\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2014 (2014), No. 31, pp. 1--10.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2014 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2014/31\hfil Solvability of a quadratic integral equation] {Solvability of a quadratic integral equation of Fredholm type in H\"older spaces} \author[J. Caballero, M. A. Darwish, K. Sadarangani\hfil EJDE-2014/31\hfilneg] {Josefa Caballero, Mohamed Abdalla Darwish, Kishin Sadarangani} % in alphabetical order \address{Josefa Caballero \newline Departamento de Matem\'aticas, Universidad de Las Palmas de Gran Canaria, Campus de Tafira Baja, 35017 Las Palmas de Gran Canaria, Spain} \email{fefi@dma.ulpgc.es} \address{Mohamed Abdalla Darwish \newline Department of Mathematics, Sciences Faculty for Girls, King Abdulaziz University, Jeddah, Saudi Arabia.\newline Department of Mathematics, Faculty of Science, Damanhour University, Damanhour, Egypt} \email{dr.madarwish@gmail.com} \address{Kishin Sadarangani \newline Departamento de Matem\'aticas, Universidad de Las Palmas de Gran Canaria, Campus de Tafira Baja, 35017 Las Palmas de Gran Canaria, Spain} \email{ksadaran@dma.ulpgc.es} \thanks{Submitted October 31, 2013. Published January 27, 2014.} \subjclass[2000]{45G10, 45M99, 47H09} \keywords{Fredholm; H\"older condition; Schauder fixed point theorem} \begin{abstract} In this article, we prove the existence of solutions of a quadratic integral equation of Fredholm type with a modified argument, in the space of functions satisfying a H\"older condition. Our main tool is the classical Schauder fixed point theorem. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{definition}[theorem]{Definition} \newtheorem{example}[theorem]{Example} \newtheorem{remark}[theorem]{Remark} \allowdisplaybreaks \section{Introduction} Differential equations with a modified arguments arise in a wide variety of scientific and technical applications, including the modelling of problems from the natural and social sciences such as physics, biological and economics sciences. A special class of these differential equations have linear modifications of their arguments, and have been studied by several authors, see \cite{Ba}--\cite{Mu2} and their references. The aim of this article is to investigate the existence of solutions of the following integral equation of Fredholm type with a modified argument, \begin{equation}\label{e1.1} x(t)=p(t)+x(t)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau,\quad t\in[0,1]. \end{equation} Our solutions are placed in the space of functions satisfying the H\"older condition. A sufficient condition for the relative compactness in these spaces and the classical Schauder fixed point theorem are the main tools in our study. \section{Preliminaries} Our starting point in this section is to introduce the space of functions satisfying the H\"older condition and some properties in this space. These properties can be found in \cite{BaNa}. Let $[a,b]$ be a closed interval in $\mathbb{R}$, by $C[a,b]$ we denote the space of continuous functions on $[a,b]$ equipped with the supremum norm; i.e., $\|x\|_\infty=\sup\{|x(t)|:t\in[a,b]\}$ for $x\in C[a,b]$. For $0<\alpha\leq 1$ fixed, by $H_\alpha[a,b]$ we will denote the space of the real functions $x$ defined on $[a,b]$ and satisfying the H\"older condition; that is, those functions $x$ for which there exists a constant $H_x^\alpha$ such that \begin{equation}\label{e2.1} |x(t)-x(s)|\leq H_x^\alpha|t-s|^\alpha, \end{equation} for all $t,s\in[a,b]$. It is easily proved that $H_\alpha[a,b]$ is a linear subspace of $C[a,b]$. In the sequel, for $x\in H_\alpha[a,b]$, by $H_x^\alpha$ we will denote the least possible constant for which inequality \eqref{e2.1} is satisfied. More precisely, we put \begin{equation}\label{e2.2} H_x^\alpha=\sup\big\{\frac{|x(t)-x(s)|}{|t-s|^\alpha}:t\in[a,b],\;t\neq s\big\}. \end{equation} The spaces $H_\alpha[a,b]$ with $0<\alpha\leq 1$ can be equipped with the norm $$ \|x\|_\alpha=|x(a)|+\sup\big\{\frac{|x(t)-x(s)|}{|t-s|^\alpha}:t\in[a,b], \;t\neq s\big\}, $$ for $x\in H_\alpha[a,b]$. In \cite{BaNa}, the authors proved that $(H_\alpha[a,b],\|\cdot\|_\alpha)$ with $0<\alpha\leq 1$ is a Banach space. The following lemmas appear in \cite{BaNa}. \begin{lemma}\label{l2.1} For $x\in H_\alpha[a,b]$ with $0<\alpha\leq 1$, the following inequality is satisfied \begin{equation}\label{e2.3} \|x\|_\infty\leq \max\big(1,(b-a)^\alpha\big)\|x\|_\alpha. \end{equation} \end{lemma} \begin{lemma}\label{l2.2} For $0<\alpha<\gamma\leq 1$, we have \begin{equation}\label{e2.4} H_\gamma[a,b]\subset H_\alpha[a,b]\subset C[a,b]. \end{equation} Moreover, for $x\in H_\gamma[a,b]$ the following inequality holds \begin{equation}\label{e2.5} \|x\|_\alpha\leq \max\big(1,(b-a)^{\gamma-\alpha}\big)\|x\|_\gamma. \end{equation} \end{lemma} Now, we present the following sufficient condition for relative compactness in the spaces $H_\alpha[a,b]$ with $0<\alpha\leq 1$ which appears in Example $6$ of \cite{BaNa} and it is an important result for our study. \begin{theorem}\label{t2.3} Suppose that $0<\alpha<\beta\leq 1$ and that $A$ is a bounded subset in $H_\beta[a,b]$ (this means that $\|x\|_\beta\leq M$ for certain constant $M>0$, for any $x\in A$) then $A$ is a relatively compact subset of $H_\alpha[a,b]$. \end{theorem} \section{Main results} In this section, we will study the solvability of \eqref{e1.1} in the H\"older spaces. We will use the following assumptions: \begin{enumerate} \item[(i)] $p\in H_\beta[0,1]$, $0<\beta\leq 1$. \item[(ii)] $k:[0,1]\times[0,1]\to\mathbb{R}$ is a continuous function such that it satisfies the H\"older condition with exponent $\beta$ with respect to the first variable, that is, there exists a constant $K_\beta$ such that $$ |k(t,\tau)-k(s,\tau)|\leq K_\beta\;|t-s|^\beta, $$ for any $t,s,\tau\in [0,1]$. \item[(iii)] $r:[0,1]\to[0,1]$ is a measurable function. \item[(iv)] The following inequality is satisfied $$ \|p\|_\beta(2K+K_\beta)<\frac{1}{4}, $$ where the constant $K$ is defined by $$ K=\sup\big\{\int_0^1 |k(t,\tau)|\,d\tau:t\in[0,1]\big\}, $$ which exists by (ii). \end{enumerate} \begin{theorem}\label{t3.1} Under assumptions {\rm (i)--(iv)}, Equation \eqref{e1.1} has at least one solution belonging to the space $H_\alpha[0,1]$, where $\alpha$ is arbitrarily fixed number satisfying $0<\alpha<\beta$. \end{theorem} \begin{proof} Consider the operator $\mathcal{T}$ defined on $H_\beta[0,1]$ by $$ (\mathcal{T}x)(t)=p(t)+x(t)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau,\quad t\in[0,1]. $$ In the sequel, we will prove that $\mathcal{T}$ transforms the space $H_\beta[0,1]$ into itself. In fact, we take $x\in H_\beta[0,1]$ and $t,s\in[0,1]$ with $t\neq s$. Then, by assumptions (i) and (ii), we obtain \begin{align*} &\frac{|(\mathcal{T}x)(t)-(\mathcal{T}x)(s)|}{|t-s|^\beta}\\ &=\frac{\big|p(t)+x(t)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau-p(s) -x(s)\int_0^1 k(s,\tau)\;x(r(\tau))\,d\tau\big|}{|t-s|^\beta} \\ &\leq \frac{|p(t)-p(s)|}{|t-s|^\beta} +\frac{\big|x(t)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau -x(s)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau\big|}{|t-s|^\beta}\\ &\quad +\frac{\big|x(s)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau -x(s)\int_0^1 k(s,\tau)\;x(r(\tau))\,d\tau\big|}{|t-s|^\beta}\\ &\leq \frac{|p(t)-p(s)|}{|t-s|^\beta}+\frac{|x(t)-x(s)|}{|t-s|^\beta} \int_0^1 |k(t,\tau)|\;|x(r(\tau))|\,d\tau\\ &\quad +\frac{|x(s)|\int_0^1 |k(t,\tau)-k(s,\tau)|\, |x(r(\tau))|\,d\tau}{|t-s|^\beta}\\ &\leq \frac{|p(t)-p(s)|}{|t-s|^\beta} +\frac{|x(t)-x(s)|}{|t-s|^\beta}\|x\|_\infty\int_0^1 |k(t,\tau)|\,d\tau\\ &\quad +\frac{\|x\|_\infty\cdot\|x\|_\infty\int_0^1 |k(t,\tau)-k(s,\tau)| \,d\tau}{|t-s|^\beta}\\ &\leq \frac{|p(t)-p(s)|}{|t-s|^\beta} +K\|x\|_\infty\frac{|x(t)-x(s)|}{|t-s|^\beta} +\frac{\|x\|_\infty^2\int_0^1 K_\beta|t-s|^\beta\,d\tau}{|t-s|^\beta}\\ &\leq H_p^\beta+K\|x\|_\infty H_x^\beta+K_\beta\|x\|_\infty^2. \end{align*} By Lemma \ref{l2.1}, since $\|x\|_\infty\leq\|x\|_\beta$ and, as $H_x^\beta\leq\|x\|_\beta$, we infer that $$ \frac{|(\mathcal{T}x)(t)-(\mathcal{T}x)(s)|}{|t-s|^\beta} \leq H_p^\beta+(K+K_\beta)\|x\|_\beta^2. $$ Therefore, \begin{equation} \begin{aligned} \label{e3.1} \|\mathcal{T}x\|_\beta &=|(\mathcal{T}x)(0)|+\sup\big\{\frac{|(\mathcal{T}x)(t)-(\mathcal{T}x)(s)|} {|t-s|^\beta}:t,s\in[0,1],t\neq s\big\}\\ &\leq |(\mathcal{T}x)(0)|+H_p^\beta+(K+K_\beta)\|x\|_\beta^2\\ &\leq |p(0)|+|x(0)|\int_0^1 |k(0,\tau)|\;|x(r(\tau))|\,d\tau +H_p^\beta+(K+K_\beta)\|x\|_\beta^2\\ &\leq \|p\|_\beta+\|x\|_\infty\cdot\|x\|_\infty\int_0^1 |k(0,\tau)|\,d\tau +(K+K_\beta)\|x\|_\beta^2\\ &\leq \|p\|_\beta+K\|x\|_\beta^2+(K+K_\beta)\|x\|_\beta^2\\ &=\|p\|_\beta+(2K+K_\beta)\|x\|_\beta^2<\infty. \end{aligned} \end{equation} This proves that the operator $\mathcal{T}$ maps $H_\beta[0,1]$ into itself. Taking into account that the inequality $$ \|p\|_\beta+(2K+K_\beta)r^20$. Suppose that $y\in B_{r_0}^\beta$ and $\|x-y\|_\alpha\leq\delta$, where $\delta$ is a positive number such that $\delta<\frac{\varepsilon}{2(2K+3K_\beta)r_0}$. Then, for any $t,s\in[0,1]$ with $t\neq s$, we have \begin{align*} &\frac{|[(\mathcal{T}x)(t)-(\mathcal{T}y)(t)]-[(\mathcal{T}x)(s)- (\mathcal{T}y)(s)]|}{|t-s|^\alpha}\\ &=\Big|\frac{\big[x(t)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau -y(t)\int_0^1 k(t,\tau)\;y(r(\tau))\,d\tau\big]}{|t-s|^\alpha}\\ &\quad -\frac{\big[x(s)\int_0^1 k(s,\tau)\;x(r(\tau))\,d\tau -y(s)\int_0^1 k(s,\tau)\;y(r(\tau))\,d\tau\big]}{|t-s|^\alpha} \Big|\\ &\leq\Big| \frac{\big[x(t)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau -y(t)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau\big]}{|t-s|^\alpha}\\ &\quad +\frac{\big[y(t)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau -y(t)\int_0^1 k(t,\tau)\;y(r(\tau))\,d\tau\big]}{|t-s|^\alpha}\\ &\quad -\frac{\big[x(s)\int_0^1 k(s,\tau)\;x(r(\tau))\,d\tau -y(s)\int_0^1 k(s,\tau)\;x(r(\tau))\,d\tau\big]}{|t-s|^\alpha}\\ &\quad -\frac{\big[y(s)\int_0^1 k(s,\tau)\;x(r(\tau))\,d\tau -y(s)\int_0^1k(s,\tau)\;y(r(\tau))\,d\tau\big]}{|t-s|^\alpha}\Big|\\ &=\frac{1}{|t-s|^\alpha}\Big|(x(t)-y(t))\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau +y(t)\int_0^1 k(t,\tau)(x(r(\tau)-y(r(\tau)))\,d\tau \\ &\quad -(x(s)-y(s))\int_0^1 k(s,\tau)\;x(r(\tau))\,d\tau -y(s)\int_0^1 k(s,\tau)\;(x(r(\tau))-y(r(\tau)))\,d\tau\Big|\\ &\leq\frac{1}{|t-s|^\alpha}\Big\{ |(x(t)-y(t))-(x(s)-y(s))|\cdot\Big|\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau\Big|\\ &\quad +|x(s)-y(s)|\cdot\Big|\int_0^1(k(t,\tau)-k(s,\tau))\;x(r(\tau)) \,d\tau\dot|\\ &\quad +\Big|y(t)\int_0^1 k(t,\tau)(x(r(\tau)-y(r(\tau)))\,d\tau -y(s)\int_0^1 k(s,\tau)(x(r(\tau)-y(r(\tau)))\,d\tau\Big|\Big\} \\ &\leq\frac{|(x(t)-y(t))-(x(s)-y(s))|}{|t-s|^\alpha}\|x\|_\infty \int_0^1 |k(t,\tau)|\,d\tau\\ &\quad +\big[|(x(s)-y(s))-(x(0)-y(0))|+|x(0)-y(0)|\big]\|x\|_\infty\\ &\quad\times \int_0^1\frac{|k(t,\tau)-k(s,\tau)|}{|t-s|^\alpha}d\tau +\frac{1}{|t-s|^\alpha}\Big|y(t)\\ &\quad\times \int_0^1 k(t,\tau)(x(r(\tau) -y(r(\tau)))\,d\tau-y(s)\int_0^1 k(t,\tau)(x(r(\tau)-y(r(\tau)))\,d\tau\Big|\\ &\quad +\frac{1}{|t-s|^\alpha}\Big| y(s)\int_0^1 k(t,\tau)(x(r(\tau)-y(r(\tau)))\,d\tau\\ &\quad -y(s)\int_0^1 k(s,\tau) (x(r(\tau)-y(r(\tau)))\,d\tau\Big| \\ &\leq K \|x-y\|_\alpha\|x\|_\infty+\sup_{p,q\in[0,1]}|(x(p)-y(p))-(x(q)-y(q))|\\ &\quad\times \|x\|_\infty\int_0^1\frac{K_\beta|t-s|^\beta}{|t-s|^\alpha}d\tau +|x(0)-y(0)|\|x\|_\infty \int_0^1\frac{K_\beta|t-s|^\beta}{|t-s|^\alpha}d\tau\\ &\quad +\frac{|y(s)-x(s)|}{|t-s|^\alpha} \int_0^1 |k(t,\tau)|\;|x(r(\tau)-y(r(\tau))|\,d\tau\\ &\quad +|y(s)|\int_0^1\frac{|k(t,\tau)-k(s,\tau)|}{|t-s|^\alpha} |x(r(\tau)-y(r(\tau))|\,d\tau \\ &\leq K\|x\|_\infty\|x-y\|_\alpha +\|x\|_\infty K_\beta|t-s|^{\beta-\alpha}\\ &\quad\times \sup_{p,q\in[0,1],\;p\neq q} \big\{\frac{|(x(p)-y(p))-(x(q)-y(q))|}{|p-q|^\alpha}|p-q|^\alpha\big\}\\ &\quad +K_\beta\|x\|_\beta|t-s|^{\beta-\alpha}|x(0)-y(0)|\\ &\quad +K H_y^\alpha\|x-y\|_\infty+\|y\|_\infty\;\|x-y\|_\infty \int_0^1\frac{K_\beta\;|t-s|^\beta}{|t-s|^\alpha}d\tau\\ &\leq K\|x\|_\beta\|x-y\|_\alpha+2K_\beta\|x\|_\beta\|x-y\|_\alpha+ K\|y\|_\alpha\|x-y\|_\alpha+K_\beta\|y\|_\alpha\|x-y\|_\alpha\\ &\leq\big( K\|x\|_\beta+2K_\beta\|x\|_\beta+K\|y\|_\alpha+K_\beta\|y\|_\alpha\big) \|x-y\|_\alpha. \end{align*} Since $\|y\|_\alpha\leq\|y\|_\beta$ (see, Lemma \ref{l2.2}) and $x,y\in B_{r_0}^\beta$, from the above inequality we infer that \begin{equation} \label{e3.2} \begin{aligned} \frac{|[(\mathcal{T}x)(t)-(\mathcal{T}y)(t)]-[(\mathcal{T}x)(s)- (\mathcal{T}y)(s)]|}{|t-s|^\alpha} &\leq(2Kr_0+3K_\beta r_0)\|x-y\|_\alpha\\ &\leq(2Kr_0+3K_\beta r_0)\delta<\frac{\varepsilon}{2}. \end{aligned} \end{equation} On the other hand, \begin{equation} \label{e3.3} \begin{aligned} |(\mathcal{T}x)(0)-(\mathcal{T}y)(0)| &= \Big|x(0)\int_0^1k(0,\tau)\;x(r(\tau))\,d\tau-y(0)\int_0^1k(0,\tau) y(r(\tau))\,d\tau\Big|\\ &\leq \Big|x(0)\int_0^1k(0,\tau)\;x(r(\tau))\,d\tau -x(0)\int_0^1k(0,\tau)\;y(r(\tau))\,d\tau\Big|\\ &\quad +\Big|x(0)\int_0^1k(0,\tau)\;y(r(\tau))\,d\tau-y(0) \int_0^1k(0,\tau)\;y(r(\tau))\,d\tau\Big|\\ &\leq \Big| x(0)\int_0^1k(0,\tau)(x(r(\tau))-y(r(\tau)))\,d\tau\Big|\\ &\quad +\Big| (x(0)-y(0))\int_0^1k(0,\tau)\;y(r(\tau)))\,d\tau\Big|\\ &\leq K\|x\|_\infty\|x-y\|_\infty+K\|y\|_\infty\|x-y\|_\alpha\\ &\leq K\|x\|_\beta\|x-y\|_\alpha+K\|y\|_\beta\|x-y\|_\alpha\\ &\leq 2Kr_0\|x-y\|_\alpha<2Kr_0\delta<\frac{\varepsilon}{2}. \end{aligned} \end{equation} From \eqref{e3.2} and \eqref{e3.3}, it follows that \begin{align*} &\|\mathcal{T}x-\mathcal{T}y\|\\ &=|(\mathcal{T}x)(0)-(\mathcal{T}y)(0)|\\ &\quad +\sup\big\{ \frac{|((\mathcal{T}x)(t)-(\mathcal{T}y)(t))- ((\mathcal{T}x)(s)-(\mathcal{T}y)(s))|}{|t-s|^\alpha}:t,s\in[0,1], \;t\neq s\big\}\\ &< \frac{\varepsilon}{2}+\frac{\varepsilon}{2}=\varepsilon. \end{align*} This proves that the operator $\mathcal{T}$ is continuous at the point $x\in B_{r_0}^\delta$ for the norm $\|\|_\alpha$. Since $B_{r_0}^\delta$ is compact in $H_\alpha[0,1]$, applying the classical Schauder fixed point theorem we obtain the desired result. \end{proof} \section{Example} To present an example illustrating our result we need some previous results. \begin{definition}\label{d3.2} \rm A function $f:\mathbb{R}_+\to \mathbb{R}_+$ is said to be subadditive if $f(x+y)\leq f(x)+f(y)$ for any $x,y\in\mathbb{R}_+$. \end{definition} \begin{lemma}\label{l3.3} Suppose that $f:\mathbb{R}_+\to \mathbb{R}_+$ is subadditive and $y\leq x$ then $f(x)-f(y)\leq f(x-y)$. \end{lemma} \begin{proof} Since $f(x)=f(x-y+y)\leq f(x-y)+f(y)$ the result follows. \end{proof} \begin{remark}\label{r3.4} \rm From Lemma \ref{l3.3}, we infer that if $f:\mathbb{R}_+\to \mathbb{R}_+$ is subadditive then $|f(x)-f(y)|\leq f(|x-y|)$ for any $x,y\in\mathbb{R}_+$. \end{remark} \begin{lemma}\label{l3.5} Let $f:\mathbb{R}_+\to \mathbb{R}_+$ be a concave function with $f(0)=0$. Then $f$ is subadditive. \end{lemma} \begin{proof} For $x,y\in\mathbb{R}_+$ and, since $f$ is concave and $f(0)=0$, we have \begin{align*} f(x)&=f\left(\frac{x}{x+y}(x+y)+\frac{y}{x+y}\cdot 0\right)\\ &\geq\frac{x}{x+y}f(x+y)+\frac{y}{x+y}f(0)\\ &= \frac{x}{x+y}f(x+y) \end{align*} and \begin{align*} f(y)&=f\left(\frac{x}{x+y}\cdot 0+\frac{y}{x+y}(x+y)\right)\\ &\geq \frac{x}{x+y}f(0)f(x+y)+\frac{y}{x+y}f(x+y)\\ &= \frac{y}{x+y}f(x+y). \end{align*} Adding these inequalities, we obtain $$ f(x)+f(y)\geq \frac{x}{x+y}f(x+y)+\frac{y}{x+y}f(x+y)=f(x+y). $$ This completes the proof. \end{proof} \begin{remark}\label{r3.6} \rm Let $f:\mathbb{R}_+\to \mathbb{R}_+$ be the function defined by $f(x)=\sqrt[p]{x}$, where $p>1$. Since this function is concave (because $f''(x)\leq 0$ for $x>0$) and $f(0)=0$, Lemma \ref{l3.5} says us that $f$ is subadditive. By Remark \ref{r3.4}, we have $$ |f(x)-f(y)|=|\sqrt[p]{x}-\sqrt[p]{y}|\leq \sqrt[p]{|x-y|} $$ for any $x,y\in\mathbb{R}_+$. \end{remark} \begin{example}\rm Let us consider the quadratic integral equation \begin{equation}\label{e3.4} x(t)=\arctan \sqrt[5]{q\sin t+\hat{q}}+x(t)\int_0^1 \sqrt[4]{m t^2 +\tau}\;x\big(\frac{\tau}{\tau+1}\big)\,d\tau,\quad t\in[0,1], \end{equation} where, $q$, $\hat{q}$ and $m$ are nonnegative constants. Notice that \eqref{e3.4} is a particular case of \eqref{e1.1}, where $p(t)=\arctan \sqrt[5]{q\sin t+\hat{q}}$, $k(t,\tau)=\sqrt[4]{m t^2+\tau}$ and $r(\tau)=\frac{\tau}{\tau+1}$. In what follows, we will prove that assumptions (i)-(iv) of Theorem \ref{t3.1} are satisfied. Since the inverse tangent function is concave (because its second derivative is nonpositive) and its value at zero is zero, taking into account Remarks \ref{r3.4} and \ref{r3.6}, we have \begin{align*} |p(t)-p(s)| &= \Big|\arctan \sqrt[5]{q\sin t+\hat{q}} -\arctan \sqrt[5]{q\sin s+\hat{q}}\Big|\\ &\leq \arctan\Big(\big|\sqrt[5]{q\sin t+\hat{q}}-\sqrt[5]{q\sin s+\hat{q}}\big| \Big)\\ &\leq \big|\sqrt[5]{q\sin t+\hat{q}}-\sqrt[5]{q\sin s+\hat{q}}\big|\\ &\leq \sqrt[5]{q|\sin t-\sin s|}\\ &\leq \sqrt[5]{q}\;|t-s|^{1/5}, \end{align*} where we have used that $\arctan x\leq x$ for $x\geq 0$ and $|\sin x-\sin y| \leq |x-y|$ for any $x,y\in\mathbb{R}$. This says that $p\in H_{\frac{1}{5}}[0,1]$ and, moreover, $H_p^{1/5}=\sqrt[5]{q}$. Therefore, assumptions (i) of Theorem \ref{t3.1} is satisfied. Note that \begin{align*} \|p\|_{\frac{1}{5}} &=|p(0)|+\sup\big\{ \frac{|p(t)-p(s)|}{|t-s|^{1/5}}: t,s\in[0,1],\;t\neq s\big\}\\ &\leq \arctan\sqrt[5]{\hat{q}}+H_p^{1/5}\\ &=\arctan\sqrt[5]{\hat{q}}+\sqrt[5]{q}. \end{align*} Since for any $t,s,\tau\in[0,1]$, we have (see, Remark \ref{r3.6}) \begin{align*} |k(t,\tau)-k(s,\tau)| &=\big|\sqrt[4]{m t^2+\tau}-\sqrt[4]{m s^2+\tau}\big|\\ &\leq \sqrt[4]{|m t^2-m s^2|}\\ &=\sqrt[4]{m}\sqrt[4]{|t^2-s^2|}\\ &=\sqrt[4]{m}\sqrt[4]{t+s}\sqrt[4]{|t-s|}\\ &\leq \sqrt[4]{m}\sqrt[4]{2}\;|t-s|^{1/4}\\ &=\sqrt[4]{m}\sqrt[4]{2}\;|t-s|^{1/20}\;|t-s|^{1/5}\\ &\leq \sqrt[4]{2m}\;|t-s|^{1/5}, \end{align*} assumption $(ii)$ of Theorem \ref{t3.1} is satisfied with $K_\beta=K_{\frac{1}{5}}=\sqrt[4]{2m}$. It is clear that $r(\tau)=\frac{\tau}{\tau+1}$ satisfies assumption (iii). In our case, the constant $K$ is given by \begin{align*} K&=\sup\big\{\int_0^1 |k(t,\tau)|\,d\tau:t\in[0,1]\big\}\\ &=\sup\big\{\int_0^1 \sqrt[4]{m t^2+\tau}\,d\tau:t\in[0,1]\big\}\\ &=\sup\big\{\frac{4}{5}\Big(\sqrt[4]{(m t^2+1)^5}-\sqrt[4]{m^5 t^{10}}\Big)\big\}\\ &= \frac{4}{5}\Big(\sqrt[4]{(m+1)^5}-\sqrt[4]{m^5}\Big). \end{align*} Therefore, the inequality appearing in assumption $(iv)$ takes the form $$ \|p\|_{\frac{1}{5}}(2K+K_\beta) =\Big(\arctan\sqrt[5]{\hat{q}}+\sqrt[5]{q}\Big) \Big(\frac{8}{5}\big[\sqrt[4]{(m+1)^5}-\sqrt[4]{m^5}\big] +\sqrt[4]{2m}\Big)<\frac{1}{4}. $$ It is easily seen that the above inequality is satisfied when, for example, $\hat q=0$, $q=\frac{1}{2^{20}}$ and $m=1$. Therefore, using Theorem \ref{t3.1}, we infer that \eqref{e3.4} for $\hat q=0$, $q=\frac{1}{2^{20}}$ and $m=1$ has at least one solution in the space $H_\alpha[0,1]$ with $0<\alpha<1/5$. \end{example} Note that in \eqref{e3.4}, we can take as $r(\tau)$ a particular functions such as $r(\tau)=\{e^\tau\}$, where $\{\cdot\}$ denotes the fractional part. \begin{remark}\label{r3.8} \rm Note that any solution $x(t)$ of \eqref{e1.1}, i.e., $$ x(t)=p(t)+x(t)\int_0^1 k(t,\tau)\;x(r(\tau))\,d\tau,\quad t\in[0,1], $$ satisfies that its zeroes are also zeroes of $p(t)$. From this, we infer that if $p(t)\neq 0$ for any $t\in[0,1]$ then $x(t)\neq 0$ for any $t\in[0,1]$. By Bolzano's theorem, this means that the solution $x(t)$ of Eq\eqref{e1.1} does not change of sign on $[0,1]$ when $p(t)\neq 0$ for any $t\in[0,1]$. These questions seem to be interesting from a practical standpoint. \end{remark} \section{Appendix} Suppose that $0<\alpha<\beta\leq 1$ and by $B_r^\beta$ we denote the ball centered at $\theta$ and radius $r$ in the space $H_\beta[a,b]$; i.e., $B_r^\beta=\{x\in H_\beta[a,b]:\|x\|_\beta\leq r\}$. Then $B_r^\beta$ is a compact subset in the space $H_\alpha[a,b]$. In fact, by Theorem \ref{t2.3}, since $B_r^\beta$ is a bounded subset in $H_\beta[a,b]$, $B_r^\beta$ is a relatively compact subset of $H_\alpha[a,b]$. In the sequel, we will prove that $B_r^\beta$ is a closed subset of $H_\alpha[a,b]$. Suppose that $(x_n)\subset B_r^\beta$ and $x_n\stackrel{\|\|_\alpha}{\longrightarrow}x$ with $x\in H_\alpha[a,b]$. We have to prove that $x\in B_r^\beta$. Since $x_n\stackrel{\|\cdot\|_\alpha}{\longrightarrow}x$, for $\varepsilon>0$ given we can find $n_0\in \mathbb{N}$ such that $\|x_0-x\|_\alpha\leq \varepsilon$ for any $n\geq n_0$, or, equivalently, \begin{equation}\label{e3.5} |x_n(a)-x(a)|+\sup\big\{\frac{|(x_n(t)-x(t))-(x_n(s)-x(s))|}{|t-s|^\alpha} :t,s\in[a,b],\;t\neq s\big\}<\varepsilon, \end{equation} for any $n\geq n_0$. Particularly, this implies that $x_n(a)\to x(a)$. Moreover, if in \eqref{e3.5} we put $s=a$ then we get $$ \sup\big\{\frac{|(x_n(t)-x(t))-(x_n(a)-x(a))|}{|t-a|^\alpha} :t,s\in[a,b],\;t\neq a\big\}<\varepsilon,\;{\rm for}\;{\rm any}\;n\geq n_0.$$ This says that \begin{equation}\label{e3.6} |(x_n(t)-x(t))-(x_n(a)-x(a))|<\varepsilon|t-a|^\alpha, \quad\text{for any $n\geq n_0$ and for any $t\in[a,b]$}. \end{equation} Therefore, for any $n\geq n_0$ and any $t\in[a,b]$ by \eqref{e3.5} and \eqref{e3.6}, we have \begin{align*} |x_n(t)-x(t)| &\leq |(x_n(t)-x(t))-(x_n(a)-x(a))|+|x_n(a)-x(a)|\\ &<\varepsilon(t-a)^\alpha+\varepsilon\\ &=\varepsilon(1+(b-a)^\alpha). \end{align*} From this, it follows that \begin{equation}\label{e3.7} \|x_n-x\|_\infty\to 0. \end{equation} Next, we will prove that $x\in B_r^\beta$. In fact, as $(x_n)\subset B_r^\beta\subset H_\beta[a,b]$, we have that $$ \frac{|x_n(t)-x_n(s)|}{|t-s|^\beta}\leq r $$ for any $t,s\in[a,b]$ with $t\neq s$. Consequently, $$ |x_n(t)-x_n(s)|\leq r|t-s|^\beta $$ for any $t,s\in[a,b]$. Letting $n\to\infty$ and taking into account \eqref{e3.7}, we obtain $$ |x(t)-x(s)|\leq r|t-s|^\beta $$ for any $t,s\in[a,b]$. Therefore, $$ \frac{|x(t)-x(s)|}{|t-s|^\beta}\leq r $$ for any $t,s\in[a,b]$ with $t\neq s$, and this means that $x\in B_r^\beta$. This completes the proof. \begin{thebibliography}{0} \bibitem{Ba} C.~Baco\c{t}iu; Volterra-Fredholm nonlinear systems with modified argument via weakly Picard operators theory, \emph{Carpath. J. Math.} 24(2) (2008), 1--19. \bibitem{BaNa} J.~Bana\'s, R.~Nalepa; On the space of functions with growths tempered by a modulus of continuity and its applications, \emph{J. Func. Spac. Appl.}, (2013), Article ID 820437, 13 pages. \bibitem{BeDa} M.~Benchohra, M.A.~Darwish; On unique solvability of quadratic integral equations with linear modification of the argument, \emph{Miskolc Math. Notes} 10(1) (2009), 3--10. \bibitem{CaLoSa} J.~Caballero, B.~L\'{o}pez, K.~Sadarangani; Existence of nondecreasing and continuous solutions of an integral equation with linear modification of the argument, \emph{Acta Math. Sin. (English Series)} 23 (2007), 1719--1728. \bibitem{Do} M.~Dobri\c{t}oiu; Analysis of a nonlinear integral equation with modified argument from physics, \emph{Int. J. Math. Models and Meth. Appl. Sci.} 3(2) (2008), 403--412. \bibitem{KaMc} T.~Kato, J.B.~Mcleod; The functional-differential equation $y'(x)=ay(\lambda x)+by(x)$, \emph{Bull. Amer. Math. Soc.}, 77 (1971), 891--937. \bibitem{La} M.~Lauran; Existence results for some differential equations with deviating argument, \emph{Filomat} 25(2) (2011), 21--31. \bibitem{Mu1} V.~Mure\c{s}an; A functional-integral equation with linear modification of the argument, via weakly Picard operators, \emph{Fixed Point Theory}, 9(1) (2008), 189--197. \bibitem{Mu2} V.~Mure\c{s}an; A Fredholm-Volterra integro-differential equation with linear modification of the argument, \emph{J. Appl. Math.}, 3(2) (2010), 147--158. \end{thebibliography} \end{document}