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, 2023, Volume 87, Issue 5, Pages 1062–1077
DOI: https://doi.org/10.4213/im9348e
(Mi im9348)
 

On non-trivial solvability of one system of non-linear integral equations on the real axis

Kh. A. Khachatryanab, H. S. Petrosyancb

a Yerevan State University
b Lomonosov Moscow State University
c National Agrarian University of Armenia
References:
Abstract: A system of singular integral equations with monotonic and convex non-linearity on the entire real line is considered. System of this form have applications in many areas of natural science. In particular, such systems arise in the theory of $p$-adic open-closed strings, in the mathematical theory of spatial-temporal epidemic spread within the framework of the well known Diekmann–Kaper model, in the kinetic theory of gases, in the radiative transfer theory. An existence theorem for a non-trivial and bounded solution is proved. The asymptotic behaviour of the constructed solution at $\pm\infty$ is also studied. Specific examples of non-linearities and kernel functions having an applied character are given.
Keywords: convexity, monotonicity, spectral radius, non-linearity, bounded solution.
Funding agency Grant number
Russian Science Foundation 19-11-00223
The research was supported by the Russian Science Foundation (project no. 19-11-00223, https://rscf.ru/en/project/19-11-00223/).
Received: 01.06.2022
Bibliographic databases:
Document Type: Article
UDC: 517.968.48
MSC: 45G05, 45G15
Language: English
Original paper language: Russian

§ 1. Introduction

Consider the system of non-linear integral equations on $\mathbb{R}:=(-\infty,+\infty)$

$$ \begin{equation} Q_i(\varphi_i(x))=\sum_{j=1}^n\mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)\lambda_{ij}(|t|)\varphi_j(t)\, dt,\qquad i=1,\dots, n,\quad x\in \mathbb{R}, \end{equation} \tag{1.1} $$
for the unknown bounded measurable vector function $\varphi(x)=(\varphi_1(x),\dots,\varphi_n(x))^\top$ ($\top$ denotes transposition). In system (1.1), $\{\mu_j(u)\}_{j=1}^n$, $\{\lambda_{ij}(u)\}_{i,j=1}^{n\times n}$ are measurable functions $\mathbb{R}^+:=[0,+\infty)$ satisfying the following conditions:

a) there exist numbers $\varepsilon_j\in(0,1)$, $j=1,\dots,n$, such that

$$ \begin{equation*} 0<\varepsilon_j\leqslant\mu_j(u)\leqslant1,\qquad \mu_j(u)\not\equiv1,\quad u\in \mathbb{R}^+, \quad j=1,\dots,n; \end{equation*} \notag $$

b) $\lambda_{ij}(u)\geqslant1$, $u\in \mathbb{R}^+$, $i,j=1,\dots, n$, $\lim_{u\to +\infty}\mu_j(u)=\lim_{u\to +\infty}\lambda_{ij}(u)=1$, $j=1,\dots, n$;

c) $1-\mu_j\in L_1(0,+\infty)\cap C(\mathbb{R}^+)$, $\lambda_{ij}-1\in L_1(0,+\infty)$, $\lambda_{ij}(t)=\lambda_{ji}(t)$, $ t\in\mathbb{R}$, $i,j=1,\dots, n$.

Elements of the matrix function $K(x):=(K_{ij}(x))_{i,j=1}^{n\times n}$ are defined on the set $\mathbb{R}$ and are such that:

A) $K_{ij}(x)>0$, $K_{ij}(x)=K_{ji}(x)$, $K_{ij}(-u)=K_{ij}(u)$, $x\in\mathbb{R}$, $u\in \mathbb{R}^+$, $i,j=1,\dots, n$;

B) $K_{ij}(x)$ decrease monotonically in $x$ on $\mathbb{R}^+$, $i,j=1,\dots, n$;

C) $K_{ij}\in L_1(\mathbb{R})\cap M(\mathbb{R})$, $r(A)=1$, where $r(A)$ is the spectral radius of the matrix $A:=\bigl(\int_{-\infty}^\infty K_{ij}(x)\, dx\bigr)_{i,j=1}^{n\times n}$, i.e., the largest absolute value of the eigenvalues of the matrix $A$; $M(\mathbb{R})$ is the space of essentially bounded functions on $\mathbb{R}$.

According to the Perron–Frobenius theorem (see [1]), there exists a vector $\eta:=(\eta_1, \dots, \eta_n)^\top$ with positive coordinates $\eta_j$, $j=1,\dots,n$, such that

$$ \begin{equation} \sum_{j=1}^n a_{ij}\eta_j=\eta_i,\qquad i=1,\dots,n, \end{equation} \tag{1.2} $$
where
$$ \begin{equation} a_{ij}:= \int_{-\infty}^\infty K_{ij}(x)\, dx,\qquad i,j=1,\dots,n. \end{equation} \tag{1.3} $$
We set
$$ \begin{equation} \eta_j^*:= \frac{\eta_j}{\min_{1\leqslant i\leqslant n}\{\eta_i\}},\qquad j=1,\dots,n. \end{equation} \tag{1.4} $$
The non-linearities $\{Q_i(u)\}_{i=1}^n$ (see Fig. 1) are assumed to satisfy the following conditions:

1) $Q_i(u)$ is monotone increasing in $u$ on $\mathbb{R}$, $i=1,\dots,n$;

2) $Q_i\in C(\mathbb{R})$, $Q_i(-u)=-Q_i(u)$, $u\in \mathbb{R}^+$, $i=1,\dots,n$;

3) $Q_i(u)$ is convex on $\mathbb{R}^+$, $i=1,\dots,n$;

4) $\eta_j^*$ is a fixed point of the mapping $y=Q_j(u)$, i.e. $Q_j(\eta_j^*)=\eta_j^*$, $j=1,\dots,n$;

5) the equation $Q_j(u)=\varepsilon^2 u$ has a positive root $\xi_j^*$, $ j=1,\dots,n$, where $\varepsilon:= \min \{\varepsilon_1, \dots, \varepsilon_n\}\in (0,1)$.

The study of system (1.1), in addition to theoretical interest, is also of applied interest in various areas of natural science. In particular, such systems of non-linear integral equations arise in the dynamic theory of $p$-adic open-closed strings, in the mathematical theory of the spatio-temporal (geographical) spread of an epidemic, in the kinetic theory of gases, and in the theory of radiative transfer in spectral lines (see [2]–[9]).

In the particular case $\lambda_{ij}\equiv1$ and $ \mu_j\equiv1$, $i,j=1,\dots,n$, system (1.1) was studied in detail in [10], which is mainly devoted to existence of a sign alternating bounded monotone solution, as well as to the study of the asymptotic behaviour of the constructed solutions at $\pm\infty$.

The corresponding scalar integral equation $(n=1)$ for various constraints on the kernel and on the non-linearity was studied in [2]–[7] and [10]–[14].

In the present paper, under conditions a)–c), A)–C), 1)–5), we prove the existence of a bounded solution for system (1.1) and study its asymptotic behaviour at $\pm\infty$. In particular, it will be proved that system (1.1) has a non-negative non-trivial bounded solution, and $\eta_j^*-\varphi_j\in L_1^0(\mathbb{R})$, $j=1,\dots,n$, where $L_1^0(\mathbb{R})$ is the space of Lebesgue summable functions on $\mathbb{R}$ with zero limit at $\pm\infty$. Specific examples of functions $\{\mu _j (u)\}_{j=1}^n$, $\{\lambda_{ij}(u)\}_{i,j=1}^{n\times n}$, $\{K_{ij}(x)\}_{i,j=1}^{n\times n}$ and $\{Q _i (u)\}_{i=1}^n$ will be given at the end of the paper to illustrate the importance of the results obtained.

§ 2. Notation and auxiliary facts

Along with system (1.1) we consider an auxiliary system of non-linear integral equations with sum–difference kernel on the half-axis

$$ \begin{equation} \begin{gathered} \, Q_i(\psi_i(x))=\sum_{j=1}^n\mu_j(x)\int_{0}^\infty (K_{ij}(x-t)-K_{ij}(x+t))\psi_j(t)\, dt, \\ i=1,\dots, n,\quad x\in \mathbb{R}^+, \end{gathered} \end{equation} \tag{2.1} $$
for the unknown vector function $\psi(x)=(\psi_1(x),\dots,\psi_n(x))^\top$.

For system (2.1), we introduce successive approximations

$$ \begin{equation} \begin{gathered} \, Q_i(\psi_i^{(m+1)}(x))=\sum_{j=1}^n\mu_j(x)\int_{0}^\infty (K_{ij}(x-t)-K_{ij}(x+t))\psi_j^{(m)}(t)\, dt, \\ \psi_i^{(0)}(x)=\eta_i^*,\quad i=1,\dots, n,\quad x\in \mathbb{R}^+,\quad m=0,1,\dots\,. \end{gathered} \end{equation} \tag{2.2} $$
It can be easily proved by induction that
$$ \begin{equation} \psi_i^{(m)}(x)\geqslant0,\quad \psi_i^{(m)}\in C(\mathbb{R}^+),\quad x\in \mathbb{R}^+, \quad i=1,\dots,n,\quad m=0,1,\dots, \end{equation} \tag{2.3} $$
$$ \begin{equation} \psi_i^{(m)}(x)\text{ is monotone decreasing in } m,\qquad i=1,\dots,n,\quad x\in \mathbb{R}^+. \end{equation} \tag{2.4} $$
According to [10],
$$ \begin{equation} \begin{gathered} \, \sum_{j=1}^n \eta_j^*\int_{0}^\infty (K_{ij}(x-t)-K_{ij}(x+t))(1-e^{-p^* t})\, dt\geqslant \eta_i^*\varepsilon_i (1-e^{-p^* x}), \\ x\in \mathbb{R}^+,\quad i=1,\dots,n, \end{gathered} \end{equation} \tag{2.5} $$
where $p^*=\min \{p_1,\dots, p_n\}$, while the numbers $\{p_i\}_{i=1}^n$ are uniquely defined from the characteristic equations
$$ \begin{equation} \sum_{j=1}^n\eta_j^*\int_{0}^\infty K_{ij}(t)e^{-pt}\, dt= \frac{\varepsilon_i\eta_i^*}{2},\qquad i=1,\dots,n. \end{equation} \tag{2.6} $$
Using estimate (2.5) and since the functions $\{Q_i(u)\}_{i=1}^n$ are convex on $\mathbb{R}^+$, it can be easily proved by induction in $m $ that the successive approximations $\{\psi_i^{(m)}(x)\}_{m=0}^\infty$, $i=1,\dots,n$, obey the lower estimate
$$ \begin{equation} \psi_i^{(m)}(x)\,{\geqslant} \min_{1\leqslant j\leqslant n}\biggl( \frac{\xi_j^*}{\eta_j^*} \biggr)\eta_i^* (1-e^{-p^* x}),\quad x \in \mathbb{R}^+,\quad i\,{=}\,1,\dots,n,\quad m\,{=}\,0,1,\dots\,. \end{equation} \tag{2.7} $$
From the above properties (2.3), (2.4) and (2.7), it follows that the sequence of continuous vector functions $\psi^{(m)}(x)=(\psi_1^{(m)}(x),\dots, \psi_n^{(m)}(x))^\top$, $m=0,1,\dots$, has a pointwise limit $\lim_{m\to \infty}\psi^{(m)}(x)=\psi(x)$, $x\in\mathbb{R}^+$, as $m\to \infty$, and that the limit vector function $\psi(x)=(\psi_1(x),\dots, \psi_n(x))^\top$ satisfies system (2.1) by the Beppo Levi theorem (see [15]). From (2.4) and (2.7), we also have the double estimate for the limit vector function $\psi(x)$
$$ \begin{equation} \min_{1\leqslant j\leqslant n}\biggl( \frac{\xi_j^*}{\eta_j^*} \biggr)\eta_i^* (1-e^{-p^* x})\leqslant \psi_i(x)\leqslant \eta_i^*,\qquad x\in \mathbb{R}^+,\quad i=1,\dots,n. \end{equation} \tag{2.8} $$
Using the technique developed in [10] and arguing as [10], it can be shown that
$$ \begin{equation} \eta_i^*-\psi_i\in L_1(0,+\infty),\qquad i=1,\dots,n. \end{equation} \tag{2.9} $$
Taking into account inequality (2.8), conditions c), C), 1), 2), and the fact that the convolution of a summable function with a bounded function is a continuous function (see [16]), from (2.1) we also find that
$$ \begin{equation} \psi_i\in C(\mathbb{R}^+),\qquad i=1,\dots,n. \end{equation} \tag{2.10} $$
Therefore, the convergence of the sequence of vector functions $\{\psi_{m}(x)\}_{m=0}^\infty$ to $\psi(x)=(\psi_1(x),\dots, \psi_n(x))^\top$ is uniform on each compact subset of $\mathbb{R}^+$.

A direct verification shows that the functions

$$ \begin{equation} f_i(x)= \begin{cases} \psi_i(x) &\text{if}\quad x\in \mathbb{R}^+, \\ -\psi_i(-x) &\text{if}\quad x\in (-\infty,0), \end{cases} \qquad i=1,\dots,n, \end{equation} \tag{2.11} $$
satisfy the following system of non-linear integral equations on the whole real axis:
$$ \begin{equation} Q_i(f_i(x))=\sum_{j=1}^n\mu_j(|x|)\int_{\mathbb{R}} K_{ij}(x-t) f_j(t)\, dt,\qquad i=1,\dots,n,\quad x\in\mathbb{R}. \end{equation} \tag{2.12} $$
From (2.8)(2.11) it follows directly that
$$ \begin{equation} f_i(-x)=-f_i(x),\qquad x\in \mathbb{R}^+,\quad f_i\in C(\mathbb{R}),\quad i=1,\dots,n, \end{equation} \tag{2.13} $$
$$ \begin{equation} \min_{1\leqslant j\leqslant n}\biggl( \frac{\xi_j^*}{\eta_j^*} \biggr)\eta_i^* (1-e^{-p^* x})\leqslant f_i(x)\leqslant \eta_i^*,\qquad x\in \mathbb{R}^+,\quad i=1,\dots,n, \end{equation} \tag{2.14} $$
$$ \begin{equation} \eta_i^*\pm f_i\in L_1(\mathbb{R}^{\mp}),\qquad \mathbb{R}^-:=\mathbb{R}\setminus\mathbb{R}^+, \quad i=1,\dots,n. \end{equation} \tag{2.15} $$
We set
$$ \begin{equation} b_{ij}:=2\sup_{x\in\mathbb{R}}(K_{ij}(x))\int_0^\infty (\lambda_{ij}(t)-1)\, dt,\qquad i,j=1,\dots,n. \end{equation} \tag{2.16} $$
Consider the auxiliary system of non-linear algebraic equations
$$ \begin{equation} Q_i(\xi_i)= \sum_{j=1}^n (a_{ij}+b_{ij})\xi_j,\qquad i=1,\dots,n, \end{equation} \tag{2.17} $$
for the sought-for vector $\xi:=(\xi_1,\dots,\xi_n)^\top$, where $a_{ij}$ and $b_{ij}$, $i,j=1,\dots,n$, are given by (1.3) and (2.16), respectively.

Our next aim is to prove the existence and uniqueness of a solution to system (2.17).

The following result holds.

Lemma 1. Let the elements of the matrix $A=(a_{ij})_{i,j=1}^{n\times n}$ be positive, $r(A)=1$, and the elements of the matrix $B=(b_{ij})_{i,j=1}^{n\times n}$ be non-negative. Assume that the functions $\{Q_i(u)\}_{i=1}^n$ satisfy conditions 1)–5). Then system (2.17) has a componentwise positive solution $\xi:=(\xi_1,\dots,\xi_n)^\top$. In addition, $\xi_i\geqslant \eta_i^*$, $i=1,\dots,n$.

Proof. Consider the successive approximations
$$ \begin{equation} Q_i(\xi_i^{(p+1)})= \sum_{j=1}^n (a_{ij}+b_{ij})\xi_j^{(p)},\quad i=1,\dots,n, \quad \xi_i^{(0)}=\eta_i^*,\quad p=0,1,\dots\,. \end{equation} \tag{2.18} $$
Let us prove by induction in $p$ that
$$ \begin{equation} \xi_i^{(p)}\uparrow \text{in}\,\, p,\qquad i=1,\dots,n. \end{equation} \tag{2.19} $$
Using (1.2), since the elements of the matrix $B$ are non-negative, and since the functions $\{Q_i(u)\}_{i=1}^n$ are monotone in $u$, it follows from 4) and (2.18) that
$$ \begin{equation*} \begin{aligned} \, &Q_i(\xi_i^{(1)})= \sum_{j=1}^n (a_{ij}+b_{ij})\eta_j^*\geqslant \sum_{j=1}^n a_{ij}\eta_j^*=\eta_i^*=Q_i(\eta_i^*) \\ &\qquad \Longrightarrow \quad \xi_i^{(1)}\geqslant \eta_i^*=\xi_i^{(0)},\qquad i=1,\dots,n. \end{aligned} \end{equation*} \notag $$
Assuming that $\xi_i^{(p)}\geqslant \xi_i^{(p-1)}$, $i=1,\dots,n$, for some natural $p$, since the elements of the matrix $A+B$ are positive and since the functions $\{Q_i(u)\}_{i=1}^n$ are monotone, it follows from (2.18) that $\xi_i^{(p+1)}\geqslant \xi_i^{(p)}$. Let us now show that there exists a number $c>1$ such that
$$ \begin{equation} \xi_i^{(p)}\leqslant c\eta_i^*,\qquad i=1,\dots,n,\quad p=0,1,\dots\,. \end{equation} \tag{2.20} $$
To this end, we first consider the characteristic functions of the set $[1,+\infty)$:
$$ \begin{equation} \chi_i(u):=\frac{Q_i(u\eta_i^*)}{u\eta_i^*}-1-\frac{\alpha_i}{\eta_i^*},\qquad i=1,\dots,n, \end{equation} \tag{2.21} $$
where
$$ \begin{equation} \alpha_i:= \sum_{j=1}^n b_{ij}\eta_j^*,\qquad i=1,\dots,n. \end{equation} \tag{2.22} $$
By condition 4), $\chi_i(1)=-\alpha_i/\eta_i^*<0$, $i=1,\dots, n$. On the other hand, since the functions $\{Q_i(u)\}_{i=1}^n$ are convex and monotone on $\mathbb{R}^+$, we have
$$ \begin{equation} \chi_i(u)\uparrow \text{ in } u \text{ on } [1,+\infty),\qquad i=1,\dots, n, \end{equation} \tag{2.23} $$
$$ \begin{equation} \chi_i(+\infty)=+\infty,\qquad i=1,\dots, n. \end{equation} \tag{2.24} $$
Therefore, for any $i\in \{1,\dots, n\}$, there exists a number $c_i>1$ such that $\chi_i(c_i)=0$. We set $c:=\max\{c_1,\dots,c_n\}$. Now by (2.23) and (2.24) we have
$$ \begin{equation} \frac{Q_i(c\eta_i^*)}{c\eta_i^*}\geqslant \frac{Q_i(c_i\eta_i^*)}{c_i\eta_i^*}=1+\frac{\alpha_i}{\eta_i^*},\qquad i=1,\dots, n. \end{equation} \tag{2.25} $$
Let us use (2.25) to prove (2.20). For $p=0$, inequality (2.20) is immediate from the definition of the zeroth approximation and since $c\geqslant c_i>1$, $i=1,\dots, n$. Suppose that $\xi_i^{(p)}\leqslant c\eta_i^*$, $i=1,\dots, n$ for some $p\in \mathbb{N}$. By (2.25), we have
$$ \begin{equation*} \begin{aligned} \, Q_i(\xi_i^{(p+1)}) &\leqslant c\biggl(\sum_{j=1}^n a_{ij}\eta_j^*+ \sum_{j=1}^n b_{ij}\eta_j^*\biggr) \\ &=c(\eta_i^*+\alpha_i)= c\eta_i^*\biggl(1+\frac{\alpha_i}{\eta_i^*}\biggr)\leqslant Q_i(c\eta_i^*),\qquad i=1,\dots, n. \end{aligned} \end{equation*} \notag $$
This immediately implies that $\xi_i^{(p+1)}\leqslant c\eta_i^*$, $i=1,\dots, n$, because $Q_i(u)\uparrow$ in $u$ on $\mathbb{R}$, $i=1,\dots, n$.

So, the sequence of vectors $\xi^{(p)}:=(\xi_1^{(p)}, \dots, \xi_n^{(p)})^\top$, $p=0,1,\dots$, converges as $p\to \infty$, i.e., $\lim_{p\to \infty}\xi_i^{(p)}=\xi_i$, $i=1,\dots,n$, and, by continuity of the functions $\{Q_i(u)\}_{i=1}^n$, the limit vector is a solution to system (2.17), and $\xi_i\in [\eta_i^*, c\eta_i^*]$, $i=1,\dots,n$. Lemma 1 is proved.

We have the following result on uniqueness of the solution to system (2.17).

Lemma 2. Under the conditions of Lemma 1, if the matrix $A+B$ is symmetric, then system (2.17) has at most one solution in the class of vectors

$$ \begin{equation*} \mathfrak{M}:=\{\mathrm{x}=(x_1,\dots,x_n)^\top \colon x_i\geqslant0,\,i=1,\dots,n,\, \mathrm{x}\neq0 \}. \end{equation*} \notag $$

Proof. Assume on the contrary that system (2.17) has two solutions $\xi, \widetilde{\xi}\in \mathfrak{M}$. The elements of the matrix $A+B$ are positive, and hence by (2.17) we have
$$ \begin{equation} |Q_i(\xi_i)-Q_i(\widetilde{\xi}_i)|\leqslant \sum_{j=1}^n (a_{ij}+b_{ij})|\xi_j-\widetilde{\xi}_j|, \qquad i=1,\dots,n, \end{equation} \tag{2.26} $$
where $\xi=(\xi_1,\dots,\xi_n)^\top$, $\widetilde{\xi}=(\widetilde{\xi}_1,\dots,\widetilde{\xi}_n)^\top$. Multiplying both parts of each inequality in (2.26) by the corresponding $\xi_i$ and summing the resulting inequalities over $i=1,\dots,n$, we have, since the matrix $A+B$ is symmetric,
$$ \begin{equation*} \begin{aligned} \, \sum_{i=1}^n \xi_i |Q_i(\xi_i)-Q_i(\widetilde{\xi}_i)| &\leqslant \sum_{i=1}^n \xi_i \sum_{j=1}^n (a_{ij}+b_{ij})|\xi_j-\widetilde{\xi}_j| \\ &=\sum_{j=1}^n |\xi_j-\widetilde{\xi}_j|\sum_{i=1}^n (a_{ij}+b_{ij})\xi_i =\sum_{j=1}^n |\xi_j-\widetilde{\xi}_j| Q_j(\xi_j), \end{aligned} \end{equation*} \notag $$
and hence
$$ \begin{equation*} J:=\sum_{i=1}^n \bigl(\xi_i |Q_i(\xi_i)-Q_i(\widetilde{\xi}_i)|-|\xi_i-\widetilde{\xi}_i| Q_i(\xi_i)\bigr)\leqslant 0. \end{equation*} \notag $$
From 1)–5) it is immediate that
$$ \begin{equation} J=\sum_{i\in\mathcal{P}} (\xi_i |Q_i(\xi_i)-Q_i(\widetilde{\xi}_i)|-|\xi_i-\widetilde{\xi}_i| Q_i(\xi_i))\leqslant0, \end{equation} \tag{2.27} $$
where
$$ \begin{equation} \mathcal{P}:=\bigl\{i\in \{ 1,\dots,n\}\colon \xi_i\neq0,\, \xi_i\neq\widetilde{\xi}_i\bigr\}. \end{equation} \tag{2.28} $$
By definition of $\mathcal{P}$, we can write inequality (2.27) as
$$ \begin{equation} \sum_{i\in\mathcal{P}}\xi_i |\xi_i-\widetilde{\xi}_i| \biggl(\frac{|Q_i(\xi_i)-Q_i(\widetilde{\xi}_i)|}{|\xi_i-\widetilde{\xi}_i|} -\frac{Q_i(\xi_i)}{\xi_i} \biggr)\leqslant 0. \end{equation} \tag{2.29} $$
Since each of the functions $\{Q_i(u)\}_{i=1}^n$ is convex on $\mathbb{R}^+$, we have, for any $i\in\mathcal{P}$, the strict inequality
$$ \begin{equation} \frac{|Q_i(\xi_i)-Q_i(\widetilde{\xi}_i)|}{|\xi_i-\widetilde{\xi}_i|}>\frac{Q_i(\xi_i)}{\xi_i}. \end{equation} \tag{2.30} $$
But this contradicts (2.29), proving Lemma 2.

§ 3. On solvability of system (1.1)

In this section, we will show that the original system (1.1) has a non-trivial non-negative bounded solution. Note that the above auxiliary Lemmas 1, 2, as well as properties (2.13)(2.15) will play an important role in achieving our goal of finding a solution to system (2.12).

The following theorem holds.

Theorem 1. Let conditions a)–c), A)–C) and 1)–5) be met. Then the system of integral equations (1.1) has a non-negative non-trivial bounded solution.

Proof. For system (1.1), consider the following successive approximations:
$$ \begin{equation} \begin{gathered} \, Q_i(\varphi_i^{(p+1)}(x))=\sum_{j=1}^n\mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)\lambda_{ij}(|t|)\varphi_j^{(p)}(t)\, dt, \\ \varphi_i^{(0)}(x)=\xi_i,\quad x\in \mathbb{R},\quad i=1,\dots, n,\quad p=0,1,\dots, \end{gathered} \end{equation} \tag{3.1} $$
where the numbers $\{\xi_i\}_{i=1}^n$ are uniquely determined from the system of non-linear algebraic equations (2.17) (see Lemmas 1 and 2).

We prove by induction in $p$ that

i) $\varphi_i^{(p)}(x)\downarrow$ in $p$, $i=1,\dots, n$, $x\in \mathbb{R}$,

ii) $\varphi_i^{(p)}(x)\geqslant |f_i(x)|$, $i=1,\dots, n$, $p=0,1,\dots$, $x\in \mathbb{R}$, where $f(x)=(f_1(x),\dots, f_n(x))^\top$ is a solution to the auxiliary system of non-linear integral equations (2.12) satisfying (2.13)(2.15).

Let us first show that

$$ \begin{equation*} |f_i(x)|\leqslant \varphi_i^{(1)}(x)\leqslant \xi_i,\qquad i=1,\dots, n,\quad x\in \mathbb{R}. \end{equation*} \notag $$
Indeed, using (1.2), (2.16), conditions a)–c), and Lemmas 1 and 2, we have by (3.1)
$$ \begin{equation*} \begin{aligned} \, Q_i(\varphi_i^{(1)}(x)) &\leqslant \sum_{j=1}^n\xi_j\int_{-\infty}^\infty K_{ij}(x-t)\lambda_{ij}(|t|)\, dt \\ &= \sum_{j=1}^n\xi_j\int_{-\infty}^\infty K_{ij}(x-t)(\lambda_{ij}(|t|)-1)\, dt+ \sum_{j=1}^n\xi_j\int_{-\infty}^\infty K_{ij}(x-t)\, dt \\ &\leqslant \sum_{j=1}^n\xi_j \sup_{x\in \mathbb{R}}(K_{ij}(x)) \int_{-\infty}^\infty (\lambda_{ij}(|t|)-1)\, dt+ \sum_{j=1}^n a_{ij}\xi_j \\ &= \sum_{j=1}^n\xi_j \biggl(2\sup_{x\in \mathbb{R}}(K_{ij}(x)) \int_{0}^\infty (\lambda_{ij}(t)-1)\, dt +a_{ij} \biggr) \\ &=\sum_{j=1}^n(a_{ij}+b_{ij})\xi_j=Q_i(\xi_i),\qquad i=1,\dots, n. \end{aligned} \end{equation*} \notag $$
Hence, since $\{Q_i(u)\}_{i=1}^n$ are monotone,
$$ \begin{equation*} \varphi_i^{(1)}(x)\leqslant \xi_i=\varphi_i^{(0)}(x),\qquad i=1,\dots, n,\quad x\in \mathbb{R}. \end{equation*} \notag $$
Now from Lemma 1, (2.13), and (2.14) it is immediate that $\xi_i\geqslant |f_i(x)|$, $ i=1,\dots, n$, $ x\in \mathbb{R}$.

Using again conditions a)–c), (2.12), and the monotonicity and oddness of the functions $\{Q_i(u)\}_{i=1}^n$, from (3.1) we have

$$ \begin{equation*} \begin{aligned} \, &Q_i(\varphi_i^{(1)}(x)) \geqslant \sum_{j=1}^n\mu_j(|x|) \xi_j\int_{-\infty}^\infty K_{ij}(x-t)\, dt \\ &\qquad\geqslant \sum_{j=1}^n\mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)|f_j(t)|\,dt \geqslant \sum_{j=1}^n\mu_j(|x|)\biggl|\int_{-\infty}^\infty K_{ij}(x-t)f_j(t)\, dt\biggr| \\ &\qquad\geqslant \biggl|\sum_{j=1}^n\mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)f_j(t)\, dt\biggr| = |Q_i(f_i(x))|=Q_i(|f_i(x)|) \\ &\qquad\Longrightarrow\quad \varphi_i^{(1)}(x)\geqslant |f_i(x)|,\qquad i=1,\dots, n,\quad x\in \mathbb{R}. \end{aligned} \end{equation*} \notag $$
Assuming that $\varphi_i^{(p)}(x)\leqslant \varphi_i^{(p-1)}(x)$, $\varphi_i^{(p)}(x)\geqslant |f_i(x)|$, $ i=1,\dots, n$, $ x\in \mathbb{R}$, for some $p\in \mathbb{N}$, since the functions $\{Q_i(u)\}_{i=1}^n$ are monotone and odd, and since the kernels $K_{ij}(x)$, $i,j=1,\dots, n$, are positive, a similar argument with the use of (3.1) shows that
$$ \begin{equation*} \varphi_i^{(p+1)}(x)\leqslant \varphi_i^{(p)}(x),\quad \varphi_i^{(p+1)}(x)\geqslant |f_i(x)|,\qquad i=1,\dots, n,\quad x\in \mathbb{R}. \end{equation*} \notag $$
Using the continuity of the function $\{\mu_j(u)\}_{j=1}^n$, $\{Q_j(u)\}_{j=1}^n$, condition $1)$, and since the convolution of a summable function with a bounded function is a continuous function, it can be easily shown by induction in $p$ that $\varphi_i^{(p)}\in C(\mathbb{R})$, $ i=1,\dots, n$, $ p=0,1,\dots$ . Now from i) and ii) we conclude that the sequence of continuous vector functions $\varphi^{(p)}(x)=(\varphi_1^{(p)}(x),\dots, \varphi_n^{(p)}(x))^\top$, $p=0,1,\dots$, has a pointwise limit as $p\to \infty$, i.e., $\lim_{p\to \infty}\varphi_i^{(p)}(x)=\varphi_i(x)$, $ i=1,\dots, n$, $ x\in \mathbb{R}$. By the Beppo Levi theorem, the limit vector function $\varphi(x)=(\varphi_1(x),\dots, \varphi_n(x))^\top$ satisfies system (1.1). From i), ii) it also follows that
$$ \begin{equation} |f_i(x)|\leqslant \varphi_i(x)\leqslant \xi_i,\qquad i=1,\dots, n,\quad x\in \mathbb{R}. \end{equation} \tag{3.2} $$
In view of (2.13)(2.15), from (3.2) we have $\varphi_i(x)\geqslant0$, $\varphi_i(x)\not\equiv0$, $ i=1,\dots, n$, $ x\in \mathbb{R}$. This completes the proof of Theorem 1.

Remark 1. From estimate (3.2)), using (2.14) we also have

$$ \begin{equation} \varphi_i(x)\geqslant \min_{1\leqslant j\leqslant n} \biggl(\frac{\xi_j^*}{\eta_j^*}\biggr)\eta_i^*(1-e^{-p^*|x|})>0,\qquad x\in \mathbb{R}\setminus\{0\}, \quad i=1,\dots, n. \end{equation} \tag{3.3} $$

§ 4. Asymptotic behaviour of the solution at $\pm \infty$

Let $\varphi(x)=(\varphi_1(x),\dots,\varphi_n(x))^\top$ be any solution of system (1.1) satisfying the double inequality

$$ \begin{equation} \min_{1\leqslant j\leqslant n}\biggl(\frac{\xi_j^*}{\eta_j^*}\biggr)\eta_i^*(1-e^{-p^*|x|})\leqslant \varphi_i(x)\leqslant \xi_i,\qquad i=1,\dots, n,\quad x\in \mathbb{R}. \end{equation} \tag{4.1} $$
The existence of such a solution is secured Theorem 1.

Let us show that in this case $\eta_i^*-\varphi_i\in L_1(\mathbb{R}), i=1,\dots, n$.

The following result holds.

Theorem 2. Let the conditions of Theorem 1 be met. Then any solution $\varphi(x)=(\varphi_1(x),\dots,\varphi_n(x))^\top$ of system (1.1) satisfying the double inequality (4.1) has the integral asymptotics $\eta_i^*-\varphi_i\in L_1(\mathbb{R})$, $i=1,\dots, n$.

Proof. We first claim that $\varphi_i\in C(\mathbb{R})$, $i=1,\dots,n$. Indeed, this follows from the continuity of the convolution of a summable function with a bounded function in view of the fact that $\mu_j\in C(\mathbb{R}^+)$, $Q_j\in C(\mathbb{R})$ and $Q_j(u)\uparrow$ in $u$ on $\mathbb{R}$, $j=1,\dots,n$.

Let us now use (1.2) to estimate the difference

$$ \begin{equation*} \begin{aligned} \, &|\eta_i^*-Q_i(\varphi_i(x))|=\biggl|\sum_{j=1}^n a_{ij}\eta_j^*-\sum_{j=1}^n\mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)\lambda_{ij}(|t|)\varphi_j(t)\, dt\biggr| \\ &\qquad =\biggl|\sum_{j=1}^n\biggl(\eta_j^*\int_{-\infty}^\infty K_{ij}(x-t)\, dt- \mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)\lambda_{ij}(|t|)\varphi_j(t)\, dt\biggr)\biggr| \\ &\qquad =\biggl|\sum_{j=1}^n \eta_j^*\mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)\, dt+\sum_{j=1}^n a_{ij}\eta_j^*(1-\mu_j(|x|)) \\ &\qquad \qquad -\sum_{j=1}^n\mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)\lambda_{ij}(|t|)\varphi_j(t)\, dt\biggr| \\ &\qquad\leqslant \sum_{j=1}^n a_{ij}\eta_j^*(1-\mu_j(|x|))+ \biggl| \sum_{j=1}^n \eta_j^*\mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)\, dt \\ &\qquad\qquad -\sum_{j=1}^n\mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)\varphi_j(t)\, dt \\ &\qquad\qquad- \sum_{j=1}^n\mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t)(\lambda_{ij}(|t|)-1)\varphi_j(t)\, dt\biggr| \\ &\qquad\leqslant \sum_{j=1}^n a_{ij}\eta_j^*(1-\mu_j(|x|))+ \sum_{j=1}^n\xi_j\int_{-\infty}^\infty K_{ij}(x-t)(\lambda_{ij}(|t|)-1)\, dt \\ &\qquad\qquad+ \biggl|\sum_{j=1}^n \mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t) (\eta_j^*-\varphi_j(t))\, dt\biggr| \\ &\qquad\leqslant \sum_{j=1}^n a_{ij}\eta_j^*(1-\mu_j(|x|))+ \sum_{j=1}^n\xi_j\int_{-\infty}^\infty K_{ij}(x-t)(\lambda_{ij}(|t|)-1)\, dt \\ &\qquad\qquad+ \sum_{j=1}^n \mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t) |\eta_j^*-\varphi_j(t)|\, dt \\ &\qquad=g_i(x)+\sum_{j=1}^n \mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t) |\eta_j^*-\varphi_j(t)|\, dt, \\ &\qquad\qquad\qquad\qquad\qquad\qquad i=1,\dots, n,\quad x\in \mathbb{R}, \end{aligned} \end{equation*} \notag $$
where
$$ \begin{equation} \begin{gathered} \, g_i(x):=\sum_{j=1}^n a_{ij}\eta_j^*(1-\mu_j(|x|))+ \sum_{j=1}^n\xi_j\int_{-\infty}^\infty K_{ij}(x-t)(\lambda_{ij}(|t|)-1)\, dt, \\ i=1,\dots, n,\quad x\in \mathbb{R}. \end{gathered} \end{equation} \tag{4.2} $$
From conditions c) and C) and using the Fubini theorem (see [15]) we obtain
$$ \begin{equation} g_i\in L_1(\mathbb{R}),\qquad i=1,\dots, n. \end{equation} \tag{4.3} $$
Thus we have the upper estimate
$$ \begin{equation} \begin{gathered} \, |\eta_i^*-Q_i(\varphi_i(x))|\leqslant g_i(x)+\sum_{j=1}^n \mu_j(|x|)\int_{-\infty}^\infty K_{ij}(x-t) |\eta_j^*-\varphi_j(t)|\, dt, \\ i=1,\dots, n,\quad x\in \mathbb{R}. \end{gathered} \end{equation} \tag{4.4} $$
Let $R>1$ be an arbitrary number. Consider the measurable sets
$$ \begin{equation} \begin{aligned} \, E_R^i &:=\{x\in [1,R]\colon \varphi_i(x)\leqslant \eta_i^*\}, \\ \widetilde{E}_R^i &:=\{x\in [1,R]\colon \varphi_i(x)> \eta_i^*\}, \end{aligned} \qquad i=1,\dots, n, \end{equation} \tag{4.5} $$
where $\varphi(x)=(\varphi_1(x),\dots,\varphi_n(x))^\top$ is a solution to system (1.1) satisfying (4.1).

We integrate both parts of inequality (4.4) with respect to $x$ from $1$ to $R$, take into account (4.1), and use the following easily verifiable estimates for $\{\varphi_i(x)\}_{i=1}^n$ (they follow from the left-hand side of inequalities (4.1) since each of the functions $\{Q_i(u)\}_{i=1}^n$ is convex on $\mathbb{R}^+$):

$\bullet$ for $x\in E_R^i$

$$ \begin{equation*} \eta_i^*-Q_i(\varphi_i(x))\geqslant \alpha_i(\eta_i^*-\varphi_i(x)), \end{equation*} \notag $$

$\bullet$ for $x\in \widetilde{E}_R^i$

$$ \begin{equation*} Q_i(\varphi_i(x))-\eta_i^*\geqslant \beta_i(\varphi_i(x)-\eta_i^*),\quad i=1,\dots, n, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{gathered} \, \alpha_i:= \frac{\eta_i^*-Q_i(\omega_i)}{\eta_i^*-\omega_i},\qquad \omega_i:=\min_{1\leqslant j\leqslant n}\biggl(\frac{\xi_j^*}{\eta_j^*}\biggr)\eta_i^*(1-e^{-p^*}), \\ \beta_i:=\frac{2(\eta_i^*-Q_i(\eta_i^*/2))}{\eta_i^*},\qquad i=1,\dots, n. \end{gathered} \end{equation*} \notag $$
Now from (4.4) we have
$$ \begin{equation*} \begin{aligned} \, &\alpha_i\int_{E_R^i}\bigl(\eta_i^*-\varphi_i(x)\bigr)\, dx+ \beta_i\int_{\widetilde{E}_R^i}\bigl(\varphi_i(x)-\eta_i^*\bigr)\, dx \\ &\ \leqslant \int_{E_R^i}\bigl(\eta_i^*-Q_i(\varphi_i(x))\bigr)\, dx+ \int_{\widetilde{E}_R^i}\bigl(Q_i(\varphi_i(x))-\eta_i^*\bigr)\, dx =\int_1^R|\eta_i^*-Q_i(\varphi_i(x))|\, dx \\ &\ \leqslant \int_1^R g_i(x)\, dx+ \sum_{j=1}^n \int_1^R \int_{-\infty}^{\infty} K_{ij}(x-t)|\eta_j^*-\varphi_j(t)|\, dt\, dx \\ &\ \leqslant \int_1^\infty g_i(x)\, dx+ \sum_{j=1}^n (\eta_j^*+\xi_j) \int_1^R \int_{-\infty}^1 K_{ij}(x-t)\, dt\, dx \\ &\ \qquad+\sum_{j=1}^n (\eta_j^*+\xi_j) \int_1^R \int_R^{\infty} K_{ij}(x-t)\, dt\, dx \\ &\ \qquad + \sum_{j=1}^n \int_1^R \int_1^R K_{ij}(x-t)|\eta_j^*-\varphi_j(t)|\, dt\, dx \\ &\ \leqslant \int_1^\infty g_i(x)\, dx+ \sum_{j=1}^n (\eta_j^*+\xi_j) \int_1^R \int_{x-1}^{\infty} K_{ij}(y)\, dy\, dx \\ &\ \qquad + \sum_{j=1}^n (\eta_j^*+\xi_j) \int_1^R \int_{-\infty}^{x-R} K_{ij}(y)\, dy\, dx \\ &\ \qquad+\sum_{j=1}^n \int_1^R |\eta_j^*-\varphi_j(t)|\biggl( \int_{-\infty}^{\infty} K_{ij}(x-t)\, dx\biggr)\, dt \\ &\ \leqslant \int_1^\infty g_i(x)\, dx+ \sum_{j=1}^n (\eta_j^*+\xi_j) \int_1^\infty \int_{x-1}^{\infty} K_{ij}(y)\, dy\, dx \\ &\ \qquad + \sum_{j=1}^n (\eta_j^*+\xi_j) \int_0^R \int_{-\infty}^{x-R} K_{ij}(y)\, dy\, dx + \sum_{j=1}^n a_{ij} \int_1^R |\eta_j^*-\varphi_j(t)|\, dt \\ &\ =\int_1^\infty g_i(x)\, dx+\sum_{j=1}^n (\eta_j^*+\xi_j)\int_0^\infty xK_{ij}(x)\, dx \\ &\ \qquad +\sum_{j=1}^n (\eta_j^*+\xi_j)\int_{-R}^0\int_{-\infty}^x K_{ij}(z)\, dz\, dx+ \sum_{j=1}^n a_{ij} \int_1^R |\eta_j^*-\varphi_j(t)|\, dt \\ &\ \leqslant \int_1^\infty g_i(x)\, dx+\sum_{j=1}^n (\eta_j^*+\xi_j)\int_0^\infty xK_{ij}(x)\, dx \\ &\ \qquad + \sum_{j=1}^n (\eta_j^*+\xi_j)\int_{-\infty}^0 \int_{-\infty}^x K_{ij}(z)\, dz \, dx + \sum_{j=1}^n a_{ij} \int_1^R |\eta_j^*-\varphi_j(t)|\, dt \\ &\ =\int_1^\infty g_i(x)\, dx+ \sum_{j=1}^n (\eta_j^*+\xi_j)\int_{-\infty}^{\infty} |x|K_{ij}(x)\, dx \\ &\ \qquad + \sum_{j=1}^n a_{ij} \int_1^R |\eta_j^*-\varphi_j(t)|\, dt,\qquad i=1,\dots,n. \end{aligned} \end{equation*} \notag $$
So, we have the inequality
$$ \begin{equation} \begin{aligned} \, &\alpha_i\int_{E_R^i}(\eta_i^*-\varphi_i(x))\, dx+ \beta_i\int_{\widetilde{E}_R^i}(\varphi_i(x)-\eta_i^*)\, dx \nonumber \\ &\ \leqslant \int_1^\infty g_i(x)\, dx+ \sum_{j=1}^n m_{ij}(\eta_j^*+\xi_j) +\sum_{j=1}^n a_{ij} \int_1^R |\eta_j^*-\varphi_j(t)|\, dt,\qquad i=1,\dots,n, \end{aligned} \end{equation} \tag{4.6} $$
where $m_{ij}:= \int_{-\infty}^{\infty} |x|K_{ij}(x)\, dx,\quad i,j=1,\dots,n$.

Multiplying both parts of each inequality in (4.6) by the corresponding $\eta_i^*$, summing over all $i=1,\dots,n$, taking into account the symmetry of the matrix $A$, and employing (1.2), we obtain

$$ \begin{equation*} \begin{aligned} \, &\sum_{i=1}^n \eta_i^* \biggl( \alpha_i\int_{E_R^i}(\eta_i^*-\varphi_i(x))\, dx+ \beta_i\int_{\widetilde{E}_R^i}(\varphi_i(x)-\eta_i^*)\, dx \biggr) \\ &\leqslant \sum_{i=1}^n \eta_i^* \!\int_1^\infty g_i(x)\, dx\,{+} \sum_{i=1}^n \eta_i^*\sum_{j=1}^n m_{ij}(\eta_j^*\,{+}\,\xi_j)\,{+}\sum_{j=1}^n \int_1^R |\eta_j^*\,{-}\,\varphi_j(t)|\, dt \biggl(\sum_{i=1}^n a_{ji}\eta_i^*\biggr) \\ &=\sum_{i=1}^n \eta_i^*\!\int_1^\infty g_i(x)\, dx+ \sum_{i=1}^n \eta_i^*\sum_{j=1}^n m_{ij}(\eta_j^*+\xi_j)+\sum_{j=1}^n \eta_j^* \int_1^R |\eta_j^*-\varphi_j(t)|\, dt, \end{aligned} \end{equation*} \notag $$
from which it immediately follows that
$$ \begin{equation} \begin{aligned} \, &\sum_{i=1}^n \eta_i^* \biggl( (\alpha_i-1)\int_{E_R^i}(\eta_i^*-\varphi_i(x))\, dx+ (\beta_i-1)\int_{\widetilde{E}_R^i}(\varphi_i(x)-\eta_i^*)\, dx \biggr) \nonumber \\ &\qquad\leqslant \sum_{i=1}^n \eta_i^*\int_1^\infty g_i(x)\, dx+ \sum_{i=1}^n \eta_i^*\sum_{j=1}^n m_{ij}(\eta_j^*+\xi_j). \end{aligned} \end{equation} \tag{4.7} $$
Since the functions $\{Q_i(u)\}_{i=1}^n$ are convex on $\mathbb{R}^+$, we have
$$ \begin{equation*} \alpha_i-1=\frac{\omega_i-Q_i(\omega_i)}{\eta_i^*-\omega_i}>0,\quad \beta_i-1=\frac{\eta_i^*-2Q_i(\eta_i^*/2)}{\eta_i^*}>0,\qquad i=1,\dots,n. \end{equation*} \notag $$
Now from (4.7) we have the estimate
$$ \begin{equation} \sum_{i=1}^n \eta_i^* \int_1^R |\eta_i^*-\varphi_i(x)|\, dx\leqslant \frac{\sum_{i=1}^n \eta_i^*\int_1^\infty g_i(x)\, dx+ \sum_{i=1}^n \eta_i^*\sum_{j=1}^n m_{ij}(\eta_j^*+\xi_j)}{\min\{\min_{1\leqslant i\leqslant n}(\alpha_i-1), \min_{1\leqslant i\leqslant n}(\beta_i-1)\}}. \end{equation} \tag{4.8} $$
Since $\eta_i^*\geqslant 1$, $i=1,\dots,n$, it follows from (4.8) that
$$ \begin{equation} \begin{aligned} \, &\int_1^R |\eta_i^*-\varphi_i(x)|\, dx \nonumber \\ &\ \leqslant \frac{\sum_{i=1}^n \eta_i^*\int_1^\infty g_i(x)\, dx+ \sum_{i=1}^n \eta_i^*\sum_{j=1}^n m_{ij}(\eta_j^*+\xi_j)}{\min\{\min_{1\leqslant i\leqslant n}(\alpha_i-1), \min_{1\leqslant i\leqslant n}(\beta_i-1)\}},\qquad i=1,\dots,n. \end{aligned} \end{equation} \tag{4.9} $$
Making $R\to +\infty$ in (4.9), we find that $\eta_i^*-\varphi_i\in L_1(1,+\infty)$ and
$$ \begin{equation*} \begin{aligned} \, &\int_1^\infty |\eta_i^*-\varphi_i(x)|\, dx \\ &\ \leqslant \frac{\sum_{i=1}^n \eta_i^*\int_1^\infty g_i(x)\, dx+ \sum_{i=1}^n \eta_i^*\sum_{j=1}^n m_{ij}(\eta_j^*+\xi_j)}{\min\{\min_{1\leqslant i\leqslant n}(\alpha_i-1), \min_{1\leqslant i\leqslant n}(\beta_i-1)\}},\qquad i=1,\dots,n. \end{aligned} \end{equation*} \notag $$
A similar argument shows that $\eta_i^*-\varphi_i\in L_1(-\infty,-1)$, $i=1,\dots,n$. Since $\varphi_i\in C(\mathbb{R})$, we have $\eta_i^*-\varphi_i\in L_1(-1,1)$, $i=1,\dots,n$. As a result, $\eta_i^*-\varphi_i\in L_1(\mathbb{R})$, $i=1,\dots,n$, proving Theorem 2.

Remark 2. It easily follows that

$$ \begin{equation} \lim_{x\to \pm\infty}\varphi_i(x)=\eta_i^*,\qquad i=1,\dots,n. \end{equation} \tag{4.10} $$
Indeed, by the well known limit formula for the convolution operation (see [17]),
$$ \begin{equation} \begin{aligned} \, \lim_{x\to \pm\infty}g_i(x) &=\sum_{j=1}^n a_{ij}\eta_j^* \lim_{x\to \pm\infty}(1-\mu_j(|x|)) \nonumber \\ &\qquad +\sum_{j=1}^n \xi_j\lim_{x\to \pm\infty} \int_{-\infty}^{\infty} K_{ij}(x-t)(\lambda_{ij}(|t|)-1)\, dt \nonumber \\ &=\sum_{j=1}^n a_{ij} \xi_j\lim_{t\to \pm\infty}(\lambda_{ij}(|t|)-1)=0,\qquad i=1,\dots,n. \end{aligned} \end{equation} \tag{4.11} $$
Second, if $f, g\in L_1(\mathbb{R})\cap M(\mathbb{R})$, then
$$ \begin{equation} \lim_{x\to \pm\infty}(f*g)(x)=\lim_{x\to \pm\infty}\int_{-\infty}^{\infty} f(x-t)g(t)\, dt=0 \end{equation} \tag{4.12} $$
(see [18]). Now using (4.12), Theorem 2, condition C), and estimate (4.1), we have
$$ \begin{equation*} \lim_{x\to \pm\infty}\sum_{j=1}^n \mu_j(|x|)\int_{-\infty}^{\infty} K_{ij}(x-t) |\eta_j^*-\varphi_j(t)|\, dt=0,\qquad i=1,\dots,n. \end{equation*} \notag $$
Therefore, $\lim_{x\to\pm\infty}|\eta_i^*-Q_i(\varphi_i(x))|\,{=}\,0$, $i\,{=}\,1,\dots,n$. Since $Q_i(\eta_i^*)\,{=}\,\eta_i^*$, $Q_i\in C(\mathbb{R})$, and since $Q_i(u)\uparrow$ in $u$ on $\mathbb{R}$, $i=1,\dots,n$, from the last limit relation we obtain $\lim_{x\to\pm\infty}\varphi_i(x)=\eta_i^*$, $i=1,\dots,n$.

Remark 3. Unfortunately, the question of uniqueness of the solution to system (1.1) in the class of bounded functions on $\mathbb{R}$ is still open.

§ 5. Examples

In this concluding section of this paper, we will give some examples of applied nature of functions $\{K_{ij}(x)\}_{i,j=1}^{n\times n}$, $\{\mu_{j}(t)\}_{j=1}^{n}$, $\{\lambda_{ij}(t)\}_{i,j=1}^{n\times n}$ and $\{Q_{i}(u)\}_{i=1}^{n}$ satisfying all the conditions of the above results.

Examples of kernels $\{K_{ij}(x)\}_{i,j=1}^{n\times n}$. In the dynamical theory of $p$-adic open-closed strings, for scalar field of tachyons, and in mathematical theory of the spatio-temporal pandemic spread, there appears system (1.1) with kernels of the form (see [2]–[6])

$$ \begin{equation} K_{ij}(x)=\frac{a_{ij}}{\sqrt{\pi\delta}}\, e^{-x^2/\delta},\qquad x\in \mathbb{R},\quad i,j=1,\dots,n, \end{equation} \tag{5.1} $$
where $\delta>0$ is an arbitrary numeric parameter, and $a_{ij}>0$ are elements of a matrix $A$ of unit spectral radius.

In the kinetic theory of gases and in the theory of radiative transfer, the kernels $\{K_{ij}(x)\}_{i,j=1}^{n\times n}$ have the form (see [7]–[9])

$$ \begin{equation} K_{ij}(x)=\int_a^b e^{-|x|s}G_{ij}(s)\, ds,\qquad x\in \mathbb{R},\quad i,j=1,\dots,n, \end{equation} \tag{5.2} $$
where $\{G_{ij}(s)\}_{i,j=1}^{n\times n}$ are continuous positive functions on $[a,b)$, $ 0<a<b\leqslant+\infty$, the spectral radius of the matrix
$$ \begin{equation*} G:=\biggl(2\int_a^b \frac{G_{ij}(s)}{s}\, ds\biggr)_{i,j=1}^{n\times n} \end{equation*} \notag $$
is 1, and $G_{ij}(s)=G_{ji}(s)$, $s\in[a,b)$.

It is easy to check that conditions A)–C) for kernels (5.1) and (5.2) are automatically satisfied.

Examples of functions $\{\mu_{i}(t)\}_{i=1}^{n}$. The corresponding examples are given by:

$\mathrm{p}_1)$ $\mu_{i}(t)=1-(1-\varepsilon_i)e^{-t}$, $t\in\mathbb{R}^+$, $i=1,\dots,n$;

$\mathrm{p}_2)$ $\mu_{i}(t)=1-(1-\varepsilon_i)e^{-t^2/\delta}$, $\delta>0$, $t\in\mathbb{R}^+$, $i=1,\dots,n$, where $\varepsilon_i\in(0,1)$ are some numerical parameters.

Note that examples of the form $\mathrm{p}_1)$ occur in applications of the kinetic theory of gases and transfer theory, and examples of the form $\mathrm{p}_2)$ occur in mathematical biology.

Examples of functions $\{\lambda_{ij}(t)\}_{i,j=1}^{n\times n}$. Examples of singular (at zero) functions $\{\lambda_{ij}(t)\}_{i,j=1}^{n\times n}$ are given by:

$\mathrm{q}_1)$ $\lambda_{ij}(t)=1+(d_{ij}/\sqrt{t})e^{-t}$, $i,j=1,\dots,n$, $t>0$, where $d_{ij}=d_{ji}>0$ are arbitrary parameters;

$\mathrm{q}_2)$ $\lambda_{ij}(t)=1+(c_{ij}/t^\alpha)e^{-t^2}$, $i,j=1,\dots,n$, $t>0$, where $c_{ij}=c_{ji}>0$, $\alpha\in(0,1)$ are numerical parameters.

For the sake of completeness, we also give some examples of non-linearities $\{Q_{i}(u)\}_{i=1}^{n}$ of applied nature.

Examples of functions $\{Q_{i}(u)\}_{i=1}^{n}$. In mathematical epidemiology, the following specific non-linearities $\{Q_{i}(u)\}_{i=1}^{n}$ arise frequently (see [5], [6]):

$$ \begin{equation*} Q_j^{-1}(u)= \begin{cases} \gamma_j(1-e^{-u}), &u\geqslant0, \\ \gamma_j(e^u-1), &u<0, \end{cases} \qquad j=1,\dots,n, \end{equation*} \notag $$
where $Q_j^{-1}$ is the inverse of the function $Q_j$ and $\gamma_j>1$ are numeric parameters. In this theory, the inequalities $\gamma_j>1$, which are known as the threshold conditions, represent the critical values of the number of infected persons above which the epidemic cannot be stopped without a serious medical intervention. Note also that the inequalities $\gamma_j>1$, $j=1,\dots,n$, guarantee the fulfillment of conditions 1)–5).

In the theory of $p$-adic open-closed strings and in kinetic theory of gases, the non-linearity $\{Q_{i}(u)\}_{i=1}^{n}$ has the following form:

$\mathrm{r}_1)$ $Q_i(u)=\delta_i u^p$, $u\in \mathbb{R}$, $j=1,\dots,n$, where $\delta_i>0$ are numerical parameters, and $p>2$ is an arbitrary odd number;

$\mathrm{r}_2)$ $Q_i(u)=q_iu^p/(\eta_i^*)^{p-1}+(1-q_i)u$, $u\in \mathbb{R}$, $i=1,\dots,n$, where $q_i\in (0,1]$ are parameters, and the numbers $\{\eta_i^*\}_{i=1}^n$ can be found from (1.2)(1.4).


Bibliography

1. P. Lancaster, Theory of matrices, Academic Press, New York–London, 1969  mathscinet  zmath
2. V. S. Vladimirov and Ya. I. Volovich, “Nonlinear Dynamics Equation in $p$-Adic String Theory”, Theoret. and Math. Phys., 138:3 (2004), 297–309  crossref  adsnasa
3. V. S. Vladimirov, “Solutions of $p$-adic string equations”, Theoret. and Math. Phys., 167:2 (2011), 539–546  crossref  adsnasa
4. Kh. A. Khachatryan, “On the solvability of a boundary value problem in $ p$-adic string theory”, Trans. Moscow Math. Soc., 2018, 101–115  crossref  mathscinet
5. O. Diekmann and H. G. Kaper, “On the bounded solutions of a nonlinear convolution equation”, Nonlinear Anal., 2:6 (1978), 721–737  crossref  mathscinet  zmath
6. O. Diekmann, “Thresholds and travelling waves for the geographical spread of infection”, J. Math. Biol., 6:2 (1978), 109–130  crossref  mathscinet  zmath
7. A. Kh. Khachatryan and Kh. A. Khachatryan, “Solvability of a nonlinear model Boltzmann equation in the problem of a plane shock wave”, Theoret. and Math. Phys., 189:2 (2016), 1609–1623  crossref  adsnasa
8. C. Cercignani, The Boltzmann equation and its applications, Appl. Math. Sci., 67, Springer-Verlag, New-York, 1988  crossref  mathscinet  zmath
9. N. B. Engibaryan, “On a problem in nonlinear radiative transfer”, Astrophysics, 2:1 (1966), 12–14  crossref  adsnasa
10. Kh. A. Khachatryan, “The solvability of a system of nonlinear integral equations of Hammerstein type on the whole line”, Izv. Saratov Univ. Math. Mech. Inform., 19:2 (2019), 164–181 (Russian)  mathnet  crossref  mathscinet  zmath
11. 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
12. Kh. A. Khachatryan, “On the solubility of certain classes of non-linear integral equations in $p$-adic string theory”, Izv. Math., 82:2 (2018), 407–427  crossref  adsnasa
13. Kh. A. Khachatryan, “Existence and uniqueness of solution of a certain boundary-value problem for a convolution integral equation with monotone non-linearity”, Izv. Math., 84:4 (2020), 807–815  crossref  adsnasa
14. Kh. A. Khachatryan and H. S. Petrosyan, “Integral equations on the whole line with monotone nonlinearity and difference kernel”, J. Math. Sci. (N.Y.), 255:6 (2021), 790–804  crossref  mathscinet  zmath
15. A. N. Kolmogorov and S. V. Fomin, Introductory real analysis, Prentice-Hall, Inc., Englewood Cliffs, NJ, 1970  mathscinet  zmath
16. W. Rudin, Functional analysis, McGraw-Hill Series in Higher Mathematics, McGraw-Hill Book Co., New York–Düsseldorf–Johannesburg, 1973  mathscinet  zmath
17. N. B. Engibaryan, “Conservative systems of integral convolution equations on the half-line and the entire line”, Sb. Math., 193:6 (2002), 847–867  crossref
18. L. G. Arabadzhyan and A. S. Khachatryan, “A class of integral equations of convolution type”, Sb. Math., 198:7 (2007), 949–966  crossref  adsnasa

Citation: Kh. A. Khachatryan, H. S. Petrosyan, “On non-trivial solvability of one system of non-linear integral equations on the real axis”, Izv. Math., 87:5 (2023), 1062–1077
Citation in format AMSBIB
\Bibitem{KhaPet23}
\by Kh.~A.~Khachatryan, H.~S.~Petrosyan
\paper On non-trivial solvability of one system of non-linear integral equations on the real axis
\jour Izv. Math.
\yr 2023
\vol 87
\issue 5
\pages 1062--1077
\mathnet{http://mi.mathnet.ru//eng/im9348}
\crossref{https://doi.org/10.4213/im9348e}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4666687}
\zmath{https://zbmath.org/?q=an:1527.45004}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2023IzMat..87.1062K}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001101882800012}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85177190766}
Linking options:
  • https://www.mathnet.ru/eng/im9348
  • https://doi.org/10.4213/im9348e
  • https://www.mathnet.ru/eng/im/v87/i5/p215
  • 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
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024