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, 2024, Volume 215, Issue 9, Pages 1249–1268
DOI: https://doi.org/10.4213/sm10066e
(Mi sm10066)
 

This article is cited in 1 scientific paper (total in 1 paper)

Molchanov's criterion for compactness of the resolvent for a nonselfadjoint Sturm–Liouville operator

S. N. Tumanovab

a Moscow Center of Fundamental and Applied Mathematics, Moscow, Russia
b Faculty of Mechanics and Mathematics, Lomonosov Moscow State University, Moscow, Russia
References:
Abstract: A Molchanov-type condition is considered in applications to ordinary differential operators of arbitrary order with complex-valued coefficients. It is proved to be a necessary condition for the compactness of the resolvent for a wide class of operators of this type. A counterexample is given showing that this condition does not suffice for the compactness of the resolvent for a Sturm–Liouville operator with nonnegative real part of the potential. Molchanov's criterion is generalized to potentials taking values in a sector bounded away from the negative half-axis and more narrow than a half-plane.
Bibliography: 18 titles.
Keywords: nonselfadjoint Sturm–Liouville operator, discreteness of spectrum, compactness of resolvent, Molchanov's criterion.
Funding agency Grant number
Russian Science Foundation 20-11-20261
The study was supported by the Russian Science Foundation, project no. 20-11-20261, https://rscf.ru/en/project/20-11-20261/.
Received: 17.01.2024 and 12.06.2024
Published: 10.12.2024
Bibliographic databases:
Document Type: Article
Language: English
Original paper language: Russian

§ 1. Introduction

A significant result in the theory of qualitative spectral analysis for ordinary differential operators is Molchanov’s theorem, which provides a criterion for the complete continuity of the resolvent [1] for a singular Sturm–Liouville operator of the form

$$ \begin{equation*} Ly=-y''+qy. \end{equation*} \notag $$

The criterion, formulated originally for operators with real potentials $q(x)\geqslant\mathrm{const}$, $x\in\mathbb{R}$, was generalized in [2]–[8], mainly to relax the condition that $q$ is bounded below or to extend to the vector case when the role of $q$ is played by a Hermitian matrix, $q\geqslant0$. We do not consider partial differential operators.

Generalizations concerning complex-valued $q$ are few, and the key work here is Lidskii’s paper [3], where Molchanov’s criterion was generalized to potentials taking values in the sector $0\leqslant\arg q(x)\leqslant\pi/2$ or $-\pi/2\leqslant\arg q(x)\leqslant0$ in the complex plane.

Throughout what follows,

$$ \begin{equation} q\in L_{1,\mathrm{loc}}(\mathbb{R}_+) \text{ is a complex-valued function}. \end{equation} \tag{1.1} $$

We consider the case of operators on a half-axis (in the space $L_2(\mathbb{R}_+)$), whereas the results in the above papers concern operators in the space $L_2(\mathbb{R})$ on the whole axis.

This difference is insignificant for our results: the theorems below can easily be extended to $L_2(\mathbb{R})$ with just one correction: the limit $x\to+\infty$ in Molchanov’s condition formulated below must be replaced by $x\to\infty$.

We say that $q$ satisfies Molchanov’s condition if

$$ \begin{equation*} \lim_{x\to+\infty}\int_x^{x+a}|q(\xi)|\,d\xi=\infty \end{equation*} \notag $$
for each $a>0$.

Even for real $q$, if we abandon boundedness below, Molchanov’s condition is not a criterion for the complete continuity of the resolvent any longer, for example, for ${q(x)=-x^2}$; see [9].

Nevertheless, it remains to be necessary even when no restrictions other than (1.1) are imposed on $q$. We prove this result in the more general case of ordinary differential operators of an arbitrary order $n\geqslant 2$.

Supplementing it with sufficiency theorems and a counterexample, we draw a picture of possible generalizations of Molchanov’s theorem to the complex-valued case in terms of Molchanov’s condition itself.

We consider the differential expression

$$ \begin{equation} l(y)=-y''+qy \end{equation} \tag{1.2} $$
and the linear sets
$$ \begin{equation*} \begin{gathered} \, \mathscr{D}=\bigl\{y\in L_2(\mathbb{R}_+)\mid y,y'\in \mathrm{AC}_{\mathrm{loc}}(\mathbb{R}_+),\ l(y)\in L_2(\mathbb{R}_+)\bigr\}, \\ \mathscr{D}_0=\bigl\{y\in\mathscr{D}\mid y(0)=y'(0)=0,\ \exists\, x_0>0\ \forall\, x\geqslant x_0\ y(x)=0\bigr\} \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} \mathscr{D}_U=\bigl\{y\in\mathscr{D}\mid U(y)=0\bigr\}, \end{equation*} \notag $$
where $U$ is some form of a boundary condition at $x=0$:
$$ \begin{equation*} U(y)=A y(0)+B y'(0), \qquad A,B\in\mathbb{C}, \quad |A|+|B|>0. \end{equation*} \notag $$

Following [10], we define differential operators $L_0\subset L_U$ in $L_2(\mathbb{R}_+)$ on relevant domains $\mathscr D_0 \subset \mathscr D_U$ by the differential expression (1.2).

Theorem 1. For $L_0$ to have an extension with compact resolvent, it is necessary that $q$ satisfy Molchanov’s condition.

This is a consequence of our general result in Theorem 5, which we state and prove in § 3.

Definition 1. A potential $q$ is said to satisfy the $\mathbb{R}^-$-condition if for all sufficiently large $x>x_0\geqslant0$ the values of $q(x)$ lie in the sector $\alpha\leqslant\arg (q(x)-q_0)\leqslant\beta$ for some $-\pi<\alpha\leqslant\beta<\pi$ and $q_0\in\mathbb{C}$.

In other words, the $\mathbb{R}^-$-condition means such that for some $q_0\in\mathbb{C}$ the difference $q-q_0$ asymptotically takes values outside a small sector containing $\mathbb{R}^-$.

Definition 2. A potential $q$ is said to be sectorial if it satisfies the $\mathbb{R}^-$-condition for $\beta-\alpha<\pi$.

Theorem 2. Assume that a potential $q\in L_{1,\mathrm{loc}}(\mathbb{R}_+)$ is sectorial. Then the operator $L_U$ has a compact resolvent if and only if $q$ satisfies Molchanov’s condition.

As the following theorem shows, the condition $\beta-\alpha<\pi$ cannot be relaxed.

Theorem 3. There is a potential $q$ taking purely imaginary values $q(x)\in i\mathbb{R}$ for $x\in\mathbb{R}_+$ such that $|q|\to+\infty$ as $x\to+\infty$ but the operator $L_0$ has no extension with compact resolvent.

For this potential we have $\beta-\alpha=\pi$, and it obviously satisfies Molchanov’s condition. The operator $L_U$ with the Dirichlet boundary condition $U(y)=y(0)$ has a bounded resolvent at least in the left-hand half-plane (see [3], Lemma 2, and [11]), but this resolvent is not a completely continuous operator.

The following theorem gives a sufficient condition for the compactness of the resolvents of operators with potentials satisfying the $\mathbb{R}^-$-condition for $\beta - \alpha > \pi$. In this case the sectorial property of the operators under consideration is lost; in particular, the numerical image of $L_U$ can cover the whole complex plane [12].

Theorem 4. Assume that for some $x_0>0$ $|q(x)| \geqslant 1$ for $x \geqslant x_0$ and

Under these conditions, for the resolvent of $L_U$ to be compact it is sufficient that

$$ \begin{equation} \lim_{x\to+\infty}\int_x^{x+a}|q(\xi)|^{1/2}\,d\xi=\infty \end{equation} \tag{1.3} $$
for any $a>0$.

The further presentation of is as follows: we devote § 2 to proving Theorems 24; in § 3 we return to Theorem 1, which is stated and proved there for general differential operators of order $n\geqslant 2$.

§ 2. Proof of Theorems 24

Since neither Molchanov’s condition nor the complete continuity of the resolvent depend on the shift of the potential by a constant or its values on a finite interval $[0,x_0]$, we assume without loss of generality that $q_0=0$ and $x_0=0$.

The proof of Molchanov’s criterion [1] is based on Rellich’s criterion of the compactness of the resolvent of a positive definite selfadjoint operator ([13], Ch. II, § 24, Theorem 11), which we generalize to $\mathrm m$-sectorial operators. For relevant definitions, see [14], Ch. V, § 3.10.

Lemma 1. Let $A$ be an $\mathrm m$-sectorial operator in a Hilbert space $\mathfrak{H}$ with domain $\mathscr{D}_A\subset\mathfrak{H}$.

Then $A$ has a compact resolvent if and only if the set of vectors $\varphi\in\mathscr{D}_A$ satisfying the condition

$$ \begin{equation*} \operatorname{Re}(A\varphi,\varphi)\leqslant1 \end{equation*} \notag $$
is compact.

Proof. We associate with $A$ a densely defined closed sectorial sesquilinear form $\mathfrak{a}$ ([14], Ch. VI, § 2, Theorem 2.7) such that its generating kernel is $\mathscr{D}_A$.

Recall that a linear set $\mathscr{D}_A$ is called a generating kernel of a closed form $\mathfrak{a}$ if the closure of the restriction of $\mathfrak{a}$ to $\mathscr{D}_A$ coincides with form $\mathfrak{a}$ itself.

Given $\mathfrak{a}$, we construct the symmetric form $\mathfrak{t}=\operatorname{Re}\mathfrak{a}$: we set

$$ \begin{equation*} \mathfrak{t}[u,v]=\frac{1}{2}(\mathfrak{a}[u,v]+\overline{\mathfrak{a}[v,u]}) \end{equation*} \notag $$
for $u,v\in\mathscr{D}_\mathfrak{a}$ (in the domain of $\mathfrak{a}$).

The form $\mathfrak{t}$ is closed on the domain $\mathscr{D}_\mathfrak{t}=\mathscr{D}_\mathfrak{a}$, densely defined, and nonnegative. According to the second representation theorem ([14], Ch. VI, § 2, Theorem 2.23), it is associated with a selfadjoint operator $T\geqslant0$, $\mathscr{D}_\mathfrak{t}=\mathscr{D}(T^{1/2})\supset\mathscr{D}_T$, where the domain $\mathscr{D}_T$ of $T$ is a generating kernel of $\mathfrak{t}$, and

$$ \begin{equation*} \mathfrak{t}[u,v]=(T^{1/2}u,T^{1/2}v), \qquad u,v\in\mathscr{D}_\mathfrak{t}. \end{equation*} \notag $$
The domain $\mathscr{D}_A$ is also a kernel of $\mathfrak{t}$. In view of the fact that $\mathscr{D}_A$ is a kernel of $\mathfrak{a}$, for any $u\in\mathscr{D}_\mathfrak{t}=\mathscr{D}_\mathfrak{a}$ there is a sequence $u_n\in\mathscr{D}_A$ such that $u_n\to u$ and $\mathfrak{a}[u_n,u_n]\to\mathfrak{a}[u,u]$ as $n\to\infty$. Clearly, the sequence $\mathfrak{t}[u_n,u_n]=\operatorname{Re}\mathfrak{a}[u_n,u_n]$ is convergent; owing to the closedness of $\mathfrak{t}$, the limit $\lim\mathfrak{t}[u_n,u_n]$ is equal to $\mathfrak{t}[u,u]$.

The resolvents of both $A$ and $T$ are compact or not simultaneously ([14], Ch. VI, § 3, Theorem 3.3).

The subsequent reasoning uses Rellich’s criterion for $T$.

We prove the necessity. Let $\Phi=\bigl\{\varphi\in\mathscr{D}_A \mid\operatorname{Re}(A\varphi,\varphi)\leqslant1\bigr\}$. Since $\mathscr{D}_T$ is a kernel of $\mathfrak{t}$, for any $\varepsilon>0$ and $\varphi\in\Phi$ there exists $u\in\mathscr{D}_T$ such that $|\mathfrak{t}[u,u]-\mathfrak{t}[\varphi,\varphi]|<\varepsilon$ and $|u-\varphi|<\varepsilon$. Since the resolvent of $A$ is compact, so is also the resolvent of $T$; since $(Tu,u)=\mathfrak{t}[u,u]\leqslant1+\varepsilon$, the whole of the set $\{u\}$ is compact and, as $\varepsilon$ is arbitrary, $\Phi$ is compact.

We prove the sufficiency. Let $U=\bigl\{u\in\mathscr{D}_T\mid (Tu,u)\leqslant1\bigr\}$. Since $\mathscr{D}_A$ is the kernel of $\mathfrak{t}$, for any $\varepsilon>0$ and any $u\in U$ there exists $\varphi\in\mathscr{D}_A$ such that $|\mathfrak{t}[u,u]-\mathfrak{t}[\varphi,\varphi]|<\varepsilon$ and $|u-\varphi|<\varepsilon$. Since $\operatorname{Re}(A\varphi,\varphi)=\mathfrak{t}[\varphi,\varphi]\leqslant1+\varepsilon$, we conclude that the whole system $\{\varphi\}$ is compact and, as $\varepsilon>0$ is arbitrary, $U$ is compact. Therefore, $T$ has a compact resolvent; thus, $A$ also does.

The lemma is proved.

Proof of Theorem 2. The necessity follows from Theorem 1; we dwell only on the sufficiency.

Without loss of generality we assume that $-\pi<\alpha\leqslant0\leqslant\beta<\pi$ (otherwise we enlarge the sector in which the potential ranges so that the assumptions of the theorem still hold).

We show that $L_U$ has a resolvent in some domain.

Under the assumptions of the theorem the limit-point case ([11], Corollary to Theorem 4) takes place for the homogeneous equation $l(y)=\lambda y$, and the operators $L_{D}$ and $L_N$ with the Dirichlet and Neumann boundary conditions ($U(y)=y(0)$ and $U(y)=y'(0)$) have bounded resolvents $R_{D,\lambda}$ and $R_{N,\lambda}$, respectively, in some sector $\Lambda\subset\mathbb{C}$ ([15], Theorem 4.1).

The explicit expressions for the corresponding resolvents in terms of Green’s functions imply that $L_U-\lambda$ for $\lambda\in\Lambda$ has at any rate a right inverse

$$ \begin{equation*} R_\lambda=w_D(\lambda)R_{D,\lambda}+w_N(\lambda)R_{N,\lambda}, \qquad w_D(\lambda)+w_N(\lambda)=1, \end{equation*} \notag $$
where $w_D$ and $w_N$ are meromorphic functions in $\Lambda$. This means that the operator $L_U-\lambda$ is surjective, at least in some subdomain $\lambda\in\Lambda_0\subset\Lambda$ not containing poles of $w_D$ and $w_N$. In view of the fact that the limit-point case occurs, the conjugate differential expression and the conjugate boundary condition specify the conjugate operator $(L_U-\lambda)^*$, which turns out to be surjective in a similar way. Hence $L_U-\lambda$ maps $\mathscr{D}_U$ bijectively onto the whole of $L_2(\mathbb{R}_+)$ for $\lambda\in\Lambda_0$, and the resolvent is well defined.

For each $\lambda\in\Lambda$ the operator $L_0-\lambda$ has defect 1 (the image of $L_0-\lambda$ is orthogonal to the only solution in $L_2(\mathbb{R}_+$) of the equation $y''=(\overline{q(x)}-\overline{\lambda})y$), and the resolvents of two arbitrary extensions of $L_0$ differ at most by a one-dimensional operator at each point $\lambda\in\Lambda$ where they exist simultaneously. Therefore, they are compact or noncompact simultaneously. So it suffices to carry out the rest of the proof only for the operator $L_D$ with the Dirichlet boundary condition at zero.

As the resolvent exists for $\lambda\in\Lambda$, $L_D$ is closed.

We denote the domain of $L_D$ by $\mathscr{D}_D$ and also consider the linear manifold $\mathscr{D}_{D0}$ of elements of $\mathscr{D}_D$ with compact support:

$$ \begin{equation} \mathscr{D}_D=\bigl\{y\in\mathscr{D}\mid y(0)=0\bigr\}\quad\text{and} \quad \mathscr{D}_{D0}=\bigl\{y\in\mathscr{D}_D\mid \exists\, x_0>0\ \forall\, x\geqslant x_0\ y(x)=0\bigr\}. \end{equation} \tag{2.1} $$
The domain $\mathscr{D}_{D0}$ is a generating kernel of $L_D$ in the following sense: the closure of the restriction of $L_D$ to $\mathscr{D}_{D0}$ coincides with $L_D$ ([3] Lemma 6).

We set $\theta=-(\alpha+\beta)/2$. The operator $M=e^{i\theta}L_D$ is $\mathrm m$-sectorial. Since the resolvent exists, this operator is maximal closed and

$$ \begin{equation*} \begin{gathered} \, (My,y)=e^{i\theta}\int_0^{+\infty}|y'(x)|^2\,dx+\int_0^{+\infty}e^{i\theta}q(x)|y(x)|^2\,dx, \\ |{\arg(My,y)}|\leqslant\frac{\beta-\alpha}{2}<\frac{\pi}{2} \end{gathered} \end{equation*} \notag $$
for any $y\in\mathscr{D}_{D0}$.

The further argument essentially repeats Molchanov’s reasoning in [1]; nevertheless, we present it here: it will be used in the proof of Theorem 4.

Assume that Molchanov’s condition holds but the resolvent of $M$ is not compact. In view of Lemma 1 we can take a noncompact sequence $Y=\{y_n\}_{n=1}^{\infty}\subset\mathscr{D}_{D}$ such that $\operatorname{Re}(My_n,y_n)\leqslant1$, $n\in\mathbb{N}$. Since $\mathscr{D}_{D0}$ is a generating kernel of $M$, we assume without loss of generality that all the $y_n$ are in $\mathscr{D}_{D0}$. In fact, for every $\varepsilon>0$ it suffices to find for each $y_n\in Y$ an element $\widehat y_n\in\mathscr{D}_{D0}$ such that

$$ \begin{equation*} \|y_n-\widehat y_n\|<\varepsilon\quad\text{and} \quad |{\operatorname{Re}(M\widehat y_n,\widehat y_n)-\operatorname{Re}(My_n,y_n)}|<\varepsilon. \end{equation*} \notag $$
The system $\{\widehat y_n\}_{n=1}^{\infty}$ cannot be compact for all $\varepsilon>0$: there is $\varepsilon_0>0$ for which $\{\widehat y_n\}_{n=1}^{\infty}$ is not compact. In this case $\operatorname{Re}(M\widehat y_n,\widehat y_n)\leqslant1+ \varepsilon_0$. The suitable system $\{\widehat y_n/\sqrt{1+\varepsilon_0}\}_{n=1}^{\infty}$ consists of normalized elements.

There is $\varepsilon_0>0$ such that for any $T>0$ there exists $y_T\in Y$ such that

$$ \begin{equation} \int_T^{+\infty} |y_T(x)|^2\,dx\geqslant\varepsilon_0. \end{equation} \tag{2.2} $$
This follows from the compactness of the sequence of truncated functions $Y_T=\{y_{n,T}\}_{n=1}^{\infty}$, where $y_{n,T}(x)=y_n(x)$ for $x\in[0,T]$ and $y_{n,T}(x)=0$ for $x>T$ ([3], Lemma 8). If for any $\varepsilon>0$ there were $T_0=T_0(\varepsilon)$ such that
$$ \begin{equation*} \int_{T_0}^{+\infty} |y_n(x)|^2\,dx<\varepsilon \end{equation*} \notag $$
for each $n\in\mathbb{N}$, then the sequence $Y$ itself would be compact (see the final part of the proof of Theorem 4 in [3]).

We take $d=(\varepsilon_0^{1/2}\cos^{1/2}\theta)/4>0$, an arbitrary $T>0$ and the corresponding $y_T$ and partition the ray $[T,+\infty)$ into intervals $D_n$ of equal length $d$, $n\in\mathbb{N}$; then on at least one of them, which we denote by $D_T$, it is true that

$$ \begin{equation} \operatorname{Re}\biggl(e^{i\theta}\int_{D_T}|y_T'(x)|^2\,dx+\int_{D_T}e^{i\theta}q(x)|y_T(x)|^2\,dx\biggr) \leqslant\frac{1}{\varepsilon_0}\int_{D_T} |y_T(x)|^2\,dx \end{equation} \tag{2.3} $$
and
$$ \begin{equation} \int_{D_T} |y_T(x)|^2\,dx>0. \end{equation} \tag{2.4} $$
If the inequalities reverse to (2.3) were valid on all intervals $D_n$, $n\in\mathbb{N}$, such that (2.4) holds, then, summing them up and using (2.2), we would obtain the estimate $\operatorname{Re}(My_T,y_T)>1$, which is a contradiction.

We normalize $y_T$ by setting

$$ \begin{equation} v_T=y_T \frac{d^{1/2}}{\bigl(\int_{D_T} |y_T(x)|^2\,dx\bigr)^{1/2}}. \end{equation} \tag{2.5} $$
In view of the fact that the length $|D_T|$ is $d$, we have
$$ \begin{equation} \begin{gathered} \, \operatorname{Re}\biggl( e^{i\theta}\int_{D_T}|v_T'(x)|^2\,dx+\int_{D_T}e^{i\theta}q(x)|v_T(x)|^2\,dx\biggr)\leqslant\frac{d}{\varepsilon_0}, \\ \frac{1}{|D_T|}\int_{D_T}|v_T(x)|^2\,dx=1. \end{gathered} \end{equation} \tag{2.6} $$
In particular,
$$ \begin{equation*} \operatorname{Re}\biggl(e^{i\theta}\int_{D_T}|v_T'(x)|^2\,dx\biggr)\leqslant\frac{d}{\varepsilon_0}, \quad\text{that is,}\quad \int_{D_T}|v_T'(x)|^2\,dx\leqslant\frac{d}{\varepsilon_0\cos\theta}. \end{equation*} \notag $$

For any two points $x_1,x_2\in D_T$,

$$ \begin{equation*} \begin{aligned} \, \bigl||v_T(x_1)|-|v_T(x_2)|\bigr| &\leqslant\bigl|v_T(x_1)-v_T(x_2)\bigr| \\ &\leqslant\int_{D_T}|v_T'(x)|\,dx\leqslant d^{1/2}\biggl(\int_{D_T}|v_T'(x)|^2\,dx\biggr)^{1/2}\leqslant \frac{1}{4}, \end{aligned} \end{equation*} \notag $$
since $|v_T|$ is continuous on $D_T$ and the integral mean of $|v_T|^2$ is $1$ (see (2.6)). Therefore, there exists $x_0\in D_T$ such that $|v_T(x_0)|=1$. We conclude that ${|v_T(x)|\geqslant3/4}$ for all $x\in D_T$.

We turn to (2.6) again, which yields

$$ \begin{equation*} \int_{D_T}\operatorname{Re}(e^{i\theta}q(x))|v_T(x)|^2\,dx \leqslant\frac{d}{\varepsilon_0}. \end{equation*} \notag $$
However, $|\arg(e^{i\theta}q(x))|\leqslant(\beta-\alpha)/2$; hence
$$ \begin{equation*} \operatorname{Re}(e^{i\theta}q(x))\geqslant|q|\cos\frac{\beta-\alpha}{2} \end{equation*} \notag $$
and, consequently,
$$ \begin{equation*} \int_{D_T}|q(x)|\,dx\leqslant\frac{d}{\varepsilon_0}\frac{4^2}{3^2}\sec\frac{\beta-\alpha}{2}, \end{equation*} \notag $$
which contradicts Molchanov’s condition for $q$ since $T>0$ can be arbitrary. This completes the proof.

Theorem 2 is proved.

Proof of Theorem 3. We construct a counterexample, which will give us a proof.

We partition the half-axis into half-open intervals of length $\pi$ by setting $I_k=[\pi(k-1);\pi k)$, $k\in\mathbb{N}$, and we set $q_k(x)=ikr_{n_k}(x-\pi(k-1))$, $x\in I_k$, where

$$ \begin{equation*} r_n(t)=(-1)^{j+1}, \qquad t\in\biggl[\pi\frac{j-1}{n},\pi\frac{j}{n}\biggr), \quad j=1,\dots,n. \end{equation*} \notag $$
The procedure for choosing $\{n_k\}_{k=1}^{\infty}$ will become clear below.

Setting $q|_{I_k}=q_k$, we define $q$ for all $x\in\mathbb{R}_+$. It is obvious that $|q(x)|\to+\infty$ as ${x\to+\infty}$.

As the boundary condition, we take the Dirichlet condition $U(y)=y(0)$. To simplify the notation, we denote the operator itself by $L$.

The limit-point case takes place for the homogeneous equation $y''(x)=(q(x)- \lambda)y(x)$; as already noted above, the operator $L$ has a bounded resolvent in the left half-plane (see [3], Lemma 2, and [11]). We will show how, by choosing $\{n_k\}_{k=1}^{\infty}$ properly, we can ensure that the resolvent is not a compact operator.

We show that for each $k\in\mathbb{N}$ there exists $n_k\in\mathbb{N}$ and a function $y_k\in\mathscr{D}_{U}$ vanishing everywhere outside $I_k$ such that $\|y_k\|=1$ and $\|Ly_k\|<2^{13}$. This will imply the noncompactness of the resolvent of the operator $L$ corresponding to the set $\{n_k\}_{k=1}^{\infty}$.

This problem is local for each $I_k$. In this connection it suffices to show that for each $u>0$ there is $n\in\mathbb{N}$ such that for the operator

$$ \begin{equation*} L_n(u)y=-y''+iur_ny \end{equation*} \notag $$
in $L_2[0,\pi]$ defined in the domain
$$ \begin{equation*} \mathscr{E}=\bigl\{y\in L_2[0,\pi]\mid y,y'\in \mathrm{AC}[0,\pi],\ y''\in L_2[0,\pi],\ y(0)=y(\pi)=0\bigr\} \end{equation*} \notag $$
there exists $y\in\mathscr{E}$, $y(0)=y'(0)=y(\pi)=y'(\pi)=0$, $\|y\|=1$, such that ${\|L_n(u)y\|<2^{13}}$. Here and in what follows the norms in $L_2[0,\pi]$ are meant.

Along with the operators $L_n(u)$, we consider another operator $L_0$ with the same domain $\mathscr{E}$:

$$ \begin{equation*} L_0y=-y''; \end{equation*} \notag $$
we let $\mu_j=j^2$, $j\in\mathbb{N}$, denote the eigenvalues of $L_0$.

The operator $B_ny=r_ny$ is bounded in $L_2[0,\pi]$, $\|B_n\|=1$, and $B_n\xrightarrow{w}0$ as ${n \to \infty}$ in the sense of weak operator convergence, which can easily be verified using the indicator functions $\chi_{[a,b]}$ of the intervals $[a,b] \subset [0,\pi]$, whose linear combinations are dense in $L_2[0,\pi]$.

We show that, given $u>0$, the eigenvalues of the $L_n(u)$ converge pointwise to $\{\mu_j\}_{j=1}^{\infty}$ as $n\to\infty$: for any $\mu_j$ there is a sequence $\lambda_{j,n}\to \mu_j$ of eigenvalues of the operators $L_n(u)$, and for each compact set $\mathcal{C}$ not containing $\{\mu_j\}_{j=1}^{\infty}$ there is $n_0>0$ such that the compact set $\mathcal{C}$ does not contain eigenvalues of $L_n(u)$ for ${n>n_0}$.

Fix $u>0$ and an integer $j_0>u-1/2$. We let $\Gamma_j$, $j\in\mathbb{N}$, denote the closed disc with centre $\mu_j$ and radius $u$. Set $\Gamma=\Gamma_1\cup\cdots\cup\Gamma_{j_0}$.

The compact sets $\Gamma$, $\Gamma_{j}$, $j>j_0$, are pairwise disjoint; $\Gamma$ contains $j_0$ eigenvalues of the operator $L_0$, while each disc $\Gamma_{j}$ for $j>j_0$ contains one eigenvalue.

For $n\in\mathbb{N}$ the eigenvalues of $L_n(u)=L_0+iuB_n$ lie in the union of $\Gamma$ and the $\Gamma_{j}$, $j>j_0$; $\Gamma$ contains $j_0$ eigenvalues of the perturbed operator $L_n(u)$ (taking account of multiplicities), while each disc $\Gamma_{j}$, $j>j_0$, contains one eigenvalue of it; see [14], Ch. V, § 4.3.

We show that the limit points of the eigenvalues $\lambda_{j,n}$ of $L_n(u)$ as $n\to\infty$ coincide with the $\mu_j$, $j\in\mathbb{N}$ (the limit points are considered in the relevant compact sets $\Gamma$ and $\Gamma_{j}$, $j>j_0$).

Assume the converse: as $l\to\infty$, a subsequence $\lambda_{n_l}$ converges to some ${\lambda_0\ne\mu_j}$, $j\in\mathbb{N}$. We let $f_{n_l}$ denote the eigenfunctions of $L_{n_l}(u)$ corresponding to the eigenvalues $\lambda_{n_l}$, $\|f_{n_l}\|=1$. Then

$$ \begin{equation*} (L_0-\lambda_{n_l})f_{n_l}=-iuB_{n_l}f_{n_l}\quad\text{and} \quad \|{-iuB_{n_l}f_{n_l}}\|=u. \end{equation*} \notag $$

Set $h_{n_l}=(L_0-\lambda_0)f_{n_l}$; the norms $\|h_{n_l}\|\leqslant u+|\lambda_{n_l}-\lambda_0|$ are bounded uniformly in $n_l$. Since $\lambda_0\ne\mu_j$, the operators $(L_0-\lambda_0)^{-1}$ and $(L_0-\overline{\lambda_0})^{-1}$ are defined on the whole of $L_2[0,\pi]$ and are compact. The sequence $f_{n_l}$ is also compact and, without loss of generality, can be assumed to be convergent: $f_{n_l}\to f$, $\|f\|=1$. Thus, $B_{n_l}f_{n_l}\xrightarrow{w}0$.

We set $g=(L_0-\overline{\lambda_0})^{-1}f$; then

$$ \begin{equation} \begin{aligned} \, \notag \|f\|^2 &=\lim_{l\to\infty}(f_{n_l},(L_0-\overline{\lambda_0})g)=\lim_{l\to\infty}((L_0-\lambda_0)f_{n_l},g) \\ &=\lim_{l\to\infty}\bigl\{((L_0-\lambda_{n_l})f_{n_l},g)+iu(B_{n_l}f_{n_l},g)\bigr\} \notag \\ &=\lim_{l\to\infty}((L_{n_l}(u)-\lambda_{n_l})f_{n_l},g)=0. \end{aligned} \end{equation} \tag{2.7} $$
We have used the relations $(\lambda_{n_l}-\lambda_0)f_{n_l}\to0$ and $(B_{n_l}f_{n_l},g)\to0$.

However, $\|f\|\,{=}\,1$, and we arrive at a contradiction; hence $\lambda_n\,{\to}\,\mu_j$ for some ${j\,{\in}\,\mathbb{N}}$.

For $j>j_0$ the disc $\Gamma_j$ contains a unique eigenvalue $\lambda_{j,n}$ for each $n\in\mathbb{N}$; thus, $\lambda_{j,n}\to\mu_j$ as $n\to\infty$. However, for $j=1,\dots,j_0$ the following inconvenient cases are potentially possible:

By eliminating these possibilities we will prove that for each $\mu_j$ there is a sequence of eigenvalues $\lambda_{j,n}$ converging to $\mu_j$.

Assume that some sequence of eigenvalues $\lambda_n$ of $L_{n}(u)$ converges to $\mu_j$ as $n\to\infty$. As before, $f_n$, $\|f_n\|=1$, is the eigenfunction of $L_{n}(u)$ corresponding to $\lambda_n$. We show that we can extract a subsequence $f_{n_l}$ converging to the eigenfunction $f_0$, $\|f_0\|=1$, of $L_0$ that corresponds to $\mu_j$.

We consider the orthogonal decomposition $L_2[0,\pi]=\mathfrak{M}\oplus\mathfrak{N}$, where $\mathfrak{M}$ is the one-dimensional eigenspace of $L_0$ corresponding to $\mu_j$ and $\mathfrak{N}$ is its orthogonal complement.

Setting $f_n=f_{0n}+f_n^\perp$, where $f_{0n}\in\mathfrak{M}$ and $f_n^\perp\in\mathfrak{N}$, we write

$$ \begin{equation*} (L_0-\mu_j)f_n^\perp=-iuB_nf_n+(\lambda_n-\mu_j)f_n. \end{equation*} \notag $$
Again, we extract a convergent subsequence $f_{n_l}^\perp\to f_0^\perp\in\mathfrak{N}$. Using the boundedness of the $f_{0n}$ and the one-dimensionality of $\mathfrak{M}$, without loss of generality we can assume that $f_{0n_l}\to f_0\in\mathfrak{M}$. Repeating the reasoning in (2.7) for an arbitrary $g\in\mathfrak{N}$, we conclude that $f_0^\perp=0$. In other words, we have found a subsequence $f_{n_l}$ such that $f_{n_l}=f_{0n_l}+f_{n_l}^\perp\to f_0\in\mathfrak{M}$.

The problem with inconvenient potential cases is dealt with as follows.

$\bullet$ Assume that $\lambda_{j,n}\ne\lambda_{k,n}$ converge to the same $\mu_j$ as $n\to\infty$. Then we consider the normalized eigenfunctions $f_{n}$ and $g_n$ of $L_{n}(u)$ corresponding to $\lambda_{j,n}$ and $\lambda_{k,n}$. We extract subsequences $f_{n_l}\to f_0\in\mathfrak{M}$ and $g_{n_l}\to e^{i\gamma}f_0\in \mathfrak{M}$, $\gamma\in\mathbb{R}$ (because $\|f_0\|=\|g_0\|=1$). Since $\lambda_{j,n_l}\ne\lambda_{k,n_l}$, we have

$$ \begin{equation*} 0=\int_0^\pi f_{n_l}g_{n_l}\,dx\to e^{i\gamma}\int_0^\pi f^2_0\,dx\ne0, \end{equation*} \notag $$
which is a contradiction.

$\bullet$ A multiple eigenvalue $\lambda_n$ (corresponding to a Jordan cell) converges to $\mu_j$. Let $f_n$, $\|f_n\|=1$, be eigenfunctions and $g_n\ne0$ be associated vectors (generally, not normalized): $g_n{\perp}\, f_n$ and $(L_n(u) - \lambda_n)g_n = f_n$, so that

$$ \begin{equation} (L_0-\mu_j)g_n=-iuB_n g_n+(\lambda_n-\mu_j)g_n+f_n. \end{equation} \tag{2.8} $$
We show that the norms $\|g_n\|$ are above some $C>0$ for all $n\in \mathbb{N}$. Otherwise, there is a subsequence $g_{n_l}\to 0$. Without loss of generality we assume that $f_{n_l}\to f_0\in\mathfrak{M}$, $\|f_0\|=1$. We arrive at a contradiction:
$$ \begin{equation*} (L_0-\mu_j)g_{n_l}=f_0+o(1), \qquad l\to\infty, \end{equation*} \notag $$
so
$$ \begin{equation*} 0=\lim_{l\to\infty}(g_{n_l}, (L_0-\mu_j)f_0)=\lim_{l\to\infty}((L_0-\mu_j)g_{n_l}, f_0)=(f_0,f_0)=1. \end{equation*} \notag $$
We normalize $g_n$ in (2.8) by setting $\widehat g_n=g_n/\|g_n\|$ and obtain
$$ \begin{equation*} (L_0-\mu_j)\widehat g_n=-iuB_n \widehat g_n+(\lambda_n-\mu_j)\widehat g_n+\frac{f_n}{\|g_n\|}. \end{equation*} \notag $$
The right-hand side is bounded for all $n$; we introduce the representation $\widehat g_n=\widehat g_{0n}+\widehat g_n^\perp$, $\widehat g_{0n}\in \mathfrak{M}$, $\widehat g_n^\perp\in\mathfrak{N}$, and extract subsequences $\widehat g_{0n_l}\to \widehat g_0\in\mathfrak{M}$ and $\widehat g_{n_l}^\perp\to \widehat g_0^\perp \in\mathfrak{N}$ similarly to the construction for eigenfunctions. Without loss of generality we assume that $f_{n_l}\to f_0\in\mathfrak{M}$, $\|f_0\|=1$.

For any $h\in\mathfrak{N}$, $h=(L_0-\mu_j)v$, $v\perp f_0$, we have

$$ \begin{equation*} (\widehat g_0^\perp, h)=\lim_{l\to\infty}(\widehat g_n^\perp,(L_0-\mu_j)v)=\lim_{l\to\infty}((L_0-\mu_j)\widehat g_n,v)=0, \end{equation*} \notag $$
which yields the convergence $\widehat g_{n_l}\to \widehat g_0=e^{i\gamma}f_0\in\mathfrak{M}$, $\gamma\in\mathbb{R}$ (since $\|g_0\|=1$). The elimination of the inconvenient case under consideration follows from the contradiction
$$ \begin{equation*} 0=(\widehat g_{n_l},f_{n_l})\to e^{i\gamma}(f_0,f_0)\ne0. \end{equation*} \notag $$

Thus, we have proved that for each $\mu_j$ there exists a sequence of eigenvalues such that $\lambda_{j,n}\to \mu_j$. This fact and the localization of eigenvalues in $\Gamma$ and $\Gamma_{j}$, $j>j_0$, imply, among other things, that for any compact set $\mathcal{C}$ not containing the points $\{\mu_j\}_{j=1}^{\infty}$ there is $n_0>0$ such that $\mathcal{C}$ does not contain eigenvalues of $L_n(u)$ for $n>n_0$, which completes the proof of pointwise convergence.

Set $\mathscr{E}_0=\bigl\{y\in \mathscr{E}\mid y'(0)=y'(\pi)=0\bigr\}$.

We show that for each $u>0$ there are $n_0>0$ and $y_{n_0} \in \mathscr{E}_0$, $\|y_{n_0}\| = 1$, such that $\|L_{n_0}(u)y_{n_0}\|<2^{13}$, which will complete the construction of the counterexample.

We introduce the notation $H_n=L_n(u)\mathscr{E}_0=L_2[0,\pi]\ominus \langle y_{n,1},y_{n,2}\rangle$, where $y_{n,1}$ and $y_{n,2}$ form a fundamental system of solutions of the homogeneous equation

$$ \begin{equation*} -y''-iur_ny=0. \end{equation*} \notag $$
Here $\langle y_{n,1},y_{n,2}\rangle$ is the 2-plane spanned by $y_{n,1}$ and $y_{n,2}$.

We let $R_n(u)$ denote the inverse of $L_n(u)$. Its existence and the estimate $\|R_n(u)\|\leqslant 1$ follow from the straightforward calculation:

$$ \begin{equation*} \|My\|\|y\|\geqslant\bigl|(My,y)\bigr|=|(Ly,y)\pm i(uB_ny,y)\bigr|\geqslant(Ly,y)\geqslant \|y\|^2, \end{equation*} \notag $$
where $M=L_n(u)$ or $M=L_n(u)^*$.

We denote the restriction of $R_n(u)$ to $H_n$ by $\mathring{R}_n(u)$ and estimate its norm as an operator from $H_n$ to $L_2[0,\pi]$ as follows:

$$ \begin{equation*} \|\mathring{R}_n(u)\|^2=\sup_{f\perp\langle y_{n,1},y_{n,2}\rangle}\frac{\|R_n(u)f\|^2}{\|f\|^2}\geqslant \min_{\mathscr{L},\,\dim\mathscr{L}=2}\ \max_{f\perp\mathscr{L}} \frac{\bigl((R_n(u))^*R_n(u)f,f\bigr)}{\|f\|^2}=s_3^2, \end{equation*} \notag $$
where the minimum is taken over all possible two-dimensional subspaces of $L_2[0,\pi]$ and $s_3$ is the third singular number of the compact operator $R_n(u)$. The last equality follows from [16], Ch. II, § 1.

In view of the inequalities $1\geqslant \|R_n(u)\|=s_1\geqslant s_2$, using Weyl’s lemma we deduce that

$$ \begin{equation*} s_3s_2s_1\geqslant|\lambda_1|\,|\lambda_2|\,|\lambda_3|, \quad \text{which yields } s_3\geqslant|\lambda_1|\,|\lambda_2|\,|\lambda_3|, \end{equation*} \notag $$
where the $\lambda_j$, $j=1,2,3$, are the three largest (in modulus) eigenvalues of $R_n(u)$, which are the inverses of the three smallest (in modulus) eigenvalues of $L_n(u)$. The latter ones converge to the $\mu_j$ as $n\to\infty$, which implies that each $|\lambda_j|$ is above $1/\mu_4=1/2^4$ for some large $n_0$. Hence for $R_{n_0}(u)$ we have $s_3>1/2^{12}$ and
$$ \begin{equation*} 1\geqslant\|\mathring{R}_{n_0}(u)\|>\frac{1}{2^{12}}. \end{equation*} \notag $$
Then there exists $f_{n_0}\in H_n$, $f_{n_0}\ne0$, such that $\|R_{n_0}(u)f_{n_0}\|\geqslant (1/2^{13})\|f_{n_0}\|$.

Setting $y_{n_0}=R_{n_0}(u)f_{n_0}/\|R_{n_0}(u)f_{n_0}\|$, we find the required element of $\mathscr{E}_0$.

Theorem 3 is proved.

Proof of Theorem 4. We extract a square root of $q$ by setting $p(x)=\sqrt{q(x)}$ for ${x\geqslant0}$ and choosing the branch so that $|{\arg p(x)}|\leqslant \pi/2-\varkappa/2$. Following [17], we set
$$ \begin{equation*} \rho(x)=\operatorname{Re} p(x)-\frac{1}{2}\biggl|\frac{p'(x)}{p(x)} \biggr|, \qquad x\geqslant0. \end{equation*} \notag $$

Setting $C_0=(1-\delta)\sin\varkappa/2>0$, we derive the estimates

$$ \begin{equation} \rho(x)=\operatorname{Re} p(x)-\frac{|p(x)|}{4}\biggl| \frac{q'(x)}{q^{3/2}(x)} \biggr|\geqslant\operatorname{Re} p(x)-|p(x)|\delta\sin\frac{\varkappa}{2}\geqslant C_0|p(x)|\geqslant C_0>0, \end{equation} \tag{2.9} $$
since $\operatorname{Re} p(x)\geqslant |p(x)|\sin\varkappa/2$ in view of the inequality $|{\arg p(x)}|\leqslant \pi/2-\varkappa/2$ and since $|p(x)|\geqslant1$.

By [17], Theorem 1, $L_U$ has a bounded resolvent in some neighbourhood of zero $\lambda\in\Omega\subset\mathbb{C}$. The definite case (an analogue of the limit-point case) takes place for $L_U$. Reasoning like in the proof of Theorem 2, we assume that $L_U$ is specified by the Dirichlet boundary condition $U(y)=y(0)$ and defined on the domain $\mathscr{D}_D$ introduced in (2.1). We also derive from Theorem 1 in [17] that $\mathscr{D}_{D0}$ is a generating kernel for $L_U$.

The operator $R_0=L_U^{-1}$ is bounded and can be extended to a bounded operator $\widetilde{R_0}\colon L_2(\mathbb{R}_+,1/|q|)\to L_2(\mathbb{R}_+)$ ([17], Theorem 4). This is equivalent to the boundedness of $Rg=\widetilde{R_0}(pg)$ as an ordinary operator in $L_2(\mathbb{R}_+)$. Since $R$ being compact implies that $R_0$ is too, we aim to prove the complete continuity of the operator $R$.

The boundedness of $R$ means that

$$ \begin{equation*} M=\frac{1}{p}L_U \end{equation*} \notag $$
has a closure as an operator in $L_2(\mathbb{R}_+)$. Since $L_2(\mathbb{R}_+)\subset L_2(\mathbb{R}_+,1/|q|)$ is dense in the metric of the weighted space, this closure $\overline M$ is generated by the kernel $\mathscr{D}_D$, and $\overline M^{-1}=R$.

From the facts that $\mathscr{D}_{D0}$ is a generating kernel of $L_U$ and $|p|>1$ we conclude that $\mathscr{D}_{D0}$ is a generating kernel for $\overline M$.

We turn to the proof of the compactness of the resolvent of $\overline M$.

We show that $\overline M$ is an $\mathrm m$-sectorial operator. To do this, it suffices to show merely that for $y\in\mathscr{D}_{D0}$ the form $(My,y)$ is sectorial (see similar arguments in the proof of Theorem 2):

$$ \begin{equation} \begin{aligned} \, \notag (My,y) &=\int_0^{+\infty}\biggl(y'(x)\biggl(-\frac{p'(x)}{p^2(x)}\overline{y(x)} +\frac{1}{p(x)}\overline{y'(x)}\biggr)+p(x)|y(x)|^2\biggr)\,dx \\ &=\int_0^{+\infty}\biggl(p(x)|y(x)|^2+\overline{p(x)}\biggl|\frac{y'(x)}{p(x)}\biggr|^2 -\frac{p'}{p}\biggl(\frac{y'(x)}{p(x)}\biggr)\overline{y(x)}\biggr)\,dx. \end{aligned} \end{equation} \tag{2.10} $$

We set

$$ \begin{equation*} \xi(x)=-\frac{p'}{p}\biggl(\frac{y'(x)}{p(x)}\biggr)\overline{y(x)}, \qquad |\xi(x)|\leqslant\frac{1}{2}\biggl|\frac{p'}{p}\biggr|\biggl( \biggl|\frac{y'(x)}{p(x)}\biggr|^2+|y(x)|^2\biggr); \end{equation*} \notag $$
for convenience, we write
$$ \begin{equation*} \xi(x)=m(x)\frac{1}{2}\biggl|\frac{p'}{p}\biggr|\biggl( \biggl|\frac{y'(x)}{p(x)}\biggr|^2+|y(x)|^2 \biggr)e^{i\varphi(x)}, \end{equation*} \notag $$
where $0\leqslant m(x)\leqslant1$ and $\varphi(x)=\arg\xi(x)$. As a result, (2.10) takes the following form:
$$ \begin{equation} \begin{aligned} \, \notag (My,y) &=\int_0^{+\infty}\biggl(\biggl(p(x)+m(x)\frac{1}{2}\biggl|\frac{p'}{p}\biggr|e^{i\varphi(x)}\biggr)|y(x)|^2 \\ &\qquad +\biggl(\overline{p(x)}+m(x)\frac{1}{2}\biggl|\frac{p'}{p}\biggr|e^{i\varphi(x)} \biggr)\biggl|\frac{y'(x)}{p(x)}\biggr|^2\biggr)\,dx. \end{aligned} \end{equation} \tag{2.11} $$

We have

$$ \begin{equation*} \biggl|m(x)\frac{1}{2}\biggl|\frac{p'}{p}\biggr|e^{i\varphi(x)} \biggr|\leqslant\frac{1}{4}\biggl|\frac{q'(x)}{q^{3/2}(x)}\biggr|\,|p(x)|\leqslant \delta\sin\frac{\varkappa}{2}|p(x)|. \end{equation*} \notag $$

We turn to the terms in (2.11). In view of the inequality $|{\arg p(x)}|\leqslant\pi/2-\varkappa/2$ and the above estimate it follows from geometric considerations that

$$ \begin{equation*} \biggl|\arg\biggl( p(x)+m(x)\frac{1}{2}\biggl|\frac{p'}{p}\biggr|e^{i\varphi(x)} \biggr)\biggr|\leqslant\frac{\pi}2-\epsilon\quad\text{and} \quad \sin\biggl(\frac{\varkappa}{2}-\epsilon\biggr)\leqslant\delta\sin\frac{\varkappa}{2} \end{equation*} \notag $$
for some $0<\epsilon\leqslant\varkappa/2$; hence $|\arg(My,y)|\leqslant\pi/2-\epsilon$, and the $\mathrm m$-sectoriality of $\overline M$ is proved.

In what follows we need an estimate from below for $y\in\mathscr{D}_{D0}$, which is immediately derived from (2.11) in view of (2.9):

$$ \begin{equation*} \operatorname{Re} (My,y)\geqslant\int_0^{+\infty}C_0|p(x)|\biggl(|y(x)|^2+\biggl|\frac{y'(x)}{p(x)}\biggr|^2\biggr)\,dx. \end{equation*} \notag $$

The reasoning that follows repeats essentially the proof of Theorem 2. We dwell only on the most important points.

In a similar way, assuming that the resolvent of $\overline M$ is noncompact, we find a noncompact sequence $Y=\{y_n\}_{n=1}^{\infty}\subset\mathscr{D}_{D0}$ such that $\operatorname{Re} (My_n,y_n)<1$ for all ${n\in\mathbb{N}}$.

We find $\varepsilon_0>0$ and, for each $T>0$, consider $y_T\in Y$ such that (2.2) holds.

For $d=C_0\varepsilon_0/4$ we find an interval $D_T$ of length $d$ such that

$$ \begin{equation*} \begin{gathered} \, C_0\int_{D_T}\biggl(|p(x)|\,|y_T(x)|^2+\frac{|y_T'(x)|^2}{|p(x)|}\biggr)\,dx\leqslant\frac{1}{\varepsilon_0}\int_{D_T} |y_T(x)|^2\,dx, \\ \int_{D_T} |y_T(x)|^2\,dx>0. \end{gathered} \end{equation*} \notag $$

We perform the normalization (2.5) and finally arrive at the relations

$$ \begin{equation} C_0\int_{D_T}\biggl(|p(x)|\,|v_T(x)|^2+\frac{|v_T'(x)|^2}{|p(x)|}\biggr)\,dx\leqslant \frac{d}{\varepsilon_0}\quad\text{and} \quad \frac{1}{|D_T|} \int_{D_T}|v_T(x)|^2\,dx=1. \end{equation} \tag{2.12} $$

For any $x_1,x_2\in D_T$ we have

$$ \begin{equation*} \begin{aligned} \, &\bigl||v_T(x_1)|^2-|v_T(x_2)|^2\bigr|\leqslant \bigl|v_T^2(x_1)-v_T^2(x_2)\bigr| \\ &\qquad\leqslant2\int_{D_T}|v_T(x)|\,|v_T'(x)|\,dx =2\int_{D_T}|p(x)|^{1/2}|v_T(x)|\frac{|v_T'(x)|}{|p(x)|^{1/2}}\,dx \\ &\qquad \leqslant\int_{D_T}\biggl(|p(x)|\,|v_T(x)|^2+\frac{|v_T'(x)|^2}{|p(x)|}\biggr)\,dx \leqslant \frac{d}{C_0\varepsilon_0}=\frac{1}{4}. \end{aligned} \end{equation*} \notag $$
Taking account of (2.12) we conclude that $|v_T(x)|^2\geqslant3/4$ for all $x\in D_T$.

Turning to (2.12) again, we obtain the estimate

$$ \begin{equation*} C_0\frac{3}{4}\int_{D_T}|p(x)|\,dx\leqslant C_0\int_{D_T}|p(x)|\,|v_T(x)|^2\,dx\leqslant\frac{d}{\varepsilon_0}, \end{equation*} \notag $$
or
$$ \begin{equation*} \int_{D_T}|q(x)|^{1/2}\,dx=\int_{D_T}|p(x)|\,dx\leqslant\frac{1}{3}, \end{equation*} \notag $$
which leads to a contradiction and completes the proof.

Theorem 4 is proved.

The conditions in Theorem 4 are also necessary for the compactness of the resolvent of the operator $\overline M$, which can be proved just as in [1]. This means that the method of the proof we propose does not make it possible to relax condition (1.3).

§ 3. Necessary condition for the compactness of the resolvent

We establish a necessary condition for the compactness of the resolvent of an ordinary differential operator of an arbitrary order $n\geqslant2$. Since the results from the previous sections are not needed here, we use the same notation as there for similar objects.

Assume we are given complex-valued functions $p_1\equiv C_0=\mathrm{const}$ and $p_j\in L_{1,\mathrm{loc}}(\mathbb{R}_+)$, $j=2,\dots,n$. We introduce the differential expression

$$ \begin{equation} l(y)=\frac{d^n}{dx^n}y+\sum_{j=1}^n p_j\frac{d^{n-j}}{dx^{n-j}}y \end{equation} \tag{3.1} $$
and the linear manifold
$$ \begin{equation*} \begin{aligned} \, \mathscr{D}_0 &=\bigl\{y\in L_2(\mathbb{R}_+)\mid y,y',\dots,y^{(n-1)}\in \mathrm{AC}_{\mathrm{loc}}(\mathbb{R}_+),\ l(y)\in L_2(\mathbb{R}_+), \\ &\qquad y(0)=\dots=y^{(n-1)}(0)=0,\ \exists\, x_0>0\ \forall\, x\geqslant x_0\ y(x)=0\bigr\}. \end{aligned} \end{equation*} \notag $$

We consider the differential operator $L_0$ in $L_2(\mathbb{R}_+)$ with domain $\mathscr{D}_0$ that is specified by the differential expression (3.1).

Theorem 5. For $L_0$ to have an extension with compact resolvent, it is necessary that

$$ \begin{equation*} \lim_{x\to+\infty}\int_x^{x+a}\sum_{j=2}^n|p_j(\xi)|\,d\xi=\infty \end{equation*} \notag $$
for each $a>0$.

Before we prove the theorem, we state a lemma.

Lemma 2. Assume that continuous complex-valued functions $\phi_\nu$ and $W_\nu$, $\nu=1,\dots,n$, $n\in\mathbb{N}$, defined on a interval $[0,d]$ are such that $W_n(0)=1$ and the system of functions $\{\phi_\nu\}_{\nu=1}^{n}$ is linearly independent on $[0,d_1]$ for each $d_1>0$, $0<d_1\leqslant d$.

Then the operator $R$ in $L_2[0,d]$ defined by

$$ \begin{equation*} (Rg) (x)=\sum_{\nu=1}^n\phi_\nu(x)\int_0^x W_\nu(\xi)g(\xi)\,d\xi \end{equation*} \notag $$
is infinite-dimensional (that is, its image is not a finite-dimensional linear set).

Proof. We construct a sequence $g_l\in L_2[0,d]$, $l\in\mathbb{N}$, such that the sequence of images $f_l=Rg_l$ is linearly independent.

With this aim in view, first we construct disjoint intervals

$$ \begin{equation*} I_l=(\alpha_l,\beta_l)\subset[0,d], \qquad l=0,1,\dots, \quad 0<\dots<\alpha_1<\beta_1<\alpha_0<\beta_0, \end{equation*} \notag $$
such that the system $\{\phi_\nu\}_{\nu=1}^{n}$ is linearly independent on each $I_l$. It suffices to find one interval $I_0=(\alpha_0,\beta_0)$ of this kind and to construct further using recursion.

We choose an arbitrary $\beta_0$, $0<\beta_0<d$. If the system $\{\phi_\nu\}_{\nu=1}^{n}$ is linearly independent on $(\beta_0/2,\beta_0)$, then the construction is complete. Otherwise, let ${\mathscr{L}\subset\mathbb{C}^n}$ be a maximal linear subspace such that $\sum_\nu A_\nu\phi_\nu(x)\equiv0$ for all $A=(A_\nu)\in\mathscr{L}$ and all $x\in(\beta_0/2,\beta_0)$. Clearly, $1\leqslant\dim\mathscr{L}\leqslant n$.

Let $\alpha'$ be the infimum of $\alpha>0$ such that for the interval $(\alpha,\beta_0)$ there is $A=(A_\nu)\in\mathscr{L}$ such that $\sum_\nu A_\nu\phi_\nu(x)\equiv0$ for $x\in(\alpha,\beta_0)$. It is obvious that ${0\leqslant\alpha'\leqslant\beta_0/2}$.

We show that $\alpha'>0$. Otherwise, there exists $A_k=(A_{k\nu})\in\mathscr{L}$, $\|A_k\|=1$, $k\in \mathbb{N}$, such that $\sum_\nu A_{k\nu}\phi_\nu(x)\equiv0$ for $x\in(\beta_0/(k+1),\beta_0)$. We extract a convergent subsequence $A_{k_s}\to A_0=(A_{0\nu})\in\mathscr{L}$ as $s\to\infty$, $\|A_0\|=1$.

For $s\geqslant s_0\geqslant1$ we have

$$ \begin{equation*} \sum_\nu A_{0\nu}\phi_\nu(x)=\sum_\nu(A_{0\nu}-A_{k_s\nu})\phi_\nu(x) \quad\text{for all } x\in\biggl(\frac{\beta_0}{k_{s_0}+1},\beta_0\biggr). \end{equation*} \notag $$
Letting $s\to\infty$, as $s_0$ is arbitrary and all the $\phi_\nu$ are continuous, we infer that $\sum_\nu A_{0\nu}\phi_\nu(x)=0$ for all $x\in[0,\beta_0]$, which contradicts the independence of the system $\{\phi_\nu\}_{\nu=1}^{n}$ on the interval $[0,\beta_0]$. Thus, $\alpha'>0$.

Set $\alpha_0\!=\!\alpha'/2$. Then $\{\mkern-1mu\phi_\nu\mkern-1mu\}_{\nu=1}^{n}$ turns out to be linearly independent on ${I_0\!=\!(\mkern-1mu\alpha_0,\beta_0)}$. Otherwise, for some $B=(B_\nu)\in\mathbb{C}^n$ we have

$$ \begin{equation*} \sum_\nu B_\nu\phi_\nu(x)\equiv0 \quad\text{for } x\in(\alpha_0,\beta_0) \end{equation*} \notag $$
and therefore for all $x\in(\beta_0/2,\beta_0)$, that is, $B\in\mathscr{L}$, which contradicts the choice of $\alpha'$.

Let $I_l=(\alpha_l,\beta_l)$, $l=0,1,\dots$, be a system of disjoint intervals of linear independence of $\{\phi_\nu\}_{\nu=1}^{n}$. Let $\beta_0$ be chosen so that $W_n(x)\ne0$ for $x\in[0,\beta_0]$.

For $l\in\mathbb{N}$ we construct $g_l\in L_2[0,d]$ as follows: it is identically $0$ outside $I_l$, and the following procedure is used on $I_l$.

On $I_1$ the system of functions $\{W_\nu\}_{\nu=1}^{n}$ can be expressed in terms of $1\leqslant k\leqslant n$ independent functions $W_{s_j}$ (at any rate, $W_n(x)\not\equiv0$ on $I_l$):

$$ \begin{equation*} W_\nu=\sum_{j=1}^k A_{\nu j}W_{s_j}. \end{equation*} \notag $$
We choose $g_l\in L_2(I_l)$ so that $\overline{g_l}$ is not orthogonal to $W_{s_1}$; if $k>1$, then we additionally require that $\overline{g_l}$ is orthogonal to all $\{W_{s_j}\}_{j=2}^{n}$ in the sense of the metric in $L_2(I_l)$.

We set $f_l=Rg_l$, $l\in\mathbb{N}$. Then

$$ \begin{equation*} f_l(x)= \begin{cases} 0, & x\in[0,\alpha_l], \\ \displaystyle \sum_{\nu=1}^nA_{\nu 1}\phi_\nu(x)\int_{\alpha_l}^{\beta_l} W_{s_1}(\xi)g_l(\xi)\,d\xi, & x\in[\beta_l,d]. \end{cases} \end{equation*} \notag $$

It is significant for us that the integral in the above formula is not zero and

$$ \begin{equation*} \sum_\nu A_{\nu 1}\phi_\nu(x)\not\equiv0 \quad\text{for } x\in I_j \quad\text{for all } 0\leqslant j<l. \end{equation*} \notag $$

Thus, $f_l(x)\equiv0$ for $x\in I_j$ when $j>l$, but $f_l(x)\not\equiv0$ for $x\in I_j$ when $0\leqslant j<l$. The sequence $f_l$ is linearly independent.

Lemma 2 is proved.

Proof of Theorem 5. Assume the converse: there exists an extension $L\supset L_0$ with compact resolvent and
$$ \begin{equation} \int_{D_k}|p_j(x)|\,dx<C_j, \qquad j=1,\dots,n, \quad k\in\mathbb{N}, \end{equation} \tag{3.2} $$
for some system $D_k=[a_k,b_k]\subset\mathbb{R}_+$ of disjoint intervals of equal length $d=|D_k|$.

Since $p_1\equiv \mathrm{const}$, it is convenient to write (3.2) for all $p_j$. Without loss of generality we assume that $C_j>0$.

If necessary, we reduce the lengths of the intervals so that

$$ \begin{equation} \eta=\sum_{j=1}^nC_jd^{j-1}<1. \end{equation} \tag{3.3} $$

We arrive at a contradiction if for each $k\in\mathbb{N}$ there exists a function $y_k\in\mathscr{D}_0$ vanishing outside $D_k$ and such that $\|y_k\|=1$ and $\|L_0y_k\|<C$.

To find such functions we consider the operators $L_{k,0}$ in $L_2[0,d]$ specified by the differential expressions $l_k(y(x))=l(y(x+a_k))$, $x\in[0,d]$, on the domains

$$ \begin{equation*} \begin{aligned} \, \widetilde{\mathscr{D}_{0k}} &=\bigl\{y\in L_2[0,d]\mid y,y',\dots,y^{(n-1)}\in \mathrm{AC}[0,d],\ l_k(y)\in L_2[0,d], \\ &\qquad y(0)=\dots=y^{(n-1)}(0)=y(d)=\dots=y^{(n-1)}(d)=0\bigr\}. \end{aligned} \end{equation*} \notag $$

Once we find a sequence $f_k\in\widetilde{\mathscr{D}_{0k}}$ such that $\|f_k\|=1$ and $\|L_{k,0}f_k\|<C$ for ${k\in\mathbb{N}}$, we solve the problem by setting $y_k(x)=f_k(x-a_k)$ for $x\in D_k$ and extending $y_k$ by zero outside $D_k$.

For each homogeneous equation $l_k(y)=0$, $k\in\mathbb{N}$, we consider the special solution $\psi_{k,\nu}$ of the Cauchy problem that is specified by the initial conditions

$$ \begin{equation*} \psi_{k,\nu}^{(j-1)}(0)=\delta_\nu^j, \qquad j,\nu=1,\dots,n, \quad k\in\mathbb{N}, \end{equation*} \notag $$
where $\delta_\nu^j$ is the Kronecker symbol.

The existence of these solutions follows from [13], Ch. V, § 16, Theorem 1. All the $\psi_{k,\nu}$ and their derivatives up to the $(n-1)$st order inclusive are absolutely continuous on $[0,d]$.

We fix $k\in\mathbb{N}$. For any $g\in L_2[0,d]$ the equation $l_k(f)=g$ has a unique solution such that $f_k^{(j)}(0)=0$, $j=0,\dots,n-1$ ([13], Ch. V, § 16, Theorem 1). It can be represented as $f=R_kg$, where $R_k$ (the Cauchy operator) is a bounded Volterra operator in $L_2[0,d]$, namely,

$$ \begin{equation} f=(R_kg) (x)= \int_0^x\frac{1}{W(\xi)} \begin{vmatrix} \psi_{k,1}(\xi) & \psi_{k,2}(\xi) & \cdots & \psi_{k,n}(\xi)\\ \psi_{k,1}'(\xi) & \psi_{k,2}'(\xi) & \cdots & \psi_{k,n}'(\xi)\\ \vdots & \vdots & \ddots & \cdots \\ \psi_{k,1}^{(n-2)}(\xi) & \psi_{k,2}^{(n-2)}(\xi) & \cdots & \psi_{k,n}^{(n-2)}(\xi)\\ \psi_{k,1}(x) & \psi_{k,2}(x) & \cdots & \psi_{k,n}(x) \end{vmatrix} g(\xi)\,d\xi, \end{equation} \tag{3.4} $$
where $W(\xi)$ is the Wronskian:
$$ \begin{equation*} \begin{aligned} \, W(\xi)&= \begin{vmatrix} \psi_{k,1}(\xi) & \psi_{k,2}(\xi) & \cdots & \psi_{k,n}(\xi)\\ \psi_{k,1}'(\xi) & \psi_{k,2}'(\xi) & \cdots & \psi_{k,n}'(\xi)\\ \vdots & \vdots & \ddots & \cdots \\ \psi_{k,1}^{(n-2)}(\xi) & \psi_{k,2}^{(n-2)}(\xi) & \cdots & \psi_{k,n}^{(n-2)}(\xi)\\ \psi_{k,1}^{(n-1)}(\xi) & \psi_{k,2}^{(n-1)}(\xi) & \cdots & \psi_{k,n}^{(n-1)}(\xi)\\ \end{vmatrix} \\ &=\exp\biggl(-\int_0^\xi p_1(a_k{+}\,t)\,dt\biggr)=e^{-\xi C_0}. \end{aligned} \end{equation*} \notag $$

Differentiating (3.4) $n-1$ times, we see that $f\in\widetilde{\mathscr{D}_{0k}}$ if and only if

$$ \begin{equation*} \int_0^d e^{\xi C_0} \begin{vmatrix} \psi_{k,1}(\xi) & \psi_{k,2}(\xi) & \cdots & \psi_{k,n}(\xi)\\ \psi_{k,1}'(\xi) & \psi_{k,2}'(\xi) & \cdots & \psi_{k,n}'(\xi)\\ \vdots & \vdots & \ddots & \cdots \\ \psi_{k,1}^{(n-2)}(\xi) & \psi_{k,2}^{(n-2)}(\xi) & \cdots & \psi_{k,n}^{(n-2)}(\xi)\\ \psi_{k,1}^{(s)}(d) & \psi_{k,2}^{(s)}(d) & \cdots & \psi_{k,n}^{(s)}(d) \end{vmatrix} g(\xi)\,d\xi=0, \qquad s=0,\dots,n-1, \end{equation*} \notag $$
or, in other words, if $g\in\mathring{\mathfrak{H}}_k=L_2[0,d]\ominus\mathfrak{K}_k$, where $\mathfrak{K}_k$ is the finite-dimensional subspace of $L_2[0,d]$ spanned by the complex conjugates of the functions
$$ \begin{equation*} W_{ks}(x)= e^{x C_0} \begin{vmatrix} \psi_{k,1}(x) & \psi_{k,2}(x) & \cdots & \psi_{k,n}(x)\\ \psi_{k,1}'(x) & \psi_{k,2}'(x) & \cdots & \psi_{k,n}'(x)\\ \vdots & \vdots & \ddots & \cdots \\ \psi_{k,1}^{(n-2)}(x) & \psi_{k,2}^{(n-2)}(x) & \cdots & \psi_{k,n}^{(n-2)}(x)\\ \psi_{k,1}^{(s)}(d) & \psi_{k,2}^{(s)}(d) & \cdots & \psi_{k,n}^{(s)}(d) \end{vmatrix}, \qquad s=0,\dots,n-1; \end{equation*} \notag $$
thus, $\dim\mathfrak{K}_k\leqslant n$.

The operator $L_{k,0}$ maps $\widetilde{\mathscr{D}_{0k}}$ bijectively onto $\mathring{\mathfrak{H}}_k$. We let $\mathring{R}_k\subset R_k$ denote its inverse, which acts bijectively from $\mathring{\mathfrak{H}}_k$ onto $\widetilde{\mathscr{D}_{0k}}$.

To find the required sequence $f_k$, it suffices to show that the norms $\|\mathring{R}_k$ are uniformly bounded below for all $k\in\mathbb{N}$: $\|\mathring{R}_k\|\geqslant C>0$. Then for each $k$ there exists $u_k\in\mathring{\mathfrak{H}}_k$, $u_k\not\equiv 0$, such that $\|v_k\|\geqslant (C/2)\|u_k\|$ for $v_k=\mathring{R}_ku_k$. The required $f_k$ is $v_k/\|v_k\|$; we have $\|L_{k,0}f_k\|\leqslant 2/C$, and the following estimate holds:

$$ \begin{equation*} \|\mathring{R}_k\|^2=\sup_{g\in\mathring{\mathfrak{H}}_k}\frac{\|R_kg\|^2}{\|g\|^2} \geqslant\min_{\substack{\mathfrak{L}\subset L_2[0,d]\\ \dim\mathfrak{L}\leqslant n}}\ \max_{\substack{g\in L_2[0,d]\\ g\perp\mathfrak{L}}} \frac{(R_k^*R_k g,g)}{(g,g)}=s_{n+1}^2(R_k). \end{equation*} \notag $$
The minimum is taken over all possible linear sets $\mathfrak{L}\subset L_2[0,d]$ of dimension at most $n$, and $s_{n+1}(R_k)$ is the $(n+1)$st singular number of the operator $R_k$ The last equality follows from [16], Ch. II, § 1.

Next, we show that from the sequence of operators $R_k$ we can extract a subsequence $R_{k_m}\to R$, $m\to\infty$, converging uniformly to an infinite-dimensional operator $R$, all of whose $s$-numbers are thus positive. The singular numbers $s_l(R_{k_m})$ converge to $s_l(R)>0$, $l\in\mathbb{N}$ (see [16], Ch. II, § 2, Corollary 2.3). In particular, for ${m>m_0}$ $\|\mathring{R}_{k_m}\|\geqslant s_{n+1}(R_{k_m})\geqslant s_{n+1}(R)/2=C>0$.

Considering the $D_k$ with indices $k_m$, we can assume without loss of generality that $\|\mathring{R}_k\|\geqslant C>0$ for all $k\in\mathbb{N}$ . This completes the proof.

We show how to extract a uniformly convergent subsequence from $R_k$.

We integrate the identity $l_k(\psi_{k,\nu})=0$ $s$ times (for $s=1,\dots,n$):

$$ \begin{equation} \begin{aligned} \, \notag \psi_{k,\nu}^{(n-s)}(x) &=\sum_{r=0}^{s-1}\frac{x^r}{r!}\psi_{k,\nu}^{(n-s+r)}(0) \\ &\qquad-\int_0^x d\xi_1\int_0^{\xi_1}\cdots\,d\xi_{s-1}\int_0^{\xi_{s-1}} \sum_{j=1}^n \widetilde{p_{jk}}(\xi_s)\psi_{k,\nu}^{(n-j)}(\xi_s)\,d\xi_s, \end{aligned} \end{equation} \tag{3.5} $$
where $\widetilde{p_{jk}}(x)=p_j(x+a_k)$. There are no inner integrals for $s=1$.

We introduce the notation $M_{k,\nu,l}=\max|\psi_{k,\nu}^{(l)}(x)|$, $x\in[0,d]$, $l=0,\dots,n-1$. Using (3.5) and (3.2) we conclude that

$$ \begin{equation*} M_{k,\nu,n-s}\leqslant e^d+d^{s-1}\sum_{j=1}^nC_j M_{k,\nu,n-j}. \end{equation*} \notag $$
Multiplying both sides by $C_s$ and summing over $s\,{=}\,1,\dots,n$ we obtain the inequality
$$ \begin{equation*} \sum_{s=1}^{n}C_sM_{k,\nu,n-s}\leqslant e^d\sum_{s=1}^{n}C_s+\eta\sum_{j=1}^nC_j M_{k,\nu,n-j}, \end{equation*} \notag $$
which, in view of (3.3), yields the estimate
$$ \begin{equation*} \sum_{s=1}^{n}C_sM_{k,\nu,n-s}\leqslant\frac{e^d}{1-\eta}\sum_{s=1}^{n}C_s. \end{equation*} \notag $$

Since $C_s>0$, this estimate means that all quantities $M_{k,\nu,l}$ are bounded uniformly in $k\in\mathbb{N}$, $\nu=1,\dots,n$ and $l=0,\dots,n-1$.

For fixed $\nu=1,\dots,n$ and $l=0,\dots,n-2$, we consider the sequence $\{\psi_{k,\nu}^{(l)}\}_{k\in\mathbb{N}}$. It is precompact in the uniform metric on $[0,d]$ by the Arzelà–Ascoli theorem. We choose a sequence of indices $k_m$ so that the subsequences $\{\psi_{k_m,\nu}^{(l)}\}_{m\in\mathbb{N}}$ are uniformly convergent on $[0,d]$ for all $\nu=1,\dots,n$ and $l=0,\dots,n-2$:

$$ \begin{equation*} \psi_{k_m,\nu}^{(l)}\Rightarrow\phi_\nu^l, \qquad m\to\infty, \end{equation*} \notag $$
where the $\phi_\nu^l$ are continuous functions with indices $\nu=1,\dots,n$ and $l=0,\dots,n-2$.

For convenience we set $\phi_\nu=\phi_\nu^0$. In view of uniform convergence it follows from the well-known theorem on limit transition ([18], Ch. XVI, Theorem 4) that each function $\phi_\nu$ has $ n-2 $ derivatives and $\phi_\nu^l=\phi_\nu^{(l)}$, $l=0,\dots,n-2$.

Noting that in the definition of $R_k$ the determinant depends only on the derivatives $\psi_{k,\nu}^{(l)}$ of order at most $n-2$, we conclude that $R_{k_m}\Rightarrow R$ as $m\to\infty$ in the sense of uniform operator convergence, where

$$ \begin{equation} (Rg) (x)= \int_0^xe^{\xi C_0} \begin{vmatrix} \phi_1(\xi) & \phi_2(\xi) & \cdots & \phi_n(\xi)\\ \phi_1'(\xi) & \phi_2'(\xi) & \cdots & \phi_n'(\xi)\\ \vdots & \vdots & \ddots & \cdots \\ \phi_1^{(n-2)}(\xi) & \phi_2^{(n-2)}(\xi) & \cdots & \phi_n^{(n-2)}(\xi)\\ \phi_1(x) & \phi_2(x) & \cdots & \phi_n(x) \end{vmatrix} g(\xi)\,d\xi, \qquad g\,{\in}\, L_2[0,d]. \end{equation} \tag{3.6} $$

The operator $R$ is compact as a uniform limit of compact operators. It remains to prove that $R$ is infinite dimensional.

Note that the system of functions $\{\phi_\nu\}_{\nu=1}^{n}$ is linearly independent on each interval $[0,d_1]\subset[0,d]$.

Assume the converse: there are $A_\nu\in\mathbb{C}$ such that $\sum_\nu A_\nu\phi_\nu(x) \equiv 0$ for all $x\in[0,d_1]$. We turn to (3.5) to express the linear combinations $\sum_\nu A_\nu\psi_{k,\nu}^{(n-s)}$ for ${s=n}$. We obtain

$$ \begin{equation*} \begin{aligned} \, \sum_{\nu=1}^n A_\nu\psi_{k,\nu}(x) &=\sum_{r=0}^{n-1}\frac{x^r}{r!}\sum_{\nu=1}^nA_\nu\psi_{k,\nu}^{(r)}(0) \\ &\qquad-C_0\int_0^x d\xi_1\int_0^{\xi_1}\cdots\,d\xi_{n-1}\int_0^{\xi_{n-1}} \sum_{\nu=1}^nA_\nu\psi_{k,\nu}^{(n-1)}(\xi_n)\,d\xi_n \\ &\qquad-\int_0^x d\xi_1\int_0^{\xi_1}\cdots\,d\xi_{n-1}\int_0^{\xi_{n-1}} \sum_{j=2}^n \widetilde{p_{jk}}(\xi_n)\sum_{\nu=1}^nA_\nu\psi_{k,\nu}^{(n-j)}(\xi_n)\,d\xi_n. \end{aligned} \end{equation*} \notag $$
Setting $k=k_m$ and taking the limit as $m\to\infty$, uniformly with respect to $x\in[0,d_1]$ we arrive at the identity
$$ \begin{equation*} 0\equiv\sum_{r=0}^{n-1}\frac{x^r}{r!}\sum_{\nu=1}^nA_\nu\delta_\nu^{r+1}(0) +C_0 \sum_{r=1}^{n-1}\frac{x^r}{r!}\sum_{\nu=1}^nA_\nu\delta_\nu^r, \end{equation*} \notag $$
or
$$ \begin{equation*} 0\equiv A_1+\sum_{r=1}^{n-1}\bigl(A_{r+1}+C_0A_r\bigr)\frac{x^r}{r!}, \end{equation*} \notag $$
which shows that $A_\nu=0$, $\nu=1,\dots,n$.

We represent $R$ in the form

$$ \begin{equation*} (Rg) (x)=\int_0^x\sum_{\nu=1}^n\phi_\nu(x)W_\nu(\xi)g(\xi)\,d\xi, \qquad g\in L_2[0,d], \end{equation*} \notag $$
where the $W_\nu$ are continuous functions. These functions are algebraic complements, multiplied by $e^{\xi C_0}$, in the expansion of the determinant in (3.6) with respect to the last row, and $W_n(0)=1$. Lemma 2 shows that $R$ is infinite dimensional.

Theorem 5 is proved.

Acknowledgments

The author is grateful to the corresponding member of the RAS, professor Andrei Andreevich Shkalikov for his attention, support and keen interest.


Bibliography

1. A. M. Molchanov, “On conditions for discreteness of the spectrum of self-adjoint differential equations of the second order”, Tr. Mosk. Mat. Obshch., 2, GITTL, Moscow, 1953, 169–199 (Russian)  mathnet  mathscinet  zmath
2. I. Brinck, “Self-adjointness and spectra of Sturm–Liouville operators”, Math. Scand., 7 (1959), 219–239  crossref  mathscinet  zmath
3. V. B. Lidskii, “A non-selfadjoint operator of Sturm-Liouville type with a discrete spectrum”, Tr. Mosk. Mat. Obshch., 9, GIFML, Moscow, 1960, 45–79 (Russian)  mathnet  mathscinet  zmath
4. R. S. Ismagilov, “Conditions for the semiboundedness and discreteness of the spectrum for one-dimensional differential equations”, Soviet Math. Dokl., 2 (1961), 1137–1140  mathnet  mathscinet  zmath
5. M. Sh. Birman and B. S. Pavlov, “On the complete continuity of certain imbedding operators”, Vestn. Leningrad. Univ. Ser. Mat. Mekh. Astronom., 16:1 (1961), 61–74 (Russian)  mathscinet  zmath
6. B. M. Levitan and G. A. Suvorchenkova, “Sufficient conditions for a discrete spectrum in the case of a Sturm–Liouville equation with operator coefficients”, Funct. Anal. Appl., 2:2 (1968), 147–152  mathnet  crossref  mathscinet  zmath
7. D. Fortunato, “Remarks on the non self-adjoint Schrödinger operator”, Comment. Math. Univ. Carolin., 20:1 (1979), 79–93  mathscinet  zmath
8. R. S. Ismagilov and A. G. Kostyuchenko, “On the spectrum of a vector Schrödinger operator”, Funct. Anal. Appl., 41:1 (2007), 31–41  mathnet  crossref  mathscinet  zmath
9. E. C. Titchmarsh, Eigenfunction expansions associated with second-order differential equations, Clarendon Press, Oxford, 1946, 175 pp.  mathscinet  zmath
10. M. A. Naimark, “Investigation of the spectrum and the expansion in eigenfunctions of a nonselfadjoint operator of the second order on a semi-axis”, Tr. Mosk. Mat. Obshch., 3, GITTL, Moscow, 1954, 181–270 (Russian)  mathnet  mathscinet  zmath
11. E. S. Birger and G. A. Kaljabin, “The theory of Weil limit-circles in the case of non-self-adjoint second-order differential-equation systems”, Differ. Equ., 12:9 (1977), 1077–1084  mathnet  mathscinet  zmath
12. Kh. K. Ishkin, “Spectral properties of the nonsectorial Sturm–Liouville operator on the semiaxis”, Math. Notes, 113:5 (2023), 663–679  mathnet  crossref  mathscinet  zmath
13. M. A. Naimark, Linear differential operators, 3d ed., Fizmatlit, Moscow, 2010, 528 pp.  mathscinet  zmath; English transl. of 1st ed., v. I, II, Frederick Ungar Publishing Co., New York, 1967, 1968, xiii+144 pp., xv+352 pp.  mathscinet  mathscinet  zmath
14. T. Kato, Perturbation theory for linear operators, Grundlehren Math. Wiss., 132, Springer-Verlag New York, Inc., New York, 1966, xix+592 pp.  crossref  mathscinet  zmath
15. B. M. Brown, D. K. R. McCormack, W. D. Evans and M. Plum, “On the spectrum of second-order differential operators with complex coefficients”, R. Soc. Lond. Proc. Ser. A Math. Phys. Eng. Sci., 455:1984 (1999), 1235–1257  crossref  mathscinet  zmath  adsnasa
16. I. C. Gohberg and M. G. Krein, Introduction to the theory of linear nonselfadjoint operators, Transl. Math. Monogr., 18, Amer. Math. Soc., Providence, RI, 1969, xv+378 pp.  mathscinet  mathscinet  zmath  zmath
17. S. N. Tumanov, “One condition for the discreteness of the spectrum and compactness of the resolvent of a nonsectorial Sturm–Liouville operator on a semiaxis”, Dokl. Math., 107:2 (2023), 117–119  mathnet  crossref  mathscinet  zmath
18. V. A. Zorich, Mathematical analysis, v. II, Universitext, 2nd ed., Springer, Heidelberg, 2016, xx+720 pp.  crossref  mathscinet  zmath

Citation: S. N. Tumanov, “Molchanov's criterion for compactness of the resolvent for a nonselfadjoint Sturm–Liouville operator”, Sb. Math., 215:9 (2024), 1249–1268
Citation in format AMSBIB
\Bibitem{Tum24}
\by S.~N.~Tumanov
\paper Molchanov's criterion for compactness of the resolvent for a~nonselfadjoint Sturm--Liouville operator
\jour Sb. Math.
\yr 2024
\vol 215
\issue 9
\pages 1249--1268
\mathnet{http://mi.mathnet.ru/eng/sm10066}
\crossref{https://doi.org/10.4213/sm10066e}
\mathscinet{https://mathscinet.ams.org/mathscinet-getitem?mr=4837042}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2024SbMat.215.1249T}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001375658800001}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85212503410}
Linking options:
  • https://www.mathnet.ru/eng/sm10066
  • https://doi.org/10.4213/sm10066e
  • https://www.mathnet.ru/eng/sm/v215/i9/p125
  • This publication is cited in the following 1 articles:
    Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Математический сборник Sbornik: Mathematics
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2025