\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2007(2007), No. 29, pp. 1--13.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu (login: ftp)} \thanks{\copyright 2007 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2007/29\hfil $\Psi$-conditional asymptotic stability] {On the $\Psi$-conditional asymptotic stability of the solutions of a nonlinear Volterra integro-differential system} \author[A. Diamandescu\hfil EJDE-2007/29\hfilneg]{Aurel Diamandescu} \address{Aurel Diamandescu \newline Department of Applied Mathematics, University of Craiova, 13, ``Al. I. Cuza'' st., 200585, Craiova, Romania} \email{adiamandescu@central.ucv.ro} \thanks{Submitted October 4, 2006. Published February 14, 2007.} \subjclass[2000]{45M10, 45J05} \keywords{$\Psi$-stability; $\Psi$-conditional asymptotic stability} \begin{abstract} We provide sufficient conditions for $\Psi$-conditional asymptotic stability of the solutions of a nonlinear Volterra integro-differential system. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{definition}[theorem]{Definition} \newtheorem{example}[theorem]{Example} \newtheorem{remark}[theorem]{Remark} \newtheorem{corollary}[theorem]{Corollary} \section{Introduction} The purpose of this paper is to provide sufficient conditions for $\Psi $-conditional asymptotic stability of the solutions of the nonlinear Volterra integro-differential system \begin{equation} \label{e1} x'=A(t)x+\int_0^tF(t,s,x(s))ds \end{equation} and for the linear system \begin{equation} x'=[ A(t)+B(t)] x \label{e2} \end{equation} as a perturbed systems of \begin{equation} y'=A(t)y. \label{e3} \end{equation} We investigate conditions on a fundamental matrix $Y(t)$ of the linear equation \eqref{e3} and on the functions $B(t)$ and $F(t,s,x)$ under which the solutions of \eqref{e1}, \eqref{e2} or \eqref{e3} are $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. Here, $\Psi $ is a continuous matrix function. The introduction of the matrix function $\Psi $ permits to obtain a mixed asymptotic behavior of the solutions. The problem of $\Psi $- stability for systems of ordinary differential equations has been studied by many authors, as e.g. Akinyele \cite{a1,a2}, Constantin \cite{c1,c2}, Hallam \cite{h1}, Kuben \cite{k1}, Morchalo \cite{m2}. In these papers, the function $\Psi $ is a scalar continuous function (and monotone in \cite{a2}, nondecreasing in \cite{c1}). In our papers \cite{d1,d2,d3}, we have proved sufficient conditions for various types of $\Psi $-stability of the trivial solution of the equations \eqref{e1}, \eqref{e2} and \eqref{e3}. In these papers, the function $\Psi $ is a continuous matrix function. Recent works for stability of solutions of \eqref{e1} have been by Avramescu \cite{a3}, by Hara, Yoneyama and Itoh \cite{h2}, by Lakshmikantham and Rama Mohana Rao \cite{l1}, by Mahfoud \cite{m1} and others. Coppel's paper \cite[Chapter III, Theorem 12]{c3}, \cite{c4} deal with the instability and conditional asymptotic stability of the solutions of a systems of differential equations. Sp\"{a}th's paper \cite{s2} and Weyl's paper \cite{w1} deal with the conditional stability of solutions of systems of differential equations. In our papers \cite{d4,d5}, we have proved a necessary and sufficient conditions for $\Psi $-instability and $\Psi $-conditional stability of the equation \eqref{e3} and sufficient conditions for $\Psi $-instability and $\Psi $-conditional stability of trivial solution of the equations \eqref{e1} and \eqref{e2}. \section{Definitions, notation and hypotheses} Let $\mathbb{R}^d$ denote the Euclidean $d$-space. For $x = ( x_1, x_2,\dots x_d)^T \in R^d$, let $\| x\| = \max \{ | x_1|, |x_2| ,\dots | x_d| \}$ be the norm of $x$. For a $d\times d$ matrix $A = (a_{ij})$, we define the norm $A$ by $| A| = \sup_{\| x\| \leq 1}\|Ax\| $; it is well-known that $| A| = \max_{1\leq i\leq d } \sum_{j=1}^d| a_{ij}| $. In the equations \eqref{e1}--\eqref{e3} we assume that $A(t)$ is a continuous $d\times d$ matrix on $\mathbb{R}_{+} = [0, \infty )$ and $F : D \times \mathbb{R}^{d}\to \mathbb{R}^d$, $D = \{ (t,s) \in \mathbb{R}^2: 0\leq s \leq t <\infty \}$, is a continuous $d$-vector with respect to all variables. Let $\Psi _i : \mathbb{R}_{+}\to (0,\infty )$, $i = 1, 2, \dots d$, be a continuous functions and \begin{equation*} \Psi = \mathop{\rm diag }[\Psi _1, \Psi _2,\dots \Psi_d]. \end{equation*} A matrix $P$ is said to be a projection matrix if $P^{2} = P$. If $P$ is a projection, then so is $I - P$. Two such projections, whose sum is $I$ and whose product is $0$, are said to be supplementary. \begin{definition} \label{def2.1} \rm The solution $x(t)$ of \eqref{e1} is said to be $\Psi $-stable on $\mathbb{R}_{+}$, if for every $\varepsilon > 0$ and any $t_0\geq $ 0, there exists a $\delta =\delta (\varepsilon ,t_0) > 0$ such that any solution $\widetilde{x}$(t) of \eqref{e1} which satisfies the inequality $\| \Psi (t_0)(\widetilde{x}(t_0)-x(t_0))\| < \delta (\varepsilon,t_0)$ exists and satisfies the inequality $\| \Psi (t)(\widetilde{x} (t)-x(t))\| < \varepsilon $ for all $t \geq t_0$. Otherwise, is said that the solution x(t) is $\Psi$-unstable on $\mathbb{R}_{+}$. \end{definition} \begin{definition} \label{def2.2} \rm A function $\varphi : \mathbb{R}_{+}\to \mathbb{R}^d$ is said to be $\Psi $-bounded on $\mathbb{R}_{+}$ if $\Psi (t)\varphi (t)$ is bounded on $\mathbb{R}_{+}$. \end{definition} \begin{remark} \rm For $\Psi _{i} = 1$, $i = 1, 2,\dots d$, we obtain the notion of classical stability, instability and boundedness, respectively. \end{remark} \begin{definition} \label{def2.3}\rm The solution $x(t)$ of \eqref{e1} is said to be $\Psi $-conditionally stable on $\mathbb{R}_{+}$ if it is not $\Psi $-stable on $\mathbb{R}_{+}$ but there exists a sequence ($x_{n}(t)$) of solutions of \eqref{e1} defined for all $t\geq $ 0 such that \begin{equation*} \lim_{n\to \infty }\Psi (t)x_n(t) = \Psi(t)x(t),\quad \text{uniformly on }R_{+}. \end{equation*} If the sequence $x_n(t)$ can be chosen so that \begin{equation*} \lim_{t\to \infty } \Psi (t)(x_{n}(t) -x(t)) = 0,\quad\text{for } n = 1, 2, \dots \end{equation*} then $x(t)$ is said to be $\Psi $-conditionally asymptotically stable on $R_{+}$. \end{definition} \begin{remark} \rm (1) It is easy to see that if $| \Psi (t)| $ and $| \Psi ^{-1}(t)| $ are bounded on $\mathbb{R}_{+}$, then the $\Psi $-conditional asymptotic stability is equivalent with the classical conditional asymptotic stability. (2) In the same manner as in classical conditional asymptotic stability, we can speak about $\Psi $-conditional asymptotic stability of a linear equation. Indeed, let $x(t)$, $y(t)$ be two solutions of the linear equation \eqref{e3}. We suppose that $x(t)$ is $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. Then $y(t)$ is $\Psi$-unstable on $\mathbb{R}_{+}$ (see \cite[Theorem 1]{d4}) and \begin{gather*} \lim_{n\to \infty }\Psi (t)y_n(t) = \Psi (t)y(t),\quad \text{uniformly on }\mathbb{R}_{+}, \\ \lim_{t\to \infty }\Psi (t)( y_{n}(t) -y(t) ) = 0,\quad \text{for }n = 1, 2, \dots \end{gather*} where y$_{n}(t) = x_{n}(t) - x(t) + y(t)$, $n \in N$ are solutions of the linear equation \eqref{e3}. Thus, all solutions of \eqref{e3} are $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. \end{remark} \section{$\Psi $-conditional asymptotic stability of linear equations} In this section we give necessary and sufficient conditions for the $\Psi $-conditional asymptotic stability of the linear equation \eqref{e3} and sufficient conditions for the $\Psi $-conditional asymptotic stability of the linear equations \eqref{e3} and \eqref{e2}. \begin{theorem} \label{thm3.1} The linear equation \eqref{e3} is $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$ if and only if it has a $\Psi $-unbounded solution on $\mathbb{R}_{+}$ and a non-trivial solution $y_0(t)$ such that $\lim_{t\to \infty }\Psi (t)y_0(t) = 0$. \end{theorem} \begin{proof} Let $Y(t)$ be a fundamental matrix for \eqref{e3}. Suppose that the linear equation \eqref{e3} is $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. From Definition \ref{def2.3} and \cite[Theorem 3.1]{d1}, it follows that $|\Psi $(t)Y(t)$|$ is unbounded on $\mathbb{R}_{+}$. Thus, the linear equation \eqref{e3} has at least one $\Psi $-unbounded solution on $\mathbb{R}_{+}$. In addition, there exists a sequence $(y_{n}(t))$ of non-trivial solutions of \eqref{e3} such that $\lim_{n\rightarrow \infty }\Psi (t)y_{n}(t)=0$, uniformly on $\mathbb{R}_{+}$ and $\lim_{t\rightarrow \infty }\Psi (t)y_{n}(t)=0$ for $n=1,2,\dots $. The proof of the ``only if'' part is complete. Suppose, conversely, that \eqref{e3} has at least one $\Psi $-unbounded solution on $\mathbb{R}_{+}$ and at least one non-trivial solution $y_0(t)$ such that $\lim_{t\to \infty }\Psi (t)y_0(t) =0$. It follows that the matrix $\Psi (t)Y(t)$ is unbounded on $\mathbb{R}_{+}$. Consequently, the linear equation \eqref{e3} is $\Psi $-unstable on $\mathbb{R}_{+}$ (See \cite[Theorem 1]{d4}). On the other hand, $( \frac{1}{n}y_0(t)) $ is a sequence of solutions of \eqref{e3} such that $\lim_{n\to \infty }\frac{1}{n}\Psi (t)y_0(t) = 0$, uniformly on $\mathbb{R}_{+}$ and $\lim_{t\to \infty }\frac{1}{n}\Psi (t)y_0(t) = 0$ for $n \in \mathbb{N}$. Thus, the linear equation \eqref{e3} is $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. The proof is complete. \end{proof} We remark that Theorem \ref{thm3.1} generalizes a similar result in connection with the classical conditional asymptotic stability in \cite{c3}. The conditions for $\Psi $-conditional asymptotic stability of the linear equation \eqref{e3} can be expressed in terms of a fundamental matrix for \eqref{e3}. \begin{theorem} \label{thm3.2} Let $Y(t)$ be a fundamental matrix for \eqref{e3}. Then, the linear equation \eqref{e3} is $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$ if and only if there are satisfied two following conditions: \begin{itemize} \item[(a)] There exists a projection $P_1$ such that $\Psi (t)Y(t)P_1$ is unbounded on $\mathbb{R}_{+}$; \item[(b)] there exists a projection $P_2 \neq 0$ such that $\lim_{t\to \infty }\Psi (t)Y(t)P_2 = 0$. \end{itemize} \end{theorem} \begin{proof} First, we shall prove the sufficiency. From the hypoyhesis (a) and \cite[Theorem 1]{d4}, it follows that the linear equation \eqref{e3} is $\Psi$-unstable on $\mathbb{R}_{+}$. Let $y(t)$ be a non-trivial solution on $\mathbb{R}_{+}$ of the linear equation \eqref{e3}. Let $(\lambda _{n})$ be such that $\lambda _{n}\in \mathbb{R}\setminus \{ 1\}$, $\lim_{n\to \infty }\lambda _{n} = 1$ and let $(y_{n} )$ be defined by \begin{equation*} y_{n}(t) = Y(t)P_{2} Y^{-1}(0)(\lambda _{n} y(0)) + Y(t)(I -P_{2})Y^{-1}(0)y(0), t \geq 0. \end{equation*} It is easy to see that $y_n(t)$, $n \in N$, are solutions of the linear equation \eqref{e3}. For $n \in N$ and $t \geq 0$, we have \begin{align*} \| \Psi (t)y_{n}(t) -\Psi (t)y(t)\| &= \| \Psi (t)Y(t)P_{2} Y^{-1}(0)((\lambda_{n}- 1)y(0))\| \\ & \leq | \lambda _{n}-1| | \Psi (t)Y(t)P _{2}| \| Y^{-1}(0)y(0)\| \end{align*} Thus, \begin{gather*} \lim_{n\to \infty } \Psi (t)y_n(t) = \Psi (t)y(t), \quad\text{uniformly on } \mathbb{R}_{+}, \\ \lim_{t\to \infty }\Psi (t)( y_{n}(t) - y(t) ) = 0, \quad\text{for }n = 1, 2, \dots . \end{gather*} It follows that the linear equation \eqref{e3} is $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. Now, we shall prove the necessity. From $\Psi $-conditional asymptotic stability on $\mathbb{R}_{+}$ of \eqref{e3}, it follows that $\Psi (t)Y(t)$ is unbounded on $\mathbb{R}_{+}$ (see \cite[Theorem 1]{d4}. In addition, there exists a non-trivial solution $y_0(t)$ on $\mathbb{R}_{+}$ of \eqref{e3} such that $\lim_{t\to \infty }\Psi (t)y_0(t) = 0$. Thus, there exists a constant vector $c \neq 0$ such that $\Psi (t)Y(t)c$ is such that $\lim_{t\to \infty }\Psi (t)Y(t)c = 0$. Let $c_s = \| c\| $. Let $P_2$ be the null matrix in which the $s$-th column is replaced with $\| c\| ^{-1}c$. Thus, $P_2$ is a projection and $\lim_{t\to \infty }\Psi (t)Y(t)P_2 =0$. The proof is now complete. \end{proof} A sufficient condition for $\Psi $-conditional asymptotic stability is given by the following theorem. \begin{theorem} \label{thm3.3} If there exist two supplementary projections $P_{1}$, $P_{2}$, $P_{i} \neq 0$, and a positive constant $K$ such that the fundamental matrix $Y(t)$ of the equation \eqref{e3} satisfies the condition \begin{equation*} \int_0^{{t}}| \Psi (t)Y(t)P_1 Y^{-1}(s) \Psi ^{-1}(s)| ds+\int_{t}^\infty |\Psi (t)Y(t)P_2 Y^{-1}(s)\Psi ^{-1}(s)| {ds}\leq { K} \end{equation*} for all $t \geq 0$, then, the linear equation \eqref{e3} is $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. \end{theorem} The proof of the above theorem follows from \cite[Theorem 2 and Lemmas 1, 2]{d4}. \begin{theorem} \label{thm3.4} Suppose that: \begin{enumerate} \item There exist supplementary projections $P_1$, $P_2$, $P_i \neq 0$, and a constant $K>0$ such that the fundamental matrix $Y(t)$ of \eqref{e3} satisfies the conditions \begin{gather*} | \Psi (t)Y(t)P_1 Y^{-1}(s)\Psi ^{-1}(s)| \leq K, \quad\text{for }0 \leq s \leq t, \\ | \Psi (t)Y(t)P_2 Y^{-1}(s)\Psi ^{-1}(s)| \leq K, \quad\text{for }0 \leq t \leq s. \end{gather*} \item $\lim_{t\to \infty }\Psi (t)Y(t)P_1=0$. \item $B(t)$ is a $d\times d$ continuous matrix function on $\mathbb{R}_{+}$ such that \begin{equation*} \int_0^\infty | \Psi {(t)B(t)}\Psi^{-1}{(t)}| dt \quad\text{is convergent.} \end{equation*} \item The linear equations \eqref{e2} and \eqref{e3} are $\Psi$-unstable on $\mathbb{R}_{+}$. \end{enumerate} Then \eqref{e2} is $\Psi$-conditionally asymptotically stable on $\mathbb{R}_{+}$. \end{theorem} \begin{proof} We choose t$_0\geq 0$ sufficiently large so that \begin{equation*} q = K\int_{t_0}^\infty | \Psi {(t)B(t)}\Psi ^{-1} (t)| dt < 1. \end{equation*} We put \begin{equation*} S = \{ x : t_0,\infty )\to \mathbb{R}^d:x \text{ is continuous and $\Psi$-bounded on }[t_0,\infty ) \}. \end{equation*} Define on the set $S$ a norm by \begin{equation*} ||| x||| = \sup_{t\geq t_0} \| \Psi (t)x(t)\| . \end{equation*} It is well known that $( S, ||| \cdot |||) $ is a Banach real space. For $x \in S$, we define \begin{equation*} ( {Tx}) (t) = \int_{t_0}^t{Y(t)P}_1 Y^{-1}(s)B(s)x(s)ds - \int_t^\infty Y(t)P_2 Y^{-1} (s)B(s)x(s)ds, \quad t \geq t_0. \end{equation*} It is easy to see that $(Tx)(t)$ exists and is continuous for $t \geq t_0$ (see the Proof of \cite[Theorem 3]{d5}). We have \begin{align*} \| \Psi (t)( Tx) (t)\| &\leq {\ K}\int_{t_0}^\infty | \Psi (s)B(s)\Psi ^{-1}(s)| \| \Psi (s)x(s)\| ds \\ &\leq {\ q }\sup_{t\geq t_0}\| \Psi (t)x(t)\| = q||| x||| ,\quad\text{for }t \geq t_0. \end{align*} This shows that $TS \subseteq S$. On the other hand, $T$ is linear and \begin{equation*} ||| Tx_1-Tx_2||| = ||| T(x_1-x_2)||| \leq {\ q}||| x_1-x_2||| . \end{equation*} Thus, $T$ is a contraction on the Banach space $( S, ||| \cdot ||| )$. Now, for every fixed $\Psi$- bounded solution $y$ of \eqref{e3} we define an operator $S_{ y} : S \to S$, by the relation \begin{equation} {S}_{ y}x(t) = y(t) + Tx(t), \quad t \in [t_0,\infty ). \label{e4} \end{equation} It follows by the Banach contraction principle that $S_y$ has a unique fixed point in $S$. An easy computation shows that the fixed point $x(t)= S_yx(t)$, $t \in [t_0,\infty)$, is a $\Psi$-bounded solution of \eqref{e2}. Let $S_{2}$, $S_{3}$ be the spaces of $\Psi$-bounded solutions of equations \eqref{e2} and \eqref{e3} respectively. We define the mapping $C : S_{3}\to S_{2}$ in the following way: For every $y \in S_{3}$, $Cy$ will be the fixed point of the contraction $S_y$. Now, from $x=Cy$ and $x_{0}=Cy_{0}$, we have that $x=y+Tx$, $x_{0}=y_{0}+Tx_{0}$ respectively. We obtain \begin{align*} |||{\ x-x}_{0}|||& \leq |||{\ y-y}_{0}|||+|||{\ Tx-Tx}_{0}||| \\ & \leq |||{\ y-y}_{0}|||+q|||{\ x-x}_{0}|||. \end{align*} Thus \begin{equation} |||{\ x-x}_{0}|||\leq {(1-q)}^{-1}|||{\ y-y}_{0}|||. \label{e5} \end{equation} On the other hand, \begin{align*} |||{\ y-y}_{0}|||& =|||{\ x-Tx-x}_{0}{\ +}\text{ }{Tx}_{0}||| \\ & \leq |||{\ x-x}_{0}\text{ }|||\text{ }+\text{ }|||{\ Tx-Tx}_{0}\text{ }||| \\ & \leq {\ (1+q)}|||{\ x-x}_{0}\text{ }|||. \end{align*} Thus, $C$ is homeomorfism. Now, we prove that if $x,$ $y\in S$ are $\Psi $-bounded solutions of \eqref{e2} and \eqref{e3} respectively such that $x=Cy$, then \begin{equation*} \lim_{t\rightarrow \infty }\Vert \Psi {(t)(x(t)-y(t))}\Vert =0. \end{equation*} Indeed, for a given $\varepsilon >0$, we choose $t_{1}\geq t_{0}$ so that \begin{equation*} {K}\sup_{t\geq t_{0}}\Vert \Psi {(t)x(t)}\Vert \int_{t_{1}}^{\infty }|\Psi {% (s)B(s)}\Psi ^{-1}(s)|ds<\frac{\varepsilon }{3}. \end{equation*} Thus, for $t\geq t_{1}$, we have \begin{align*} & \Vert \Psi {(t)(x(t)-y(t))}\Vert \\ & =\Vert \Psi {(t)(Tx)(t)}\Vert \\ & \leq \int_{t_{0}}^{t}\Vert \Psi (t)Y(t)P_{1}Y^{-1}{\ (s)B(s)x(s)}\Vert ds \\ & \quad +\int_{t}^{\infty }\Vert \Psi (t)Y(t)P_{2}Y^{-1}(s)\Psi ^{-1}(s)\Psi {(s)B(s)}\Psi ^{-1}(s)\Psi {(s)x(s)}\Vert {ds} \\ & \leq |\Psi (t)Y(t)P_{1}|\int_{t_{0}}^{t_{1}}\Vert {\ Y}^{-1}{(s)B(s)x(s)}% \Vert ds \\ & \quad +K\sup_{t\geq t_{0}}\Vert \Psi {(t)x(t)}\Vert \int_{t_{1}}^{\infty }|\Psi {(s)B(s)}\Psi ^{-1}(s)|ds \\ & \quad +K\sup_{t\geq t_{0}}\Vert \Psi {(t)x(t)}\Vert \int_{t}^{\infty }|\Psi {(s)B(s)}\Psi ^{-1}(s)|ds \\ & <|\Psi (t)Y(t)P_{1}|\int_{t_{0}}^{t_{1}}\Vert Y^{-1}{(s)B(s)x(s)}\Vert {ds+2}\frac{\varepsilon }{3}. \end{align*} Thus and assumption 3, \begin{equation} \lim_{t\rightarrow \infty }\Vert \Psi {(t)(x(t)-y(t))}\Vert =0. \label{e6} \end{equation} >From the hypotheses, \cite[Theorem1 and 2]{d4} it follows that the linear equation \eqref{e3} is $\Psi $-conditionally asymptotically stable on $% \mathbb{R}_{+}$. Let $x(t)$ be a $\Psi $-bounded solution on $\mathbb{R}_{+}$ of \eqref{e2}. >From the assumption 4, this solution is $\Psi $-unstable on $\mathbb{R}_{+}$. Let $y=C^{-1}x$. From Definition \ref{def2.3}, it follows that there exists a sequence $(y_{n})$ of solutions of \eqref{e3} defined on $\mathbb{R}_{+}$ such that \begin{gather*} \lim_{n\rightarrow \infty }\Psi (t)y_{n}(t)=\Psi (t)y(t),\quad \text{% uniformly on }\mathbb{R}_{+}, \\ \lim_{t\rightarrow \infty }\Psi (t)(y_{n}(t)-y(t))=0,\quad \text{for }% n=1,2,\dots . \end{gather*} Let $x_{n}=Cy_{n}$. From \eqref{e5} it follows that the sequence $(x_{n})$ of solutions of \eqref{e2} defined on $[t_{0},\infty )$ (in fact, defined on $\mathbb{R}_{+}$) satisfies the condition \begin{equation*} \lim_{n\rightarrow \infty }\Psi {(t)x}_{n}{(t)=}\Psi (t)x(t),\quad \text{% uniformly on }[{t}_{0},\infty ). \end{equation*} Clearly, \begin{equation*} \lim_{n\rightarrow \infty }{\ x}_{n}{(t}_{0})=x(t_{0}). \end{equation*} By the Dependence on initial conditions Theorem (see \cite[Chapter I, Theorem 3]{c3}), it follows that \begin{equation*} \lim_{n\rightarrow \infty }x_{n}(t)=x(t),\quad \text{uniformly on }[0,t_{0}]. \end{equation*} Hence, \begin{equation*} \lim_{n\rightarrow \infty }\Psi (t)x_{n}(t)=\Psi (t)x(t),\quad \text{uniformly on }[0,t_{0}]. \end{equation*} Thus, \begin{equation*} \lim_{n\rightarrow \infty }\Psi {(t)x}_{n}{(t)=}\Psi (t)x(t),\quad \text{uniformly on }\mathbb{R}_{+}. \end{equation*} This shows that the linear equation \eqref{e2} is $\Psi $-conditionally stable on $\mathbb{R}_{+}$. From \eqref{e6} and \begin{equation*} \Psi (t)(x_{n}(t)-x(t))=\Psi (t)(x_{n}(t)-y_{n}(t))+\Psi (t)(y_{n}(t)-y(t))+\Psi (t)(y(t)-x(t)), \end{equation*} it follows that \begin{equation*} \lim_{t\rightarrow \infty }\Psi (t)(x_{n}(t)-x(t))=0,\quad \text{for }% n=1,2,\dots . \end{equation*} This shows that the linear equation \eqref{e2} is $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. The proof is complete. \end{proof} \begin{theorem} \label{thm3.5} Suppose that: \begin{enumerate} \item There exist two supplementary projections $P_1$, $P_2$, $P_i\neq 0$, and a positive constant $K$ such that the fundamental matrix $Y(t)$ of the equation \eqref{e3} satisfies the condition \begin{equation*} \int_0^{{t}}| \Psi (t)Y(t)P_{1} Y^{-1}{ (s)}\Psi ^{-1}(s)| ds + \int_{t}^{\infty }| \Psi (t)Y(t)P_{2} Y^{-1}(s)\Psi ^{-1}(s)| {ds}\leq { K} \end{equation*} for all t $\geq 0$. \item $B(t)$ is a $d\times d$ continuous matrix function on $\mathbb{R}_{+}$ such that \begin{equation*} \lim_{t\to \infty }| \Psi {(t)B(t)}\Psi^{-1}{(t)}| = 0. \end{equation*} \end{enumerate} Then, the linear equation \eqref{e2} is $\Psi$-conditionally asymptotically stable on $\mathbb{R}_{+}$. \end{theorem} The proof of the above theorem is similar to the proof of Theorem \ref{thm3.4}. \begin{remark} \rm The first condition of the above Theorems can certainly be satisfied if A(t) = A is a d$\times $d real constant matrix which has characteristic roots with different real parts. In this case, e.g., there exists an interval $(\alpha , \beta )\subset \mathbb{R}$ such that for $\lambda \in (\alpha, \beta)$, $\Psi (t) = e^{-\lambda {t}}I_{d}$ and $Y(t)$ can satisfy the first hypotheses of Theorems. We have a similar situation if $A(t)$ is a $d\times d$ real continuous periodic matrix (See \cite[Examples 1, 2]{d5}). Thus, the above results can be considered as a generalization of a well-known result in conection with the classical conditional asymptotic stability. \end{remark} \begin{remark} \rm If in the above Theorems, the linear equation \eqref{e3} is only $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$, then the perturbed equation \eqref{e2} can not be $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. This is shown by the next example transformed after an equation due to Perron \cite{p1}. \end{remark} \begin{example} \label{exa1} \rm Let $a, b \in \mathbb{R}$ such that $0 < 4a< 1$, $b \neq 0$ and \begin{equation*} A(t)=\begin{pmatrix} \sin \ln (t+1)+\cos \ln (t+1)-4a & 0 \\ 0 & -2a \end{pmatrix}. \end{equation*} Then, a fundamental matrix for the homogeneous equation \eqref{e3} is \begin{equation*} Y(t) = \begin{pmatrix} e^{(t+1)[\sin \ln (t+1)-4a]} & 0 \\ 0 & e^{-2a(t+1)} \end{pmatrix}. \end{equation*} Let \begin{equation*} \Psi (t)=\begin{pmatrix} 1 & 0 \\ 0 & e^{a(t+1)} \end{pmatrix}. \end{equation*} We have \begin{equation*} \Psi (t)Y(t)=\begin{pmatrix} e^{(t+1)[\sin \ln (t+1)-4a]} & 0 \\ 0 & e^{-a(t+1)} \end{pmatrix}. \end{equation*} Let $t_{n}' = e^{(2n+\frac{1}{2})\pi } - 1$ for $n = 1, 2\dots $. Since $\lim_{n\to \infty }| \Psi (t_{n}')Y(t_{n}')| = \infty $, it follows that the linear equation \eqref{e3} is $\Psi$-unstable on $\mathbb{R}_{+}$ (see \cite[Theorem 1]{d4}) From Theorem \ref{thm3.1} it follows that the linear equation \eqref{e3} is $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. If we take \begin{equation*} {B(t) = }\begin{pmatrix} 0 & be^{-2a(t+1)} \\ 0 & 0 \end{pmatrix}, \end{equation*} then, a fundamental matrix for the perturbed equation \eqref{e2} is \begin{equation*} \widetilde{Y}(t)=\begin{pmatrix} be^{(t+1)[\sin \ln (t+1)-4a]}\int_1^{t+1}e^{-s \sin \ln s} {ds} & e^{(t+1)[\sin \ln (t+1)-4a]} \\ e^{-2a(t+1)} & 0 \end{pmatrix}. \end{equation*} We have \begin{equation*} \Psi (t)\widetilde{Y}(t)= \begin{pmatrix} be^{(t+1)[\sin \ln (t+1)-4a]}\int_1^{t+1}e^{-s \sin \ln s} {ds} & e^{(t+1)[\sin \ln (t+1)-4a]} \\ e^{-a(t+1)} & 0 \end{pmatrix}. \end{equation*} Since $\lim_{n\to \infty }| \Psi (t_n') \widetilde{Y}(t_n')| = \infty $, it follows that the perturbed equation \eqref{e2} is $\Psi$-unstable on $\mathbb{R}_{+}$ (see \cite[Theorem 1]{d4}). Let $\alpha \in (0,\frac \pi 2)$. Let t$_n = e^{(2n-\frac 12)\pi }$ for $n = 1, 2,\dots$. For $t_n \leq s \leq t_ne^\alpha $ we have $s\cos \alpha \leq -s \sin \ln s \leq s$ and hence \begin{align*} e^{t_ne^\pi (\sin \ln t_ne^\pi -4a)}\int_1^{t_ne^\pi }e^{-s \sin \ln s}ds &> e^{t_ne^\pi (\sin \ln t_ne^\pi -4a)}\int_{t_n}^{t_ne^\alpha }e^{-s \sin \ln s}{ds }\\ &\geq e^{t_{n}e^{\pi}(1-4a)}\int_{t_{n}}^{t_{n}e^{\alpha }}e^{s \cos \alpha }ds \\ &= e^{t_{n}[(1-4a)e^{\pi }+\cos \alpha ]} \frac{e^{t_{n}(e^{\alpha }-1)\cos \alpha }-1}{\cos \alpha }\to \infty . \end{align*} Thus, the columns of $\Psi (t)\widetilde{Y}(t)$ are unbounded on $\mathbb{R}_{+}$. It follows that the perturbed equation \eqref{e2} is not $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$ (see Theorem \ref{thm3.1}). Finally, we have $| \Psi (t)B(t)\Psi ^{-1}(t) = be^{-3a(t+1)}$. Thus, $B(t)$ satisfies the conditions: \begin{equation*} \lim_{t\to \infty }| \Psi {(t)B(t)}\Psi ^{-1} {(t)}| =0; \end{equation*} and $\int_0^\infty | \Psi {(t)B(t)}\Psi ^{-1}{(t)}| dt$ can be a sufficiently small number. \end{example} \section{$\Psi$-conditional asymptotic stability of the nonlinear equation \eqref{e1}} In this section we give sufficient conditions for the $\Psi $-conditional asymptotic stability of $\Psi$-bounded solutions of the nonlinear Volterra integro-differential system \eqref{e1}. \begin{theorem} \label{thm4.1} Suppose that: \begin{enumerate} \item There exist supplementary projections $P_1$, $P_2$, $P_i \neq 0$ and a constant $K> 0$ such that the fundamental matrix $Y(t)$ of \eqref{e3} satisfies the condition \begin{equation*} \int_0^{{t}}| \Psi (t)Y(t)P_1 Y^{-1}(s) \Psi ^{-1}(s)| ds + \int_{{t}}^\infty| \Psi (t)Y(t)P_2 Y^{-1}(s)\Psi ^{-1}(s)| {ds}\leq { K} \end{equation*} for all $t \geq 0$. \item The function $F(t,s,x)$ satisfies the inequality \begin{equation*} \| \Psi {(t)}\left( {F(t,s,x(s)) }-{ F(t,s,y(s))} \right) \| \leq { f(t,s)}\| \Psi (s)\left( {x(s) }-{ y(s)}\right) \| , \end{equation*} for $0 \leq s \leq t < \infty $ and for all continuous and $\Psi$-bounded functions $x, y : \mathbb{R}_{+}\to \mathbb{R}^d$, where $f(t,s)$ is a continuous nonnegative function on $D$ such that \begin{equation*} F(t,s,0) = 0, \quad \lim_{t\to \infty }\int_0^t f(t,s)ds = 0,\quad \sup_{t\geq 0}\int_0^t{f(t,s)ds < K}^{-1}. \end{equation*} \end{enumerate} Then, all $\Psi$-bounded solutions of \eqref{e1} are $\Psi$-conditionally asymptotically stable on $\mathbb{R}_{+}$. \end{theorem} \begin{proof} Let \begin{equation*} q = K \sup_{t\geq 0} \int_0^t f(t,s)ds < 1. \end{equation*} We put \begin{equation*} S = \{ x : \mathbb{R}_{+}\to {R}^d : x\text{ is continuous and } \Psi\text{-bounded on }\mathbb{R}_{+}\}. \end{equation*} Define on the set $S$ a norm by \begin{equation*} ||| {x}||| = \sup_{t\geq 0}\| \Psi {(t)x(t)}\| . \end{equation*} It is well-known that ($S$, $||| \cdot ||| $) is a Banach space. For $x \in S $, we define \begin{align*} \left( {Tx}\right) (t) &= \int_0^t{Y(t)P}_1 Y^{-1}(s)\int_o^s F(s,u,x(u))\,du\,ds \\ &\quad - \int_t^\infty {Y(t)P}_2 Y^{-1}(s)\int_o^s{\ F(s,u,x(u))\,du\,ds, t } \geq 0. \end{align*} For $0 \leq t \leq v$, we have \begin{align*} &\| \Psi {(t)}\int_{t}^{{v}}{Y(t)P}_{2} Y^{-1} (s)\int_{o}^{s}{ F(s,u,x(u))\,du\,ds}\| \\ &= \| \int_{t}^{{v}}\Psi (t)Y(t)P_{2} Y^{-1} (s)\Psi ^{-1}(s)\int_{o}^{s}\Psi (s)F(s,u,x(u))du\,ds\| \\ &\leq \int_{t}^{{v}}| \Psi (t)Y(t)P_{2} Y^{-1} (s)\Psi ^{-1}(s)| \int_{o}^{s}\| \Psi {(s)F(s,u,x(u))}\| \,du\,ds \\ &\leq \int_{t}^{{v}}| \Psi (t)Y(t)P_{2} Y^{-1}(s) \Psi ^{-1}(s)| \int_0^{s}{ f(s,u)}\| \Psi {\ (u)x(u)}\| \,du\,ds \\ &\leq \sup_{u\geq 0} \| \Psi {(u)x(u)}\| \int_{t}^{{v}}| \Psi (t)Y(t)P_{2} Y^{-1}(s)\Psi ^{-1}(s)| \int_0^{s} f(s,u)\,du\,ds \\ &\leq {\ qK}^{-1}\sup_{u\geq 0} \| \Psi {(u)x(u)} \| \int_{t}^{{v}}| \Psi (t)Y(t)P_{2} Y^{-1}{\ (s)}\Psi ^{-1}(s)| {ds.} \end{align*} >From assumption 1, it follows that the integral \begin{equation*} \int_t^\infty {Y(t)P}_2 Y^{-1}(s)\int_o^s{F(s,u,x(u))\,du\,ds} \end{equation*} is convergent. Thus, $(Tx)(t)$ exists and is continuous for $t \geq 0$. For $x \in S$ and $t \geq 0$, we have \begin{align*} \| \Psi {(t)(Tx)(t)}\| &= \|\int_0^t\Psi (t)Y(t)P_1 Y^{-1}(s)\Psi ^{-1}(s) \int_o^s\Psi {(s)F(s,u,x(u))\,du\,ds } \\ &\quad - \int_t^\infty \Psi (t)Y(t)P_2 Y^{-1}(s)\Psi ^{-1}(s)\int_o^s\Psi { (s)F(s,u,x(u))\,du\,ds}\| \\ & \leq \int_0^t| \Psi (t)Y(t)P_1 Y^{-1}(s)\Psi ^{-1}(s)| \int_o^s\| \Psi { (s)F(s,u,x(u))}\| \,du\,ds \\ &\quad + \int_t^\infty | \Psi (t)Y(t)P_2 Y^{-1}(s) \Psi ^{-1}(s)| \int_o^s\| \Psi {(s)F(s,u,x(u))} \| \,du\,ds \\ &\leq \int_0^t| \Psi (t)Y(t)P_1 Y^{-1}(s)\Psi ^{-1}(s)| \int_0^s{f(s,u)}\| \Psi {(u)x(u)} \| \,du\,ds \\ &\quad + \int_t^\infty | \Psi (t)Y(t)P_2 Y^{-1}(s) \Psi ^{-1}(s)| \int_o^s{ f(s,u)}\| \Psi {(u)x(u)} \| \,du\,ds \\ &\leq q \sup_{u\geq 0} \| \Psi {(u)x(u)} \| . \end{align*} This shows that $TS \subseteq S$. On the other hand, for $x, y \in S$ and $t \geq 0$, we have \begin{align*} &\| \Psi {(t)}\left( (Tx)(t) - (Ty)(t) \right) \| \\ &=\| \int_0^{t}\Psi (t)Y(t)P_{1} Y^{-1}(s) \Psi ^{-1}(s)\int_{o}^{s}\Psi (s)\left( F(s,u,x(u)) - F(s,u,y(u))\right) \,du\,ds \\ &\quad -\int_{t}^{\infty }\Psi (t)Y(t)P_{2} Y^{-1}(s)\Psi ^{-1} (s)\int_{o}^{s}\Psi (s)\left( F(s,u,x(u)) - F(s,u,y(u)) \right) \,du\,ds\| \\ &\leq \int_0^{t}| \Psi (t)Y(t)P_{1} Y^{-1}(s)\Psi ^{-1}(s)| \int_{o}^{s}\| \Psi (s)\left( F(s,u,x(u))-F(s,u,y(u))\right) \| \,du\,ds \\ &\quad +\int_{t}^{\infty }\!| \Psi (t)Y(t)P_{2} Y^{-1}(s) \Psi ^{-1}(s)| \int_{o}^{s}\!\| \Psi (s)\left( F(s,u,x(u)) - F(s,u,y(u))\right) \| \,du\,ds \\ &\leq \int_0^t| \Psi (t)Y(t)P_1 Y^{-1}(s)\Psi ^{-1} (s)| \int_0^s{f(s,u)}\| \Psi (u)(x(u) - y(u)) \| \,du\,ds \\ &\quad + \int_t^\infty | \Psi (t)Y(t)P_2 Y^{-1}(s)\Psi ^{-1}(s)| \int_0^s{ f(s,u)}\| \Psi {(u)(x(u) - y(u))}\| \,du\,ds \\ &\leq q \sup_{u\geq 0} \| \Psi {(u)(x(u) - y(u))}\| . \end{align*} It follows that \begin{equation*} \sup_{t\geq 0} \| \Psi {(t)}\left( {(Tx)(t) - (Ty)(t)}\right) \| \leq q \sup_{t\geq 0} \| \Psi {(t)(x(t) - y(t))}\| . \end{equation*} Thus, we have \begin{equation*} ||| Tx - Ty||| \leq q ||| x - y||| . \end{equation*} This shows that $T$ is a contraction of the Banach space $( S,||| \cdot |||) $. As in the Proof of Theorem \ref{thm3.4}, it follows by the Banach contraction principle that for any function $y \in S$, the integral equation \begin{equation} \label{e7} x = y + Tx \end{equation} has a unique solution $x \in S$. Furthermore, by the definition of $T$, $x(t)- y(t)$ is differentiable and \begin{equation*} \left( {x(t) - y(t)}\right) '{\ = A(t)}\left( {x(t) - y(t)}\right) + \int_0^{t}F(t,s,x(s))ds, t \geq 0. \end{equation*} Hence, if y(t) is a $\Psi$-bounded solution of \eqref{e3}, $x(t)$ is a $\Psi$-bounded solution of \eqref{e1}. Conversely, if $x(t)$ is a $\Psi$-bounded solution of \eqref{e1}, the function $y(t)$ defined by \eqref{e7} is a $\Psi$-bounded solution of \eqref{e3}. Thus, \eqref{e7} establishes a one-to-one correspondence $C$ between the $\Psi$-bounded solutions of \eqref{e1} and \eqref{e3}: $x = Cy$. Now, we consider the analogous equation \begin{equation*} x_0 = y_0 + Tx_0. \end{equation*} We get \begin{equation} {(1 - q)}||| {\ x - x}_0 ||| \leq ||| {\ y - y}_0 ||| . \label{e8} \end{equation} Now, we prove that if $x, y \in S$ are $\Psi$-bounded solutions of \eqref{e1} and \eqref{e3} respectively such that $x = Cy$, then \begin{equation} \lim_{t\to \infty }\| \Psi {(t)( x(t) - y(t) )}\| = 0. \label{e9} \end{equation} For a given $\varepsilon >0$, we can choose $t_1 \geq 0$ such that \begin{equation*} {K}||| x||| \int_0^{t }f(t,s)ds < \frac \varepsilon 2, \end{equation*} for $t \geq t_1$. Moreover, since $\lim_{t\to \infty }| \Psi (t)Y(t)P_1| = 0$ (see \cite[Lemma 1]{d4}), there exists a number $t_2\geq t_1$ such that \begin{equation*} {qK}^{-1}| \Psi (t)Y(t)P_1||| | x||| \int_0^{{t}_1}| {P}_1 Y^{-1}(s)\Psi ^{-1} (s)| ds < \frac \varepsilon 2 \end{equation*} for $t \geq t_2$. We have, for $t \geq t_2$, \begin{align*} &\| \Psi {(t)( x(t) - y(t) )}\| \\ & \leq \int_0^t| \Psi (t)Y(t)P_1 Y^{-1}(s)\Psi ^{-1}(s)| \int_o^s\| \Psi {% (s)F(s,u,x(u))}\| {\,du\,ds +} \\ &\quad + \int_t^\infty | \Psi (t)Y(t)P_2 Y^{-1}(s) \Psi ^{-1}(s)| \int_o^s\| \Psi {(s)F(s,u,x(u))} \| \,du\,ds \\ &\leq \int_0^t| \Psi (t)Y(t)P_1 Y^{-1}(s)\Psi ^{-1}(s)| \int_0^s{f(s,u)}\| \Psi {(u)x(u)} \| \,du\,ds \\ &\quad + \int_t^\infty | \Psi (t)Y(t)P_2 Y^{-1}(s) \Psi ^{-1}(s)| \int_o^s{f(s,u)}\| \Psi {(u)x(u)} \| \,du\,ds \\ &\leq {qK}^{-1}| \Psi (t)Y(t)P_1|\,|||x||| \int_0^{{t}_1}| {P}_1 Y^{-1}(s)\Psi ^{-1}(s)| ds \\ &\quad + ||| x||| \int_{t_1}^{{t}}| \Psi (t)Y(t)P_1 Y^{-1}(s)\Psi ^{-1}(s)| \Big( \int_0^s{f(s,u)du}\Big) ds \\ &\quad + ||| x||| \int_{{t}}^\infty | \Psi (t)Y(t)P_2 Y^{-1}(s)\Psi ^{-1}(s)| \Big( \int_0^s{f(s,u)du}\Big) ds < \varepsilon . \end{align*} Now, let $x(t)$ be a $\Psi$-bounded solution of \eqref{e1}. This solution is $\Psi$-unstable on $\mathbb{R}_{+}$. Indeed, if not, for every $\varepsilon $ > 0 and any $t_0\geq 0$, there exists a $\delta =\delta (\varepsilon ,t_0)> 0$ such that any solution $\widetilde{{x}}(t)$ of \eqref{e1} which satisfies the inequality $\| \Psi (t_0) ( \widetilde{{x}}(t_0) - x(t_0) ) \| < \delta (\varepsilon ,t_0)$ exists and satisfies the inequality $\| \Psi (t)(\widetilde{{x}}(t) - x(t))\| < \varepsilon $ for all $t \geq t_0$. Let $z_0\in R^{d}$ be such that $P_{1}z_0 = 0$ and $0 < \| \Psi (0)z_0\| < \delta (\varepsilon ,0)$ and let $\widetilde{x}(t)$ the solution of \eqref{e1} with the initial condition $\widetilde{x}(0) = x(0) + z_0$. Then $\| \Psi (t)z(t)\| < \varepsilon $ for all $t \geq 0$, where $z(t) = \widetilde{x}(t) - x(t)$. Now we consider the function $y(t) = z(t) - (Tz)(t)$, $t \geq 0$. Clearly, $y(t)$ is a $\Psi$-bounded solution on $\mathbb{R}_{+}$ of \eqref{e3}. Without loss of generality, we can suppose that $Y(0) = I$. It is easy to see that $P_1y(0) = 0$. If $P_2y(0) \neq 0$, from \cite[Lemma 2]{d4}, it follows that $\limsup_{t\to \infty }\| \Psi (t)y(t)\| = \infty $, which is contradictory. Thus, $P_2y(0) = 0$ and then $y(t) = 0$ for $t \geq 0 $. It follows that $z = Tz$ and then $z = 0$, which is a contradiction. This shows that the solution $x(t)$ is $\Psi$-unstable on $\mathbb{R}_{+}$. Let $y = x - Tx$. From Theorem \ref{thm3.3}, it follows that there exists a sequence $(y_{n})$ of solutions of \eqref{e3} defined on $\mathbb{R}_{+}$ such that \begin{gather*} \lim_{n\to \infty }\Psi (t)y_n (t) = \Psi (t)y(t),\quad\text{uniformly on } \mathbb{R}_{+}, \\ \lim_{t\to \infty }\Psi (t)( y_{n}(t) - y(t) )=0, \quad n = 1,2,\dots . \end{gather*} Let $x_{n} = Cy_{n}$. From \eqref{e8} it follows that the sequence $(x_{n})$ of solutions of \eqref{e1} defined on $\mathbb{R}_{+}$ is such that \begin{equation*} \lim_{n\to \infty }\Psi (t)x_n(t) = \Psi (t)x(t), \quad\text{uniformly on }% \mathbb{R}_{+}. \end{equation*} This shows that the solution $x(t)$ is $\Psi $-conditionally stable on $\mathbb{R}_{+}$. From \eqref{e9} and \begin{equation*} \Psi (t)( x_{n}(t) - x(t) ) = \Psi (t)( x_{n}(t)- y_{n}(t) ) + \Psi (t)( y_{n}(t) - y(t) ) + \Psi (t)( y(t) - x(t) ), \end{equation*} it follows that \begin{equation*} \lim_{t\to \infty }\Psi (t)( x_{n}(t) - x(t) ) = 0, \quad\text{for } n = 1, 2, \dots . \end{equation*} This shows that the solution $x(t)$ is $\Psi $-conditionally asymptotically stable on $\mathbb{R}_{+}$. The proof is now complete. \end{proof} \begin{corollary} \label{coro4.1} If in Theorem \ref{thm4.1} we assume that $f(t.s) =g(t)h(s)$, where $g$ and $h$ are nonnegative continuous functions on $\mathbb{R}_{+}$ such that \begin{gather*} \sup_{t\geq 0} {g(t)}\int_0^t h(s)ds < K^{-1} , \\ \lim_{t\to \infty }{g(t)}\int_0^t h(s)ds = 0, \end{gather*} then the conclusion of the Theorem remains valid. \end{corollary} \begin{corollary} \label{coro4.2} If in Theorem \ref{thm4.1} we assume that $f(t.s) =g(t)h(s)$, where $g$ and $h$ are nonnegative continuous functions on $\mathbb{R}_{+}$ such that \begin{gather*} I = \int_0^\infty h(s)\,ds \quad\text{is convergent}, \\ \lim_{t\to \infty }g(t) = 0, \quad \sup_{t\geq 0}g(t) < \frac 1{KI}, \end{gather*} then the conclusion of the Theorem remains valid. \end{corollary} \begin{thebibliography}{99} \bibitem{a1} Akinyele, O. \emph{On partial stability and boundedness of degree k; } Atti. Accad. Naz. Lincei Rend. Cl. Sci. Fis. Mat. Natur., (8), 65(1978), 259 - 264. \bibitem{a2} Akinyele, O. \emph{On the $\varphi $-stability for differential systems;} Atti. Accad. Naz. Lincei Rend. Cl. Sci. Fis. Mat. Natur.,(4), 68(1980), 287 - 293. \bibitem{a3} Avramescu, C. \emph{Asupra comport\~{a}rii asimptotice a solutiilor unor ecuatii functionale;} Analele Universit\~{a}tii din Timisoara, Seria Stiinte Matematice-Fizice, vol. VI, 1968, 41 - 55. \bibitem{c1} Constantin, A. \emph{Asymptotic properties of solutions of differential equations; }Analele Universit\~{a}tii din Timisoara, Seria Stiinte Matematice, vol. XXX, fasc. 2-3, 1992, 183 - 225. \bibitem{c2} Constantin, A. \emph{A note on the $\varphi $-stability for differential systems; }Boll. Un. Math. Ital., 6A(1992), 233 - 243. \bibitem{c3} Coppel, W.A. \emph{Stability and Asymptotic Behavior of Differential Equations,} Heath, Boston, 1965. \bibitem{c4} Coppel, W.A. \emph{On the stability of ordinary differential equations,} J. London Math. Soc. 38(1963), 255 - 260. \bibitem{d1} Diamandescu, A. \emph{On the $\Psi $-stability of a nonlinear Volterra integro-differential system;} Electron. J. Diff. Eqns., Vol. 2005(2005), No. 56, pp. 1 - 14; URL http://ejde.math.txstate.edu \bibitem{d2} Diamandescu, A. \emph{On the $\Psi $-asymptotic stability of a nonlinear Volterra integro-differential system;} Bull. Math. Sc. Math. Roumanie, Tome 46(94), No. 1-2, 2003, pp. 39 - 60. \bibitem{d3} Diamandescu, A. \emph{On the $\Psi $-uniform asymptotic stability of a nonlinear Volterra integro-differential system;} Analele Universit\u{a}\c{t}ii de Vest Timi\c{s}oara, Seria Matematic\u{a} - Informatic\u{a}, Vol. XXXIX, fasc. 2, 2001, pp. 35 - 62. \bibitem{d4} Diamandescu, A. \emph{On the $\Psi $-instability of a nonlinear Volterra integro-differential system;} Bull. Math. Sc. Math. Roumanie, Tome 46(94), No. 3 - 4, 2003, pp. 103 - 119.. \bibitem{d5} Diamandescu, A. \emph{On the $\Psi $-conditional stability of the solutions of a nonlinear Volterra integro-differential system;} Proceedings of the National Conference on Mathematical Analysis and Applications, Timi\c{s}oara, 12-13 Dec. 2000, pp. 89 - 106. \bibitem{h1} Halam, T.G. \emph{On asymptotic equivalence of the bounded solutions of two systems of differential equations;} Mich. Math. Journal., Vol. 16(1969), 353 - 363. \bibitem{h2} Hara, T., Yoneyama, T. and Ytoh, T. \emph{Asymptotic Stability Criteria for Nonlinear Volterra Integro - Differential Equations; } Funkcialaj Ecvacioj, 33(1990), 39 - 57. \bibitem{k1} Kuben, J. \emph{Asymptotic equivalence of second order differential equations;} Czech. Math. J. 34(109), (1984), 189 - 202. \bibitem{l1} Lakshmikantham, V. and Rama Mohana Rao, M. \emph{Stability in variation for nonlinear integro-differential equations;} Appl. Anal. 24(1987), 165 - 173. \bibitem{m1} Mahfoud, W.E. \emph{Boundedness properties in Volterra integro-differential systems;} Proc. Amer. Math. Soc., 100(1987), 37 - 45. \bibitem{m2} Morchalo, J. \emph{On $\Psi -L_{p}$-stability of nonlinear systems of differential equations;} Analele Stiintifice ale Universit\~{a}tii quotedblright Al. I. Cuzaquotedblright Iasi, Tomul XXXVI, s. I - a, Matematic\~{a}, 1990, f. 4, 353 - 360. \bibitem{p1} Perron, O. \emph{Die Stabilit\"{a}tsfrage bei Differentialgleichungen,} Math. Z., 32(1930), 703 - 728. \bibitem{s1} Sansone, G. and Conti, R. \emph{Non-linear differential equations,} Pergamon Press, 1964. \bibitem{s2} Sp\"{a}th, H. \emph{Uber das asymptotische Verhalten der L\"{o} sungenlinearer Differentialgleichungen,} Math. Z., 30(1929), 487 - 513. \bibitem{w1} Weyl, H. \emph{Comment on the preceding paper,} Amer. J. Math., 68(1946), 7 - 12. \bibitem{y1} Yoshizawa, T. \emph{Stability Theory by Liapunov's Second Method}, The Mathematical Society of Japan, 1966. \end{thebibliography} \end{document}