$. Thus, $U(\lambda)\equiv\widetilde\Delta(\lambda)$.

Theorem 1 is proved.

§ 3. Spectrum

In addition to (2.1), we assume that the boundary conditions (1.2) are regular; thus, according to (1.8), we have

$$ \begin{equation*} |J_1+iJ_2|\,|J_1-iJ_2| \ne0. \end{equation*} \notag $$

Theorem 2. For a set $\Lambda$ to be the spectrum of a Dirac operator (1.1), (1.2), (2.1) with complex-valued potential $V\in L_2(0,\pi)$, it is necessary and sufficient that it consists of two sequences of eigenvalues $\lambda_{n,j}$ satisfying the condition

$$ \begin{equation} \lambda_{n,j}=2n+\frac{\ln z_j}{i\pi}+\varepsilon_{n,j}, \end{equation} \tag{3.1} $$
where the $z_j$ are the roots of the equation
$$ \begin{equation*} (J_1+iJ_2)z^2+2J_0z+(J_1-iJ_2) =0 \end{equation*} \notag $$
the branch of the logarithm takes values in the strip $\operatorname{Im} \lambda\in(-\pi,\pi]$, and $\{\varepsilon_{n,j}\}\in l_2$, $j=1,2$, $n\in \mathbb{Z}$.

Proof. Necessity was proved in [14].

Sufficiency. Assume that two sequences $\lambda_{n,j}$ satisfy condition (3.1). Clearly, there exists a constant $M$ such that

$$ \begin{equation} \sup|\varepsilon_{n,j}|<M\quad\text{and} \quad \sum_{j=1,2,\, n\in \mathbb Z}|\varepsilon_{n,j}|^2<M. \end{equation} \tag{3.2} $$
It follows from [14] that the eigenvalues of problem (1.1), (1.2), (2.1) with zero potential $V(x)$ are given by the formulae
$$ \begin{equation*} \lambda_{n,j}^0=2n+\frac{\ln z_j}{i\pi}, \end{equation*} \notag $$
$j=1,2$, $n\in \mathbb Z$. Let $t_j={\ln z_j}/(i\pi)$. By Hadamard’s theorem the characteristic function $\Delta_0(\lambda)$ of the nonperturbed problem (1.7), (2.1) has the form
$$ \begin{equation*} \Delta_0(\lambda)=J_0-(J_{13}+J_{24})\sin\pi\lambda-J_{23}\cos\pi\lambda=c\lambda^m\prod_{(n,j)\in T_1}\frac{\lambda_{n,j}^0-\lambda}{\lambda_{n,j}^0}, \end{equation*} \notag $$
where $c\ne0$ and $T_1$ is the set of pairs $(n,j)$ such that $\lambda_{n,j}^0\ne0$. We obviously have
$$ \begin{equation} |\Delta_0(\lambda)|< c_1e^{\pi |{\operatorname{Im}\lambda}|}. \end{equation} \tag{3.3} $$
Let $T_0$ be the set of pairs $(n,j)$ such that $\lambda_{n,j}^0=0$ and $m$ be the cardinality of $T_0$. It follows from [14] that $0\leqslant m\leqslant2$. We also set $T=T_0\cup T_1$ and
$$ \begin{equation*} \Delta(\lambda)=c\prod_{(n,j)\in T_0}(\lambda-\lambda_{n,j})\prod_{(n,j)\in T_1}\frac{\lambda_{n,j}-\lambda}{\lambda_{n,j}^0}. \end{equation*} \notag $$
Let $f(\lambda)=\Delta(\lambda)-\Delta_0(\lambda)$. We investigate the properties of the function $f(\lambda)$ basing on the following results.

Proposition 1. The function $f(\lambda)$ is an entire function of exponential type not exceeding $\pi$.

Proof. We let $\Gamma$ denote the union of discs $\Gamma(2n+t_j,1/4)$, $n\in \mathbb Z$. If $\lambda\notin\Gamma$, then
$$ \begin{equation} f(\lambda)=-\Delta_0(\lambda)\biggl(1-\frac{\Delta(\lambda)}{\Delta_0}\biggr) =-\Delta_0(\lambda)(1-\phi(\lambda)), \end{equation} \tag{3.4} $$
where
$$ \begin{equation*} \phi(\lambda)=\prod_{(n,j)\in T_0}\biggl(1-\frac{\lambda_{n,j}}{\lambda}\biggr) \prod_{(n,j)\in T_1}\biggl(1+\frac{\varepsilon_{n,j}}{\lambda_{n,j}^0-\lambda}\biggr) =\prod_{(n,j)\in T}\biggl(1+\frac{\varepsilon_{n,j}}{2n+t_j-\lambda}\biggr). \end{equation*} \notag $$
We estimate the function $\phi(\lambda)$. Let $\alpha_{n,j}(\lambda)={\varepsilon_{n,j}}/(2n+t_j-\lambda)$. The inequalities (3.2) imply that
$$ \begin{equation} \sum_{(n,j)\in T}|\alpha_{n,j}(\lambda) \leqslant\sum_{(n,j)\in T}\frac{|\varepsilon_{n,j}|^2+|2n+t_j-\lambda|^{-2}}{2}<c_3. \end{equation} \tag{3.5} $$

It is straightforward to see that for all $|n|>n_0$, where $n_0$ is sufficiently large, we have

$$ \begin{equation} |\alpha_{n,j}(\lambda)|<\frac 14 \end{equation} \tag{3.6} $$
for any $\lambda\notin\Gamma$. If $|n|\leqslant n_0$, then (3.6) is valid for all sufficiently large $|\lambda|$; therefore, this inequality holds for all $|\lambda|\geqslant C_0$. We derive from (3.5), (3.6) and the elementary inequality
$$ \begin{equation} |{\ln(1+z)}|\leqslant2|z|, \end{equation} \tag{3.7} $$
which holds for $|z|\leqslant1/4$, that
$$ \begin{equation*} \sum_T|{\ln(1+\alpha_{n,j}}|\leqslant c_4. \end{equation*} \notag $$
Below we choose the branch of $\ln(1+z)$ that vanishes at $z=0$. According to [15], Ch. V, § 1.72, we can rewrite the last relation as
$$ \begin{equation} |\phi(\lambda)|\leqslant\prod_T|1+\alpha_{n,j}(\lambda)|\leqslant e^{c_4}. \end{equation} \tag{3.8} $$
We infer from (3.3), (3.4) and (3.8) that
$$ \begin{equation} |f(\lambda)|<c_5e^{\pi |{\operatorname{Im}\lambda}|} \end{equation} \tag{3.9} $$
outside the domain $\Gamma'=\Gamma\cup\{|\lambda|<C_0\}$.

We set

$$ \begin{equation*} D=\bigcup_{(n,j)\in T}\biggl[2n+\operatorname{Re}t_j-\frac 14,\, 2n+\operatorname{Re}t_j+\frac14\biggr]\quad\text{and} \quad D_0=(0,2)\setminus D. \end{equation*} \notag $$
It is easy to see that $D_0$ is a union of finitely many line segments such that the sum of their lengths is at least $1$. Let $x_0$ be the midpoint of one of these segments. Then all points $x_0 + 2k$, $k\in \mathbb Z$, lie outside $D$. In particular, (3.9) is valid for $\lambda$ on the straight lines $\operatorname{Im}\lambda=\pm \widehat C_0$, where $\widehat C_0=C_0+|t_1|+|t_2|+1$, or vertical line segments with endpoints $(x_0+2k, -\widehat C_0)$ and $(x_0+2k, \widehat C_0)$, where $|2k-1|>C_0$, $k\in\mathbb{Z}$. By the maximum principle, (3.9) holds on the whole complex plane; hence $f(\lambda)$ is an entire function of exponential type not exceeding $\pi$.

Proposition 1 is proved.

Proposition 2. The function $f$ is in the class $\mathrm{PW}_\pi$.

We set

$$ \begin{equation*} W(\lambda)=\ln\phi(\lambda)=\sum_{(n,j)\in T}\ln(1+\alpha_{n,j}(\lambda)); \end{equation*} \notag $$
then
$$ \begin{equation} f(\lambda)=-\Delta_0(\lambda)\bigl(1-e^{W(\lambda)}\bigr). \end{equation} \tag{3.10} $$
We estimate the function $W(\lambda)$ if $\lambda\notin\Gamma'$. It follows from (3.2), (3.6) and (3.7) that
$$ \begin{equation*} \begin{aligned} \, |W(\lambda)| &\leqslant\sum_{(n,j)\in T}|{\ln(1+\alpha_{n,j}(\lambda))}| \\ &\leqslant\frac{2M}{|\lambda|}+ \sum_{n=-\infty}^\infty \biggl(\frac{|\varepsilon{_n,1}|^2+|\varepsilon_{n,2}|^2}{10M}+ \frac{10M}{|2n-\lambda|^2}\biggr) \\ &\leqslant\frac{2M}{|\lambda|}+\frac{1}{10}+20M\sum_{n=0}^\infty\frac{1}{n^2+ |{\operatorname{Im}\lambda}|^2} \\ &\leqslant\frac{2M}{|\lambda|}+\frac{1}{10}+20M\biggl(\frac{2}{|{\operatorname{Im}\lambda}|^2} +\int_1^\infty\frac{dx}{x^2+|{\operatorname{Im}\lambda}|^2}\biggr) \\ &\leqslant\frac{2M}{|{\operatorname{Im}\lambda}|} +\frac{1}{10}+20M\biggl(\frac{2}{|{\operatorname{Im}\lambda}|^2} +\frac{\pi}{2|{\operatorname{Im}\lambda}|}\biggr). \end{aligned} \end{equation*} \notag $$
The last inequality yields
$$ \begin{equation} |W(\lambda)|<\frac14, \end{equation} \tag{3.11} $$
provided that $|{\operatorname{Im}\lambda}|\geqslant M_1=10(\pi+2+22M)+\widehat C_0$. From the elementary inequality
$$ \begin{equation*} \frac{|z|}{2}\leqslant |1-e^z|\leqslant2|z|, \end{equation*} \notag $$
which is valid for $|z|\leqslant1/4$, we derive that $|1-e^{W(\lambda)}|\leqslant2|W(\lambda)|$. The last inequality, (3.3) and (3.10) imply that
$$ \begin{equation} |f(\lambda)|\leqslant c_6|W(\lambda)| \end{equation} \tag{3.12} $$
for $\lambda\in l$, where $l$ is the line $\operatorname{Im}\lambda=M_1$. We prove that
$$ \begin{equation} \int_l|W(\lambda)|^2\,d\lambda<\infty. \end{equation} \tag{3.13} $$
The elementary inequality $|{\ln(1+z)-z}|\leqslant|z|^2$, which is true for $|z|\leqslant1/2$, yields the relation
$$ \begin{equation*} \ln(1+z)-z=r(z), \end{equation*} \notag $$
where $|r(z)|\leqslant|z|^2$; hence
$$ \begin{equation*} W(\lambda)=S_1(\lambda)+S_2(\lambda), \end{equation*} \notag $$
where
$$ \begin{equation*} S_1(\lambda)=\sum_{n=-\infty}^\infty (\alpha_{n,1}(\lambda)+\alpha_{n,2}(\lambda))\quad\text{and} \quad |S_2(\lambda)|\leqslant\sum_{n=-\infty}^\infty (|\alpha_{n,1}(\lambda)|^2+|\alpha_{n,2}(\lambda)|^2). \end{equation*} \notag $$
It is obvious that
$$ \begin{equation} |W(\lambda)|\leqslant|S_1(\lambda)|+|S_2(\lambda)|. \end{equation} \tag{3.14} $$

We set

$$ \begin{equation*} I_m=\int_l|S_m(\lambda)|^2\,d\lambda \end{equation*} \notag $$
$(m=1,2)$. First we consider the integral $I_1$. It follows from [16] that
$$ \begin{equation} \begin{aligned} \, \notag I_1 &=\int_l\,\biggl|\sum_{n=-\infty}^\infty \biggl(\frac{\varepsilon_{n,1}}{2n+t_1-\lambda} +\frac{\varepsilon_{n,2}}{2n+t_2-\lambda}\biggr)\biggr|^2\,d\lambda \\ \notag &\leqslant 2\biggl(\int_l\,\biggl|\sum_{n=-\infty}^\infty \frac{\varepsilon_{n,1}}{2n+t_1-\lambda}\biggr|^2\,d\lambda+ \int_l\,\biggl|\sum_{n=-\infty}^\infty \frac{\varepsilon_{n,2}}{2n+t_2-\lambda}\biggr|^2\,d\lambda\biggr) \\ &=2\biggl(\int_{l_1}\biggl|\sum_{n=-\infty}^\infty \frac{\varepsilon_{n,1}}{2n-\lambda}\biggr|^2\,d\lambda+ \int_{l_2}\biggl|\sum_{n=-\infty}^\infty \frac{\varepsilon_{n,2}}{2n-\lambda}\biggr|^2\,d\lambda\biggr) <\infty, \end{aligned} \end{equation} \tag{3.15} $$
where the $l_j$ are the lines $\operatorname{Im}\lambda=M_1-\operatorname{Im}t_j$, $j=1,2$.

We can clearly see that

$$ \begin{equation*} |S_2(\lambda)|\leqslant\sum_{n=-\infty}^\infty \frac{|\varepsilon_{n,1}|^2}{|2n+t_1-\lambda|^2}+\sum_{n=-\infty}^\infty \frac{|\varepsilon_{n,2}|^2}{|2n+t_2-\lambda|^2}\leqslant c_7; \end{equation*} \notag $$
therefore,
$$ \begin{equation} \begin{aligned} \, \notag I_2 &\leqslant c_7 \int_l\biggl(\sum_{n=-\infty}^\infty \frac{|\varepsilon_{n,1}|^2}{|2n+t_1-\lambda|^2}+\sum_{n=-\infty}^\infty \frac{|\varepsilon_{n,2}|^2}{|2n+t_2-\lambda|^2}\biggr)\,d\lambda \\ & \leqslant c_8\sum_{n=-\infty}^\infty (|\varepsilon_{n,1}|^2+|\varepsilon_{n,2}|^2) \int_{\widetilde l}\frac{d\lambda}{|2n-\lambda|^2} \leqslant c_9\sum_{n=-\infty}^\infty (|\varepsilon_{n,1}|^2+|\varepsilon_{n,2}|^2)\leqslant c_{10}, \end{aligned} \end{equation} \tag{3.16} $$
where $\widetilde l=l_1\cup l_2$. Relations (3.14)(3.16) imply (3.13). It follows from (3.12), (3.13), and [12], Ch. 3, § 3.2.2, that
$$ \begin{equation*} \int_R|f(\lambda)|^2\,d\lambda<\infty. \end{equation*} \notag $$

Proposition 2 is proved.

Thus, the function $\Delta(\lambda)$ satisfies all assumptions of Theorem 1. Hence there exists a potential $V\in L_2(0,\pi)$ such that the spectrum of the corresponding problem (1.1), (1.2), (2.1) is as specified by (3.1).

Theorem 2 is proved.

Acknowledgements

The author is grateful to the referee for useful remarks.


Bibliography

1. M. G. Gasymov and T. T. Dzhabiev, “Solution of the inverse problem by two spectra for the Dirac equation on a finite interval”, Dokl. Akad. Nauk Azerb. SSR, 22:7 (1966), 3–7 (Russian)  mathscinet  zmath
2. S. Albeverio, R. Hryniv and Y. Mykytyuk, “Inverse spectral problems for Dirac operators with summable potentials”, Russ. J. Math. Phys., 12:4 (2005), 406–423  mathscinet  zmath
3. T. V. Misyura, “Characteristics of the spectra of the periodic and antiperiodic boundary value problems that are generated by the Dirac operator. II”, Teor. Funktsii Funktsional. Anal. i Prilozhen., 31, Khar'kov University Publishing House, Khar'kov, 1979, 102–109 (Russian)  mathscinet  zmath
4. I. M. Nabiev, “Solution of the inverse quasiperiodic problem for the Dirac system”, Mat. Zametki, 89:6 (2011), 885–893  mathnet  crossref  mathscinet  zmath; English transl. in Math. Notes, 89:6 (2011), 845–852  crossref
5. V. A. Yurko, “An inverse spectral problem for singular non-self-adjoint differential systems”, Mat. Sb., 195:12 (2004), 123–156  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 195:12 (2004), 1823–1854  crossref
6. V. Yurko, “Inverse spectral problems for differential systems on a finite interval”, Results Math., 48:3–4 (2005), 371–386  crossref  mathscinet  zmath
7. A. S. Makin, “On the spectrum of two-point boundary value problems for the Dirac operator”, Differ. Uravn., 57:8 (2021), 1023–1031  mathscinet  zmath; English transl. in Differ. Equ., 57:8 (2021), 993–1002  crossref
8. V. Tkachenko, “Non-self-adjoint periodic Dirac operators”, Operator theory, system theory and related topics (Beer-Sheva/Rehovot 1997), Oper. Theory Adv. Appl., 123, Birkhäuser, Basel, 2001, 485–512  mathscinet  zmath
9. V. A. Marchenko, Sturm-Liouville operators and applications, Naukova Dumka, Kiev, 1977, 331 pp.  mathscinet  zmath; English transl., Oper. Theory Adv. Appl., 22, Birkhäuser Verlag, Basel, 1986, xii+367 pp.  crossref  mathscinet  zmath
10. V. Tkachenko, “Non-selfadjoint periodic Dirac operators with finite-band spectra”, Integral Equations Operator Theory, 36:3 (2000), 325–348  crossref  mathscinet  zmath
11. B. Ya. Levin, Lectures on entire functions, Moscow State University, Moscow, 1971, 124 pp.; English transl., Transl. Math. Monogr., 150, Amer. Math. Soc., Providence, RI, 1996, xvi+248 pp.  mathscinet  zmath
12. S. M. Nikol'skiĭ, Approximation of functions of several variables and imbedding theorems, 2nd ed., Nauka, Moscow, 1977, 455 pp.  mathscinet  zmath; English transl. of 1st ed., Grundlehren Math. Wiss., 205, Springer-Verlag, New York–Heidelberg, 1975, viii+418 pp.  crossref  mathscinet  zmath
13. B. J. Levin and I. V. Ostrovskii, “On small perturbations of the set of zeros of functions of sine type”, Izv. Akad. Nauk SSSR Ser. Mat., 43:1 (1979), 87–110  mathnet  mathscinet  zmath; English transl. in Math. USSR-Izv., 14:1 (1980), 79–101  crossref
14. P. Djakov and B. Mityagin, “Unconditional convergence of spectral decompositions of 1D Dirac operators with regular boundary conditions”, Indiana Univ. Math. J., 61:1 (2012), 359–398  crossref  mathscinet  zmath
15. M. A. Lavrentiev and B. V. Shabat, Methods of the theory of functions of complex variable, 4th ed., Nauka, Moscow, 1973, 736 pp.  zmath; German transl. of 3d ed., M. A. Lawrentjew and B. W. Schabat, Methoden der komplexen Funktionentheorie, Math. Naturwiss. Tech., 13, VEB Deutscher Verlag der Wissenschaften, Berlin, 1967, x+846 pp.  mathscinet  zmath
16. J.-J. Sansuc and V. Tkachenko, “Characterization of the periodic and anti-periodic spectra of nonselfadjoint Hill's operators”, New results in operator theory and its applications, Oper. Theory Adv. Appl., 98, Birkhäuser, Basel, 1997, 216–224  crossref  mathscinet  zmath

Citation: A. S. Makin, “Structure of the spectrum of a nonselfadjoint Dirac operator”, Mat. Sb., 214:1 (2023), 43–60; Sb. Math., 214:1 (2023), 39–57
Citation in format AMSBIB
\Bibitem{Mak23}
\by A.~S.~Makin
\paper Structure of the spectrum of a~nonselfadjoint Dirac operator
\jour Mat. Sb.
\yr 2023
\vol 214
\issue 1
\pages 43--60
\mathnet{http://mi.mathnet.ru/sm9709}
\crossref{https://doi.org/10.4213/sm9709}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4619860}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2023SbMat.214...39M}
\transl
\jour Sb. Math.
\yr 2023
\vol 214
\issue 1
\pages 39--57
\crossref{https://doi.org/10.4213/sm9709e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001037692200003}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85171368685}
Linking options:
  • https://www.mathnet.ru/eng/sm9709
  • https://doi.org/10.4213/sm9709e
  • https://www.mathnet.ru/eng/sm/v214/i1/p43
  • Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Математический сборник Sbornik: Mathematics
    Statistics & downloads:
    Abstract page:314
    Russian version PDF:39
    English version PDF:48
    Russian version HTML:173
    English version HTML:90
    References:39
    First page:18
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024