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

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Izv. RAN. Ser. Mat.:
Year:
Volume:
Issue:
Page:
Find






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


Izvestiya: Mathematics, 2024, Volume 88, Issue 1, Pages 114–132
DOI: https://doi.org/10.4213/im9438e
(Mi im9438)
 

Interpolating asymptotic integration methods for second-order differential equations

S. A. Stepin

Lomonosov Moscow State University, Faculty of Mechanics and Mathematics
References:
Abstract: The problem of asymptotic behaviour at infinity of solutions to second-order differential equation can be reduced via the Liouville transform to that of an equation with almost constant coefficients. In the present paper, we compare various methods of asymptotic integration in application to the reduced equation $u''-(\lambda^2+\varphi(t))u=0$ and interpolate the corresponding results in the case $\operatorname{Re}\lambda>0$, provided that a complex-valued function $\varphi(t)$ is in a certain sense small for large values of the argument.
Keywords: asymptotic integration, comparison equation, retraction principle, Lyapunov type function.
Received: 17.11.2022
Russian version:
Izvestiya Rossiiskoi Akademii Nauk. Seriya Matematicheskaya, 2024, Volume 88, Issue 1, Pages 121–140
DOI: https://doi.org/10.4213/im9438
Bibliographic databases:
Document Type: Article
UDC: 517.9
Language: English
Original paper language: Russian

§ 1. Introduction and statement of the results

A wide class of second-order differential equations

$$ \begin{equation} \frac{d^2y}{dx^2}-Q(x)y=0 \end{equation} \tag{1} $$
can be reduced (cf. [1]) to
$$ \begin{equation} u''-(\lambda^2+\varphi(t))u=0, \end{equation} \tag{2} $$
with $\lambda=1$ after passing to the new independent variable
$$ \begin{equation*} t(x)=\int_{x_0}^x\sqrt{Q(s)}\,ds \end{equation*} \notag $$
and substituting $u(t)=(Q(x(t)))^{1/4}y(x(t))$. Here, $-2\varphi(t)$ is the Schwarzian derivative $S[x(t)]$, so that
$$ \begin{equation*} \varphi(t(x))=-\frac14\, \frac{Q''(x)}{Q(x)^2}+\frac5{16}\, \frac{Q'(x)^2}{Q(x)^3}. \end{equation*} \notag $$

The above two substitutions are known as the Liouville transform. If $Q(x)$ behaves quite regularly at infinity and if the integral $\int^{\infty}\!\!\sqrt{Q(s)}\,ds$ diverges, the function $\varphi$ turns out to be small in a certain sense for large $t$. Sometimes several transformations of such a type are required to achieve the desired smallness of $\varphi$.

From the perturbation theory point of view it is natural to expect that solutions to the transformed equation (2) and to the associated comparison equation corresponding to the case $\varphi(t)\equiv 0$ are asymptotically close at infinity, and, therefore, the original equation (1) has a fundamental system of solutions

$$ \begin{equation*} y_{1,2}(x)\sim Q(x)^{-1/4} \exp\biggl(\mp\int_{x_0}^x\sqrt{Q(s)}\,ds\biggr),\qquad x\to\infty. \end{equation*} \notag $$
The above asymptotic formulas are known as Liouville–Green approximations (or WKB-approximations; see [2]). Note that under the condition
$$ \begin{equation*} \int^{\infty}\biggl(\frac{|Q''(x)|}{Q(x)^{3/2}} +\frac{Q'(x)^2}{Q(x)^{5/2}}\biggr)\,dx<\infty \end{equation*} \notag $$
the integral $\int^{\infty}\!\!\sqrt{Q(s)}\,ds$ diverges (see [3]) and the function $\varphi(t)$ is integrable on the half-axis $\mathbb R_+$. Below, the function $\varphi(t)$ in (2) is assumed to be continuous.

In order to study the asymptotic behaviour of solutions to equation (2) in the case $\varphi\in\mathrm{L}_1(\mathbb R_+)$, we reduce the corresponding first-order system

$$ \begin{equation*} \begin{pmatrix}u\\u'\end{pmatrix}'=\begin{pmatrix}0&1 \\ \lambda^2+\varphi(t)&0\end{pmatrix} \begin{pmatrix}u\\u'\end{pmatrix} \end{equation*} \notag $$
to the $L$-diagonal form
$$ \begin{equation} Y'=\biggl\{\begin{pmatrix} \lambda &0 \\ 0 &-\lambda \end{pmatrix}+\frac{\varphi(t)}{2\lambda}\begin{pmatrix} 1 &1 \\ -1 &-1 \end{pmatrix}\biggr\}Y \end{equation} \tag{3} $$
via the substitution $\left(\begin{smallmatrix}u\\u'\end{smallmatrix}\right) =\left(\begin{smallmatrix}1&1\\ \lambda&-\lambda\end{smallmatrix}\right)Y$, where $Y=\left(\begin{smallmatrix}\xi\\\eta\end{smallmatrix}\right)$. Now an application of the Levinson’s fundamental theorem (see [4]) produces the following result.

Proposition 1. If $\varphi\in\mathrm{L}_1(\mathbb R_+)$, then there exists a pair of linearly independent solutions to equation (2) such that

$$ \begin{equation*} \begin{aligned} \, u_1(t)&=\exp(-\lambda t)\biggl\{1+O\biggl(\int_t^{\infty}|\varphi(s)|\,ds\biggr)\biggr\}, \\ u_2(t)&=\exp(\lambda t)\biggl\{1+O\biggl(\int_t^{\infty}|\varphi(s)|\,ds +\int_0^t|e^{2\lambda(s-t)}\varphi(s)|\,ds\biggr)\biggr\}. \end{aligned} \end{equation*} \notag $$

In the case $\varphi\in\mathrm{L}_2(\mathbb R_+)$, one can apply to equation (2) another method of asymptotic integration based on the substitution $u(t)=\exp\bigl(\mp\lambda t+\int_0^tw(s)\,ds\bigr)$ and transformation of (2) to the Riccati equation

$$ \begin{equation*} w'\mp2\lambda w=\varphi(t)-w^2, \end{equation*} \notag $$
after which this equation is reduced to an integral equation with quadratic non-linear term (see [4]). The following result is obtained within this approach.

Proposition 2. Let $\varphi\in\mathrm{L}_2(\mathbb R_+)$ and let $\varphi(t)\to 0$ as $t\to\infty$. Then equation (2) has a fundamental system of solutions

$$ \begin{equation*} u_{1,2}(t)=\exp\left(\mp\lambda t\mp\frac1{2\lambda}\int_0^t\varphi(s)\,ds\right) \bigl(1+\varepsilon_{1,2}(t)\bigr), \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{aligned} \, \varepsilon_1(t)&= O\biggl(\int_t^{\infty}|\varphi(s)|^2\,ds +\int_t^{\infty}|e^{2\lambda(t-s)}\varphi(s)|\,ds\biggr), \\ \varepsilon_2(t)&= O\biggl(\int_t^{\infty}|\varphi(s)|^2\, ds +\int_0^t|e^{2\lambda(s-t)}\varphi(s)|\,ds\biggr). \end{aligned} \end{equation*} \notag $$

The similarity between the asymptotic formulas for the solutions $u_{1,2}(t)$ with sharp estimates of the remainder terms, as given by Propositions 1 and 2, becomes even more vivid if the formula for $u_1(t)$ in Proposition 1 is written as

$$ \begin{equation*} u_1(t)=\exp(-\lambda t)\biggl\{1+O\biggl(\int_t^{\infty}|\varphi(s)|\,ds +\int_t^{\infty}|e^{2\lambda(t-s)}\varphi(s)|\,ds\biggr)\biggr\}. \end{equation*} \notag $$
The natural question here is whether the above formulas can be derived within a unified approach capable of interpolating the specified summability conditions for the function $\varphi$. It turns out that the desired interpolation in the classes $\mathrm{L}_p(\mathbb R_+)$, $p\in[1,2]$, can be achieved via an approach based on the known principle (see [5]) involving the topological notions of a retract and retraction (see § 2).

Theorem 1. If $\varphi\in\mathrm{L}_p(\mathbb R_+)$, $p\in[1,2]$, then equation (2) has solutions

$$ \begin{equation} u_{1,2}(t)=\exp\biggl(\mp\lambda t\mp\frac1{2\lambda} \int_0^t\varphi(s)\,ds\biggr) \bigl(1+\varepsilon_{1,2}(t)\bigr), \end{equation} \tag{4} $$
where
$$ \begin{equation*} \begin{aligned} \, \varepsilon_1(t)&= O\biggl(\int_t^{\infty}|\varphi(s)|^p\,ds +\int_t^{\infty}|e^{2\lambda(t-s)}\varphi(s)|\,ds\biggr), \\ \varepsilon_2(t)&= O\biggl(\int_t^{\infty}|\varphi(s)|^p\,ds +\int_0^t|e^{2\lambda(s-t)}\varphi(s)|\,ds\biggr). \end{aligned} \end{equation*} \notag $$

The estimates for the remainder terms $\varepsilon_{1,2}(t)$ in formulas (4) coincide with those in Propositions 1 and 2 for $p=1$ and $p=2$, respectively, and can be actually considered as their deformations with respect to the parameter $p\in[1,2]$. The example $\varphi(t)=\sin t/\sqrt{t}$ shows that the condition of Theorem 1, which provides the asymptotic formulas

$$ \begin{equation*} u_{1,2}(t)\sim\exp\biggl(\mp\lambda t \mp\frac1{2\lambda}\int_0^t\varphi(s)\,ds\biggr),\qquad t\to\infty, \end{equation*} \notag $$
for the solutions to equation (2), is sharp in the scale of space of integrable functions $\mathrm{L}_p(\mathbb R_+)$. Indeed, the corresponding equation (2) for $\lambda=1$ turns out to have (see § 5) solutions behaving asymptotically as $u_{1,2}(t)\sim t^{\pm1/20}e^{\mp t}$.

From the point of view of comparative analysis of various methods of asymptotic integration it is worth comparing the above assumptions on the function $\varphi(t)$ that provide the asymptotic equivalence as $t\to\infty$ of the solutions to equation (2) and to the corresponding model comparison equation with Hartman and Wintner type conditions (whose approach was developed in [6]–[9]). In [10], a number of results of this type (heterogeneous, at first glance) were unified by a method based on Leninson’s fundamental theorem. To be exact, we formulate here two such results (see [7], [8]), which we combine into a single proposition. For the sake of simplicity, we restrict to the case $\lambda>0$, which is non-elliptic in the language of [5].

Proposition 3. Let the integral $\int^{\infty}\varphi(t)\,dt$ converge (no matter absolutely or not) and let either of the conditions

$$ \begin{equation*} \begin{aligned} \, \Phi_0(t)&= \int_t^{\infty} \biggl|\int_s^{\infty}\varphi(r)\,dr\biggr|\mu(s)\,ds<\infty, \\ \Phi_1(t)&= \int_t^{\infty} \biggl|\int_s^{\infty}\varphi(r)e^{2\lambda(s-r)}\,dr\biggr|\mu(s)\,ds<\infty, \end{aligned} \end{equation*} \notag $$
be met with $\mu(t)=\sup_{s\geqslant t}\bigl|\int_s^{\infty}\varphi(r)\,dr\bigr|$. Then equation (2) has the fundamental system of solutions
$$ \begin{equation*} \begin{aligned} \, u_1(t)&= \exp(-\lambda t)\{1+O(\mu(t))+O(\Phi_{\kappa}(t))\}, \\ u_2(t)&= \exp(\lambda t) \biggl\{1+O\biggl(\int_0^te^{2\lambda(s-t)}\mu(s)\,ds\biggr) +O(\Phi_{\kappa}(t))\biggr\}, \end{aligned} \end{equation*} \notag $$
where $\kappa=0$ or $\kappa=1$, respectively.

The complementary convergence conditions for $\Phi_0(t)$ and $\Phi_1(t)$ are actually obtained one from another by deformation and are included into a one-parameter family of conditions, which guarantee the equivalence of solutions to equation (2) and to the comparison equation as $t\to\infty$. The desired interpolation between $\Phi_0(t)$ and $\Phi_1(t)$ is secured by the method based on the asymptotic factorization of the fundamental matrix for the corresponding first-order system (see § 6).

Theorem 2. Assuming that the integral $\int^{\infty}\varphi(t)\,dt$ converges, let the condition

$$ \begin{equation} \Phi_{\kappa}(t)=\int_t^{\infty} \biggl|\int_s^{\infty}\varphi(r)e^{2\kappa\lambda(s-r)}\, dr\biggr| \mu(s)\,ds <\infty, \end{equation} \tag{5} $$
be met for some parameter $\kappa\in[0,1]$, where $\mu(t)=\sup_{s\geqslant t}\bigl|\int_s^{\infty}\varphi(r)\,dr\bigr|$. Then equation (2) has solutions
$$ \begin{equation*} u_{1,2}(t)=\exp\bigl(\mp\lambda t\bigr)\{1+\delta_{1,2}(t)+O(\Phi_{\kappa}(t))\}, \end{equation*} \notag $$
and the remainder terms behave as
$$ \begin{equation*} \delta_1(t)=O(\mu(t)),\qquad \delta_2(t)=O\biggl(\int_0^te^{2\lambda(s-t)}\mu(s)\,ds\biggr). \end{equation*} \notag $$

The present paper is organized as follows. Lyapunov type functions are defined in § 2, where we also give an appropriate version of the retraction principle which applies directly to system (3). In § 3, we derive some auxiliary integral estimates, which will be used below. Further, in § 4 and § 5, we obtain a priori estimates for the solutions to equation (2). These estimates are then used for asymptotic integration of this equation (see the proof of Theorem 1). Corresponding counterexamples are also given. The sketch of the reduction method based on a transformation of a first-order system to the $L$-diagonal form is outlined in § 6. Finally, this method is applied in § 7 to the asymptotic integration problem under consideration (see the proof of Theorem 2). As a result, we obtain new conditions for asymptotic equivalence for solutions to equation (2) and to the corresponding comparison equation.

Examples show that the classes of functions $\varphi(t)$ satisfying the hypotheses of Theorems 1 and 2 are in general position. The corresponding results can be also given for equation (1) in the original representation (before application of the Liouville transform). Under proper conditions, these results provide effective estimates for the accuracy of Liouville–Green approximation for the corresponding solutions. Some of the results of the present paper were announced in [11].

§ 2. Lyapunov type functions and their properties

We set

$$ \begin{equation*} \sigma(t)=\int_0^te^{2\alpha(s-t)}|\varphi(s)|\,ds,\qquad \tau(t)= \int_t^{\infty}e^{2\alpha(t-s)}|\varphi(s)|\,ds, \end{equation*} \notag $$
where $\alpha=\operatorname{Re}\lambda>0$, and, for $a\geqslant 2/|\lambda|$, define the Lyapunov type functions by
$$ \begin{equation*} v^{(+)}(t,Y)=|\eta|^2-a^2\sigma(t)^2|\xi|^2,\qquad v^{(-)}(t,Y)=|\xi|^2-a^2\tau(t)^2|\eta|^2. \end{equation*} \notag $$

Statement 1. Let $Y(t)=\left(\begin{smallmatrix}\xi(t)\\\eta(t)\end{smallmatrix}\right)$ be a solution to system (3). Then

$$ \begin{equation*} \frac{d}{dt}v^{(+)}(t,Y(t))\leqslant 0 \end{equation*} \notag $$
on the null-level of the function $v^{(+)}(t,Y)$ for $t\geqslant T$ such that $\sigma(t)<1/a$.

Proof. Let us evaluate the derivative
$$ \begin{equation*} \frac12\,\frac{d}{dt}\, v^{(+)}(t,Y(t)) =\operatorname{Re}(\eta'\overline{\eta}) -a^2\sigma(t)\sigma'(t)|\xi|^2-a^2\sigma(t)^2\operatorname{Re}(\xi'\overline{\xi}), \end{equation*} \notag $$
where
$$ \begin{equation*} \xi'=\lambda\xi+\frac{\varphi(t)}{2\lambda}(\xi+\eta),\qquad \eta'=-\lambda\eta-\frac{\varphi(t)}{2\lambda}(\xi+\eta), \end{equation*} \notag $$
and, in addition, $\sigma'=|\varphi(t)|-2\alpha\sigma$. As a result, on the null-level of the function $v^{(+)}(t,Y)$, that is, for $|\eta|=a\sigma(t)|\xi|$, we have
$$ \begin{equation*} \begin{aligned} \, &\frac12\,\frac{d}{dt}\, v^{(+)}(t,Y(t)) =\operatorname{Re}\biggl(-\lambda|\eta|^2 -\frac{\varphi(t)}{2\lambda}(\xi+\eta)\overline{\eta}\biggr) \\ &\quad\qquad- a^2\sigma(t)\bigl(|\varphi(t)|-2\alpha\sigma(t)\bigr)|\xi|^2 -a^2\sigma(t)^2\operatorname{Re}\biggl(\lambda|\xi|^2 +\frac{\varphi(t)}{2\lambda}(\xi+\eta)\overline{\xi}\biggr) \\ &\quad= -a^2\sigma(t)|\varphi(t)||\xi|^2 -\operatorname{Re}\biggl(\frac{\varphi(t)}{2\lambda}(\xi+\eta)\overline{\eta}\biggr) -a^2\sigma(t)^2\operatorname{Re} \biggl(\frac{\varphi(t)}{2\lambda}(\xi+\eta)\overline{\xi}\biggr) \\ &\quad\leqslant a^2\sigma(t)|\xi|^2\biggl(\frac1{2a|\lambda|} \bigl(1+a\sigma(t)\bigr)^2-1\biggr)|\varphi(t)|\leqslant 0, \end{aligned} \end{equation*} \notag $$
where $a\geqslant 2/|\lambda|$ and $\sigma(t)<1/a$ for $t\geqslant T$. This proves Statement 1.

Statement 2. If $Y(t)=\left(\begin{smallmatrix}\xi(t)\\\eta(t)\end{smallmatrix}\right)$ is a solution to system (3), then on the null-level of the function $v^{(-)}(t,Y)$ one has $dv^{(-)}(t,Y(t))/dt\geqslant 0$ for $t\geqslant T$ such that $\tau(t)< 1/a$.

Proof. Proceeding similarly and using the relation $\tau'=2\alpha\tau-|\varphi(t)|$, we estimate the derivative
$$ \begin{equation*} \frac12\frac{d}{dt}\, v^{(-)}(t,Y(t))=\operatorname{Re} (\xi'\overline{\xi}) -a^2\tau(t)\tau'(t)|\eta|^2 -a^2\tau(t)^2\operatorname{Re}(\eta'\overline{\eta}) \end{equation*} \notag $$
on the null-level of the function $v^{(-)}(t,Y)$. As a result, under the condition $|\xi|=a\tau(t)|\eta|$, we have
$$ \begin{equation*} \begin{aligned} \, &\frac12\,\frac{d}{dt}\, v^{(-)}(t,Y(t)) =\operatorname{Re} \biggl(\lambda|\xi|^2 +\frac{\varphi(t)}{2\lambda}(\xi+\eta)\overline{\xi}\biggr) \\ &\quad\qquad-a^2\tau(t)\bigl(2\alpha\tau(t)-|\varphi(t)|\bigr)|\eta|^2 + a^2\tau(t)^2\operatorname{Re} \biggl(\lambda|\eta|^2 +\frac{\varphi(t)}{2\lambda}(\xi+\eta)\overline{\eta}\biggr) \\ &\quad= a^2\tau(t)|\varphi(t)||\eta|^2 +\operatorname{Re} \biggl(\frac{\varphi(t)}{2\lambda}(\xi+\eta)\overline{\xi}\biggr) + a^2\tau(t)^2\operatorname{Re} \biggl(\frac{\varphi(t)}{2\lambda}(\xi+\eta)\overline{\eta}\biggr) \\ &\quad\geqslant a^2\tau(t)|\eta|^2 \biggl(1-\frac1{2a|\lambda|}\bigl(1+a\tau(t)\bigr)^2\biggr)|\varphi(t)|\geqslant 0, \end{aligned} \end{equation*} \notag $$
where $a\geqslant 2/|\lambda|$ and $\tau(t)<1/a$ for $t\geqslant T$. This proves Statement 2.

Below, in application to system (3), we give an appropriate version of a topological principle required for our purposes. This principle involves the notions of a retraction, that is, a continuous projection (an idempotent) in a topological space, and a retract — the range of such projection. A continuous vector field associated with system (3) defines the flow and the corresponding fibering $\Psi\colon X\mapsto\Psi(X)$ of the extended phase space, where $X=(t,Y)$, and the fibre $\Psi(X)$ is the integral trajectory passing through $X$. Consider the subsets of the extended phase space

$$ \begin{equation*} \Omega=\{(t,Y)\colon v^{(-)}(t,Y)\leqslant 0,\, t\geqslant T\},\qquad \Gamma=\{(t,Y)\colon |\xi|=a\tau(t)|\eta|,\, t\geqslant T\}, \end{equation*} \notag $$
where $a=2/|\lambda|$ and $\tau(t)<1/a$ as $t\geqslant T$. By Proposition 2, for $t\geqslant T$ we have $dv^{(-)}(t,Y(t))/dt\geqslant 0$, and hence the field corresponding to (3) on $\Gamma$ is directed outside $\Omega$. Below, in § 4, we will define a set $\Sigma\subset\Omega|_{t=T}$ such that $\Sigma\cap\Gamma$ is a retract of $\Psi(\Sigma)\cap\Gamma$, and is not a retract of $\Sigma$ itself. Due to this fact, there exist initial data $\widehat{X}\in \Sigma$ such that $\Psi(\widehat{X})\subset\Omega$.

While applying the construction outlined above (see § 4) we will follow the general scheme developed in [5], which in fact goes back to [12]. However, for our purposes, a certain modification of this scheme is required to be able to achieve the interpolation estimates for the remainder terms $\varepsilon_{1,2}(t)$ as in Theorem 1.

§ 3. Auxiliary estimates

We set

$$ \begin{equation*} \delta(t)=\sup_{s\geqslant t}\frac1{1+s-t}\int_t^s|\varphi(r)|\,dr. \end{equation*} \notag $$
If $\varphi\in\mathrm{L}_p(\mathbb R_+)$, $p\geqslant 1$, then by Hölder’s inequality,
$$ \begin{equation*} \frac1{1+s-t}\int_t^s|\varphi(r)|\,dr \leqslant \frac{(p-1)^{1/q}}{p}\biggl( \int_t^{\infty}|\varphi(r)|^p\, dr\biggr)^{1/p}, \end{equation*} \notag $$
where $q=p/(p-1)$, and therefore, $\delta(t)\to 0$ as $t\to\infty$.

Lemma 1. The functions $\tau(t)$ and $\delta(t)$ satisfy the two-sided estimate

$$ \begin{equation*} \frac{2\alpha}{1+2\alpha}\tau(t)\leqslant \delta(t)\leqslant (1+2\alpha)\sup_{r\geqslant t}\tau(r). \end{equation*} \notag $$

Indeed, on the one hand, integrating by parts, we get

$$ \begin{equation*} \begin{aligned} \, \tau(t) &= 2\alpha\int_t^{\infty} \biggl(\int_t^s|\varphi(r)|\, dr\biggr)e^{2\alpha(t-s)}\,ds \\ &\leqslant 2\alpha\delta(t)\int_t^{\infty}(1+s-t)e^{2\alpha(t-s)}\,ds =\biggl(1+\frac1{2\alpha}\biggr)\delta(t) \end{aligned} \end{equation*} \notag $$
and, on the other hand, since $\tau'(r)=2\alpha\tau(r)-|\varphi(r)|$, we have, for arbitrary $s\geqslant t$,
$$ \begin{equation*} \frac1{1+s-t}\int_t^s|\varphi(r)|\, dr =\frac{2\alpha}{1+s-t}\int_t^s\tau(r)\, dr + \frac{\tau(t)-\tau(s)}{1+s-t}\leqslant 2\alpha\sup_{r\geqslant t}\tau(r)+\tau(t). \end{equation*} \notag $$

Lemma 2. $\delta(t)\leqslant (1+2\alpha)\sup_{r\geqslant t}\sigma(r)$. In addition, if $t\geqslant A$, then

$$ \begin{equation*} \sigma(t)\leqslant e^{2\alpha(A-t)}\int_0^t|\varphi(s)|\, ds + \biggl(1+\frac1{2\alpha}\biggr)\sup_{s\geqslant A}\delta(s). \end{equation*} \notag $$

Indeed, since $\sigma'(r)=|\varphi(r)|-2\alpha\sigma(r)$, it follows that

$$ \begin{equation*} \frac1{1+s-t}\int_t^s|\varphi(r)|\, dr=\frac{2\alpha}{1+s-t}\int_t^s\sigma(r)\, dr+ \frac{\sigma(s)-\sigma(t)}{1+s-t}\leqslant 2\alpha\sup_{r\geqslant t}\sigma(r)+\sigma(s) \end{equation*} \notag $$
for arbitrary $s\geqslant t$. On the other hand, integrating by parts for $t\geqslant A$, we obtain
$$ \begin{equation*} \begin{aligned} \, \sigma(t) &= \int_0^Ae^{2\alpha(s-t)}|\varphi(s)|\, ds +e^{2\alpha(A-t)}\int_A^t|\varphi(r)|\, dr \\ &\qquad+ 2\alpha\int_A^t\biggl(\int_s^t|\varphi(r)|\, dr\biggr)e^{2\alpha(s-t)}\, ds \\ &\leqslant e^{2\alpha(A-t)}\int_0^t|\varphi(r)|\, dr + 2\alpha\sup_{s\geqslant A} \delta(s)\int_A^t(1+t-s)e^{2\alpha(s-t)}\, ds \\ &\leqslant e^{2\alpha(A-t)}\int_0^t|\varphi(s)|\, ds + \biggl(1+\frac1{2\alpha}\biggr)\sup_{s\geqslant A} \delta(s). \end{aligned} \end{equation*} \notag $$

Lemma 3. Assume that $\varphi\in\mathrm{L}_p(\mathbb R_+),p\in[1,2]$. Then $\tau\in\mathrm{L}_p(\mathbb R_+)$, $\tau\varphi\in\mathrm{L}_1(\mathbb R_+)$, $\tau(t)\to 0$ as $t\to\infty$, and for sufficiently large $t$,

$$ \begin{equation*} \int_t^{\infty}|\varphi(s)|\tau(s)\, ds \leqslant (2\alpha)^{1-p}\int_t^{\infty}|\varphi(s)|^p\, ds. \end{equation*} \notag $$

Proof. Multiplying the equality $2\alpha\tau(s)=|\varphi(s)|+\tau'(s)$ by $\tau(s)^{p-1}$ and integrating over $[t,A]$, we have
$$ \begin{equation*} 2\alpha\int_t^A\tau(s)^p\, ds =\int_t^A|\varphi(s)|\tau(s)^{p-1}\, ds+ \frac{\tau(A)^p-\tau(t)^p}{p}. \end{equation*} \notag $$
Since $\varphi\in\mathrm{L}_p(\mathbb R_+)$, it follows that $\tau(t)\to 0$ as $t\to\infty$ by Lemma 1. As a result,
$$ \begin{equation*} 2\alpha\int_t^A\tau(s)^p\, ds \leqslant \int_t^A|\varphi(s)|\tau(s)^{p-1}\, ds \leqslant \biggl(\int_t^A|\varphi(s)|^pds\biggr)^{1/p} \biggl(\int_t^A\tau(s)^p\, ds\biggr)^{1/q} \end{equation*} \notag $$
for sufficiently large $A$, where $q=p/(p-1)$. Consequently, $\tau\in\mathrm{L}_p(\mathbb R_+)$, and
$$ \begin{equation*} \int_t^{\infty}\tau(s)^p\,ds \leqslant (2\alpha)^{-p}\int_t^{\infty}|\varphi(s)|^p\, ds. \end{equation*} \notag $$
Note that in this setting $q\geqslant p$, and, in addition, $\tau(t)\to 0$ as $t\to\infty$. Hence $\tau\in\mathrm{L}_q(\mathbb R_+)$. In view of the condition $\varphi\in\mathrm{L}_p(\mathbb R_+)$, this implies integrability of the function $\tau\varphi$ on the half-axis $\mathbb R_+$. Therefore, by Hölder’s inequality, we have, for sufficiently large $t$,
$$ \begin{equation*} \begin{aligned} \, &\int_t^{\infty}|\varphi(s)|\tau(s)\, ds \leqslant \biggl(\int_t^{\infty}|\varphi(s)|^p\, ds\biggr)^{1/p} \biggl(\int_t^{\infty}\tau(s)^q\, ds\biggr)^{1/q} \\ &\qquad\leqslant \biggl(\int_t^{\infty}|\varphi(s)|^p\, ds\biggr)^{1/p} \biggl(\int_t^{\infty}\tau(s)^p\, ds\biggr)^{1/q} \leqslant (2\alpha)^{-p/q}\int_t^{\infty}|\varphi(s)|^p\, ds. \end{aligned} \end{equation*} \notag $$
This proves Lemma 3.

Lemma 4. Let $\varphi\in\mathrm{L}_p(\mathbb R_+)$, $p\in[1,2]$. Then $\sigma\in\mathrm{L}_p(\mathbb R_+)$, $\sigma\varphi\in\mathrm{L}_1(\mathbb R_+)$, $\sigma(t)\to 0$ as $t\to\infty$, and

$$ \begin{equation*} \int_t^{\infty}|\varphi(s)|\sigma(s)\, ds \leqslant \bigl((2\alpha)^{1-p}+1\bigr) \int_t^{\infty}|\varphi(s)|^p\, ds+ \frac{\sigma(t)^{p}}{4\alpha} \end{equation*} \notag $$
for sufficiently large $t$.

Proof. Multiplying the equality $2\alpha\sigma(s)=|\varphi(s)|-\sigma'(s)$ by $\sigma(s)^{p-1}$ and integrating over $[0,t]$, we have
$$ \begin{equation*} \begin{aligned} \, 2\alpha\int_0^t\sigma(s)^p\, ds &=\int_0^t|\varphi(s)|\sigma(s)^{p-1}\, ds- \frac{\sigma(t)^p}{p} \\ &\leqslant \biggl(\int_0^t|\varphi(s)|^p\, ds\biggr)^{1/p} \biggl(\int_0^t\sigma(s)^p\, ds\biggr)^{1/q}, \end{aligned} \end{equation*} \notag $$
where $q=p/(p-1)\geqslant 2$. Consequently,
$$ \begin{equation*} \int_0^t\sigma(s)^p\, ds\leqslant (2\alpha)^{-p}\int_0^t|\varphi(s)|^p\, ds. \end{equation*} \notag $$
Therefore, $\sigma\in\mathrm{L}_p(\mathbb R_+)$. By Lemma 2, in our setting $\sigma(t)\to 0$ as $t\to\infty$, and hence, $\sigma\in\mathrm{L}_q(\mathbb R_+)$. Since $\varphi\in\mathrm{L}_p(\mathbb R_+)$, the function $\sigma\varphi$ is integrable on the half-axis $\mathbb R_+$. Next, multiplying the equation $2\alpha\sigma(s)=|\varphi(s)|-\sigma'(s)$ by $\sigma(s)^{p-1}$ and integrating over $[t,\infty)$, we find that
$$ \begin{equation*} \begin{aligned} \, 2\alpha\int_t^{\infty}\sigma(s)^p\, ds &=\int_t^{\infty}|\varphi(s)|\sigma(s)^{p-1}\, ds+ \frac{\sigma(t)^p}{p} \\ &\leqslant \biggl(\int_t^{\infty}|\varphi(s)|^p\, ds\biggr)^{1/p} \biggl(\int_t^{\infty}\sigma(s)^p\, ds\biggr)^{1/q}+ \sigma(t)^p. \end{aligned} \end{equation*} \notag $$
Therefore,
$$ \begin{equation*} \biggl(\int_t^{\infty}\sigma(s)^p\, ds\biggr)^{1/q} \leqslant (2\alpha)^{-p/q} \biggl(\int_t^{\infty}|\varphi(s)|^p\, ds\biggr)^{1/q} +(2\alpha)^{-1/q}\sigma(t)^{p/q}. \end{equation*} \notag $$
Here, we used the fact that if $a,b,c\geqslant 0$ and $a^q\leqslant ab+c$, where $q\geqslant 2$, then $a\leqslant b^{1/(q-1)}+c^{1/q}$. Finally, applying Hölder’s and Young’s inequalities, we have, for sufficiently large $t$,
$$ \begin{equation*} \begin{aligned} \, \int_t^{\infty}|\varphi(s)|\sigma(s)\, ds &\leqslant \biggl(\int_t^{\infty}|\varphi(s)|^p\, ds\biggr)^{1/p} \biggl(\int_t^{\infty}\sigma(s)^q\, ds\biggr)^{1/q} \\ &\leqslant (2\alpha)^{-p/q}\int_t^{\infty}|\varphi(s)|^p\, ds + (2\alpha)^{-1/q}\sigma(t)^{p/q} \biggl(\int_t^{\infty}|\varphi(s)|^p\, ds\biggr)^{1/p} \\ &\leqslant \biggl((2\alpha)^{-p/q}+\frac1{p}\biggr) \int_t^{\infty}|\varphi(s)|^p\, ds+ \frac{\sigma(t)^{p}}{2\alpha q}. \end{aligned} \end{equation*} \notag $$
This proves Lemma 4.

§ 4. Asymptotic integration

Statement 3. Let $\sigma(t)<|\lambda|/2$ if $t\geqslant T$ and let $Y(t)=\left(\begin{smallmatrix}\xi(t)\\\eta(t)\end{smallmatrix}\right)$ be any solution to system (3) such that $|\eta(T)|\leqslant (2/|\lambda|)\sigma(T)|\xi(T)|$. Then, for all $t\geqslant T$,

$$ \begin{equation*} |\eta(t)|\leqslant\frac2{|\lambda|}\sigma(t)|\xi(t)|. \end{equation*} \notag $$

In the definition of the Lyapunov type function $v^{(+)}(t,Y)$, we set $a=2/|\lambda|$, and note that, for an arbitrary trajectory of system (3) with initial data satisfying the condition $v^{(+)}(T,Y(T))\leqslant 0$, from Statement 1 we have, for all $t\geqslant T$,

$$ \begin{equation*} v^{(+)}(t,Y(t))\leqslant 0. \end{equation*} \notag $$

Statement 4. Let $\tau(t)<|\lambda|/2$ if $t\geqslant T$. Then system (3) has a solution $\widehat{Y}(t)=\left(\begin{smallmatrix}\widehat{\xi}(t) \\ \widehat{\eta}(t)\end{smallmatrix} \right)$ such that $\widehat{\eta}(t)\ne 0$ and, for all $t\geqslant T$,

$$ \begin{equation*} |\widehat{\xi}(t)|\leqslant \frac2{|\lambda|}\tau(t)|\widehat{\eta}(t)|. \end{equation*} \notag $$

Proof. Proceeding as in § 2, we define the set
$$ \begin{equation*} \Sigma=\biggl\{\biggl(T,\xi,\frac{R}{a\tau(T)}\biggr),|\xi|\leqslant R\biggr\} \subset\Omega|_{t=T}, \end{equation*} \notag $$
where $ R > 0$. The set $\Psi(\Sigma)\cap\,\Gamma$ consists of segments of the integral curves $\Psi(X)\cap \,\Gamma$, where $X\in\Sigma$ so that the corresponding mapping $\gamma\colon\Sigma\to\Psi(\Sigma)\cap\Gamma$ is continuous, and the transform
$$ \begin{equation*} \pi\colon (t,\xi,\eta) \longmapsto\biggl(T,\frac{\xi R}{a\tau(t)\overline{\eta}},\frac{R}{a\tau(T)}\biggr) \end{equation*} \notag $$
is single-valued on $\Psi(X)\cap\Gamma$. Indeed, if a segment of the integral curve $(t,Y(t))\subset\Psi(X)$ lies in $\Gamma$, then, on this segment,
$$ \begin{equation*} \frac{d}{dt}\, v^{(-)}(t,Y(t)) = 2a^2\tau(t)|\eta|^2 \biggl(1-\frac1{2a|\lambda|}(1+a\tau(t))^2\biggr)|\varphi(t)|=0, \end{equation*} \notag $$
so that $\varphi(t)\equiv 0$. Hence, $\arg (\xi/\overline{\eta})$ is constant on this segment. Therefore, we have $\pi(\Psi(X)\cap\Gamma)=\mathrm{const}$, as far as $|\xi|=a\tau(t)|\eta|$ on $\Gamma$.

Now let us show that there exist initial data $\widehat{X}\in \Sigma$ such that $\Psi(\widehat{X})\subset\Omega$, and so, $v^{(-)}(t,\widehat{Y}(t))\leqslant 0$ for $t\geqslant T$. Assume that $\Psi(X)\cap\Gamma\ne\varnothing$ for arbitrary $X\in\Sigma$. Then the mapping

$$ \begin{equation*} \pi\circ\gamma\colon \Sigma\to\Sigma\cap\Gamma \end{equation*} \notag $$
is continuous on $\Sigma$ and is identical on $\partial\Sigma$. Hence the composition $\pi\circ\gamma$ is a retraction. However, $\Sigma\cap\Gamma=\partial\Sigma$ is a sphere, which by no means can be a retract of the ball $\Sigma$ according to the Brouwer’s fixed-point theorem (see, for example, [13]). Next, $\Psi(\Sigma)$ does not contain the stationary point of system (3), and hence $\widehat{\eta}(t)\ne 0$ for $t\geqslant T$. This proves Statement 4.

Proof of Theorem 1. Let $\tau(t)\to 0$ as $t\to\infty$ and let $\tau\varphi\in\mathrm{L}_1(\mathbb R_+)$. Integrating the equality
$$ \begin{equation*} \widehat{\eta}'(t)=-\lambda\widehat{\eta}(t) -\frac{\varphi(t)}{2\lambda} \bigl(\widehat{\xi}(t)+\widehat{\eta}(t)\bigr), \end{equation*} \notag $$
and normalizing, we have
$$ \begin{equation*} \begin{aligned} \, \widehat{\eta}(t) &=\exp\biggl(-\int_0^t\biggl(\lambda +\frac{\varphi(s)}{2\lambda}\biggr)\, ds\biggr) \exp\biggl(\frac1{2\lambda}\int_t^{\infty}\varphi(s) \frac{\widehat{\xi}(s)}{\widehat{\eta}(s)}\, ds\biggr) \\ &=\exp\biggl(-\lambda t-\frac1{2\lambda} \int_0^t\varphi(s)\, ds\biggr)\biggl\{1 + O\biggl(\int_t^{\infty}|\varphi(s)|\tau(s)\, ds\biggr)\biggr\} \end{aligned} \end{equation*} \notag $$
for the solution $\widehat{Y}(t)=\left(\begin{smallmatrix}\widehat{\xi}(t) \\ \widehat{\eta}(t) \end{smallmatrix} \right)$ of system (3), which was constructed in Statement 4. Here, we used the estimate $\widehat{\xi}(t)=O(\tau(t)\widehat{\eta}(t))$. The solution $u_1(t)=\widehat{\xi}(t)+\widehat{\eta}(t)$ to the original equation (2) obtained in this way behaves asymptotically as
$$ \begin{equation*} \begin{aligned} \, u_1(t) &=\widehat{\eta}(t)\bigl(1+O(\tau(t))\bigr) \\ &=\exp\biggl(-\lambda t-\frac1{2\lambda} \int_0^t\varphi(s)\, ds\biggr)\biggl\{1 + O\biggl(\tau(t)+\int_t^{\infty}|\varphi(s)|\tau(s)\, ds\biggr)\biggr\}. \end{aligned} \end{equation*} \notag $$

Now let $\sigma(t)\to 0$ as $t\to\infty$ and $\sigma\varphi\in\mathrm{L}_1(\mathbb R_+)$. Consider the solution $\widetilde{Y}(t)=\left(\begin{smallmatrix}\widetilde{\xi}(t) \\ \widetilde{\eta}(t) \end{smallmatrix}\right)$ to system (3) specified in Proposition 3 such that $\widetilde{\xi}(T)\ne 0$. Hence $\widetilde{\xi}(t)\ne 0$ for all $t\geqslant T$. Proceeding as above and integrating the equality

$$ \begin{equation*} \widetilde{\xi}'(t)=\lambda\widetilde{\xi}(t) + \frac{\varphi(t)}{2\lambda}\bigl(\widetilde{\xi}(t)+\widetilde{\eta}(t)\bigr), \end{equation*} \notag $$
we have, after a suitable normalization,
$$ \begin{equation*} \begin{aligned} \, \widetilde{\xi}(t) &=\exp\biggl(\int_0^t\biggl(\lambda +\frac{\varphi(s)}{2\lambda}\biggr)\, ds\biggr) \exp\biggl(-\frac1{2\lambda} \int_t^{\infty}\varphi(s) \frac{\widetilde{\eta}(s)}{\widetilde{\xi}(s)}\, ds\biggr) \\ &=\exp\biggl(\lambda t+\frac1{2\lambda}\int_0^t\varphi(s)\, ds\biggr) \biggl\{1+ O\biggl(\int_t^{\infty}|\varphi(s)|\sigma(s)\, ds\biggr)\biggr\}, \end{aligned} \end{equation*} \notag $$
where $\widetilde{\eta}(t)=O(\sigma(t)\widetilde{\xi}(t))$. Thus we have constructed the second desired solution $u_2(t)=\widetilde{\xi}(t)+\widetilde{\eta}(t)$ to equation (2) with the asymptotics
$$ \begin{equation*} \begin{aligned} \, u_2(t) &=\widetilde{\xi}(t)\bigl(1+O(\sigma(t))\bigr) \\ &=\exp\biggl(\lambda t+\frac1{2\lambda}\int_0^t\varphi(s)\, ds\biggr) \biggl\{1+ O\biggl(\sigma(t)+\int_t^{\infty}|\varphi(s)|\sigma(s)\, ds\biggr)\biggr\}. \end{aligned} \end{equation*} \notag $$

If $\varphi\in\mathrm{L}_p(\mathbb R_+)$, then $\tau(t)\to0$ and $\sigma(t)\to0$ as $t\to\infty$ by Lemmas 1 and 2. In turn, from Lemmas 3 and 4 we have the conditions $\tau\varphi\in\mathrm{L}_1(\mathbb R_+)$ and $\sigma\varphi\in\mathrm{L}_1(\mathbb R_+)$ with sharp estimates for the corresponding integrals. Consequently, under the assumption $\varphi\in\mathrm{L}_p(\mathbb R_+),p\in[1,2]$, equation (2) has the fundamental system of solutions

$$ \begin{equation*} u_{1,2}(t)=\exp\biggl(\mp\lambda t\mp\frac1{2\lambda} \int_0^t\varphi(s)\, ds\biggr) \bigl(1+\varepsilon_{1,2}(t)\bigr), \end{equation*} \notag $$
where
$$ \begin{equation*} \varepsilon_1(t)= O\biggl(\tau(t) +\int_t^{\infty}|\varphi(s)|^p\, ds\biggr),\qquad \varepsilon_2(t)= O\biggl(\sigma(t)+\int_t^{\infty}|\varphi(s)|^p\, ds\biggr). \end{equation*} \notag $$
This proves Theorem 1.

The set of functions $\varphi(t)$ satisfying the conditions

$$ \begin{equation*} \begin{alignedat}{5} \tau\varphi &\in \mathrm{L}_1(\mathbb R_+), &\qquad \tau(t)&\to 0, &\quad t&\to\infty, \\ \sigma\varphi &\in \mathrm{L}_1(\mathbb R_+), &\qquad \sigma(t) &\to 0, &\quad t&\to\infty, \end{alignedat} \end{equation*} \notag $$
under which equation (2) has solutions with the asymptotics
$$ \begin{equation} u_{1,2}(t)\sim\exp\biggl(\mp\lambda t\mp\frac1{2\lambda}\int_0^t\varphi(s)\, ds\biggr), \end{equation} \tag{6} $$
is somewhat wider than the class $\mathrm{L}_p(\mathbb R_+)$, $p\in[1,2]$. Let us construct a function $\varphi\notin\mathrm{L}_p(\mathbb R_+)$ such that $\tau(t)\to 0$ as $t\to\infty$ and $\tau\varphi\in\mathrm{L}_1(\mathbb R_+)$. To this end, we set $\lambda=1/2$, $a_n=n^{3/2}$, $h_n=(\ln n)^{-1}, \varepsilon_n=n^{-4/5}$, where $n=2,3,\dots$, and define
$$ \begin{equation*} \varphi(t)= \begin{cases} h_n, &t\in(a_n-\varepsilon_n,a_n), \\ 0, &t\notin(a_n-\varepsilon_n,a_n). \end{cases} \end{equation*} \notag $$
Then $\lim_{t\to\infty}\tau(t)=0$ since $\varphi(t)\to 0$ as $t\to\infty$, and, at the same time,
$$ \begin{equation*} \int_0^{\infty}\varphi(t)^p\, dt=\sum_{n=2}^{\infty}\frac1{n^{4/5}(\ln n)^p}=\infty \end{equation*} \notag $$
for arbitrary $p\geqslant 1$. Now let us show that the integral
$$ \begin{equation*} \begin{aligned} \, \int_1^{\infty}\tau(t)^2\, dt &= \sum_{n=2}^{\infty}\int_{a_{n-1}}^{a_n}e^{2t} \biggl(\int_t^{\infty}e^{-s}\varphi(s)\, ds\biggr)^2\, dt \\ &\leqslant \sum_{n=2}^{\infty}e^{2a_n} \biggl(\int_{a_n-\varepsilon_n}^{\infty}e^{-s}\varphi(s)\, ds\biggr)^2 (a_n-a_{n-1}) \end{aligned} \end{equation*} \notag $$
converges, where $(a_n-a_{n-1})\sim3\sqrt{n}/2$. Indeed, since $\sum_{k=n}^{\infty}e^{-a_k}\sim e^{-a_n}$, we have
$$ \begin{equation*} \int_{a_n-\varepsilon_n}^{\infty}e^{-s}\varphi(s)\, ds \leqslant \sum_{k=n}^{\infty}e^{-a_k+\varepsilon_k}h_k\varepsilon_k \leqslant h_n\varepsilon_ne^{\varepsilon_n}\sum_{k=n}^{\infty}e^{-a_k} =O\bigl(h_n\varepsilon_ne^{-a_n}\bigr). \end{equation*} \notag $$
Hence
$$ \begin{equation*} e^{2a_n}\biggl(\int_{a_n-\varepsilon_n}^{\infty} e^{-s}\varphi(s)\, ds\biggr)^2 (a_n-a_{n-1}) =O\bigl(n^{-11/10}(\ln n)^{-2}\bigr) \end{equation*} \notag $$
and, therefore, $\int^{\infty}\tau(t)^2\, dt<\infty$. Now an integration by parts
$$ \begin{equation*} \int_0^{\infty}\varphi(t)\tau(t)\,dt =\frac12\biggl(\int_0^{\infty}e^{-t}\varphi(t)\, dt\biggr)^2 + \int_0^{\infty}\tau(t)^2\, dt, \end{equation*} \notag $$
where $\tau(\infty)=0$, shows that $\tau\varphi\in\mathrm{L}_1(\mathbb R_+)$ in our setting.

§ 5. Bellman asymptotic solutions

Under the condition $\varphi\in\mathrm{L}_3(\mathbb R_+)$, the method based on the transformation of (2) to the Riccati equation can be applied (see [4]) to construct solutions with the asymptotics

$$ \begin{equation*} u_{1,2}(t)\sim\exp\biggl(\mp\lambda t\mp\frac1{2\lambda} \int_0^t\varphi(s)\, ds\pm \frac1{4\lambda^2} \int_0^t\varphi(s)\int_0^se^{2\lambda(r-s)}\varphi(r)\, dr\, ds\biggr). \end{equation*} \notag $$

Lemma 5. If $\varphi(t)=t^{-\kappa}\sin t$, where $\kappa\in(0,1)$, then the limit

$$ \begin{equation*} \lim_{t\to\infty}\biggl\{\int_0^t\varphi(s) e^{-2s}\biggl(\int_0^se^{2r}\varphi(r)\, dr\biggr)\, ds - \frac25\int_0^t\varphi(s)^2\, ds\biggr\} \end{equation*} \notag $$
exists and is finite.

Proof. Integrating by parts two times, we have, after an appropriate regularization,
$$ \begin{equation*} \begin{aligned} \, I(s) &:=\int_0^se^{2r}\varphi(r)\, dr=\frac12\,\varphi(s)e^{2s} + \frac{\kappa}2\int_0^s\frac{\sin r}{r^{\kappa+1}}\,e^{2r}\, dr \\ &\,\qquad-\frac14\lim_{\varepsilon\to 0} \biggl\{e^{2r}\frac{\cos r}{r^{\kappa}} \bigg|_{\varepsilon}^s + \int_{\varepsilon}^s e^{2r}\biggl(\frac{\sin r}{r^{\kappa}} +\kappa\frac{\cos r}{r^{\kappa+1}}\biggr)\, dr\biggr\} \\ &\,=\frac12\,\varphi(s)e^{2s}{+}\, \frac{\kappa}2\int_0^s\frac{\sin r}{r^{\kappa+1}}\, e^{2r}\, dr -\frac{I(s)}4+\frac{1\,{-}\,e^{2s}\cos s}{4s^{\kappa}} +\frac{\kappa}4\int_0^s \! \frac{1\,{-}\,e^{2r}\cos r}{r^{\kappa+1}}\, dr. \end{aligned} \end{equation*} \notag $$
Hence
$$ \begin{equation*} 5I(s)=2\varphi(s)e^{2s}+A(s)+B(s)+C(s), \end{equation*} \notag $$
where
$$ \begin{equation*} A(s)=\frac{1-e^{2s}\cos s}{s^{\kappa}},\quad B(s) =2\kappa\int_0^s\frac{\sin r}{r^{\kappa+1}}\, e^{2r}\, dr,\quad C(s)=\kappa\int_0^s\frac{1-e^{2r}\cos r}{r^{\kappa+1}}\, dr. \end{equation*} \notag $$
By the Dirichlet’s test, the integral $\int^{\infty}\varphi(s)e^{-2s}A(s)\, ds$ converges conditionally, and $\varphi(s)e^{-2s}B(s)$ and $\varphi(s)e^{-2s}C(s)$ possess absolutely integrable majorants, since
$$ \begin{equation*} \int_1^s\frac{e^{2r}}{r^{\kappa+1}}\, dr \sim \frac{e^{2s}}{2s^{\kappa+1}},\qquad s\to\infty. \end{equation*} \notag $$
Therefore,
$$ \begin{equation*} \int_0^t\varphi(s)e^{-2s}I(s)\, ds =\frac25\int_0^t\varphi(s)^2\, ds+\mathrm{const} +o(1),\qquad t\to\infty. \end{equation*} \notag $$
This proves Lemma 5.

Corollary 1. If $\varphi(t)=t^{-\kappa}\sin t$, where $\kappa>1/3$, then equation (2) for $\lambda=1$ has a fundamental system of solutions which behave asymptotically as

$$ \begin{equation*} u_{1,2}(t)\sim\exp\biggl(\mp t\pm \frac1{10}\int_0^t\varphi(s)^2\, ds\biggr),\qquad t\to\infty; \end{equation*} \notag $$
this asymptotic formula is inconsistent with (6) for $\kappa\leqslant 1/2$.

On the other hand, the following result holds (see [14]).

Statement 5. Let $\varphi(t)=\varphi_1(t)+\varphi_2(t)$, where $\varphi_1'(t)\in\mathrm{L}_1(\mathbb R_+)$ and $\varphi_2'(t)=O(t^{-\alpha})$, $\alpha>1/2$. Then the limit

$$ \begin{equation*} \lim_{t\to\infty}\biggl\{\int_0^t\varphi(s)\int_0^se^{2(r-s)}\varphi(r)\, dr\, ds - \frac12\int_0^t\varphi(s)^2\, ds\biggr\} \end{equation*} \notag $$
exists and is finite, provided that $\varphi(t)\to 0$ as $t\to\infty$.

§ 6. The asymptotic reduction method

In parallel with the equation

$$ \begin{equation} (r(t)x')'+f(t)x=0, \end{equation} \tag{7} $$
where the function $f(t)$ is (in general)complex-valued and $r(t)>0$, we consider the comparison equation
$$ \begin{equation} (r(t)y')'+g(t)y=0 \end{equation} \tag{8} $$
with a real-valued function $g(t)$. Equation (8) is regarded as an (unperturbed) model with respect to (7). It is assumed that (8), being non-oscillatory, possesses a fundamental system of solutions $\{y_1(t),y_2(t)\}$ which are positive for sufficiently large $t$ and such that
$$ \begin{equation*} y_1(t)y_2'(t)-y_1'(t)y_2(t)=\frac1{r(t)},\qquad \rho(t):=\frac{y_2(t)}{y_1(t)}\nearrow\infty, \quad t\to\infty. \end{equation*} \notag $$
The solution $y_1(t)$ is called principal (or subdominant); this solution is determined uniquely up to a constant factor (see [5]).

Let us write equation (7) as the system

$$ \begin{equation} X'=\begin{pmatrix} 0 &\dfrac1{r} \\ -f &0 \end{pmatrix}X, \end{equation} \tag{9} $$
where $X=\left(\begin{smallmatrix} x\\ rx'\end{smallmatrix}\right)$, which can be reduced to the form
$$ \begin{equation*} Y'=(g-f)\begin{pmatrix} -y_1y_2&-y_2^2\\ y_1^2&y_1y_2\end{pmatrix}Y \end{equation*} \notag $$
by the substitution $X=\left(\begin{smallmatrix}y_1&y_2\\ ry_1'&ry_2'\end{smallmatrix}\right)Y$. The transformation $Y=\left(\begin{smallmatrix}1&0\\ 0&1/\rho\end{smallmatrix}\right)\widetilde{Y}$ enables one to balance the off-diagonal matrix entries of the system
$$ \begin{equation*} \widetilde{Y}'=\biggl\{\frac{\rho'}{\rho}\begin{pmatrix} 0&0\\ 0&1\end{pmatrix} +(g-f)y_1y_2\begin{pmatrix} -1&-1\\ 1&1\end{pmatrix}\biggr\} \widetilde{Y} =\{\Lambda(t)+\widetilde{V}(t)\}\widetilde{Y}. \end{equation*} \notag $$
The goal of the current procedure (cf. [6]) is to reduce system (9) to the $L$-diagonal form. If $(g-f)y_1y_2\notin\mathrm{L}_1(\mathbb R_+)$, then one can try to achieve integrability of the perturbation $\widetilde{V}(t)$ by making an appropriate invertible transformation $\widetilde{Y}=(I+ \widetilde{Q}(t))\widehat{Y}$ so that
$$ \begin{equation*} \widehat{Y}'=\bigl\{\Lambda+(I+\widetilde{Q})^{-1}\bigl([\Lambda,\widetilde{Q}] +\widetilde{V}(I+\widetilde{Q})-\widetilde{Q}'\bigr)\bigr\}\widehat{Y}, \end{equation*} \notag $$
where $[\Lambda,\widetilde{Q}]=\Lambda \widetilde{Q}-\widetilde{Q}\Lambda$. If $\widetilde{Q}(t)=q(t)\left(\begin{smallmatrix} -1&-1\\ 1&1\end{smallmatrix}\right)$, then $(I+\widetilde{Q})^{-1}=I-\widetilde{Q}$, and so the reduced system takes the form
$$ \begin{equation*} \widehat{Y}'=\biggl\{\frac{\rho'}{\rho}\begin{pmatrix} 0&0\\ 0&1\end{pmatrix} +q\frac{\rho'}{\rho}\begin{pmatrix} 0&1\\ 1&0\end{pmatrix} +\biggl[(g-f)y_1y_2-q'-q^2\frac{\rho'}{\rho}\biggr] \begin{pmatrix} -1&-1\\ 1&1\end{pmatrix} \biggr\}\widehat{Y}. \end{equation*} \notag $$
As the functional parameter of this transformation we take $q(t)$ from the interpolation (one-parameter) family
$$ \begin{equation*} q(t)=\rho(t)^{\kappa}\int_t^{\infty}(f-g)y_1y_2\rho^{-\kappa}\, ds,\qquad 0\leqslant \kappa\leqslant 1, \end{equation*} \notag $$
so that $q'=(g-f)y_1y_2+\kappa q\rho'/\rho$. As a result, we obtain the system
$$ \begin{equation*} \begin{aligned} \, \widehat{Y}' &=\biggl\{\frac{\rho'}{\rho}\begin{pmatrix} 0&0\\ 0&1\end{pmatrix} +q\frac{\rho'}{\rho}\begin{pmatrix} \kappa&1+\kappa\\ 1-\kappa &-\kappa\end{pmatrix} +q^2\frac{\rho'}{\rho}\begin{pmatrix} 1&1\\ -1&-1\end{pmatrix}\biggr\}\widehat{Y} \\ &=\{\Lambda(t)+\widehat{V}(t)+R(t)\}\widehat{Y}. \end{aligned} \end{equation*} \notag $$

Finally, in order to provide the desired decay of $\widehat{V}(t)$ at infinity, we carry out the additional transformation

$$ \begin{equation*} \widehat{Y}=(I+\widehat{Q}(t))Z=\begin{pmatrix} 1 +\widehat{q}_{11}&\widehat{q}_{12}\\ \widehat{q}_{21}&1+\widehat{q}_{22} \end{pmatrix}Z. \end{equation*} \notag $$
After a corresponding substitution, the system takes the form
$$ \begin{equation*} Z'=\bigl\{\Lambda+(I+\widehat{Q})^{-1}\bigl([\Lambda,\widehat{Q}] + \widehat{V}(I+\widehat{Q})-\widehat{Q}'+R(I+\widehat{Q})\bigr)\bigr\}Z, \end{equation*} \notag $$
where
$$ \begin{equation*} \widehat{V}(I+\widehat{Q})=q\frac{\rho'}{\rho}\begin{pmatrix} \kappa(1+\widehat{q}_{11}) +(1+\kappa)\widehat{q}_{21} &\kappa\widehat{q}_{12}+(1+\kappa)(1+\widehat{q}_{22}) \\ (1-\kappa)(1+\widehat{q}_{11})-\kappa\widehat{q}_{21} &(1-\kappa)\widehat{q}_{12}-\kappa(1+\widehat{q}_{22}) \end{pmatrix} \end{equation*} \notag $$
and $[\Lambda,\widehat{Q}]\,{=}\,(\rho'/\rho) \left(\begin{smallmatrix} 0&-\widehat{q}_{12}\\ \widehat{q}_{21}&0 \end{smallmatrix}\right)$. We next set $L(t) = \int_t^{\infty}q(s)\rho'(s)/\rho(s)\, ds$ and define $\widehat{q}_{11}=e^{-\kappa L}-1$, so that $(\widehat{q}_{11})'=\kappa(1+\widehat{q}_{11})q\rho'/\rho$. As the off-diagonal entries
$$ \begin{equation*} \widehat{q}_{12}=(\kappa+1)\frac1{\rho(t)}\, e^{-\kappa L(t)}\int_0^tq\rho'e^{\kappa L}\, ds, \qquad \widehat{q}_{21}=(\kappa-1)\rho(t)e^{\kappa L(t)} \int_t^{\infty}q\frac{\rho'}{\rho^2}\, e^{-\kappa L}\,ds \end{equation*} \notag $$
we take the solutions to the equations
$$ \begin{equation*} (\widehat{q}_{12})'=(\kappa q-1)\widehat{q}_{12}\frac{\rho'}{\rho} +(1+\kappa)q\frac{\rho'}{\rho},\qquad (\widehat{q}_{21})' =(1-\kappa q)\widehat{q}_{21}\frac{\rho'}{\rho} +(1-\kappa)q\frac{\rho'}{\rho}, \end{equation*} \notag $$
and $\widehat{q}_{22}=e^{\kappa L}+\widehat{q}_{12}\widehat{q}_{21}-1$ is chosen according to the relation
$$ \begin{equation*} (\widehat{q}_{12}\widehat{q}_{21})' =\bigl((1-\kappa)\widehat{q}_{12}+ (1+\kappa)\widehat{q}_{21}\bigr)q\frac{\rho'}{\rho}. \end{equation*} \notag $$
Proceeding in this manner, we have
$$ \begin{equation*} \begin{aligned} \, Z' &=\biggl\{\Lambda+q\frac{\rho'}{\rho}(I+\widehat{Q})^{-1} \begin{pmatrix} (1+\kappa)\widehat{q}_{21} &(1+\kappa)\widehat{q}_{22} \\ (1-\kappa)\widehat{q}_{11} &-(1+\kappa(1+\widehat{q}_{12}))\widehat{q}_{21} \end{pmatrix} \\ &\qquad+(I+\widehat{Q})^{-1}R(I+\widehat{Q})\biggr\}Z= \{\Lambda(t)+V(t)\}Z. \end{aligned} \end{equation*} \notag $$

Lemma 6. If $q(t)\to 0$ as $t\to\infty$ and the integral

$$ \begin{equation*} L(t)=\int_t^{\infty}q(t)\frac{\rho'(t)}{\rho(t)}\, dt \end{equation*} \notag $$
converges, then
$$ \begin{equation*} \widehat{Q}(t)\to0,\qquad t\to\infty, \end{equation*} \notag $$
so that the matrix entries $\widehat{q}_{ij}(t)$ behave as
$$ \begin{equation} \widehat{q}_{11}(t)=O(\kappa|L(t)|),\qquad\widehat{q}_{21}(t)=O(M(t)),\qquad \widehat{q}_{22}(t)=O(\kappa|L(t)|+M(t)), \end{equation} \tag{10} $$
where $M(t)=\sup_{s\geqslant t}|q(s)|$, and
$$ \begin{equation} \widehat{q}_{12}(t)=O\biggl(\frac1{\rho(t)}\int_0^t|q(s)|\rho'(s)\, ds\biggr). \end{equation} \tag{11} $$

Note that $\widehat{q}_{11}(t)=O(\kappa|L(t)|)$ as $t\to\infty$ and, besides, for sufficiently large $t$, we have

$$ \begin{equation*} |\widehat{q}_{21}(t)|\leqslant 2\rho(t)M(t)\int_t^{\infty}\frac{\rho'(s)}{\rho(s)^2}\, ds =2M(t). \end{equation*} \notag $$
We next derive the estimate
$$ \begin{equation*} |\widehat{q}_{12}(t)| \leqslant \frac2{\rho(t)}\max_{s\geqslant 0}e^{\kappa L(s)} \int_0^t|q(s)|\rho'(s)\, ds, \end{equation*} \notag $$
where the right-hand side tends to zero as $t\to\infty$, since $\rho(t)\nearrow\infty$ and $q(t)\to 0$. Finally, by the above, $\widehat{q}_{22}(t)=O(\kappa|L(t)|+M(t))$ as $t\to\infty$.

Corollary 2. Under the hypothesis of Lemma 6,

$$ \begin{equation*} \|V(t)\|=O\biggl(\bigl(\kappa|L(t)| +M(t)\bigr)|q(t)|\frac{\rho'(t)}{\rho(t)}\biggr), \qquad t\to\infty. \end{equation*} \notag $$

The above reduction scheme was employed earlier (see [10]) in the cases $\kappa=0$ and $\kappa=1$ for the derivation of certain well-known (cf. Proposition 3) conditions for asymptotic equivalence for the solutions to equations (7) and (8).

§ 7. Hartman and Wintner type conditions

If $V\in\mathrm{L}_1(\mathbb R_+)$, then by Levinson’s theorem the reduced $L$-diagonal system

$$ \begin{equation*} Z'= \{\Lambda(t)+V(t)\}Z \end{equation*} \notag $$
has the fundamental matrix
$$ \begin{equation*} \begin{pmatrix} 1+\varepsilon_{11}(t)&\varepsilon_{12}(t) \\ \varepsilon_{21}(t)&1+\varepsilon_{22}(t) \end{pmatrix}\begin{pmatrix} 1&0 \\ 0&\rho(t) \end{pmatrix}, \end{equation*} \notag $$
where
$$ \begin{equation} \varepsilon_{ii}(t)= O\biggl(\int_t^{\infty}\|V(s)\|\, ds\biggr),\qquad i=1,2, \end{equation} \tag{12} $$
and
$$ \begin{equation} \varepsilon_{12}(t)=O\biggl(\frac1{\rho(t)}\int_0^t\rho(s)\|V(s)\|\, ds\biggr),\qquad \varepsilon_{21}(t)=O\biggl(\rho(t)\int_t^{\infty}\frac1{\rho(s)}\|V(s)\|\, ds\biggr). \end{equation} \tag{13} $$

After inverse transformations, we obtain an asymptotic factorization of the fundamental matrix for the initial system (9), and, as a result, we arrive at the conclusion that the solutions to equations (7) and (8) are asymptotically equivalent, with sharp estimates for the accuracy of the corresponding approximations (cf. [14] and [15]).

Proposition 4. Let $q(t)\to 0$ as $t\to\infty$, let the integral

$$ \begin{equation*} L(t)=\int_t^{\infty}q(t)\frac{\rho'(t)}{\rho(t)}\, dt \end{equation*} \notag $$
converge, and let the function $(\kappa|L(t)|+M(t))|q(t)|\rho'(t)/\rho(t)$ be integrable on the half-axis $\mathbb R_+$. Then equation (7) has the fundamental system of solutions
$$ \begin{equation} x_j(t)=y_j(t)\{1+\widehat{q}_{1j}(t)+\widehat{q}_{2j}(t) + O(\varepsilon_{1j}(t))+O(\varepsilon_{2j}(t))\},\qquad j=1,2, \end{equation} \tag{14} $$
where $\widehat{q}_{ij}(t)$ and $\varepsilon_{ij}(t)$ satisfy estimates (10), (11) and (12), (13).

For a proof, it suffices to evaluate the product

$$ \begin{equation*} \begin{pmatrix} x_1 &x_2 \\ rx_1' &rx_2' \end{pmatrix}=\begin{pmatrix} y_1 &y_2 \\ ry_1' &ry_2' \end{pmatrix}\begin{pmatrix} 1 &0 \\ 0 &\dfrac1{\rho} \end{pmatrix}\begin{pmatrix} 1+q_{11} &q_{12} \\ q_{21} &1+q_{22} \end{pmatrix}\begin{pmatrix} 1+\varepsilon_{11} &\varepsilon_{12} \\ \varepsilon_{21} &1+\varepsilon_{22} \end{pmatrix}\begin{pmatrix} 1 &0 \\ 0 &\rho \end{pmatrix}, \end{equation*} \notag $$
where the entries $q_{ij}$ of the matrix $Q=(I+\widetilde{Q})(I+\widehat{Q})-I$ are given by
$$ \begin{equation*} \begin{alignedat}{3} q_{11} &=\widehat{q}_{11}-q(1+\widehat{q}_{11}+\widehat{q}_{21}), &\qquad q_{12} &=\widehat{q}_{12}-q(1+\widehat{q}_{12}+\widehat{q}_{22}), \\ q_{21} &=\widehat{q}_{21}+q(1+\widehat{q}_{11}+\widehat{q}_{21}), &\qquad q_{22} &=\widehat{q}_{22}+q(1+\widehat{q}_{12}+\widehat{q}_{22}), \end{alignedat} \end{equation*} \notag $$
so that $q_{11}+q_{21}=\widehat{q}_{11}+\widehat{q}_{21}$ and $q_{21}+q_{22}=\widehat{q}_{21}+\widehat{q}_{22}$. Multiplying the above matrices, we find that
$$ \begin{equation*} \begin{aligned} \, x_1(t) &=y_1(t)\bigl(1+q_{11}(t)+q_{21}(t)\bigr)\bigl(1+\varepsilon_{11}(t)\bigr)+ y_1(t)\bigl(1+q_{12}(t)+q_{22}(t)\bigr)\varepsilon_{21}(t) \\ &=y_1(t)\{1+\widehat{q}_{11}(t)+\widehat{q}_{21}(t) +O(\varepsilon_{11}(t)) +O(\varepsilon_{21}(t))\}, \\ x_2(t) &=y_2(t)\bigl(1+q_{11}(t)+q_{21}(t)\bigr)\varepsilon_{12}(t)+ y_2(t)\bigl(1+q_{12}(t)+q_{22}(t)\bigr)\bigl(1+\varepsilon_{22}(t)\bigr) \\ &=y_2(t)\{1+\widehat{q}_{12}(t)+\widehat{q}_{22}(t) +O(\varepsilon_{12}(t)) +O(\varepsilon_{22}(t))\}. \end{aligned} \end{equation*} \notag $$
This gives us the desired fundamental system of solutions $\{x_1(t),x_2(t)\}$ to equation (7) with asymptotics (14) and with estimates (10)(13) for the remainder terms.

Proof of Theorem 2. Equation (2) with $\varphi(t)\equiv 0$ will be chosen in our context as an unperturbed comparison equation within the framework of the above scheme, so that $y_{1,2}(t)=e^{\mp\lambda t}$ and $\rho(t)=e^{2\lambda t}$.

We claim first that if $\int^{\infty}\varphi(t)\, dt<\infty$, then both integrals

$$ \begin{equation*} q(t)=\int_t^{\infty}\varphi(s)e^{2\kappa\lambda(t-s)}\, ds,\qquad \kappa L(t)=2\kappa\lambda\int_t^{\infty}q(s)\, ds \end{equation*} \notag $$
also converge for $\kappa>0$ and behave as $O(\mu(t))$, where $\mu(t)=\sup_{s\geqslant t} \bigl|\int_s^{\infty}\varphi(r)\, dr\bigr|$. Indeed, an integration by parts
$$ \begin{equation*} \int_t^{\infty}\varphi(s)e^{-2\kappa\lambda s}\, ds =e^{-2\kappa\lambda t} \int_t^{\infty}\varphi(s)\, ds -2\kappa\lambda\int_t^{\infty}e^{-2\kappa\lambda s} \biggl(\int_s^{\infty}\varphi(r)\, dr\biggr)\, ds \end{equation*} \notag $$
shows that
$$ \begin{equation*} \biggl|\int_t^{\infty}\varphi(s)e^{-2\kappa\lambda s}\, ds\biggr| \leqslant 2\mu(t)e^{-2\kappa\lambda t}, \end{equation*} \notag $$
and, hence, $|q(t)|\leqslant 2\mu(t)$. Similarly we apply integration by parts to deal with the expression
$$ \begin{equation*} \int_t^{\infty}q(s)\, ds=\int_t^{\infty}e^{2\kappa\lambda s} \biggl(\int_s^{\infty}\varphi(r)e^{-2\kappa\lambda r}\, dr\biggr)\, ds =\frac1{2\kappa\lambda}\biggl(\int_t^{\infty}\varphi(s)\, ds-q(t)\biggr), \end{equation*} \notag $$
so that $\kappa|L(t)|\leqslant 3\mu(t)$.

We next note that condition (5) guarantees that the function

$$ \begin{equation*} \bigl(\kappa|L(t)|+M(t)\bigr)|q(t)|\frac{\rho'(t)}{\rho(t)} \end{equation*} \notag $$
is integrable on the half-axis $\mathbb R_+$. Indeed, in our setting, $\rho'(t)/\rho(t)=2\lambda$, and
$$ \begin{equation*} \kappa|L(t)q(t)|\leqslant 3|q(t)|\mu(t),\qquad |M(t)q(t)|\leqslant 2|q(t)|\mu(t). \end{equation*} \notag $$
Now an application of Proposition 4 shows that equation (2) has required solutions with asymptotics
$$ \begin{equation*} \begin{aligned} \, u_1(t)&= \exp(-\lambda t)\{1+\widehat{q}_{11}(t)+\widehat{q}_{21}(t) + O(\varepsilon_{11}(t))+O(\varepsilon_{21}(t))\}, \\ u_2(t)&= \exp(\lambda t)\{1+\widehat{q}_{12}(t)+\widehat{q}_{22}(t) + O(\varepsilon_{12}(t))+O(\varepsilon_{22}(t))\}, \end{aligned} \end{equation*} \notag $$
where, by Lemma 6,
$$ \begin{equation*} \begin{aligned} \, \widehat{q}_{11}(t) &= O\bigl(\kappa|L(t)|\bigr)=O(\mu(t)),\qquad \widehat{q}_{21}(t)=O(M(t))=O(\mu(t)), \\ \widehat{q}_{12}(t) &= O\biggl(\frac1{\rho(t)}\int_0^t|q(s)|\rho'(s)\, ds\biggr) = O\biggl(\int_0^te^{2\lambda(s-t)}\mu(s)\, ds\biggr), \\ \widehat{q}_{22}(t) &= O\bigl(\kappa|L(t)|+M(t)\bigr) = O\biggl(\int_0^te^{2\lambda(s-t)}\mu(s)\, ds\biggr). \end{aligned} \end{equation*} \notag $$
In addition, we have
$$ \begin{equation*} \begin{gathered} \, \begin{aligned} \, |\varepsilon_{11}(t)|+|\varepsilon_{21}(t)|+|\varepsilon_{22}(t)| &= O\biggl(\int_t^{\infty}\|V(s)\|\, ds\biggr) \\ &=O\biggl(\int_t^{\infty}\bigl(\kappa|L(s)|+M(s)\bigr)|q(s)|\, ds\biggr) = O\bigl(\Phi_{\kappa}(t)\bigr), \end{aligned} \\ \varepsilon_{12}(t) =O\biggl(\frac1{\rho(t)}\int_0^t\rho(s)\|V(s)\|\, ds\biggr) =O\biggl(\int_0^te^{2\lambda(s-t)}\mu(s)\, ds\biggr). \end{gathered} \end{equation*} \notag $$
This proves Theorem 2.

To conclude, we give an illustrative example of equation (2) to which Theorem 2 applies, but Theorem 1 does not. Setting $\varphi(t)=\sin t^2$, we have

$$ \begin{equation*} \tau(t)=\int_t^{\infty}e^{2\lambda(t-s)}|\varphi(s)|\, ds \geqslant \int_t^{\infty}e^{2\lambda(t-s)}\varphi(s)^2\, ds =\frac1{4\lambda}+O(t^{-1}). \end{equation*} \notag $$
Hence $\tau\varphi\notin\mathrm{L}_1(\mathbb R_+)$, and so the hypotheses of Theorem 1 are not met. At the same time,
$$ \begin{equation*} \frac12|q(t)|\leqslant \mu(t)=\sup_{s\geqslant t} \biggl|\int_s^{\infty}\sin r^2\, dr\biggr|\leqslant \frac1{t}, \end{equation*} \notag $$
and, therefore, $\Phi_{\kappa}(t)\leqslant 2/t$. We also have $\int_0^te^{2\lambda(s-t)}\mu(s)\, ds=O(t^{-1})$, and so by Theorem 2 the corresponding equation (2) has linearly independent solutions of the form
$$ \begin{equation*} u_{1,2}(t)=\exp(\mp\lambda t)\{1+O(t^{-1})\}. \end{equation*} \notag $$


Bibliography

1. M. V. Fedoryuk, Asymptotic analysis. Linear ordinary differential equations, Springer-Verlag, Berlin, 1993  crossref  mathscinet  zmath
2. J. Heading, An introduction to phase-integral methods, Methuen & Co., Ltd., London; John Wiley & Sons, Inc., New York, 1962  mathscinet  zmath
3. S. A. Stepin, “The WKB method and dichotomy for ordinary differential equations”, Dokl. Math., 72:2 (2005), 783–786
4. R. Bellman, Stability theory of differential equations, McGraw-Hill Book Co., Inc., New York–Toronto–London, 1953  mathscinet  zmath
5. Ph. Hartman, Ordinary differential equations, John Wiley & Sons, Inc., New York–London–Sydney, 1964  mathscinet  zmath
6. W. A. Harris, Jr., and D. A. Lutz, “A unified theory of asymptotic integration”, J. Math. Anal. Appl., 57:3 (1977), 571–586  crossref  mathscinet  zmath
7. W. F. Trench, “Linear perturbations of a nonoscillatory second-order equation”, Proc. Amer. Math. Soc., 97:3 (1986), 423–428  crossref  mathscinet  zmath
8. J. Šimša, “Asymptotic integration of a second order ordinary differential equation”, Proc. Amer. Math. Soc., 101:1 (1987), 96–100  crossref  mathscinet  zmath
9. Shao Zhu Chen, “Asymptotic integrations of nonoscillatory second order ordinary differential equations”, Trans. Amer. Math. Soc., 327:2 (1991), 853–865  crossref  mathscinet  zmath
10. S. Bodine and D. A. Lutz, “Asymptotic integration of nonoscillatory differential equations: a unified approach”, J. Dyn. Control Syst., 17:3 (2011), 329–358  crossref  mathscinet  zmath
11. S. A. Stepin, “Interpolation in the asymptotic integration of nonoscillatory differential equations”, Dokl. Math., 85:2 (2012), 174–177  crossref  zmath
12. V. V. Nemytskii and V. V. Stepanov, Qualitative theory of differential equations, Princeton Math. Ser., 22, Princeton Univ. Press, Princeton, NJ, 1960  mathscinet  zmath
13. L. A. Lusternik and V. J. Sobolev, Elements of functional analysis, Internat. Monogr. Adv. Math. Phys., Hindustan Publishing Corp., Delhi; Halsted Press [John Wiley & Sons, Inc.], New York, 1974  mathscinet  zmath
14. S. A. Stepin, “Asymptotic integration of nonoscillatory second-order differential equations”, Dokl. Math., 82:2 (2010), 751–754  crossref  zmath
15. S. Bodine and D. A. Lutz, “Asymptotic solutions and error estimates for linear systems of difference and differential equations”, J. Math. Anal. Appl., 290:1 (2004), 343–362  crossref  mathscinet  zmath

Citation: S. A. Stepin, “Interpolating asymptotic integration methods for second-order differential equations”, Izv. RAN. Ser. Mat., 88:1 (2024), 121–140; Izv. Math., 88:1 (2024), 114–132
Citation in format AMSBIB
\Bibitem{Ste24}
\by S.~A.~Stepin
\paper Interpolating asymptotic integration methods
for second-order differential equations
\jour Izv. RAN. Ser. Mat.
\yr 2024
\vol 88
\issue 1
\pages 121--140
\mathnet{http://mi.mathnet.ru/im9438}
\crossref{https://doi.org/10.4213/im9438}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4727544}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2024IzMat..88..114S}
\transl
\jour Izv. Math.
\yr 2024
\vol 88
\issue 1
\pages 114--132
\crossref{https://doi.org/10.4213/im9438e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001202734300007}
Linking options:
  • https://www.mathnet.ru/eng/im9438
  • https://doi.org/10.4213/im9438e
  • https://www.mathnet.ru/eng/im/v88/i1/p121
  • Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Известия Российской академии наук. Серия математическая Izvestiya: Mathematics
    Statistics & downloads:
    Abstract page:151
    Russian version PDF:5
    English version PDF:27
    Russian version HTML:7
    English version HTML:55
    References:9
    First page:7
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024