\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2012 (2012), No. 137, pp. 1--12.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2012 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2012/137\hfil Existence of solutions] {Existence of solutions for mixed Volterra-Fredholm integral equations} \author[A. Aghajani, Y. Jalilian, K. Sadarangani \hfil EJDE-2012/137\hfilneg] {Asadollah Aghajani, Yaghoub Jalilian, Kishin Sadarangani} % in alphabetical order \address{Asadollah Aghajani \newline School of Mathematics, Iran University of Science and Technology, Narmak, Tehran 16846-13114, Iran} \email{aghajani@iust.ac.ir} \address{Yaghoub Jalilian \newline Department of Mathematics, Razi University, Kermanshah 67149, Iran} \email{yajalilian@iust.ac.ir} \address{Kishin Sadarangani \newline Department of Mathematics, University of Las Palmas de Gran Canaria, 35017 Las Palmas de Gran Canaria, Spain} \email{ksadaran@dma.ulpgc.es} \thanks{Submitted May 11, 2012. Published August 19, 2012.} \subjclass[2000]{45G10, 47H08} \keywords{Functional integral equation; fixed-point; Nemytski operator; \hfill\break\indent measure of weak noncompactness; Carath\'eodory conditions} \begin{abstract} In this article, we give some results concerning the continuity of the nonlinear Volterra and Fredholm integral operators on the space $L^{1}[0,\infty)$. Then by using the concept of measure of weak noncompactness, we prove an existence result for a functional integral equation which includes several classes of nonlinear integral equations. Our results extend some previous works. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \newtheorem{definition}[theorem]{Definition} \newtheorem{example}[theorem]{Example} \allowdisplaybreaks \section{Introduction} Integral Equations occur in mechanics and many related fields of engineering and mathematical physics \cite{b3,b4,b5,b8,b9,b10,c1,d2,k2,m2,m3,t1,z1}. They also form one of useful mathematical tools in many branches of pure analysis such as functional analysis \cite{k1,t1}. Recently many papers have been devoted to the existence of solutions of nonlinear functional integral equations \cite{a1,a2,b1,b2,b5,b8}. Our main purpose is to prove an existence theorem for a class of functional integral equations which contains many integral or functional integral equations. For example, we can mention the nonlinear Volterra integral equations, mixed Volterra-Fredholm integral equations and Fredholm integral equations on the unbounded interval $[0,\infty)$. The concept of measure of weak noncompactness was developed by De Blasie \cite{d1}. Bana\'s and Knap \cite{b3} introduced a measure of weak noncompactness in the space of real Lebesgue integrable functions on an interval which is convenient for our purpose. In the proof of main result we will use a measure of weak noncompactness given by Bana\'s and Knap to find a special subset of $L^{1}[0,\infty)$ and also by applying the Schauder fixed point theorem on this set, the existence result which generalizes several previous works \cite{a3,b4,b5,b6,b8,b10,d2,z1} will be proven. Organization of this article: Section 2 gives some definitions and preliminary results about continuous operator on $L^{1}(\mathbb{R}_{+})$, Section 3 describes the concept of measure of weak noncompactness and weakly compact sets in $L^{1}(\mathbb{R}_{+})$ and finally in Section 4 we give our main result and some examples. \section{Notations and auxiliary results} In this paper, $\mathbb{R}_{+}$ indicates the interval $[0,\infty)$ and for the Lebesgue measurable subset $D$ of $\mathbb{R}$, $m(D)$ implies the Lebesgue measure of $D$. Also, let $L^{1}(D)$ be the space of all Lebesgue integrable functions $y$ on $D$ equipped with the standard norm $\|y\|_{L^{1}(D)}=\int_{D}|y(x)|dx$. \begin{lemma}[\cite{d5}] \label{lem2.1} Let $\Omega$ be a Lebesgue measurable subset of $\mathbb{R}$ and $1\leq p\leq\infty$. If $\{f_{n}\}$ is convergent to $f\in L^{p}(\Omega)$ in the $L^{p}$-norm, then there is a subsequence $\{f_{n_{k}}\}$ which converges to $f$ a.e., and there is $g\in L^{p}(\Omega)$, $g\geq 0$, such that \[ |f_{n_{k}}(x)|\leq g(x),\quad \text{a.e. } x\in \Omega. \] \end{lemma} Let $I\subset\mathbb{R}$ be an interval. A function $f:I\times\mathbb{R}\to\mathbb{R}$ is said to have the Carath\'{e}odory property if \begin{itemize} \item[(M)] for all $x\in\mathbb{R}$ the function $t\mapsto f(t,x)$ is Lebesgue measurable on $I$; \item[(C)] for almost all $t\in I$ the function $x\mapsto f(t,x)$ is continuous on $\mathbb{R}$. \end{itemize} One of the most important nonlinear mappings is the so-called Nemytski operator which is also called the substitution (or superposition) operator \cite{b3,b5,d5,z1}. By substituting the function $x:I\to\mathbb{R}$ into the function $f$ the Nemytski operator $F:x\to f(.,x(.))$ has been obtained which acts on a space containing functions $x$. Krasnosel'skii \cite{k2} and Appell and Zabreiko \cite{a3} have proven the following assertion when $I$ is a bounded and an unbounded domain respectively. \begin{theorem} \label{thm2.1} The superposition operator $F$ generated by function $f(t, x)$ maps the space $L^{1}(I)$ continuously into itself if and only if $|f (t,x)|\leq g(t)+c|x|$ for all $t$ in an interval $I$, and $x\in\mathbb{R}$, where $g$ is a function from the space $L^{1}(I)$ and $c$ is a nonnegative constant. \end{theorem} \begin{remark} \label{rmk2.1}\rm The Carath\'{e}odory property can be generalized to functions $f:\Omega\times\mathbb{R}^{m}\to\mathbb{R}$ where $\Omega$ is a measurable subset of $\mathbb{R}^{n}$. Theorem \ref{thm2.1} holds similarly if and only if there exist $c\in\mathbb{R}$ and $g\in L^{1}(\Omega)$ such that \begin{equation}\label{} |f(x,y)|\leq g(x)+c\sum_{i=1}^{m}|y_i|, \end{equation} for almost all $x\in\Omega$ and all $y=(y_1,\dots ,y_{m})\in\mathbb{R}^{m}$. \end{remark} Now we are going to review a theorem from \cite{b3} about the continuity of the linear Volterra integral operator on the space $L^{1}=L^{1}(\mathbb{R}_{+})$. Let $\Delta=\{(t,s):0\leq s\leq t\}$ and $k:\bigtriangleup\to \mathbb{R}$ be a measurable function with respect to both variables. Consider \begin{eqnarray*} (Kx)(t)=\int_0^{t}k(t,s)x(s)ds,\quad t\in\mathbb{R}_{+},\; x\in L^{1}(\mathbb{R}_{+}). \end{eqnarray*} We notice that $K$ is a linear Volterra integral operator generated by $k$. \begin{theorem} \label{thm2.2} Let $k$ be measurable on $\Delta$ and such that \begin{equation}\label{e} \operatorname{ess\,sup}_{s\geq0}\int_{s}^{\infty}|k(t,s)|dt<\infty. \end{equation} Then the Volterra integral operator $K$ generated by $k$ maps (continuously) the space $L^{1}(\mathbb{R}_{+})$ into itself and the norm $\|K\|$ of this operator is majorized by the number $\operatorname{ess\,sup}_{s\geq0}\int_{s}^{\infty}|k(t,s)|dt$. \end{theorem} Assume that $A$ is a measurable subset of $\mathbb{R}_{+}$, we denote by $\|K\|_{A}$ the norm of linear Volterra operator $K:L^{1}(A)\to L^{1}(A)$. Now we give a result concerning the continuity of the nonlinear Volterra operator on $L^{1}(\mathbb{R}_{+})$. In what follows we suppose that $u:\mathbb{R}_{+}\times\mathbb{R}_{+}\times\mathbb{R}\to\mathbb{R}$ is a function which satisfies: \begin{itemize} \item[(a)] $t\to u(t,s,x)$ is measurable for all $s\in \mathbb{R}_{+}$ and $x\in\mathbb{R}$; \item[(b)] $(s,x)\to u(t,s,x)$ is continuous for almost all $t\in\mathbb{R}_{+}$. \end{itemize} \begin{theorem} \label{thm2.3} Let $u:\mathbb{R}_{+}\times\mathbb{R}_{+}\times\mathbb{R}\to\mathbb{R}$ be a function such that \begin{equation} \label{eq4} |u(t,s,x)|\leq k_1(t,s)+k_2(t,s)|x|,\quad t,s\in \mathbb{R}_{+}\; x\in\mathbb{R}, \end{equation} where $k_i:\mathbb{R}_{+}\times\mathbb{R}_{+}\to\mathbb{R}_{+}$ (i=1,2) are measurable functions. Moreover, the integral operator $K_2$ generated by $k_2$ is a continuous map from $L^{1}(\mathbb{R}_{+})$ into itself and $\int_0^{t}k_1(t,s)ds\in L^{1}(\mathbb{R}_{+})$. Then the operator \[ (Ux)(t)=\int_0^{t}u(t,s,x(s))ds, \] maps $L^{1}(\mathbb{R}_{+})$ continuously into itself \end{theorem} \begin{proof} Let $\{x_{n}\}$ be an arbitrary sequence in $L^{1}=L^{1}(\mathbb{R}_{+})$ which converges to $x\in L^{1}$ in the $L^{1}$-norm. By using Lemma \ref{lem2.1} there is a subsequence $\{x_{n_{k}}\}$ which converges to $x$ a.e., and there is $g\in L^{1}$, $g\geq 0$, such that \begin{equation} \label{eq6} |x_{n_{k}}(s)|\leq g(s),\quad \text{a.e. on } \mathbb{R}_{+}. \end{equation} Since $x_{n_{k}}\to x$ almost everywhere in $\mathbb{R}_{+}$, it readily follows from (b) that \begin{equation}\label{eq7} u(t,s,x_{n_{k}}(s))\to u(t,s,x(s)),\quad \text{for almost all } s,t\in \mathbb{R}_{+}. \end{equation} From inequalities \eqref{eq4} and \eqref{eq6}, we infer that \begin{equation} \label{eq8} |u(t,s,x_{n_{k}}(s))|\leq k_1(t,s)+k_2(t,s)g(s),\quad ~ \text{for almost all }s, t\in \mathbb{R}_{+}. \end{equation} As a consequence of the Lebesgue's Dominated Convergence Theorem, \eqref{eq7} and \eqref{eq8} yield \[ \int_0^{t} u(t,s,x_{n_{k}}(s))ds\to\int_0^{t}u(t,s,x(s))ds, \] for almost all $t\in\mathbb{R}_{+}$. Inequality \eqref{eq8} implies that \begin{equation} \label{eq88} |(Ux_{n_{k}})(t)|\leq\int_0^{t}|u(t,s,x_{n_{k}}(s))|ds \leq \int_0^{t}k_1(t,s)ds+ \int_0^{t}k_2(t,s)g(s)ds, \end{equation} for almost all $t\in \mathbb{R}_{+}$. Regarding the assumptions on $k_1$ and $k_2,$ we obtain \begin{equation} \label{eq888} \int_0^{\infty} \int_0^{t}k_1(t,s)\,ds\,dt+\int_0^{\infty} \int_0^{t}k_2(t,s)g(s)\,ds\,dt<\infty. \end{equation} Then inequalities \eqref{eq88}-\eqref{eq888} and the Lebesgue's Dominated Convergence Theorem imply \begin{eqnarray*} \|Ux_{n_{k}}-Ux\|_{L^{1}}\to0. \end{eqnarray*} Since any sequence $\{x_{n}\}$ converging to $x$ in $L^{1}$ has a subsequence $\{x_{n_{k}}\}$ such that $Ux_{n_{k}}\to Ux$ in $L^{1}$, we can conclude that $U:L^{1}(\mathbb{R}_{+})\to L^{1}(\mathbb{R}_{+})$ is a continuous operator. \end{proof} Similar to the above theorem, we can prove the following theorem for the nonlinear Fredholm integral operators. \begin{theorem} \label{thm2.4} Let $v:\mathbb{R}_{+}\times\mathbb{R}_{+}\times\mathbb{R}\to\mathbb{R}$ be a function satisfying {\rm (a)--(b)} such that \begin{equation} \label{e.29} |v(t,s,x)|\leq k_1(t,s)+k_2(t,s)|x|,\quad t,s\in \mathbb{R}_{+}\; x\in\mathbb{R}, \end{equation} where $k_i:\mathbb{R}_{+}\times\mathbb{R}_{+}\to\mathbb{R}_{+}$ ($i=1,2$) are measurable functions. Moreover, the integral operator $(K_2x)(t)=\int_0^{\infty} k_2(t,s)|x(s)|ds$ maps $L^{1}(\mathbb{R}_{+})$ continuously into itself and $k_1(t,s)\in L^{1}(\mathbb{R}_{+}\times\mathbb{R}_{+})$. Then the operator \begin{eqnarray*} (Vx)(t)=\int_0^{\infty}v(t,s,x(s))ds, \end{eqnarray*} maps $L^{1}(\mathbb{R}_{+})$ continuously into itself. \end{theorem} \section{Measure of weak noncompactness} Let $(E,\|\cdot\|)$ be an infinite dimensional Banach space with zero element $\theta$. We write $B(x,r)$ to denote the closed ball centered at $x$ with radius $r$ and $\operatorname{conv}X$ to denote the closed convex hull of $X$. Further let: $\textbf{m}_{E}$ be the family of all nonempty bounded subsets of $E$, $\textbf{n}_{E}^{w}:$ the subfamily of $\textbf{m}_{E}$ consisting of all relatively weakly compact sets, and $\overline{X}^{w}:$ the weak closure of a set $X$. In this paper, we use the following definition of the measure of weak noncompactness \cite{b6}. \begin{definition} \label{def3.1} A mapping $\mu :\textbf{m}_{E}\to \mathbb{R}_{+}$ is said to be a measure of weak noncompactness if it satisfies the following conditions: \begin{itemize} \item[(1)] The family $\ker \mu=\{X\in \textbf{m}_{E}:\mu(X)=0\}$ is nonempty and $\ker \mu\subset \textbf{n}_{E}^{w}$, \item[(2)] $X\subset Y\Rightarrow \mu(X)\leq\mu(Y)$, \item[(3)] $\mu (\operatorname{conv}X)=\mu(X)$, \item[(4)] $\mu(\lambda X+(1-\lambda)Y)\leq\lambda \mu(X)+(1-\lambda)\mu(Y)$ for $\lambda\in[0,1]$, \item[(5)] If $X_{n}\in \textbf{m}_{E}$, $X_{n}=\overline{{X}}_{n}^{w}$ for $n=1,2,\dots $ and if $\lim_{n\to\infty} \mu(X_{n})=0,$ then the intersection set $X_{\infty}=\bigcap_{n=1}^{\infty} X_{n}$ is nonempty. \end{itemize} \end{definition} In the sequel, we will use a measure of weak noncompactness represented by a convenient formula in the space $L^{1}(\mathbb{R}_{+})$ \cite{b7}. For $X\in\textbf{m}_{L^{1}(\mathbb{R}_{+})}$ define: \begin{gather*} c(X)=\lim_{\varepsilon\to 0}\{\sup_{x\in X}\{\sup[\int_{D}|x(t)|dt:D\subset\mathbb{R}_{+},~m(D)\leq\varepsilon]\}\},\\ d(X)=\lim_{T\to\infty}\{\sup[\int_{T}^{\infty}|x(t)|dt:x\in X]\}, \\ \mu(X)=c(X)+d(X). \end{gather*} In \cite{b7}, it is shown that $\mu$ is a measure of weak noncompactness on $L^{1}(\mathbb{R}_{+})$. By using the following theorem \cite{d4}, we can infer that $\ker\mu=\textbf{n}_{L^{1}(\mathbb{R}_{+})}^{w}$. \begin{theorem} \label{thm3.1} A bounded set $X$ is relatively weakly compact in $L^{1}(\mathbb{R}_{+})$ if and only if the following two conditions are satisfied: \begin{itemize} \item[(1)] for any $\varepsilon>0$ there exists $\delta>0$ such that if $m(D)\leq \delta$ then $\int_{D}|x(t)|dt\leq\varepsilon$ for all $x\in X$, \item[(2)] for any $\varepsilon>0$ there exists $T>0$ such that $\int_{T}^{\infty}|x(t)|dt\leq\varepsilon$ for any $x\in X$. \end{itemize} \end{theorem} \section{Main results} In this section, we study the existence of solutions to the functional integral equation \begin{equation}\label{eq} x(t)=f\Big(t,\int_0^{t}u(t,s,x(s))ds,\int_0^{\infty}a_2(t)v(s,x(s))ds\Big), \quad t\geq0. \end{equation} This equation is a general form of many integral equations, such as the mixed Volterra-Fredholm integral equation \begin{equation}\label{eq1} x(t)=g(t)+\int_0^{t}k(t,s)u(s,x(s))ds+a(t)\int_0^{\infty}v(s,x(s))ds, \quad t\geq0. \end{equation} Equations like \eqref{eq1} have been considered by many authors; see for example \cite{b9,c2,d3,m1,m3} and references cited therein. Moreover, \eqref{eq} contains the nonlinear Volterra and Fredholm integral equations on $\mathbb{R}_{+}$ such as: \begin{gather*} x(t)=g(t)+\int_0^{t}u(t,s,x(s))ds,\quad t\geq0, \\ x(t)=f(t)+a(t)\int_0^{\infty}v(s,x(s))ds,\quad t\geq0. \end{gather*} We consider equation \eqref{eq} under the following assumptions: \begin{itemize} \item[(i)] The function $f:\mathbb{R}_{+}\times\mathbb{R}^{2}\to\mathbb{R}$ satisfies Carath\'{e}dory conditions and there exist constant $B\in\mathbb{R}_{+}$ and function $a_1\in L^{1}(\mathbb{R}_{+})$ such that \begin{equation} \label{eq2} |f(t,x,y)|\leq a_1(t)+B(|x|+|y|),\quad t\in\mathbb{R}_{+},\; x,y\in\mathbb{R}. \end{equation} \item[(ii)] $u:\mathbb{R}_{+}\times\mathbb{R}_{+}\times\mathbb{R}\to\mathbb{R}$ satisfies (a)--(b) and $|u(t,s,x)|\leq k_1(t,s)+k_2(t,s)|x|$ for $(t,s,x)\in \mathbb{R}_{+}\times\mathbb{R}_{+}\times\mathbb{R}$, where $k_i:\mathbb{R}_{+}\times\mathbb{R}_{+}\to\mathbb{R}_{+}$ (i=1,2) satisfies Carath\'{e}odory conditions. Moreover, the integral operator $K_2$ generated by $k_2$ i.e. \begin{equation}\label{eq3} (K_2x)(t)=\int_0^{t}k_2(t,s)x(s)ds, \end{equation} is a continuous map from $L^{1}(\mathbb{R}_{+})$ into itself and ${\int_0^{t}k_1(t,s)ds\in L^{1}(\mathbb{R}_{+})}$. \item[(iii)] $v:\mathbb{R}_{+}\times\mathbb{R}\to\mathbb{R}$ satisfies Carath\'{e}odory conditions and $|v(s,x)|\leq n(s)+b|x|$ for $(s,x)\in\mathbb{R}_{+}\times\mathbb{R}$ where $n\in L^{1}(\mathbb{R}_{+})$ and $b$ is a positive constant. \item[(iv)] $a_2:\mathbb{R}_{+}\to\mathbb{R}$ is a function in $L^{1}(\mathbb{R}_{+})$. \item[(v)] $B(b\|a_2\|+\|K_2\|)<1$, where $\|K_2\|$ denotes the norm of operator $K_2$. \end{itemize} To prove the main result of this paper, we need the next lemma. Let $X$ be a nonempty, closed, convex, bounded and weakly compact subset of $L^{1}=L^{1}(\mathbb{R}_{+})$ and $I=[0,a]$ where $a>0$. Moreover, we define the operator $F$ on $L^{1}=L^{1}(\mathbb{R}_{+})$ as follows: \begin{equation} \label{f} (Fx)(t)=f\Big(t,\int_0^{t}u(t,s,x(s))ds,\int_0^{\infty}a_2(t)v(s,x(s))ds\Big). \end{equation} \begin{lemma} \label{lem4.1} Suppose that assumptions (i)--(iv) hold and the operator $F$ takes $X$ into itself. Then for any $\varepsilon>0$ there exists $D_{\varepsilon}\subset I$ with $m(I \setminus D_{\varepsilon})\leq\varepsilon$ such that $F(\operatorname{conv} FX)$ on $D_{\varepsilon}$ is a relatively compact subset of $C(D_{\varepsilon})$. \end{lemma} \begin{proof} Consider an arbitrary but fixed $\varepsilon>0$. Then using Lusin theorem and generalized version of Scorza-Dragoni theorem \cite{c1} we can find the closed set $D_{\varepsilon}\subset I$ with $m(I \setminus D_{\varepsilon})\leq\varepsilon$, such that the functions $a_i\big|_{D_{\varepsilon}},$ $k\big|_{D_{\varepsilon}\times\mathbb{R}_{+}}$, $u\big|_{D_{\varepsilon}\times\mathbb{R}_{+}\times\mathbb{R}}$ and $f\big|_{D_{\varepsilon}\times\mathbb{R}_{+}\times\mathbb{R}}$ are continuous. Let us take an arbitrary $x\in X$. Then for $t\in D_{\epsilon}$ we have \begin{equation} \label{bu} \begin{split} \big|\int_0^{t}u(t,s,x(s))ds\big| & \leq \int_0^{t}k_1(t,s)+ \int_0^{t}k_2(t,s)|x(s)|ds\\ &\leq \bar{k}_1a+\bar{k}_2\|x\|\leq\bar{k}_1a+\bar{k}_2\|X\|=:U_{\varepsilon}, \end{split} \end{equation} and \begin{equation} \label{bv} \big|\int_0^{\infty}a_2(t)v(s,x(s))ds\big| \leq \bar{a}_2 (\|n\|+b\|x\|)\leq \bar{a}_2 (\|n\|+b\|X\|)=:V_{\varepsilon}, \end{equation} where $\|X\|=\sup\{\|x\|:x\in X\}$, $\bar{a}_i=\sup\{|a_i(t)|:t\in D_{\varepsilon}\}$ and $\bar{k}_i=\sup\{|k_i(t,s)|:(t,s)\in D_{\varepsilon}\times I\}$ for $i=1,2$. Now let $y\in FX$. Then there exists $x\in X$ such that $y=Fx$. Using the inequalities \eqref{bu} and \eqref{bv} for $t\in D_{\varepsilon}$ we obtain \begin{equation} \label{by} \begin{split} |y(t)|=|(Fx)(t)| &\leq a_1(t)+|\int_0^{t}u(t,s,x(s))ds|+|\int_0^{\infty}a_2(t)v(s,x(s))ds|\\ &\leq \bar{a}_1+U_{\varepsilon}+V_{\varepsilon}=:Y_{\varepsilon}. \end{split} \end{equation} We can easily deduce that the inequality \eqref{by} is true, for any $y\in Y=\operatorname{conv} FX$. Now assume that $\{y_{n}\}$ is a sequence in $Y$ and let $t_1,t_2\in D_{\varepsilon}$. Without loss of generality we can assume that $t_1\leq t_2$. Relatively weakly compactness of the set $\{y_{n}\}$ implies that for $\varepsilon_0=t_2-t_1$ there exists $0<\delta_0\leq\varepsilon_0$ such that for any measurable subset $D$ of $[0,t_1]$ with $m(D)\leq \delta_0$, we have: \begin{equation} \label{wy} \int_{D}|y_{n}(t)|dt\leq \varepsilon_0\quad \text{for } n=1,2,\dots . \end{equation} We see that the estimate \eqref{by} does not depend on the choice of $\varepsilon$. Thus for $\varepsilon=\delta_0$ there exists a closed set $D_{\delta_0}\subset [0,t_1]$ with $m([0,t_1] \setminus D_{\delta_0})\leq \delta_0$ such that \begin{equation} \label{e4.10} |y_{n}(t)|\leq Y_{\delta_0}\quad \text{for }t\in D_{\delta_0},\; n=1,2,\dots . \end{equation} Hence from \eqref{wy} and uniform continuity of $u\big|_{D_{\varepsilon}\times D_{\delta_0}\times[-Y_{\delta_0},Y_{\delta_0}]}$ and $k_i\big|_{D_{\varepsilon}\times[0,a]}$ $(i=1,2)$ we infer that \begin{equation} \label{eu} \begin{split} &\int_0^{t_1}|u(t_1,s,y_{n}(s))-u(t_2,s,y_{n}(s))|ds\\ &\leq \int_{D_{\delta_0}}|u(t_1,s,y_{n}(s))-u(t_2,s,y_{n}(s))|ds\\ &\quad + \int_{[0,t_1] \setminus D_{\delta_0}}|u(t_1,s,y_{n}(s))-u(t_2,s,y_{n}(s))|ds \\ &\leq O(|t_1-t_2|)+2\bar{k}_1m([0,t_1] \setminus D_{\delta_0}) +2\bar{k}_2\int_{[0,t_1] \setminus D_{\delta_0}} |y_{n}(t)|dt \\ &\leq O(|t_1-t_2|)+2(\bar{k}_1+\bar{k}_2)|t_1-t_2|. \end{split} \end{equation} Here $O$ is a function which $O(\eta)\to0$ as $\eta\to0$. Thus from \eqref{eu} we have: \begin{equation} \label{1a} \begin{split} &|\int_0^{t_1}u(t_1,s,y_{n}(s))ds-\int_0^{t_2}u(t_2,s,y_{n}(s))ds|\\ &\leq \int_0^{t_1}|u(t_1,s,y_{n}(s))-u(t_2,s,y_{n}(s))|ds +|\int_{t_1}^{t_2}u(t_2,s,y_{n}(s))ds| \\ &\leq O(|t_1-t_2|)+2(\bar{k}_1+\bar{k}_2)|t_1-t_2| +\int_{t_1}^{t_2}k_1(t_2,s)ds +\int_{t_1}^{t_2}k_2(t_2,s)|y_{n}(s)|ds \\ &\leq O(|t_1-t_2|)+2(\bar{k}_1+\bar{k}_2)|t_1-t_2| +\bar{k}_1|t_1-t_2|+\bar{k}_2\int_{t_1}^{t_2}|y_{n}(s)|ds. \end{split} \end{equation} Weakly compactness of the set $\{y_{n}\}$ implies that $\int_{t_1}^{t_2}|y_{n}(s)| ds$ is arbitrary small uniformly with respect to $n\in\mathbb{N}$ if $t_2-t_1$ is small enough. Then from \eqref{1a} and \eqref{bu} the sequence $\{Uy_{n}\}$ which \[ (Uy_{n})(t)=\int_0^{t}u(t,s,y_{n}(s))ds, \] is equibounded and equicontinuous on the set $D_{\varepsilon}$. Obviously from assumption (iii) and inequality \eqref{bv} we can easily infer that the sequence $\{Vy_{n}\}$ is equibounded and equicontinuous on $D_{\varepsilon}$ where \[ (Vy_{n})(t)=\int_0^{\infty}a_1(t)v(s,y_{n}(s))ds. \] Hence, uniform continuity of $f\big|_{D_{\varepsilon}\times[-U_{\varepsilon},U_{\varepsilon}]\times[-V_{\varepsilon},V_{\varepsilon}]}$ implies that the sequence $\{Fy_{n}\}$ is equibounded and equicontinuous on $D_{\varepsilon}$. Then, by Ascoli theorem the sequence $\{Fy_{n}\}$ has a convergent subsequence in the norm $C(D_{\varepsilon})$. Therefore, $F(\operatorname{conv} FX)$ is a relatively compact subset of $C(D_{\varepsilon})$. \end{proof} Now we present our main result. \begin{theorem} \label{thm4.1} Under assumptions {\rm (i)--(v)}, the functional integral equation \eqref{eq} has at least one solution $x\in L^{1}(\mathbb{R}_{+})$. \end{theorem} \begin{proof} At first we define the operator $F$ on $L^{1}=L^{1}(\mathbb{R}_{+})$ by \[ (Fx)(t)=f\Big(t,\int_0^{t}u(t,s,x(s))ds,\int_0^{\infty}a_2(t)v(s,x(s))ds\Big). \] We prove the theorem in the following steps. Step 1. $F:L^{1}(\mathbb{R}_{+})\to L^{1}(\mathbb{R}_{+})$ is a continuous operator. \\ Using Theorems \ref{thm2.3} and \ref{thm2.4}, the operators \[ (Ux)(t)=\int_0^{t}u(t,s,x(s))ds,\quad (Vx)(t)=\int_0^{\infty}v(t,s,x(s))ds, \] map $L^{1}(\mathbb{R}_{+})$ continuously into itself. Also by assumptions (i)--(iv) and Remark \ref{rmk2.1}, the Nemytski operator generated by $f$ is a continuous operator from $L^{1}(\mathbb{R}_{+})$ into $L^{1}(\mathbb{R}_{+})$. Thus the operator $F$ is continuous. Step 2. There exists a positive number $r$ such that the operator $F$ takes the ball $B(\theta,r)$ into itself. Let $x\in L^{1}(\mathbb{R}_{+})$. Then \begin{equation} \label{iq} \begin{split} \|Fx\|&=\int_0^{\infty}|(Fx)(t)|dt\\ &=\Big |\int_0^{\infty}f\big(t,\int_0^{t}u(t,s,x(s))ds, \int_0^{\infty}a_2(t)v(s,x(s))ds\big)dt\Big| \\ &\leq \|a_1\| +B\int_0^{\infty}\Big(\int_0^{t}|u(t,s,x(s))|ds +\int_0^{\infty}|a_2(t)v(s,x(s))|ds\Big)dt \\ &\leq \|a_1\|+B(K_1+\|K_2\|\|x\|)+B\|a_2\|(\|n\|+b\|x\|), \end{split} \end{equation} where $$ K_1=\int_0^{\infty}\int_0^{t}k_1(t,s)\,ds\,dt. $$ From \eqref{iq} and assumption (v), one can deduce that for ${r=\frac{\|a_1\|+B(K_1+\|a_2\|\|n\|)}{1-B(\|K_2\|+\|a_2\|b)}}$, the operator $F$ takes $B_{r}=B(\theta,r)$ into itself. Step 3. There exists a weakly compact subset $Y$ such that the operator $F$ maps $Y$ into itself. Let $X$ be a nonempty subset of $B_{r}$. Let $\varepsilon>0$ be an arbitrary number and $D\subset\mathbb{R}_{+}$ be a measurable subset with $m(D)\leq\varepsilon$. Then for $x\in X$ we have: \begin{equation} \label{c1} \begin{split} &\int_{D}|(Fx)(t)|dt\\ &\leq \int_{D}a_1(t)dt+B\int_{D}\Big(\int_0^{t}|u(t,s,x(s))|ds +\int_0^{\infty}|a_2(t)v(s,x(s))|ds\Big)dt \\ &\leq \int_{D}a_1(t)dt + B\int_{D}\int_0^{t} k_1(t,s)\,ds\,dt\\ &\quad +B \int_{D}|(K_2x)(t)|dt+B(\|n\|+br)\int_{D}|a_2(t)|dt \\ &\leq \int_{D}a_1(t)dt + B\int_{D}\int_0^{t}k_1(t,s)\,ds\,dt\\ &\quad +B \|K_2\|_{D}\int_{D}|x(t)|dt+B(\|n\|+br)\int_{D}|a_2(t)|dt \end{split} \end{equation} Further, as a simple consequence of the fact that a single set in $L^{1}(\mathbb{R}_{+})$ is weakly compact, for $\gamma(t)=a_1(t),{\int_0^{t}k_1(t,s)ds}$ or $a_2(t)$, we have: \begin{itemize} \item[(C1)] $\lim_{\varepsilon\to 0} \{\sup[\int_{D}|\gamma(t)|dt:D\subset\mathbb{R}_{+},~m(D)\leq\varepsilon]\}=0$, \item[(C2)] $\lim_{T\to\infty}\int_{T}^{\infty}|\gamma(t)|dt=0$. \end{itemize} Then from \eqref{c1} and (C1) we conclude that \begin{equation} \label{c3} c(FX)\leq B\|K_2\| c(X). \end{equation} By similar calculations we obtain: \begin{equation} \label{c2} \begin{split} \int_{T}^{\infty}|(Fx)(t)|dt &\leq \int_{T}^{\infty}a_1(t)dt+B \int_{T}^{\infty} \int_0^{t}k_1(t,s)ds\,dt\\ &\quad +B\|K_2\|\int_{T}^{\infty}|x(t)|dt+B(\|n\|+br) \int_{T}^{\infty}|a_2(t)|dt. \end{split} \end{equation} Therefore, from (C2), we have \begin{equation} \label{c4} d(FX)\leq B\|K_2\| d(X). \end{equation} Hence by adding \eqref{c3} and \eqref{c4} we obtain \begin{equation} \label{c5} \mu(FX)\leq B\|K_2\| \mu(X). \end{equation} Assumption (v) implies that $B\|K_2\|< 1$. Thus inequality \eqref{c5} yields that there exists a closed, convex and weakly compact set $X_{\infty}\subset B_{r}$ such that $FX_{\infty}\subset X_{\infty} $. Let $Y=\operatorname{conv} FX_{\infty}$. Obviously $FY\subset Y\subset X_{\infty}$. Thus $FY$ and $ Y$ are relatively weakly compact. Step 4. The set $FY$ obtained in the Step 3 is a relatively compact subset of $L^{1}(\mathbb{R}_{+})$. Suppose $\{y_{n}\}\subset Y$, and fix an arbitrary $\varepsilon>0$. Applying Theorem \ref{thm3.1} for relatively weakly compact set $FY$ implies that there exists $T>0$ such that for $m,n\in\mathbb{N}$ \begin{equation} \label{c6} \int_{T}^{\infty}|(Fy_{n})(t)-(Fy_{m})(t)|dt\leq \frac{\varepsilon}{2}. \end{equation} Further, by using Lemma \ref{lem4.1} for any $k\in\mathbb{N}$ there exists a closed set $D_{k}\subset[0,T]$ with $m([0,T]\setminus D_{k})\leq \frac{1}{k}$ such that $\{Fy_{n}\}$ is a relatively compact subset of $C(D_{k})$. So for any $k\in\mathbb{N}$ there exists a subsequence $\{y_{k,n}\}$ of $\{y_{n}\}$ which is a Cauchy sequence in $C(D_{k})$. Also these subsequences can be chosen such that $\{y_{k+1,n}\}\subseteq \{y_{k,n}\}$. Consequently the subsequence $\{y_{n,n}\}$ is a Cauchy sequence in each space $C(D_{k})$ for any $k\in\mathbb{N}$ which for simplicity we call it again $\{y_{n}\}$. From the relatively weakly compactness of $\{Fy_{n}\}$ we can find $\delta>0$ such that for each closed subset $D_{\delta}$ with $m([0,T]\setminus D_{\delta})\leq\delta$ we obtain: \begin{equation} \label{c7} \int_{[0,T]\setminus D_{\delta}}|(Fy_{n})(t)-(Fy_{m})(t)|dt\leq \frac{\varepsilon}{4},\quad ~m,n\in\mathbb{N}. \end{equation} Considering the fact $\{Fy_{n}\}$ is Cauchy in $C(D_{k})$ for each $k\in\mathbb{N}$ one can find $k_0$ such that $m([0,T]\setminus D_{k_0})\leq\delta$ and for $m,n\geq k_0$ \begin{equation} \label{c8} \|(Fy_{n})-(Fy_{m})\|_{C(D_{k_0})}\leq \frac{\varepsilon}{4(m(D_{k_0})+1)}, \end{equation} consequently \eqref{c7} and \eqref{c8} imply that \begin{equation} \label{c9} \begin{split} &\int_0^{T}|(Fy_{n})(t)-(Fy_{m})(t)|dt\\ &= \int_{D_{k_0}}|(Fy_{n})(t)-(Fy_{m})(t)|dt +\int_{[0,T]\setminus D_{k_0}}|(Fy_{n})(t)-(Fy_{m})(t)|dt\\ &\leq \frac{\varepsilon}{2}, \end{split} \end{equation} for $m,n\geq k_0$. Now by considering \eqref{c6} and \eqref{c9} for $m,n\geq k_0$ we obtain the inequality \begin{equation} \label{c10} \|(Fy_{n})-(Fy_{m})\|_{L^{1}}=\int_0^{\infty}|(Fy_{n})(t)-(Fy_{m})(t)|dt\leq \varepsilon, \end{equation} which shows that the sequence $\{Fy_{n}\}$ is a Cauchy sequence in the Banach space $L^{1}({\mathbb{R}_{+}})$. Then $\{Fy_{n}\}$ has a convergent subsequence which implies that $FY$ is a relatively compact subset of $L^{1}({\mathbb{R}_{+}})$. Step 5. By the Step 4 there exists a bounded, closed, convex set $Y\subset L^{1}({\mathbb{R}_{+}})$ such that the operator $F:Y\to Y$ is continuous and compact. Then Schauder fixed point theorem completes the proof. \end{proof} Next, by applying our theorem we prove the existence of solutions for some integral equations. \begin{example} \label{examp4.1}\rm Consider the Fredholm integral equation \begin{equation}\label{w} x(t)= \frac{t^{2/3}}{t^{3}+1}+\int_0^{\infty} a_2(t)\tanh(\frac{s+|x(s)|}{(1+s^{2})^{2}})ds,\quad t\geq0, \end{equation} where $$ a_2(t)=\frac{t\pi}{8}\chi_{[0,1]}+\frac{1}{4(1+t^{2})}\chi_{(1,\infty)}, $$ in which for $A\subset \mathbb{R}_{+}$ and $\chi_{A}(x)=\begin{cases} 1, & x\in A, \\ 0, & x\in \mathbb{R}_{+} \setminus A. \end{cases} $ Put \begin{gather*} f(t,x,y)=\frac{t^{2/3}}{1+t^{3}}+y, \quad u(t,s,x)=0,\\ v(s,x)=\tanh(\frac{s+|x|}{(1+s^{2})^{2}}),\quad a_1(t)=\frac{t^{2/3}}{1+t^{3}},\\ n(s)=\frac{s}{(1+s^{2})^{2}},\quad B=1,\; b=1. \end{gather*} We know that $\tanh(\alpha)\leq \alpha$, for $\alpha>0$. Then \begin{gather*} |v(s,x)|\leq n(s)+b|x|,\\ |f(t,x,y)|\leq\frac{t^{2/3}}{1+t^{3}}+B(|x|+|y|), \end{gather*} Further \[ \|a_2\|=\int_0^{\infty}|a_2(t)|dt=\frac{\pi}{8}. \] Since $u=0$ we can choose $k_1=k_2=0$ and then $\|K_2\|=0$. Thus, $B(b\|a_2\|+\|K_2\|)=\frac{\pi}{8}<1$. It is easy to see that assumptions (i)-(v) are fulfilled. Consequently Theorem \ref{thm4.1} ensures that the equation \eqref{w} has at least one solution in $L^{1}(\mathbb{R}_{+})$. \end{example} \begin{example} \label{examp4.2} \rm Consider the mixed Volterra-Fredholm integral equation \begin{equation}\label{w1} \begin{split} x(t) &=\frac{1+t^{2}}{\cosh(t)}+\int_0^{t}\frac{\lfloor t+s^{2}\rfloor}{2}\exp(-t)\sin( x(s))ds\\ &\quad +\int_0^{\infty}\frac{ t\ln(1+sx^{2}(s))}{3(1+t^{2})^{2}(s+1)}ds,\quad t\geq0, \end{split} \end{equation} where the symbol $\lfloor z \rfloor$ means the largest integer less than or equal to $z$. Let \begin{gather*} f(t,x,y)=\frac{1+t^{2}}{\cosh(t)}+x+y,\\ u(t,s,x)=\frac{\lfloor t+s^{2}\rfloor}{2}\exp(-t)\sin( x),\\ v(s,x)=\frac{\ln(1+sx^{2})}{s+1},\quad a_2(t)=\frac{t}{3(1+t^{2})^{2}},\\ k_2(t,s)=\frac{\lfloor t+s^{2}\rfloor}{2}\exp(-t),\quad B=1,\quad b=1. \end{gather*} We know that $\ln(1+\alpha^{2})\leq|\alpha|$ for $\alpha\in\mathbb{R}$. Then \begin{gather*} |f(t,x,y)|\leq\frac{1+t^{2}}{\cosh(t)}+B(|x|+|y|),\\ |u(t,s,x)|\leq k_2(t,s)|x|,\quad |v(s,x)|\leq b|x|,\\ \|a_2\|=\int_0^{\infty}|a_2(t)|dt=\int_0^{\infty}\frac{t}{3(1+t^{2})^{2}}dt =\frac{1}{6},\\ \int_{s}^{\infty}|k_2(t,s)|dt =\int_{s}^{\infty}\frac{\lfloor t+s^{2}\rfloor}{2}\exp(-t)dt \leq \frac{(s^{2}+s+1)}{2}\exp(-s)\leq\frac{3}{2}\exp(-1), \end{gather*} for $s,t\in\mathbb{R}_{+}$ and $x\in\mathbb{R}$. Therefore, from Theorem \ref{thm2.2}, we have that $B\|K_2\|\leq\frac{3}{2}\exp(-1)$ and then $B(b\|a_2\|+\|K_2\|)\leq \frac{1}{6}+ \frac{3}{2}\exp(-1)<1$. Using Theorem \ref{thm4.1} we deduce that the equation \eqref{w1} has at least one solution in $L^{1}(\mathbb{R}_{+})$. \end{example} \begin{thebibliography}{00} \bibitem{a1} A. Aghajani, Y. Jalilian; \emph{Existence and global attractivity of solutions of a nonlinear functional integral equation}, Communications in Nonlinear Science and Numerical Simulation 15 (2010) 3306-3312. \bibitem{a2} A. Aghajani, Y. Jalilian; \emph{ Existence of nondecreasing positive solutions for a system of singular integral equations}, Mediterranean Journal of Mathematics (2010) DOI: 10.1007/s00009-010-0095-3. \bibitem{a3} J. Appel, P. P. Zabrejko; \emph{Nonlinear Superposition Operators}, in: Cambridge Tracts in Mathematics, vol. 95, Cambridge University Press 1990. \bibitem{b1} J. Bana\'s, T. Zajac; \emph{A new approach to the theory of functional integral equations of fractional order}, J. Math. Anal. Appl. In Press. \bibitem{b2} J. Bana\'s, D. O'Regan; \emph{On existence and local attractivity of solutions of a quadratic Volterra integral equation of fractional order}, J. Math. Anal. Appl. 34 (2008) 573-582. \bibitem{b3} J. Bana\'s, W. G. El-Sayed; \emph{Measures of noncompactness and solvability of an integral equation in the class of functions of locally bounded variation}, J. Math. Anal. Appl. 167 (1992) 133-151. \bibitem{b4} J. Bana\'s, Z. Knap; \emph{Integrable solutions of a functional-integral equation}, Revista Mat. Univ. Complutense de Madrid. 2 (1989) 31-38. \bibitem{b5} J. Bana\'s, A. Chlebowicz; \emph{On existence of integrable solutions of a functional integral equation under Carath\'{e}odory conditions}, Nonlinear Analysis: Theory, Methods and Applications 70 (2009) 3172-3179. \bibitem{b6} J. Bana\'s, J. Rivero; \emph{On measures of weak noncompactness}, Ann. Mat. Pure Appl. 151 (1988) 213-224. \bibitem{b7} J. Bana\'s, Z. Knap; \emph{Measures of weak noncompactness and nonlinear integral equations of convolution type}, J. Math. Anal. Appl. 146 (1990) 353-362. \bibitem{b8} J. Baras, A. Chlebowicz; \emph{On integrable solutions of a nonlinear Volterra integral equation under Caratheodory conditions}, Bull. Lond. Math. Soc., 41 (2009) 1073-1084. \bibitem{b9} I. Bihari; \emph{Notes on a nonlinear integral equation}, Studia Sci. Math. Hungar. 2 (1967) 1-6. \bibitem{b10} T.A. Burton; \emph{Volterra Integral and Differential Equations}, Academic Press, New York 1983. \bibitem{c1} C. Castaign; \emph{Une nouvelle extension du theoreme de Dragoni-Scorza}, C. R. Acad. Sci., Ser. A, 271 (1970) 396-398. \bibitem{c2} C. Corduneanu; \emph{Nonlinear purterbed integral equations}, Rev. Roumaine Math. Pures Appl. 13 (1968) 1279-1284. \bibitem{d1} F. S. De Blasi; \emph{On a property of the unit sphere in Banach spaces}, Bull. Math. Roumanie 21 (1977) 259-262. \bibitem{d2} K. Deimling; \emph{Ordinary Differential Equations in Banach Spaces}, in: Lect. Notes in Mathematics, vol. 596, Springer Verlag 1977. \bibitem{d3} K. Deimling; \emph{Fixed points of condensing maps}, Lecture Notes in Mathematics Springer-Verlag Berlin Heidelberg 1979. \bibitem{d4} J. Dieudonn\'e; \emph{Sur les espaces de K\"othe}, J. Anal. Math. 1 (1951) 81-115. \bibitem{d5} P. Dr\'{a}bek, J. Milota; \emph{Methods of Nonlinear Analysis}, Birkh\"{a}user Velgar AG 2007. \bibitem{k1} R. P. Kanwal; \emph{Linear Integral Equations}, Academic Press, New York 1997. \bibitem{k2} M. A. Krasnosel'skii; \emph{On the continuity of the operator $Fu(x)= f (x, u(x))$}, Dokl. Akad. Nauk SSSR 77 (1951) 185-188. \bibitem{m1} J. D. Mamedov, S. A. Asirov; \emph{Nonlinear Volterra Equations}. Ylym, Aschabad 1977 (Russian). \bibitem{m2} M. Meehan, D. O'Regan; \emph{ Existence theory for nonlinear Volterra integrodifferential and integral equations}, Nonlinear Analysis 31 (1998) 317-341. \bibitem{m3} R. K. Miller, J. A. Nohel, J. S. W. Wong; \emph{A stability theorem for nonlinear mixed integral equations}. J. Math. Anal. Appl. 25 (1969) 446-449. \bibitem{t1} F. G. Tricomi; \emph{Integral Equations}, University Press Cambridge 1957. \bibitem{z1} P. P. Zabrejko, A. I. Koshelev, M. A. Krasnosel'skii, S. G. Mikhlin, L. S. Rakovshchik, V. J. Stecenko; \emph{Integral Equations, Noordhoff}, Leyden 1975. \end{thebibliography} \end{document}