\documentclass[reqno]{amsart} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2003(2003), No. 123, 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 (login: ftp)} \thanks{\copyright 2003 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE--2003/123\hfil Integral inequalities in several variables] {On integral inequalities for functions of several independent variables} \author[Hassane Khellaf\hfil EJDE--2003/123\hfilneg] {Hassane Khellaf} \address{University of Badji Mokhtar, Faculty of Science, Department of Mathematics, B. P. 12, Annaba 23000, Algeria} \email{khellafhassane@yahoo.fr} \date{} \thanks{Submitted September 15, 2003. Published December 16, 2003.} \subjclass[2000]{45H40, 45K05} \keywords{Integral inequality, subadditive and submultiplicative function} \begin{abstract} This paper presents some non-linear integral inequalities for functions of $n$ independent variables. These results extend the Gronwall type inequalities obtained for two variables by Dragomir and Kim \cite{2} \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \newtheorem{corollary}[theorem]{Corollary} \allowdisplaybreaks \section{Introduction} Integral inequalities play a significant role in the study of differential and integral equations. One of the most useful inequalities of Gronwall type is given in the following lemma (see \cite{1,2}). \begin{lemma} \label{l1} Let $u(t)$ and $k(t)$ be continuous, $a(t)$ and $b(t)$ Riemann integrable function on $J=[\alpha ,\beta ]\subset \mathbb{R}$ and $t\in \mathbb{R}$ with $b(t)$ and $k(t)$ nonnegative on $J$. If $u(t)\leq a(t)+b(t)\int_{\alpha }^{t}k(s)u(s)ds$ for $t\in J$, then \begin{equation} u(t)\leq a(t)+b(t)\int_{\alpha }^{t}a(s)k(s)\exp \Big( \int_{s}^{t}b(\tau )k(\tau )d\tau \Big) ds,\quad t\in J, \label{1.1} \end{equation} If $u(t)\leq a(t)+b(t)\int_{t}^{\beta }k(s)u(s)ds$ for $t\in J$, then \begin{equation} u(t)\leq a(t)+b(t)\int_{t}^{\beta }a(s)k(s)\exp \Big( \int_{t}^{s}b(\tau )k(\tau )d\tau \Big) ds,\quad t\in J. \label{1.2} \end{equation} \end{lemma} In the past few years, these inequalities have been generalized to more than one variable. Many authors have established Gronwall type integral inequalities in two or more independent variables; see for example \cite {3,4,5,6,7}. The results obtained have generated a lot of research interests due to its usefulness in the theory of differential and integral equations. Dragomir and Kim \cite{2} considered integral inequalities for functions with two independent variables. The purpose of this paper is to generalize their results by obtaining new integral inequalities in $n$ independent variables. In what follows we denote by $\mathbb{R}$ the set of real numbers and $% \mathbb{R}_{+}=[ 0,\infty)$. All the functions appearing in the inequalities are assumed to be real valued of $n$-variables which are nonnegative and continuous. All integrals exist on their domains of definitions. Throughout this paper, we shall assume that $x=(x_{1},x_{2},\dots x_{n})$ and $x^{0}=(x_{1}^{0},x_{2}^{0},\dots ,x_{n}^{0})$ are in $\mathbb{R}_{+}^{n} $. We shall denote \begin{equation*} \int_{x^{0}}^{x}dt=\int_{x_{1}^{0}}^{x_{1}}\int_{x_{2}^{0}}^{x_{2}}\dots \int_{x_{n}^{0}}^{x_{n}}\dots dt_{n}\dots dt_{1} \end{equation*} and $D_{i}=\frac{\partial}{\partial x_{i}}$ for $i=1,2,\dots ,n$. For $x,t\in \mathbb{R}_{+}^{n}$, we shall write $t\leq x$ whenever $t_{i}\leq x_{i}$, $i=1,2,\dots ,n$. \section{Results} \begin{lemma} \label{l2} Let $u(x),a(x)$ and $b(x)$ be nonnegative continuous functions, defined for $x\in \mathbb{R}_{+}^{n}$. \noindent (1) Assume that $a(x)$ is positive, continuous function, nondecreasing in each of the variables $x\in \mathbb{R}_{+}^{n}$. Suppose that \begin{equation} u(x)\leq a(x)+\int_{x^{0}}^{x}b(t)u(t)dt \label{2.1} \end{equation} holds for all $x\in \mathbb{R}_{+}^{n}$ with $x\geq x^{0}$, then \begin{equation} u(x)\leq a(x)\exp \Big( \int_{x^{0}}^{x}b(t)dt\Big) , \label{2.2} \end{equation} (2) Assume that $a(x)$ is positive, continuous function, non-increasing in each of the variables $x\in \mathbb{R}_{+}^{n}$. Suppose that \begin{equation} u(x)\leq a(x)+\int_{x}^{x^{0}}b(t)u(t)dt \label{2.3} \end{equation} holds for all $x\in \mathbb{R}_{+}^{n}$ with $x\leq x^{0}$, then \begin{equation} u(x)\leq a(x)\exp \Big( \int_{x}^{x^{0}}b(t)dt\Big) . \label{2.4} \end{equation} \end{lemma} \begin{proof} The proof of (1) is similar to the proof of (2), so we present the proof of (2) and refer the reader to {\cite[p. 112]{1}} for more details. \noindent (2) Since $a(x)$ is positive, non-increasing in each of the variables $x\in \mathbb{R}_{+}^{n}$, with $x\leq x^{0}$, then \begin{equation} \frac{u(x)}{a(x)}\leq 1+\int_{x}^{x^{0}}b(t)\frac{u(t)}{a(t)}dt, \label{2.5} \end{equation} Setting \begin{equation} v(x)=\frac{u(x)}{a(x)}, \label{2.6} \end{equation} we have \begin{equation} v(x)\leq 1+\int_{x}^{x^{0}}b(t)v(t)dt, \label{2.7} \end{equation} Let \begin{equation} r(x)=1+\int_{x}^{x^{0}}b(t)v(t)dt, \label{2.8} \end{equation} Then $r(x_{1}^{0},x_{2},\dots ,x_{n})=1$, and $v(x)\leq r(x)$, $r(x)$ is positive and nonincreasing in each of the variables $x_{2},\dots ,x_{n}\in \mathbb{R}_{+}$. Hence \begin{equation} \begin{aligned} D_{1}r(x) &=\int_{x_{2}}^{x_{2}^{0}}\int_{x_{3}}^{x_{3}^{0}}\dots \int_{x_{n}}^{x_{n}^{0}}b(x_{1,}t_{2},\dots ,t_{n})v(x_{1,}t_{2},\dots , t_{n})dt_{n}\dots dt_{2} \\ &\leq \int_{x_{2}}^{x_{2}^{0}}\int_{x_{3}}^{x_{3}^{0}}\dots \int_{x_{n}}^{x_{n}^{0}}b(x_{1,}t_{2},\dots ,t_{n})r(x_{1,}t_{2},\dots , t_{n})dt_{n}\dots dt_{2} \\ &\leq r(x)\int_{x_{2}}^{x_{2}^{0}}\int_{x_{3}}^{x_{3}^{0}}\dots \int_{x_{n}}^{x_{n}^{0}}b(x_{1,}t_{2},\dots ,t_{n})dt_{n}\dots dt_{2}, \end{aligned}\label{2.9} \end{equation} Dividing both sides of (\ref{2.9}) by $r(x)$ we get \begin{equation} \frac{D_{1}r(x)}{r(x)}\leq \int_{x_{2}}^{x_{2}^{0}}\int_{x_{3}}^{x_{3}^{0}}\dots \int_{x_{n}}^{x_{n}^{0}}b(x_{1,}t_{2},\dots ,t_{n})dt_{n}\dots dt_{2}. \label{2.10} \end{equation} Integrating with respect to $t_{1}$ from $x_{1}$ to $x_{1}^{0}$, we have \begin{equation} r(x)\leq \exp \Big( \int_{x^{0}}^{x^{0}}b(t)dt\Big) , \label{2.11} \end{equation} Hence \begin{equation} v(x)\leq \exp \Big( \int_{x}^{x_{0}}b(t)dt\Big) . \label{2.12} \end{equation} Substituting (\ref{2.12}) into (\ref{2.6}), we have the result (\ref{2.4}). \end{proof} \begin{theorem} \label{t1} Let $u(x)$, $a(x)$, $b(x)$, $c(x)$, $d(x)$, $f(x)$ be real-valued non-negative continuous functions defined for $x\in \mathbb{R}_{+}^{n}$. Let $W(u(x))$ be real-valued, positive, continuous, strictly non-decreasing, subadditive, and submultiplicative function for $u(x)\geq 0$, and let $H(u(x))$ be real-valued, positive, continuous, and non-decreasing function defined for $x\in \mathbb{R}_{+}^{n}$. Assume that $a(x)$, $f(x)$ are nondecreasing in the first variable $x_{1}$ for $x_{1}\in \mathbb{R}_{+}$. If \begin{equation} \begin{aligned} u(x) &\leq a(x)+b(x)\int_{\alpha}^{x_{1}}c(s,x_{2},\dots ,x_{n})u(s,x_{2},\dots , x_{n})ds \\ &\quad +f(x)H\Big( \int_{x^{_{0}}}^{x}d(t)W(u(t))dt\Big) , \end{aligned}\label{2.13} \end{equation} for $\alpha \geq 0$, $x,t\in \mathbb{R}_{+}^{n}$ with $\alpha \leq x_{1}$ and $x^{_{0}}\leq t\leq x$, then \begin{equation} u(x)\leq p(x)\Big\{ a(x)+f(x)H\Big[ G^{-1}\Big( G(A(t))+% \int_{x^{_{0}}}^{x}d(t)W(p(t)f(t))dt\Big) \Big]\Big\}, \label{2.14} \end{equation} for $\alpha \geq 0$, $x\in \mathbb{R}_{+}^{n}$ with $\alpha \leq x_{1}$, where \begin{gather} p(x)=1+b(x)\int_{\alpha }^{x_{1}}c(s,x_{2},\dots ,x_{n})\exp \Big( % \int_{\alpha }^{x_{1}}b(\tau ,x_{2},\dots ,x_{n})c(\tau ,x_{2},\dots ,x_{n})d\tau \Big) ds, \label{2.15} \\ A(t)=\int_{x^{_{0}}}^{\infty }d(t)W(a(t)p(t))dt, \label{2.16} \\ G(z)=\int_{z^{0}}^{z}\frac{ds}{W(H(s))},\quad z\geq z^{0}>0\,. \label{2.17} \end{gather} Here $G^{-1}$ is the inverse function of $G$ and \begin{equation*} G\Big( \int_{x^{_{0}}}^{\infty }d(t)W(a(t)p(t))dt\Big) % +\int_{x^{_{0}}}^{x}d(t)W(p(t)f(t))dt, \end{equation*} is in the domain of $G^{-1}$ for $x\in \mathbb{R}_{+}^{n}$. \end{theorem} \begin{proof} Define a function \begin{equation} z(x)=a(x)+f(x)H\Big( \int_{x^{_{0}}}^{x}d(t)W(u(t))dt\Big) , \label{2.18} \end{equation} Then (\ref{2.13}) can be restated as \begin{equation} u(x)\leq z(x)+b(x)\int_{\alpha }^{x_{1}}c(s,x_{2},\dots ,x_{n})u(s,x_{2},\dots ,x_{n})ds. \label{2.19} \end{equation} Clearly $z(x)$ is a nonnegative and continuous in $x_{1}\in \mathbb{R}% _{+}.\;x_{2},x_{3,}\dots x_{n}\in \mathbb{R}_{+}$fixed in (\ref{2.19}) and using (1) of lemma \ref{l1} to (\ref{2.19}), we get \begin{align*} u(x)& \leq z(x)+b(x)\int_{\alpha }^{x_{1}}z(s,x_{2},\dots ,x_{n})c(s,x_{2},\dots ,x_{n}) \\ & \quad \times \exp \Big( \int_{\alpha }^{x_{1}}b(\tau ,x_{2},\dots ,x_{n})c(\tau ,x_{2},\dots ,x_{n})d\tau \Big) ds, \end{align*} Moreover, $z(x)$ is nondecreasing in $x_{1},x_{1}\in R_{+}$, we obtain \begin{equation} u(x)\leq z(x)p(x), \label{2.20} \end{equation} where $p(x)$ is defined by (\ref{2.15}). From (\ref{2.18}) we have \begin{equation} u(x)\leq \left( a(x)+f(x)H(v(x))\right) p(x), \label{2.21} \end{equation} where $v(x)=\int_{x^{_{0}}}^{x}d(t)W(u(t))dt$. From (\ref{2.21}), we observe that \begin{equation} \label{2.22} \begin{aligned} v(x) &\leq \int_{x^{_{0}}}^{x}d(t)W\left( \left( a(t)+f(t)H(v(t))\right) p(t)\right) dt \\ &\leq \int_{x^{_{0}}}^{x}d(t)W(a(t)p(t))dt +\int_{x^{_{0}}}^{x}d(t)W\left( p(t)f(t)\right) W\left( H(v(t))\right) dt, \\ &\leq \int_{x^{_{0}}}^{\infty }d(t)W(a(t)p(t))dt +\int_{x^{_{0}}}^{x}d(t)W\left( p(t)f(t)\right) W\left( H(v(t))\right) dt, \end{aligned} \end{equation} Since $W$ is subadditive and submultiplicative function. Define $r(x)$ as the right side of (\ref{2.22}), then $r(x_{0}^{1},x_{2},\dots ,x_{n})=\int_{x^{_{0}}}^{\infty }d(t)W(a(t)p(t))dt$, $v(x)\leq r(x)$, $r(x)$ is positive nondecreasing in each of the variables $x_{2},\dots ,x_{n}\in \mathbb{R}_{+}$ and \begin{equation} \label{2.23} \begin{aligned} D_{1}r(x) &=\int_{x_{2}^{0}}^{x_{2}}\int_{x_{3}^{0}}^{x_{3}}\dots \int_{x_{n}^{0}}^{x_{n}}d(x_{1,}t_{2},\dots ,t_{n}) \ \\ &\quad\times W\left( p(x_{1,}t_{2},\dots ,t_{n})f(x_{1,}t_{2},\dots ,t_{n})\right) W\left( H(v(x_{1,}t_{2},\dots ,t_{n}))\right) dt_{n}\dots dt_{2} \\ &\leq \int_{x_{2}^{0}}^{x_{2}}\int_{x_{3}^{0}}^{x_{3}}\dots \int_{x_{n}^{0}}^{x_{n}}d(x_{1,}t_{2},\dots ,t_{n}) \\ &\quad\times W\left( p(x_{1,}t_{2},\dots ,t_{n})f(x_{1,}t_{2},\dots ,t_{n})\right) W\left( H(r(x_{1,}t_{2},\dots ,t_{n}))\right) dt_{n}\dots dt_{2} \\ &\leq W\left( H(r(x))\right) \int_{x_{2}^{0}}^{x_{2}}\int_{x_{3}^{0}}^{x_{3}}\dots \int_{x_{n}^{0}}^{x_{n}}d(x_{1,}t_{2},\dots ,t_{n}) \\ &\quad \times W\left( p(x_{1,}t_{2},\dots ,t_{n})f(x_{1,}t_{2},\dots ,t_{n})\right) dt_{n}\dots dt_{2}. \end{aligned} \end{equation} Dividing both sides of (\ref{2.23}) by $W(H(r(x)))$ we get \begin{equation} \begin{aligned} \frac{D_{1}r(x)}{W(H(r(x)))} &\leq \int_{x_{2}^{0}}^{x_{2}}\int_{x_{3}^{0}}^{x_{3}}\dots \int_{x_{n}^{0}}^{x_{n}}d(x_{1,}t_{2},\dots ,t_{n}) \\ &\quad \times W\left( p(x_{1,}t_{2},\dots ,t_{n})f(x_{1,}t_{2},\dots ,t_{n})\right) dt_{n}\dots dt_{2}, \end{aligned}\label{2.24} \end{equation} Note that for \begin{equation} G(z)=\int_{z^{0}}^{z}\frac{ds}{W(H(s))},\quad z\geq z^{0}>0 \label{2.25} \end{equation} it follows that \begin{equation} D_{1}G(r(x))=\frac{D_{1}r(x)}{W(H(r(x)))}, \label{2.26} \end{equation} From (\ref{2.25}) , (\ref{2.26}) and (\ref{2.24}), we have \begin{equation} \begin{aligned} D_{1}G(r(x)) &\leq \int_{x_{2}^{0}}^{x_{2}}\int_{x_{3}^{0}}^{x_{3}}\dots \int_{x_{n}^{0}}^{x_{n}}d(x_{1,}t_{2},\dots ,t_{n}) \\\ &\quad \times W\left( p(x_{1,}t_{2},\dots ,t_{n})f(x_{1,}t_{2},\dots ,t_{n})\right) dt_{n}\dots dt_{2}, \end{aligned} \label{2.27} \end{equation} Now setting $x_{1}=s$ in (\ref{2.27}) and then integrating with respect to $x_{1}^{0}$ to $x_{1}$, we obtain \begin{equation} G(r(x))\leq G(r(x_{1}^{0},x_{2},\dots ,x_{n}))+\int_{x_{0}}^{x}d(t)W(p(t)f(t))dt \label{2.28} \end{equation} Noting that $r(x_{1}^{0},x_{2},\dots ,x_{n})=\int_{x_{0}}^{\infty }d(t)W(a(t)p(t))dt$, we have \begin{equation} r(x)\leq G^{-1}\Big[ G\Big( \int_{x^{_{0}}}^{\infty }d(t)W(a(t)p(t))dt\Big) % +\int_{x^{_{0}}}^{x}d(t)W(p(t)f(t))dt\Big] . \label{2.29} \end{equation} The required inequality in (\ref{2.14}) follows from the fact $v(x)\leq r(x)$, (\ref{2.19}) and (\ref{2.29}) \end{proof} \begin{theorem} \label{t2} Let $u(x)$, $a(x)$, $b(x)$, $c(x)$, $d(x)$, $f(x)$, $W(u(x))$, and $H(u(x))$ be as defined in theorem \ref{t1}. Assume that $a(x),f(x)$ are non-increasing in the first variable $x_{1}$, for $x_{1}\in \mathbb{R}_{+}$. If \begin{equation} \begin{aligned} u(x) &\leq a(x)+b(x)\int_{x_{1}}^{\beta}c(s,x_{2},\dots ,x_{n})u(s,x_{2},\dots , x_{n})ds \\ &\quad +f(x)H\left( \int_{x}^{x_{0}}d(t)W(u(t))dt\right) , \end{aligned} \end{equation} for $\beta \geq 0,\;x\in \mathbb{R}_{+}^{n}$ with $\beta \geq x_{1}$ and $x\leq x^{_{0}}$. Then \begin{equation*} u(x)\leq \overline{p}(x)\Big\{ a(x)+f(x)H\Big( G^{-1}\Big[ G(\overline{A}% (t))+\int_{x}^{x_{0}}d(t)W(p(t)f(t))dt\Big] \Big) \Big\} , \end{equation*} for $\beta \geq 0,\;x\in \mathbb{R}_{+}^{n}$ with $\beta \geq x_{1}$, where \begin{gather*} \overline{p}(x)=1+b(x)\int_{x_{1}}^{\beta }c(s,x_{2},\dots ,x_{n})\exp \Big( \int_{x_{1}}^{s}b(\tau ,x_{2},\dots ,x_{n})c(\tau ,x_{2},\dots ,x_{n}) d\tau \Big) ds, \\ \overline{A}(t)=\int_{0}^{x^{_{0}}}d(t)W(a(t)\overline{p}(t))dt, \\ G(z)=\int_{z^{0}}^{z}\frac{ds}{W(H(s))},\quad z\geq z^{0}>0\,. \end{gather*} Here $G^{-1}$ is the inverse function of $G$ and \begin{equation*} G\Big( \int_{0}^{x^{_{0}}}d(t)W(a(t)p(t))dt\Big) +\int_{x}^{x^{_{0}}}d(t)W(p(t)f(t))dt, \end{equation*} is in the domain of $G^{-1}$ for $x\in \mathbb{R}_{+}^{n}$. \end{theorem} The proof is similar to the proof of Theorem \ref{t1} and so it is omitted. \begin{remark} \textrm{We note that in the special case $n=2$ (integral inequalities in two independent variables) $x\in \mathbb{R}_{+}^{2}$ and $x_{0}=(x_{1}^{0},x_{2}^{0})=(\infty ,\infty )$ in theorem \ref{t2}. our estimate reduces to Theorem 2.4 obtained by S. S. Dragomir and Y. H. Kim \cite{2}. } \end{remark} \begin{theorem} \label{t3} Let $u(x),a(x),b(x),c(x)$ and $f(x)$ be real-valued nonnegative continuous functions defined for $x\in \mathbb{R}_{+}^{n}$ and $L:\mathbb{R}% _{+}^{n+1}\to \mathbb{R}_{+}^{*}$ be a continuous functions which satisfies the condition \begin{equation} 0\leq L(x,u)-L(x,v)\leq M(x,v)\Phi ^{-1}(u-v), \label{2.30} \end{equation} for $u\geq v\geq 0$, where $M(x,v)$ is a real-valued nonnegative continuous function defined for $x\in \mathbb{R}_{+}^{n},v\in \mathbb{R}_{+}\mathbb{.}$ Assume that $\Phi :\mathbb{R}_{+}\to \mathbb{R}_{+}$ be a continuous and strictly increasing function with $\Phi (0)=0,\Phi ^{-1}$ is the inverse function of $\Phi $ and \begin{equation} \Phi ^{-1}(uv)\leq \Phi ^{-1}(u)\Phi ^{-1}(v), \label{2.31} \end{equation} for $u,v\in \mathbb{R}_{+}$, Assume that $a(x),f(x)$ are nondecreasing in the first variable $x_{1}$ for $x_{1}\in \mathbb{R}_{+}$. If \begin{equation} u(x)\leq a(x)+b(x)\int_{\alpha }^{x_{1}}c(s,x_{2},\dots ,x_{n})u(s,x_{2},\dots ,x_{n})ds+f(x)\Phi \Big( \int_{x_{0}}^{x}L(t,u(t))dt% \Big) , \label{2.32} \end{equation} for $\alpha \geq 0,\;x\in \mathbb{R}_{+}^{n}$ with $\alpha \leq x_{1}$ and $% x^{_{0}}0$, $v\geq 1$. \end{enumerate} \begin{theorem} \label{t5}Let $u(x)$, $a(x)$, $b(x)$, $c(x)$, $d(x)$, $f(x)$ be real-valued nonnegative continuous function defined for $x\in \mathbb{R}_{+}^{n}$ and let $g\in S$. Also let $W(u(x))$ be real-valued, positive, continuous, strictly nondecreasing, subadditive, and submultiplicative function for $% u(x)\geq 0$ and let $H(u(x))$ be a real-valued, continuous, positive, and nondecreasing function defined for $x\in \mathbb{R}_{+}^{n},$and $b(x)$ nonincreasing in the first variable $x_{1}$. Assume that a function $m(x)$ is nondecreasing in the first variable $x_{1}$ and $m(x)\geq 1$, which is defined by \begin{equation} m(x)=a(x)+f(x)H\Big( \int_{x^{_{0}}}^{x}d(t)W(u(t))dt\Big) , \label{3.1} \end{equation} for $x\in \mathbb{R}_{+}^{n}$, $x>x^{0}\geq 0$. If \begin{equation} u(x)\leq m(x)+b(x)\int_{\alpha }^{x_{1}}c(s,x_{2},\dots ,x_{n})g(u(s,x_{2},\dots ,x_{n}))ds, \label{3.2} \end{equation} for $\alpha \geq 0$, $x\in \mathbb{R}_{+}^{n}$ with $\alpha \leq x_{1}$, then \begin{equation} u(x)\leq F(x)\Big\{ a(x)+f(x)H\Big[ G^{-1}\Big( G(B(t))+% \int_{x^{_{0}}}^{x}d(t)W(F(t)f(t))dt\Big) \Big] \Big\}, \label{3.3} \end{equation} for $x\in \mathbb{R}_{+}^{n}$ , where \begin{gather} F(x)=\Omega ^{-1}\Big( \Omega (1)+\int_{\alpha }^{x_{1}}b(s,x_{2},\dots ,x_{n})c(s,x_{2},\dots ,x_{n})ds\Big) , \label{3.4} \\ B(t)=\int_{x^{_{0}}}^{\infty }d(t)W(a(t)F(t))dt, \label{3.5} \\ \Omega (\delta )=\int_{\varepsilon }^{\delta }\frac{ds}{g(s)},\quad \delta \geq \varepsilon >0. \label{3.6} \end{gather} Here $\Omega ^{-1}$ is the inverse function of $\Omega $, and $G,G^{-1}$ are defined in Theorem \ref{t1}, and $\Omega (1)+\int_{\alpha }^{x_{1}}b(s,x_{2},\dots ,x_{n})c(s,x_{2},\dots ,x_{n})ds$ is in the domain of $\Omega ^{-1}$, and \begin{equation*} G\Big( \int_{x^{_{0}}}^{\infty }d(t)W(a(t)F(t))dt)+\int_{x^{_{0}}}^{x}d(t)W(F(t)f(t)dt\Big), \end{equation*} is in the domain of $G^{-1}$ for $x\in \mathbb{R}_{+}^{n}$. \end{theorem} \begin{proof} We have $m(x)$ be a positive, continuous, nondecreasing in $x_{1}$ and $g\in S$, and $b(x)$ non-increasing in the first variable $x_{1}$. Then can be restated as \begin{equation} \frac{u(x)}{m(x)}\leq 1+\int_{\alpha }^{x_{1}}b(s,x_{2},x_{3},\dots ,x_{n})c(s,x_{2},x_{3},\dots ,x_{n})g(\frac{u(s,x_{2},x_{3},\dots ,x_{n})}{% m(s,x_{2},x_{3},\dots ,x_{n})})ds \label{3.7} \end{equation} The inequality (\ref{3.7}) may be treated as one-dimensional Bihari-Lasalle inequality the inequality type was given by Gyori \cite{3} (see \cite{1}), for any fixed $x_{2},x_{3},\dots ,x_{n}$, which implies \begin{equation} u(x)\leq F(x)m(x). \label{3.8} \end{equation} Here $F(x)$ is defined by (\ref{3.4}), by (\ref{3.1}) and (\ref{3.8}) we get \begin{equation} u(x)\leq F(x)\left\{ a(x)+f(x)H(v(x))\right\} , \label{3.9} \end{equation} where $v(x)$ is defined by \begin{equation*} v(x)=\int_{x^{_{0}}}^{x}d(t)W(u(t))dt. \end{equation*} Using the last argument in the proof of Theorem \ref{t1}, we obtain desired inequality in (\ref{3.3}). \end{proof} \begin{theorem} \label{t6} Let $u(x)$, $a(x)$, $c(x)$, $d(x)$, $f(x)$, $W(u(x)$, and $% H(u(x)) $ be as defined in the theorem \ref{t5} and let $g\in S$ and $b(x)$ be nonnegative continuous functions, nondecreasing in the first variable $% x_{1}$. Assume that a function $\overline{m}(x)$ is non-increasing in the first variable $x_{1}$ and $\overline{m}(x)\geq 1$, which is defined by \begin{equation} \overline{m}(x)=a(x)+f(x)H\Big( \int_{x}^{x^{0}}d(t)W(u(t))dt\Big) \label{3.10} \end{equation} for $x\in \mathbb{R}_{+}^{n}$, $x^{0}\geq x$. If \begin{equation} u(x)\leq \overline{m}(x)+b(x)\int_{x_{1}}^{\beta }c(s,x_{2},\dots ,x_{n})g(u(s,x_{2},\dots ,x_{n}))ds, \label{3.11} \end{equation} for $\beta \geq 0$, $x\in \mathbb{R}_{+}^{n}$ with $\beta \geq x_{1}$, then \begin{equation} u(x)\leq \overline{F}(x)\Big\{ a(x)+f(x)H\Big[ G^{-1}\Big( G(\overline{B}% (t))+\int_{x}^{x^{0}}d(t)W(\overline{F}(t)f(t))dt\Big) \Big]\Big\} , \label{3.12} \end{equation} for $x\in \mathbb{R}_{+}^{n}$. Here \begin{gather} \overline{F}(x)=\Omega ^{-1}\Big( \Omega (1)+\int_{x_{1}}^{\beta }b(s,x_{2},\dots ,x_{n})c(s,x_{2},\dots ,x_{n})ds\Big), \label{3.13} \\ \overline{B}(t)=\int_{0}^{x^{0}}d(t)W(a(t)\overline{F}(t))dt, \label{3.14} \end{gather} and $\Omega $ is defined in (\ref{3.6}). Here $\Omega ^{-1}$ is the inverse function of $\Omega $, and $G,G^{-1}$ are defined in theorem \ref{t1}, and $% \Omega (1)+\int_{x_{1}}^{\beta }b(s,x_{2},\dots ,x_{n})c(s,x_{2},\dots ,x_{n})ds$ is in the domain of $\Omega ^{-1}$, and \begin{equation*} G(\int_{0}^{x^{0}}d(t)W(a(t)\overline{F}(t))dt)+\int_{x}^{x^{0}}d(t)W(% \overline{F}(t)f(t))dt \end{equation*} is in the domain of $G^{-1}$ for $x\in \mathbb{R}_{+}^{n}$. \end{theorem} \begin{proof} We have $\overline{m}(x)$ positive, continuous, nonincreasing in $x_{1}$. Also $g\in S$ and $b(x)$ nondecreasing in the first variable $x_{1}$. Then (% \ref{3.11}) can be restated as \begin{equation} \frac{u(x)}{\overline{m}(x)}\leq 1+\int_{x_{1}}^{\beta }b(s,x_{2},x_{3},\dots ,x_{n})c(s,x_{2},x_{3},\dots ,x_{n})g\big(\frac{% u(s,x_{2},\dots ,x_{n})}{\overline{m}(s,x_{2},\dots ,x_{n})}\big)ds \label{3.15} \end{equation} This inequality can be treated as one-dimensional Bihari-Lasalle inequality \cite{3} for a fixed $x_{2},x_{3},\dots ,x_{n}$, which implies \begin{equation} u(x)\leq \overline{F}(x)\overline{m}(x) \label{3.16} \end{equation} where $\overline{F}(x)$ is defined by (\ref{3.13}). Now , by following last argument as in the proof of Theorem \ref{t2} , we obtain desired inequality in (\ref{3.12}) \end{proof} \begin{corollary} \label{c1} If $b(x)=1$ for $x\in R_{+}^{n}$, then from \begin{equation*} u(x)\leq \overline{m}(x)+\int_{x_{1}}^{\beta }c(s,x_{2},\dots ,x_{n})g(u(s,x_{2},\dots ,x_{n}))ds \end{equation*} with $\beta \geq x_{1}$, it follows that \begin{equation*} u(x)\leq \overline{F}(x)\Big\{ a(x)+f(x)H\Big[ G^{-1}\Big( G(\overline{B}% (t))+\int_{x}^{x^{0}}d(t)W(\overline{F}(t)f(t))dt\Big) \Big]\Big\} \end{equation*} for $x\in \mathbb{R}_{+}^{n}$ , where \begin{gather*} \overline{F}(x)=\Omega ^{-1}\Big( \Omega (1)+\int_{x_{1}}^{\beta }c(s,x_{2},\dots ,x_{n})ds\Big) \\ \overline{B}(t)=\int_{0}^{x^{0}}d(t)W(a(t)\overline{F}(t))dt \end{gather*} \end{corollary} \begin{remark} \textrm{We note that in the special case $n=2$ ,$x=(x_{1},x_{2})\in % \mathbb{R}_{+}^{2}$, and $x^{_{0}}=(\infty ,\infty )$ in corollary \ref{c1}. Our estimate reduces to Theorem 3.2 obtained by Dragomir and Kim \cite{2}. } \end{remark} \begin{theorem} \label{t7} Let $u(x)$, $a(x)$, $b(x)$, $c(x)$, $f(x)$, $L$, $M$, $\Phi $, and $\Phi ^{-1}$ be as defined in theorem \ref{t3}. Let $g\in S$ and $b(x)$ nonincreasing in the first variable $x_{1}$. Assume that a function $n(x)$ is nondecreasing in the first variable $x_{1}$ and $n(x)\geq 1$ which is defined by \begin{equation} n(x)=a(x)+f(x)\Phi \Big( \int_{x_{0}}^{x}L(t,u(t))dt\Big) \label{3.17} \end{equation} for $x\in \mathbb{R}_{+}^{n}$, $x\geq x_{0}\geq 0$. If \begin{equation} u(x)\leq n(x)+b(x)\int_{\alpha }^{x_{1}}c(s,x_{2},x_{3},\dots ,x_{n})g(u(s,x_{2},x_{3},\dots ,x_{n}))ds \label{3.18} \end{equation} for $\alpha \geq 0,\;x\in \mathbb{R}_{+}^{n}$ with $\alpha \leq x_{1}$, then \begin{equation} u(x)\leq F(x)\Big\{ a(x)+f(x)\Phi \Big[ e(x)\exp \Big( % \int_{x^{0}}^{x}M(t,a(t)F(t))\Phi ^{-1}\big( f(t)F(t)\big) dt\Big) \Big] % \Big\} \label{3.19} \end{equation} for $x\in \mathbb{R}_{+}^{n}$ , where $F(x)$ is defined in (\ref{3.4}), $% e(x) $ is defined in (\ref{2.35}), $\Omega $ is defined in (\ref{3.6}), Here $\Omega ^{-1}$ is the inverse function of $\Omega $, and \newline $\Omega (1)+\int_{\alpha }^{x_{1}}b(s,x_{2},\dots ,x_{n})c(s,x_{2},\dots ,x_{n})ds$ is in the domain of $\Omega $ for $x\in \mathbb{R}_{+}^{n}$. \end{theorem} \begin{proof} We follow an argument similar to that of Theorem \ref{t5}. We have $n(x)$ be a positive, continuous, nondecreasing in $x_{1}$ and $g\in S$, and $b(x)$ nonincreasing in the first variable $x_{1}$. Then can (\ref{3.18}) be restated as \begin{equation} \frac{u(x)}{n(x)}\leq 1+\int_{\alpha }^{x_{1}}b(s,x_{2},x_{3},\dots ,x_{n})c(s,x_{2},x_{3},\dots ,x_{n})g\big(\frac{u(s,x_{2},\dots ,x_{n})}{% n(s,x_{2},\dots ,x_{n})}\big)ds. \label{3.20} \end{equation} The inequality (\ref{3.20}) may be treated as one-dimensional Bihari-Lasalle inequality, for any fixed $x_{2},x_{3},\dots ,x_{n}$, which implies \begin{equation} u(x)\leq F(x)n(x) \label{3.21} \end{equation} where $F(x)$ is defined by (\ref{3.4}). From (\ref{3.17}) and (\ref{3.21}) we get \begin{equation} u(x)\leq F(x)\left[ a(x)+f(x)H\Big(\int_{x^{0}}^{x}L(t,u(t))dt\Big)\right] \label{3.22} \end{equation} Following the last argument in the proof of Theorem \ref{t3}, we obtain the desired inequality in (\ref{3.19}). \end{proof} \begin{theorem} Let $u(x)$, $a(x)$, $b(x)$, $c(x)$, $f(x)$, $L$, $M$, $\Phi $, and $\Phi ^{-1}$ be as defined in theorem \ref{t3}. Let $g\in S$ and $b(x)$ be nondecreasing in the first variable $x_{1}$. Assume that a function $% \overline{n}(x)$ is nonincreasing in the first variable $x_{1}$ and $% \overline{n}(x)\geq 1$, which is defined by \begin{equation} \overline{n}(x)=a(x)+f(x)\Phi \Big(\int_{x}^{x^{0}}L(t,u(t))dt\Big) \end{equation} for $x\in \mathbb{R}_{+}^{n}$, $x^{0}\geq x\geq 0$. If \begin{equation} u(x)\leq \overline{n}(x)+b(x)\int_{x_{1}}^{\beta }c(s,x_{2},\dots ,x_{n})g(u(s,x_{2},\dots ,x_{n}))ds \end{equation} for $\beta \geq 0,\;x\in \mathbb{R}_{+}^{n}$ with $\beta \geq x_{1}$, then \begin{equation*} u(x)\leq \overline{F}(x)\Big\{ a(x)+f(x)\Phi \Big[ \overline{e}(x) \exp \Big(\int_{x}^{x^{0}}M(t,a(t)\overline{F}(t))\Phi ^{-1}\big( f(t) \overline{F}(t)\big) dt\Big) \Big]\Big\} \end{equation*} for $x\in \mathbb{R}_{+}^{n}$, where $\overline{F}(x)$ is defined in (\ref {3.13}), $\overline{e}(x)$ is defined in (\ref{2.43}), $\Omega $ is defined in (\ref{3.6}). Here $\Omega ^{-1}$ is the inverse function of $\Omega $, and \newline $\Omega (1)+\int_{x_{1}}^{\beta }b(s,x_{2},\dots ,x_{n})c(s,x_{2},\dots ,x_{n})ds$ is in the domain of $\Omega $ for $x\in \mathbb{R}_{+}^{n}$. \end{theorem} The proof of this theorem follows by an argument similar to that of Theorem \ref{t7}; therefore, we omit it. \begin{corollary} \label{c2} if $b(x)=1$ for $x\in R_{+}^{n}$, then from \begin{equation*} u(x)\leq \overline{n}(x)+\int_{x_{1}}^{\beta }c(s,x_{2},\dots ,x_{n})g(u(s,x_{2},\dots ,x_{n}))ds, \end{equation*} for $\beta \geq 0$ with $\beta \geq x_{1}$, then it follows that \begin{equation*} u(x)\leq \overline{F}(x)\Big\{ a(x)+f(x)\Phi \Big[ \overline{e}(x)\exp \Big(\int_{x}^{x^{0}}M(t,a(t)\overline{F}(t))\Phi ^{-1}\big( f(t) \overline{F}(t)\big) dt\Big) \Big]\Big\} \end{equation*} for $x\in \mathbb{R}_{+}^{n}$, where \begin{gather*} \overline{F}(x)=\Omega ^{-1}\Big( \Omega (1)+\int_{x_{1}}^{\beta }c(s,x_{2},\dots ,x_{n})ds\Big),\\ \overline{e}(x)=\int_{x}^{x^{0}}L(t,\overline{p}(t)a(t))dt, \\ \overline{p}(x)=1+\int_{x_{1}}^{\beta }c(s,x_{2},\dots ,x_{n}) \exp \Big(\int_{x_{1}}^{s}c(\tau ,x_{2},\dots ,x_{n})d\tau \Big) ds, \end{gather*} for $x\in \mathbb{R}_{+}^{n}.\Omega $ is defined in (\ref{3.6}) , where $\Omega ^{-1}$ is the inverse function of $\Omega $, and $\Omega (1)+\int_{x_{1}}^{\beta }c(s,x_{2},\dots ,x_{n})ds$ is in the domain of $\Omega $ for $x\in \mathbb{R}_{+}^{n}$. \end{corollary} \begin{remark} \textrm{We note that in the special case $n=2$, $x=(x_{1},x_{2})\in % \mathbb{R}_{+}^{2}$, and $x^{0}=(\infty ,\infty )$ in corollary \ref{c2}. our estimate reduces to Theorem 3.4 obtained by Dragomir and Kim \cite{2}.} \end{remark} \begin{remark} \rm (1) All the preceding results remain valid when\newline $b(x)\int_{\alpha }^{x_{1}}c(s,x_{2},\dots ,x_{n})g(u(s,x_{2},\dots ,x_{n}))ds$ is replaced by the general function \newline $b_{i}(x)\int_{\alpha _{i}}^{x_{i}}c_{i}(x_{1,.}dots,x_{i-1},s_{i},x_{i+1},\dots ,x_{n})g(u(x_{1,.}\dots ,x_{i-1},s_{i},x_{i+1},\dots ,x_{n}))ds_{i}$,\newline with $i=2,\dots ,n$ fixed, and $\alpha _{i}\geq 0$, $x=(x_{1},\dots x_{n})\in \mathbb{R}_{+}^{n}$ and with $\alpha _{i}\leq s_{i}\leq x_{i}$, $x_{i},s_{i}\in \mathbb{R}_{+}$, \noindent (2) The above results remain valid when\newline $b(x)\int_{x_{1}}^{\beta }c(s,x_{2},\dots ,x_{n})g(u(s,x_{2},\dots ,x_{n}))ds $ is replaced by the general function\newline $b_{i}(x)\int_{x_{i}}^{\beta _{i}}c_{i}(x_{1,.}\dots ,x_{i-1},s_{i},x_{i+1},\dots ,x_{n})g(u(x_{1,.}\dots ,x_{i-1},s_{i},x_{i+1},\dots ,x_{n}))ds_{i}$,\newline with $i=2,\dots ,n$ fixed, and $\alpha _{i}\geq 0$, $x=(x_{1},\dots x_{n})\in \mathbb{R}_{+}^{n}$ and with $\alpha _{i}\leq s_{i}\leq x_{i}$, $% x_{i},s_{i}\in \mathbb{R}_{+}$, where $b_{i}(x)$ and $c_{i}(x)$ be real-valued nonnegative continuous function defined for $x\in \mathbb{R}_{+}^{n}$, for all $i=2,\dots ,n$. \end{remark} In a future work, we will present some applications for the results obtained in this work. \begin{thebibliography}{9} \bibitem{1} D. Bainov and P. Simeonov, \textit{Integral Inequalities and Applications}, Kluwer Academic Publishers, Dordrecht, (1992). \bibitem{2} S. S. Dragomir and Y.-Ho Kim, \textit{Some Integral Inequalities for Function of two Variables,} EJDE , Vol. 2003 (2003), No.10, pp.1--13. \bibitem{3} I. Gyori, \textit{A Generalization of Bellman's Inequality for Stieltjes Integrals and a Uniqueness Theorem}, Studia Sci. Math. Hungar, 6 (1971), 137--145. \bibitem{4} D. R. Snow, \textit{Gronwall's Inequality for Systems of Partial Differential Equation in tow independent Variables}, Proc. Amer. Math. Soc. 33 (1972), 46--54. \bibitem{5} C. C. Yeh and M.-H. Shim, \textit{The Gronwall-Bellman Inequality in Several Variables} , J. Math. Anal. Appl., 87 (1982), 157--1670. \bibitem{6} C. C. Yeh, \textit{Bellman-Bihari Integral Inequality in n Independent variables}, J. Math. Anal. Appl., 87 (1982), 311--321 \bibitem{7} C. C. Yeh, \textit{On Some Integral Inequalities in n Independent variables and their Applications}, J. Math. Anal. Appl., 87 (1982), 387--410. \end{thebibliography} \end{document}