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, 2022, Volume 86, Issue 5, Pages 980–991
DOI: https://doi.org/10.4213/im9211e
(Mi im9211)
 

On summable solutions of a class of nonlinear integral equations on the whole line

Kh. A. Khachatryanab, H. S. Petrosyanbc

a Yerevan State University
b Lomonosov Moscow State University
c National Agrarian University of Armenia
References:
Abstract: In this paper, we study a class of nonlinear integral equations with a noncompact monotone Hammerstein–Nemytskii operator on the whole real line. This class of equations is widely used in various fields of natural science. In particular, such equations arise in mathematical biology and in the theory of radiative transfer. A constructive existence theorem for a nonnegative nontrivial summable and bounded solution is proved. We also study the asymptotic behavior of the solution at $\pm\infty$. At the end of the paper, specific examples of the indicated equations are given, that satisfy all the conditions of the proved existence theorem. In an important particular case, it is possible to prove a uniqueness theorem in a certain class of essentially bounded functions.
Keywords: Hammerstein–Nemytskii operator, Diekmann function, iterations, monotonicity, summability.
Funding agency Grant number
Russian Science Foundation 19-11-00223
The research was supported by the Russian Science Foundation (project no. 19-11-00223).
Received: 30.05.2021
Russian version:
Izvestiya Rossiiskoi Akademii Nauk. Seriya Matematicheskaya, 2022, Volume 86, Issue 5, Pages 157–168
DOI: https://doi.org/10.4213/im9211
Bibliographic databases:
Document Type: Article
UDC: 517.968.4
MSC: 45G05
Language: English
Original paper language: Russian

§ 1. Introduction and problem statement

Consider the following class of nonlinear integral equations on the whole line with a monotone Hammerstein–Nemytskii operator:

$$ \begin{equation} f(x)\,{=}\,G_0(x,f(x))\,{+}\,\mu(x)\int_{-\infty}^\infty K(x-t)G_1(t,f(t))\,dt,\qquad x\,{\in}\,\mathbb{R}\,{:=}\,(-\infty,+\infty), \end{equation} \tag{1} $$
with respect to unknown measurable nonnegative function $f(x)$. Equation (1) arises in various areas of natural science. In particular, such equations arise in mathematical biology (in the theory of spatio-temporal spread of an epidemic), in econometrics (in theory of national income distribution), in astrophysics (in the theory of radiative transfer in spectral lines), (see [1]–[7] and references therein). A distinctive feature of equation (1) is the non-compactness of the corresponding nonlinear Hammerstein–Nemytskii operator in the space of essentially bounded on $\mathbb{R}$ functions. The latter circumstance, the criticality condition for equation (1) (i.e., the presence of trivial (constant) solutions), and unboundedness of the integration domain in (1) lead to certain rather subtle difficulties in constructing nontrivial solutions to equations (1). It should be noted that the general classical principles of fixed points do not work for equation (1) due to the properties of the Hammerstein–Nemytskii operator listed above. In the case when the kernel $K$ is an even conservative function and $0\leqslant\mu(x)\leqslant1$, $x\in\mathbb{R}$, $(1-\mu(x))x\in L_1(\mathbb{R})$, under sufficiently strong restrictions on $G_0$ and $G_1$, equation (1) was studied in [8].

In the case, when there exists $\nu(K):=\int_{-\infty}^\infty x K(x)\,dx\neq0$, and mapping $G_1(t, u)$ is contractive in $u$ uniformly with respect to the variable $t$, under certain conditions on $G_0(x, u)$, equation (1) and the corresponding vector analog (a system of nonlinear integral equations) were studied in [9] and [10]. In these papers, summable and essentially bounded solutions on $\mathbb{R}$ were constructed.

In the present paper, we study the question of constructing a nontrivial, nonnegative, essentially bounded and integrable on $\mathbb{R}$ solution of equation (1) in the case $\nu(K)>0$ under essentially different restrictions on $G_0$, $G_1$ and $\mu$, and moreover, we do not assume the condition that the mapping $G_1(t,u)$ is contractive with respect to $u$. At the end of the paper, we present specific applied examples of the kernel $K$ and the functions $G_0, G_1$ and $\mu$.

§ 2. Auxiliary facts

2.1. Basic conditions on kernel $K$ and on function $\mu$

With respect to kernel $K$, we assume that the following basic conditions are satisfied:

I) $K\in L_1(\mathbb{R})\cap M(\mathbb{R})$, $K(x)>0$, $x\in \mathbb{R}$, $\int_{-\infty}^\infty K(x)\,dx=1$;

II) there exists $\nu(K)=\int_{-\infty}^\infty x K(x)\,dx>0$;

III) there exists a number $\lambda_1>0$ such that

$$ \begin{equation*} \int_{-\infty}^\infty K(x)e^{-\lambda x}\,dx<+\infty,\qquad \lambda\in[0, \lambda_1]. \end{equation*} \notag $$

Assume that $\mu(x)$ is a continuous function defined on $\mathbb{R}$, and

$$ \begin{equation} 0\leqslant\mu(x)\leqslant1,\qquad x\in \mathbb{R},\quad \mu\not\equiv1. \end{equation} \tag{2} $$

2.2. Conditions on the function $G_1(t,u)$

Let $G(u)$ be a continuous function defined on the set $\mathbb{R}^+:=[0,+\infty)$ and satisfying the following conditions:

a) there exists a finite derivative $G'(0)>1$ such that

$$ \begin{equation*} G(u)\leqslant G'(0)u,\quad u\in \mathbb{R}^+; \end{equation*} \notag $$

b) the function $G(u)$ is upward convex on the set $\mathbb{R}^+$;

c) there exists a number $\eta>0$ such that $G(u)\uparrow$ in $u$ on the set $[0,\eta]$, $G(\eta)=\eta$;

d) there exist numbers $\varepsilon>0$ and $c>0$ such that

$$ \begin{equation*} G(u)\geqslant G'(0)u- cu^{1+\varepsilon},\qquad u\in [0,\eta]. \end{equation*} \notag $$

Regarding the nonlinearity of $G_1(t,u)$, we assume that the following conditions are hold:

1) for any fixed $t\in\mathbb{R}$ the function $G_1(t,u)\uparrow$ in $u$ on $[0,\eta]$;

2) the double inequality is satisfied

$$ \begin{equation*} 0\leqslant G_1(t,u)\leqslant \eta-G(\eta-u),\qquad u\in [0,\eta],\quad t\in \mathbb{R}; \end{equation*} \notag $$

3) the nonlinearity of $G_1(t,u)$ on the set $\mathbb{R}\times [0,\eta]$ satisfies the Caratheodory condition with respect to argument $u$, i.e., for any fixed $u\in [0,\eta]$ the function $G_1(t,u)$ is measurable in $t$ on $\mathbb{R}$ and for almost all $t\in \mathbb{R}$ this function is continuous in $u$ on the segment $[0,\eta]$.

2.3. Diekmann function

Before imposing the appropriate conditions on the function $G_0(x,u)$, we introduce the well-known Diekmann function for the kernel $K$ (see [1]):

$$ \begin{equation} L(\lambda):=G'(0) \int_{-\infty}^\infty K(t)e^{-\lambda t}\,dt,\qquad \lambda\in[0,\lambda_1]. \end{equation} \tag{3} $$
Note that due to conditions a) and I)
$$ \begin{equation*} L(0)=G'(0) \int_{-\infty}^\infty K(t)\,dt=G'(0)>1. \end{equation*} \notag $$
On the other hand, according to condition II)
$$ \begin{equation*} L'(0)=-G'(0) \int_{-\infty}^\infty K(t)t\,dt<0. \end{equation*} \notag $$
Since $L'(\lambda)$ is continuous, there exists $\lambda^*\in(0,\lambda_1]$ such that
$$ \begin{equation} L'(\lambda)<0\quad\text{for}\quad\lambda\in[0,\lambda^*]. \end{equation} \tag{4} $$
In what follows we will assume that
$$ \begin{equation} L(\lambda^*)<1. \end{equation} \tag{5} $$
Note also that
$$ \begin{equation} L''(\lambda)=G'(0) \int_{-\infty}^\infty K(t)t^2 e^{-\lambda t}\,dt>0 \end{equation} \tag{6} $$
(may be infinity).

Hence, the function $L(\lambda)$ is downward convex. It follows from listed properties of the function $L(\lambda)$ that there exists a unique $\sigma\in(0,\lambda^*)$ such that

$$ \begin{equation} L(\sigma)=1. \end{equation} \tag{7} $$
Fix a number $\sigma$ and consider the following function (see [1]):
$$ \begin{equation} \mathcal{L}(x):=\max \{\eta e^{\sigma x}-M e^{(\delta+\sigma)x}, 0\},\qquad x\in \mathbb{R}, \end{equation} \tag{8} $$
where $M>0$ and $\delta\in(0, \lambda^*-\sigma)$ are numerical parameters.

Note that the properties of the function $L(\lambda)$ immediately imply that for $\delta\in(0, \lambda^*-\sigma)$

$$ \begin{equation} L(\sigma+\delta)<1. \end{equation} \tag{9} $$

2.4. Conditions on function $G_0(x,u)$

We assume that $G_0(x,u)$ satisfies the following conditions:

$\mathrm{i}_1)$ for any fixed $x\in\mathbb{R}$, the function $G_0(x,u)\uparrow$ in $u$ on the segment $[0,\eta]$;

$\mathrm{i}_2)$ the function $G_0(x,u)$ on the set $\mathbb{R}\times[0,\eta]$ satisfies the Caratheodory condition with respect to the argument $u$;

$\mathrm{i}_3)$ there exists number $\xi\in(0,1)$ such that

$$ \begin{equation} G_0(x, q_\xi(x))\geqslant q_\xi(x),\quad G_0(x,\eta)\leqslant q_1(x),\qquad x\in\mathbb{R}, \end{equation} \tag{10} $$
where
$$ \begin{equation} q_\alpha(x):=\alpha\mathcal{L}(x)(1-\mu(x)),\qquad \alpha>0,\quad x\in\mathbb{R}. \end{equation} \tag{11} $$

§ 3. The main result

The following theorem holds.

Theorem 1. Let conditions I)–III), 1)–3), $\mathrm{i}_1)$–$\mathrm{i}_3)$, (2) and (5) be satisfied. Then equation (1) has a nonnegative nontrivial bounded and summable solution $f(x)$, and $\lim_{x\to\pm\infty}f(x)=0$.

Proof. Step I (a priori estimate for $\mathcal{L}(x)$). First we choose the parameters $M>0$ and $\delta\in(0, \lambda^*-\sigma)$ in such a way that the following inequality holds:
$$ \begin{equation} \mathcal{L}(x)\leqslant \eta-\eta e^{\sigma x},\qquad x\in (-\infty,0]. \end{equation} \tag{12} $$
To this end, consider the following auxiliary function:
$$ \begin{equation} F(x):=\eta-2\eta e^{\sigma x}+M e^{(\delta+\sigma)x},\qquad x\in \biggl(-\infty, \frac{1}{\delta}\ln\frac{\eta}{M}\biggr]. \end{equation} \tag{13} $$
Note that
$$ \begin{equation} F(-\infty)=\eta,\qquad F\biggl(\frac{1}{\delta}\ln\frac{\eta}{M}\biggr) =\eta\biggl(1-\biggl(\frac{\eta}{M}\biggr)^{\sigma/\delta}\biggr)>0, \end{equation} \tag{14} $$
when $M>\eta$.

On the other hand, for $\delta<\sigma$ the following inequality holds:

$$ \begin{equation*} \begin{aligned} \, F'(x) &=-2\eta\sigma e^{\sigma x}+M(\sigma+\delta)e^{(\delta+\sigma)x}= e^{\sigma x}\bigl(-2\eta\sigma+M(\sigma+\delta)e^{\delta x}\bigr) \\ &\leqslant e^{\sigma x}\bigl(-2\eta\sigma +M(\sigma+\delta)e^{\ln(\eta/M)}\bigr)=\eta e^{\sigma x}(\delta-\sigma)<0,\qquad x\leqslant \frac{1}{\delta}\ln\frac{\eta}{M}. \end{aligned} \end{equation*} \notag $$

Therefore, $F(x)\downarrow$ on the set $(-\infty, (1/\delta)\ln(\eta/M)]$. From the listed properties of the function $F$, it immediately follows that

$$ \begin{equation} F(x)>0,\qquad x\in \biggl(-\infty, \frac{1}{\delta}\ln\frac{\eta}{M}\biggr]. \end{equation} \tag{15} $$
By direct checking, we can verify that for $M>\eta$ the function $\mathcal{L}(x)$ reaches its maximum at the point
$$ \begin{equation*} x_{\mathrm{max}}=\frac{1}{\delta}\ln\frac{\eta\sigma}{M(\sigma+\delta)}<0\quad\text{and}\quad \mathcal{L}(x)=0\quad\text{for}\quad x\geqslant\frac{1}{\delta}\ln\frac{\eta}{M}. \end{equation*} \notag $$
Thus, by (15) for $M>\eta$ and $\delta\in(0, \min(\lambda^*-\sigma, \sigma))$, inequality (12) holds. Geometrically, inequality (12) is shown in Fig. 1.

Step II (nonlinear auxiliary equation). In addition to equation (1), consider the following auxiliary integral equation with nonlinearity $G$:

$$ \begin{equation} \varphi(x)=\int_{-\infty}^\infty K(x-t)G(\varphi(t))\,dt,\qquad x\in \mathbb{R}, \end{equation} \tag{16} $$
with respect to the required function $\varphi(x)$.

From the results of [11] it follows that when

A) $M>\max\biggl\{\eta,\dfrac{c\eta^{1+\varepsilon}L(\sigma+\delta)}{G'(0)(1-L(\delta+\sigma))}\biggr\}$,

B) $\delta\in(0, \min(\varepsilon\sigma, \sigma, \lambda^*-\sigma))$

equation (16) has a nonnegative nontrivial bounded monotonically increasing solution $\varphi(x)$ on $\mathbb{R}$, and

$$ \begin{equation} \mathcal{L}(x)\leqslant\varphi(x)\leqslant\begin{cases} \eta e^{\sigma x}, &x\leqslant0, \\ \eta, &x>0, \end{cases} \qquad x\in\mathbb{R}. \end{equation} \tag{17} $$
Moreover,
$$ \begin{equation} \lim_{x\to+\infty}\varphi(x)=\eta,\qquad \eta-\varphi\in L_1(\mathbb{R}^+). \end{equation} \tag{18} $$

It was also proved in [5] that if $h(x)$ is any continuous, monotonically nondecreasing nonnegative and convex upward function on $\mathbb{R}^+$, and for some $m\in\mathbb{N}$ the following integral converges

$$ \begin{equation} \int_{0}^\infty x h^m(x)K(x)\,dx<+\infty, \end{equation} \tag{19} $$
then
$$ \begin{equation} \int_{0}^\infty h^m(x)(\eta-\varphi(x))\,dx<+\infty. \end{equation} \tag{20} $$
The listed properties of the function $\varphi(x)$ play an important role in our further reasoning. In what follows we will assume that the parameters $M$ and $\delta$ have the properties A) and B), respectively. Note that the following lower bound holds:
$$ \begin{equation} \eta-\varphi(x)\geqslant\mathcal{L}(x),\qquad x\in \mathbb{R}. \end{equation} \tag{21} $$
Indeed, by (8), (17) and A), the inequality (21) for $x>0$ is obviously satisfied. Let $x\leqslant0$. Then taking into account (12) and (17), we get
$$ \begin{equation*} \eta-\varphi(x)\geqslant \eta-\eta e^{\sigma x}\geqslant\mathcal{L}(x). \end{equation*} \notag $$

Step III (successive approximations for equation (1)). Consider the following successive approximations for the main equation (1):

$$ \begin{equation} \begin{aligned} \, f_{n+1}(x) &=G_0(x,f_n(x)) \nonumber \\ &\qquad+\mu(x)\int_{-\infty}^\infty K(x-t)G_1(t,f_n(t))\,dt,\qquad x\in\mathbb{R},\quad n=0,1,2,\dots, \end{aligned} \end{equation} \tag{22} $$
as the zero approximation we choose
$$ \begin{equation} f_0(x)=q_\xi(x),\qquad x\in \mathbb{R}. \end{equation} \tag{23} $$
By induction on $n$, we prove that:

$\mathrm{k}_1)$ $f_n(x)\uparrow$ in $n, x\in\mathbb{R}$;

$\mathrm{k}_2)$ $f_n(x)$ are measurable on $\mathbb{R}$, $n=0,1,2,\dots$;

$\mathrm{k}_3)$ $f_n(x)\leqslant\min\{\eta-\varphi(x), \varphi(x)\}$, $n=0,1,2,\dots$, $x\in\mathbb{R}$.

The measurability of the zero approximation immediately follows from the definition of the function $q_\xi(x)$. We have from (11), (21), (17) and (2) that

$$ \begin{equation*} 0\leqslant f_0(x)\leqslant\mathcal{L}(x)\leqslant \min\{\eta-\varphi(x), \varphi(x)\},\qquad x\in\mathbb{R}. \end{equation*} \notag $$
Prove that $f_1(x)\geqslant f_0(x)$, $x\in\mathbb{R}$. Taking into account I), (2), 2) and $\mathrm{i}_3)$, we obtain from (22) that
$$ \begin{equation*} f_1(x)\geqslant G_0(x, q_\xi(x))\geqslant q_\xi(x)=f_0(x),\qquad x\in\mathbb{R}. \end{equation*} \notag $$
Assume now that for some positive integer $n$ the function $f_n(x)$ is measurable, $f_n(x)\geqslant f_{n-1}(x)$, $x\in\mathbb{R}$, and $f_n(x)\leqslant \min\{\eta-\varphi(x), \varphi(x)\}$. Then, by conditions 3), $\mathrm{i}_2)$, I) and (2), the function $f_{n+1}(x)$ will also be measurable. Since the functions $G_0(x,u)$ and $G_1(t,u)$ are monotonous in $u$ and the kernel $\mu(x)K(x-t)$ is nonnegative, by the induction assumption, we have from (22) that
$$ \begin{equation*} f_{n+1}(x)\geqslant G_0(x, f_{n-1}(x))+\mu(x)\int_{-\infty}^\infty K(x-t)G_1(t,f_{n-1}(t))\,dt=f_n(x). \end{equation*} \notag $$
We now prove that $f_{n+1}(x)\leqslant \min\{\eta-\varphi(x), \varphi(x)\}$. First we prove that $f_{n+1}(x)\leqslant\varphi(x)$, $x\in\mathbb{R}$. Indeed, taking into account the inductive assumption, formulas (2), (10), (16), (17), (21), conditions 2), 1), b) and Jensen’s inequality, we obtain from (22) that
$$ \begin{equation*} \begin{aligned} \, f_{n+1}(x) &\leqslant G_0(x, \varphi(x))+\mu(x)\int_{-\infty}^\infty K(x-t)G_1(t,\varphi(t))\,dt \\ &\leqslant G_0(x,\eta)+\mu(x)\int_{-\infty}^\infty K(x-t)\bigl(\eta-G(\eta-\varphi(t))\bigr)\,dt \\ &\leqslant q_1(x)+\mu(x)\int_{-\infty}^\infty K(x-t)\bigl(G(\eta)-G(\eta-\varphi(t))\bigr)\,dt \\ &= q_1(x)+ \mu(x)\int_{-\infty}^\infty K(x-t) \biggl(\frac{\varphi(t)}{\eta}G(\eta) +\frac{\eta-\varphi(t)}{\eta}G(0) \\ &\qquad +\frac{\eta-\varphi(t)}{\eta}G(\eta)+\frac{\varphi(t)}{\eta}G(0)-G(\eta-\varphi(t))\biggr)\,dt \\ &\leqslant q_1(x)+\mu(x)\int_{-\infty}^\infty K(x-t) \bigl(G(\varphi(t))+G(\eta-\varphi(t))-G(\eta-\varphi(t))\bigr)\,dt \\ &=q_1(x)+\mu(x)\varphi(x)\leqslant(1-\mu(x))\varphi(x)+\mu(x)\varphi(x)=\varphi(x). \end{aligned} \end{equation*} \notag $$
Now we prove that $f_{n+1}(x)\leqslant \eta-\varphi(x),\,x\in\mathbb{R}$. By conditions 2), 1), $\mathrm{i}_1)$ and formulas (2), (10), taking into consideration inductive hypothesis and relations (16), (17), (21), we have from (22) that
$$ \begin{equation*} \begin{aligned} \, f_{n+1}(x) &\leqslant G_0(x, \eta-\varphi(x))+\mu(x)\int_{-\infty}^\infty K(x-t)G_1(t,\eta-\varphi(t))\,dt \\ &\leqslant G_0(x,\eta)+\mu(x)\int_{-\infty}^\infty K(x-t)(\eta-G(\varphi(t)))\,dt \\ &\leqslant \mathcal{L}(x)(1-\mu(x))+\mu(x)(\eta-\varphi(x)) \\ &\leqslant (\eta-\varphi(x))(1-\mu(x))+\mu(x)(\eta-\varphi(x))=\eta-\varphi(x). \end{aligned} \end{equation*} \notag $$
Thus, it follows from $\mathrm{k}_1)$–$\mathrm{k}_3)$ that the sequence $\{f_n(x)\}_{n=0}^\infty$ of measurable functions has a pointwise limit when $n\to\infty$: $\lim_{n\to\infty}f_n(x)=f(x)$, and the limit function satisfies the following double inequality:
$$ \begin{equation} q_\xi(x)\leqslant f(x)\leqslant \min\{\eta-\varphi(x), \varphi(x)\},\qquad x\in\mathbb{R}. \end{equation} \tag{24} $$
From Levi’s (see [12]) and Krasnoselsky’s (see [13]) limit theorems it follows that $f(x)$ satisfies equation (1) almost everywhere on $\mathbb{R}$. Since $\varphi\in L_1(\mathbb{R}\setminus \mathbb{R}^+)$, $ \eta- \varphi\in L_1(\mathbb{R}^+)$, $\varphi(-\infty)=0$, $\varphi(+\infty)=\eta$, then (24) yields that $f(\pm\infty)=0$ and $f\in L_1(\mathbb{R})\cap M(\mathbb{R})$. Theorem 1 is completely proved.

Remark 1. By inequalities (24) and (17), the proof of the main result implies a stronger limit relation for solution $f(x)$ when $x\to-\infty$. Namely, for any number $\rho\in(0,\sigma)$

$$ \begin{equation*} \lim_{x\to-\infty}e^{-\rho x}f(x)=0. \end{equation*} \notag $$
Moreover, $e^{-\rho x}f(x)\in L_1(-\infty,0)$.

In Fig. 2, the preliminary zone for finding the constructed solution of equation (1) is shaded.

Remark 2. In the proof of the theorem at the second step, it was noted that under condition (19), the integral (20) converges. It immediately follows from the estimate (24) that under additional condition (19), the solution of equation (1) possesses the following property:

$$ \begin{equation*} \int_0^\infty h^m(x)f(x)\,dx<+\infty \end{equation*} \notag $$
(for the properties of the function $h$ see the proof of Theorem 1, Step II).

§ 4. Examples of function $K$, $\mu$, $G_0$ and $G_1$

In applications (see [1], [3], [5]) the following specific kernel representations $K(x)$ and function $\mu(x)$ occur:

$\mathrm{p}_1)$ $K(x)=(1/\sqrt{\pi}\,)e^{-(x-l)^2}$, $x\in\mathbb{R}$, where $l>0$ is an arbitrary parameter;

$\mathrm{p}_2)$ $K(x)=\int_a^b e^{-|x-l|s}Q(s)\, ds$, $x\in\mathbb{R}$, where $l>0$ is a parameter, and $Q\in C[a,b)$, $Q(s)>0$, $0<a\leqslant s\leqslant b\leqslant+\infty$, and

$$ \begin{equation*} \int_a^b\frac{Q(s)}{s}\, ds=\frac{1}{2}; \end{equation*} \notag $$

$\mathrm{\beta}_1)$ $\mu(x)=1-e^{-|x|}$, $x\in\mathbb{R}$;

$\mathrm{\beta}_2)$ $\mu(x)=e^{-x^2}$, $x\in\mathbb{R}$.

First, we provide several examples of the function $G(u)$, then, using the construction of the numbers $\eta$ and $\sigma$, we will give the corresponding examples of the nonlinearities $G_0$ and $G_1$.

In the mathematical theory of the spatio-temporal spread of an epidemic, the following particular examples of the nonlinearity of $G(u)$ occur (see [1]–[4]):

$\mathrm{g}_1)$ $G(u)=\gamma(1-e^{-u})$, $u\in\mathbb{R}^+$, $\gamma>1$ is a numerical parameter;

$\mathrm{g}_2)$ $G(u)=2u-u^2$, $u\in\mathbb{R}^+$.

The parameter $\gamma>1$ has a specific biological meaning. In the above theory, the inequality $\gamma>1$ is called the threshold condition. The latter is the critical number of infected persons, beyond which the epidemic cannot be stopped.

It is easy to check that for $\mathrm{g}_1)$ all conditions a)–d) are automatically satisfied. Take a closer look at the example $\mathrm{g}_2)$. Note that in this case $\eta=1$, $G'(0)=2$, $G\uparrow$ on the interval $[0,1]$, $G(0)=0$. We now construct the number $\sigma$ for the concrete kernel of the form $\mathrm{p}_1)$. In this case, the Diekmann function has the following form:

$$ \begin{equation*} L(\lambda)=\frac{2}{\sqrt{\pi}} \int_{-\infty}^\infty e^{-(x-l)^2}e^{-\lambda x}\,dx,\qquad \lambda\geqslant0, \end{equation*} \notag $$
or after certain calculations
$$ \begin{equation*} L(\lambda)=2e^{\lambda^2/4-l\lambda},\qquad \lambda\geqslant0. \end{equation*} \notag $$
Since $L'(\lambda)=(\lambda-2l)e^{\lambda^2/4-l\lambda}$, then $L(\lambda)\downarrow$ on the interval $[0, 2l]$.

We have

$$ \begin{equation*} L''(\lambda)=\biggl(1+\frac{1}{2}(\lambda-2l)^2 \biggr)e^{\lambda^2/4-l\lambda}>0, \end{equation*} \notag $$
which implies, in particular, that $L(\lambda)$ is downward convex. It is obvious that $L(0)\,{=}\,2$, $L(2l)=2e^{-l^2}<1$ for $l>\sqrt{\ln2}$. Therefore, when $l\in(\sqrt{\ln2},+\infty)$, there is a unique number $\sigma\in(0,2l)$ such that $L(\sigma)=1$. For example, in the case when $l=1$, after appropriate calculations, we find $\sigma\approx0.893$.

Finally, we provide concrete examples of the functions $G_0(x,u)$ and $G_1(x,u)$.

Examples. As $G_0(x,u)$ the following functions we can be considered:

$\mathrm{a}_1)$ $G_0(x,u)=q_1(x)u/(u+q_{\varepsilon_0}(x))$, $u\geqslant0$, $x\in \mathbb{R}$, where $\varepsilon_0\in (0, 1-\xi]$ is an arbitrary number;

$\mathrm{a}_2)$ $G_0(x,u)=q_1(x)u/(u+q_{\varepsilon_0}(x))+a(x)u^p$, $u\geqslant0$, $x\in \mathbb{R}$, $p>1$ is a parameter, $\varepsilon_0\in (0, 1-\xi]$, $a(x)$ is an arbitrary nonnegative continuous on $\mathbb{R}$ function and

$$ \begin{equation*} a(x)\leqslant \frac{q_1(x)q_{\varepsilon_0}(x)}{\eta^{p+1}+\eta^pq_{\varepsilon_0}(x)},\qquad x\in \mathbb{R}. \end{equation*} \notag $$

It is easy to prove that the above examples $\mathrm{a}_1)$ and $\mathrm{a}_2)$ satisfy conditions $\mathrm{i}_1)$–$\mathrm{i}_3)$.

As $G_1(t,u)$, we can consider the following functions:

$\mathrm{b}_1)$ $G_1(t,u)=b(t)(\eta-G(\eta-u))$, $u\geqslant0$, $t\in\mathbb{R}$, where $b(t)$ is any continuous function on $\mathbb{R}$ and

$$ \begin{equation*} 0\leqslant b(t)\leqslant1, \qquad t\in\mathbb{R}, \end{equation*} \notag $$

$\mathrm{b}_2)$ $G_1(t,u)=(u/(u+S(t)))(\eta-G(\eta-u))$, $u\geqslant0$, $t\in\mathbb{R}$, where $S(t)\geqslant0$, $t\in\mathbb{R}$ and $S\in C(\mathbb{R})$.

Direct checking will prove that conditions 1)–3) hold for examples $b_1)$ and $b_2)$.

For example, when $G(u)=2u-u^2$, as $G_1(t,u)$ we can choose the function $G_1(t,u)=b(t)u^2$, $u\geqslant0$, $t\in\mathbb{R}$.

§ 5. On the uniqueness of the solution of equation (1) in an important particular case

Assume that nonlinearity $G_0(x,u)$ admits a representation $a_1)$, where $\varepsilon_0=1- \xi$, $G_1(t,u)$ is defined according to $\mathrm{b}_1)$, and $\mu(x)$, in addition to (2), satisfies the following additional restriction:

$$ \begin{equation} M_0:=\sup_{x\in\mathbb{R}}\mu(x)<1. \end{equation} \tag{25} $$
In this case, equation (1) takes the following form:
$$ \begin{equation} f(x)=\frac{q_1(x)f(x)}{f(x)+q_{1-\xi}(x)}+\mu(x) \int_{-\infty}^\infty K(x-t)b(t)(\eta-G(\eta-f(t)))\,dt,\qquad x\in\mathbb{R}. \end{equation} \tag{26} $$

Below we will show that the following statement holds.

Theorem 2. Under conditions a)–d), (2) and (25), if $\xi\in(M_0,1)$, then equation (26) cannot have more than one solution in the following class of measurable and bounded functions on $\mathbb{R}$:

$$ \begin{equation} \mathfrak{M}:=\{f(x)\colon q_\xi(x)\leqslant f(x)\leqslant \Gamma(x),\ x\in\mathbb{R} \}, \end{equation} \tag{27} $$
where
$$ \begin{equation*} \Gamma(x):= \begin{cases} \eta e^{\sigma x}, &x\leqslant0, \\ \eta, &x>0. \end{cases} \end{equation*} \notag $$

Proof. Assume the opposite: equation (26) has two solutions $f$ and $\widetilde{f}$ from the class $\mathfrak{M}$. Below we verify that
$$ \begin{equation} \varkappa:=\sup_{x\in\mathbb{R}}e^{-\sigma x}|f(x)-\widetilde{f}(x)|<+\infty, \end{equation} \tag{28} $$
where the number $\sigma$ is determined from (7). Consider the following possibilities:

$\mathrm{r}_1)$ $x\geqslant(1/\delta)\ln(\eta/M)$;

$\mathrm{r}_2)$ $x<(1/\delta)\ln(\eta/M)$, where $\delta\in(0, \min\{\sigma\varepsilon, \sigma, \lambda^*-\sigma\})$.

In the case $\mathrm{r}_1)$, by the triangle inequality, we have

$$ \begin{equation*} e^{-\sigma x}|f(x)-\widetilde{f}(x)|\leqslant 2\eta e^{-\sigma x}\leqslant 2\eta e^{-(\sigma/\delta)\ln(\eta/M)}=2\eta\biggl(\frac{M}{\eta}\biggr)^{\sigma/\delta}. \end{equation*} \notag $$
Now consider $\mathrm{r}_2)$ $x<(1/\delta)\ln(\eta/M)$. Since $f, \widetilde{f}\in \mathfrak{M}$ and taking into account the definition of the function $\mathcal{L}(x)$, by (25), we obtain the following inequality:
$$ \begin{equation*} \begin{aligned} \, &-\eta\bigl(1-\xi(1-\mu(x))\bigr)e^{\sigma x} - \xi M(1-\mu(x)) e^{(\delta+\sigma) x}\leqslant f(x)-\widetilde{f}(x) \\ &\qquad \leqslant\eta\bigl(1-\xi(1-\mu(x))\bigr)e^{\sigma x} + \xi M(1-\mu(x)) e^{(\delta+\sigma) x}, \end{aligned} \end{equation*} \notag $$
which implies that
$$ \begin{equation*} e^{-\sigma x}|f(x)-\widetilde{f}(x)|\leqslant \xi M+\eta(1-\xi+\xi M_0), \qquad x<\frac{1}{\delta}\ln\frac{\eta}{M}. \end{equation*} \notag $$
Thus
$$ \begin{equation} \varkappa\leqslant \max\biggl\{2\eta\biggl(\frac{M}{\eta}\biggr)^{\sigma/\delta},\, \xi M+\eta(1-\xi+\xi M_0)\biggr\}. \end{equation} \tag{29} $$
From properties a)–d) of the function $G(u)$, it easily follows that for arbitrary $u_1, u_2\in[0,\eta]$ the following inequality holds:
$$ \begin{equation} |G(u_1)-G(u_2)|\leqslant G'(0)|u_1-u_2|. \end{equation} \tag{30} $$
Taking into account (30), the definition of the function $L(\lambda)$ and the fact that $f, \widetilde{f}\in \mathfrak{M}$, we have from (26) that
$$ \begin{equation*} \begin{aligned} \, e^{-\sigma x}|f(x)-\widetilde{f}(x)| &\leqslant \frac{q_1(x)q_{1-\xi}(x)e^{-\sigma x}}{(q_\xi(x)+q_{1-\xi}(x))^2}|f(x)-\widetilde{f}(x)| \\ &\qquad+ e^{-\sigma x}\mu(x)G'(0)\int_{-\infty}^\infty K(x-t)|f(t)-\widetilde{f}(t)|\,dt \\ &\leqslant (1-\xi)\varkappa+ e^{-\sigma x}G'(0)\varkappa M_0\int_{-\infty}^\infty K(x-t)e^{\sigma t}\,dt \\ &= (1-\xi)\varkappa+\varkappa G'(0)e^{-\sigma x}M_0\int_{-\infty}^\infty K(y)e^{-\sigma (y-x)}\, dy \\ &=(1-\xi+M_0)\varkappa, \end{aligned} \end{equation*} \notag $$
and therefore, we get
$$ \begin{equation*} \varkappa \leqslant (1-\xi+M_0)\varkappa, \end{equation*} \notag $$
or $\varkappa(\xi-M_0)\leqslant0$. Since $\xi\in(M_0, 1)$, the latter inequality yields that $\varkappa=0$. The latter means that $f(x)=\widetilde{f}(x)$ almost everywhere on $\mathbb{R}$. Theorem 2 is proved.

Remark 3. Note that Theorem 2 proved above complements the uniqueness theorem from [11].

Remark 4. Unfortunately, the problem of the uniqueness of a solution of equation (1) in $\mathfrak{M}$ under general restrictions on $G_0$ and $G_1$ is still an open problem.


Bibliography

1. O. Diekmann, “Thresholds and travelling waves for the geographical spread of infection”, J. Math. Biol., 6:2 (1978), 109–130  crossref  mathscinet  zmath
2. A. G. Sergeev and Kh. A. Khachatryan, “On the solvability of a class of nonlinear integral equations in the problem of a spread of an epidemic”, Tr. Mosk. Mat. Obs., 80, no. 1, MCCME, Moscow, 2019, 113–131  mathnet  zmath; English transl. Trans. Moscow Math. Soc., 80 (2019), 95–111  crossref  mathscinet
3. O. Diekmann, “Run for your life. A note on the asymptotic speed of propagation of an epidemic”, J. Differential Equations, 33:1 (1979), 58–73  crossref  mathscinet  zmath  adsnasa
4. A. Kh. Khachatryan and Kh. A. Khachatryan, “On the solvability of some nonlinear integral equations in problems of epidemic spread”, Mathematical physics and applications, Collected papers. In commemoration of the 95th anniversary of Academician Vasilii Sergeevich Vladimirov, Trudy Mat. Inst. Steklova, 306, Њ€ЂЌ, Moscow, 2019, 287–303  mathnet  crossref  mathscinet  zmath; English transl. Proc. Steklov Inst. Math., 306 (2019), 271–287  crossref
5. A. Kh. Khachatryana, Kh. A. Khachatryan, and H. S. Petrosyan, “Asymptotic behavior of a solution for one class of nonlinear integro-differential equations in the income distribution problem”, Tr. Inst. Mat. Mekh. Ural. Otd. Ross. Akad. Nauk, 27:1 (2021), 188–206  mathnet  crossref  mathscinet
6. J. D. Sargan, “The distribution of wealth”, Econometrica, 25:4 (1957), 568–590  crossref  mathscinet  zmath
7. N. B. Engibaryan, “On a problem in nonlinear radiative transfer”, Astrofizika, 2:1 (1966), 31–36 (https://arar.sci.am/dlibra/publication/27739/edition/24782/content); English transl. Astrophysics, 2:1 (1966), 12–14  crossref  adsnasa
8. Kh. A. Khachatryan and H. S. Petrosyan, “On the construction of an integrable solution to one class of nonlinear integral equations of Hammerstein–Nemytskii type on the whole axis”, Tr. Inst. Mat. Mekh. Ural. Otd. Ross. Akad. Nauk, 26:2 (2020), 278–287  mathnet  crossref  mathscinet
9. Kh. A. Khachatryan, “On positive solutions of one class of nonlinear integral equations of Hammerstein–Nemytskii type on the whole axis”, Tr. Mosk. Mat. Obs., 75, no. 1, MCCME, Moscow, 2014, 1–14  mathnet  mathscinet  zmath; English transl. Trans. Moscow Math. Soc., 75 (2014), 1–12  crossref
10. Kh. A. Khachatryan, “On solution of a system of Hammerstein–Nemitskii type nonlinear integral equations on whole axis”, Tr. Inst. Mat., 21:2 (2013), 154–161 (Russian)  mathnet
11. Kh. A. Khachatryan and H. S. Petrosyan, “On the solvability of a class of nonlinear Hammerstein–Stieltjes integral equations on the whole line”, Trudy Mat. Inst. Steklova, 308, Differential equations and dynamical systems (2020), 253–264  mathnet  crossref  mathscinet  zmath; English transl. Proc. Steklov Inst. Math., 308 (2020), 238–249  crossref
12. A. N. Kolmogorov and S. V. Fomin, Elements of the theory of functions and functional analysis, 5th ed., Nauka, Moscow, 1981  mathscinet  zmath; English transl. of 1st ed. v. I, II, Graylock Press, Albany, NY, 1957, 1961  mathscinet  mathscinet  zmath
13. M. A. Krasnosel'skii, P. P. Zabreyko, E. I. Pustylnik, and P. E. Sobolevski, Integral operators in spaces of summable functions, Nauka, Moscow, 1966  mathscinet  zmath; English transl. Monographs Textbooks Mech. Solids Fluids: Mech. Anal., Noordhoff International Publishing, Leiden, 1976  mathscinet  zmath

Citation: Kh. A. Khachatryan, H. S. Petrosyan, “On summable solutions of a class of nonlinear integral equations on the whole line”, Izv. RAN. Ser. Mat., 86:5 (2022), 157–168; Izv. Math., 86:5 (2022), 980–991
Citation in format AMSBIB
\Bibitem{KhaPet22}
\by Kh.~A.~Khachatryan, H.~S.~Petrosyan
\paper On summable solutions of a class of nonlinear integral equations on the whole line
\jour Izv. RAN. Ser. Mat.
\yr 2022
\vol 86
\issue 5
\pages 157--168
\mathnet{http://mi.mathnet.ru/im9211}
\crossref{https://doi.org/10.4213/im9211}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4582541}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2022IzMat..86..980K}
\transl
\jour Izv. Math.
\yr 2022
\vol 86
\issue 5
\pages 980--991
\crossref{https://doi.org/10.4213/im9211e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=000992252200007}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85165704146}
Linking options:
  • https://www.mathnet.ru/eng/im9211
  • https://doi.org/10.4213/im9211e
  • https://www.mathnet.ru/eng/im/v86/i5/p157
  • 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:1236
    Russian version PDF:48
    English version PDF:59
    Russian version HTML:411
    English version HTML:68
    References:69
    First page:17
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024