\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2007(2007), No. 57, 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 2007 Texas State University - San Marcos.} \vspace{8mm}} \begin{document} \title[\hfilneg EJDE-2007/57\hfil Nonexistence of the gradient blow-up phenomenon] {Sufficient conditions for nonexistence of gradient blow-up for nonlinear parabolic equations} \author[A. S. Tersenov\hfil EJDE-2007/57\hfilneg] {Aris S. Tersenov} \address{Aris S. Tersenov \newline Department of Mathematics and Statistics \\ University of Cyprus \\ P.O. Box 20537 \\ 1678 Nicosia, Cyprus \newline Tel.:+357 22892560, Fax:+357 22892550} \email{aterseno@ucy.ac.cy} \thanks{Submitted February 8, 2006. Published April 17, 2007.} \subjclass[2000]{35K55, 35K15, 35A05} \keywords{Bernstein-Nagumo condition; gradient blow-up; a priori estimates \hfill\break\indent nonlinear parabolic equation} \begin{abstract} In this paper we study the initial-boundary value problems for nonlinear parabolic equations without Bernstein-Nagumo condition. Sufficient conditions guaranteeing the nonexistence of gradient blow-up are formulated. In particular, we show that for a wide class of nonlinearities the Lipschitz continuity in the space variable together with the strict monotonicity with respect to the solution guarantee that gradient blow-up cannot occur at the boundary or in the interior of the domain. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{remark}[theorem]{Remark} \section{Introduction} In the present paper we consider the nonlinear equation \begin{equation}\label{1.1} u_t=F(t,x,u,u_x,u_{xx}) \quad \text{in } Q_T=(-l,l)\times (0,T), \end{equation} coupled with one of the boundary conditions \begin{gather}\label{1.2} u_x(t,-l)=u_x(t,l)=0, \\ \label{1.3} u_x+\sigma_{1}(t,x,u)\big|_{x=-l}= u_x+\sigma_{2}(t,x,u)\big|_{x=l}=0, \\ \label{1.4} u(t,-l)=u(t,l)=0 \end{gather} and the initial condition \begin{equation}\label{1.5} u(0,x)=u_0(x). \end{equation} We assume that $F(t,x,u,p,r)$ is continuously differentiable with respect to $r$ and satisfies the parabolicity condition, i.e. \begin{equation}\label{i.1} F_r(t,x,u,p,r)>0 \quad \text{for} \quad (t,x,u,p,r) \in \overline{Q}_T\times [- M,M]\times \mathbb{R}^2. \end{equation} Write equation \eqref{1.1} in the form \begin{equation}\label{i.2} u_t=F_r(t,x,u,u_x,\lambda u_{xx})u_{xx}+F(t,x,u,u_x,0),\quad \lambda\in [0,1], \end{equation} using the mean value theorem. The well known Bernstein-Nagumo condition \cite{BERN, SNB, Nag} (see also \cite{ACONST, SNK, KRYL, LSV, LIB, SER}) in the case of equation \eqref{i.2} appears as \begin{equation}\label{i.3} \frac{|F(t,x,u,p,0)|}{F_r(t,x,u,p,r)}\leq \phi (|p|) \quad \text{for} \quad (t,x,u,p,r)\in \overline{Q}_T\times [- M,M]\times \mathbb{R}^2, \end{equation} where $\phi(\rho)$ is a nondecreasing positive function such that $$ \int^{+\infty}\frac {\rho d\rho}{\phi (\rho)} = + \infty. $$ Condition \eqref{i.3} guarantees global a priori estimate for the gradient of bounded solutions. There are examples showing that a violation of the Bernstein-Nagumo condition can imply the gradient blow-up on the boundary as well as at interior points of the domain (see \cite{ANFI, ASISH, DL, Y, KUT, SOUPL, VZ}), while the solution itself remains bounded. Recently, in \cite{TT}, condition \eqref{i.3} was substituted by a less restrictive one that allows an arbitrary growth of $F(t,x,u,p,0)$ with respect to $p$ (see also \cite{souplet}). Let us recall some of the main results that were established in \cite{TT}. Suppose that the right hand side of equation \eqref{i.2} can be represented as follows \begin{equation}\label{i.4} F(t,x,u,p,0)=f_1(t,x,u,p)+f_2(t,x,u,p), \end{equation} where $f_2$ satisfies the restrictions \begin{gather}\label{i.5} f_2(t,y,u_1,p)-f_2(t,x,u_2,p)\geq 0, \\ \label{i.6} f_2(t,x,u_1,-p)-f_2(t,y,u_2,-p)\geq 0 \end{gather} for $t\in [0,T]$, $-l\leq y< x\leq l$, $-M\leq u_10$ in the case of problems \eqref{1.1}, \eqref{1.2}, \eqref{1.5} and \eqref{1.1}, \eqref{1.3}, \eqref{1.5} (see also Remark \ref{rmk1}). When $f_2$ is independent of $x$, one can easily see that \eqref{i.5}, \eqref{i.6} mean that $f_2(t,u,p)$ is a nonincreasing function with respect to $u$. Unfortunately if $f_2$ depends also on $x$ and satisfies \eqref{i.5}, \eqref{i.6}, its behavior becomes rather complicated. The goal of this paper is to show that under some additional assumptions the strict monotonicity of $f_2(t,x,u,p)$ in $u$ is sufficient for nonexistence of the gradient blow-up of a bounded solution. In order to motivate these additional assumptions we will recall some facts from the theory of viscosity solutions. One can easily see that in the case where $f_2$ is independent of $x$, conditions \eqref{i.5}, \eqref{i.6} reminds us one of the main assumptions (the properness, see \cite{CLI}) under which the notion of viscosity solution is introduced. For example, if we assume in \eqref{i.2} that $F_r$ is independent of $u$ and $F(t,x,u,p,0)=f_2(t,u,p)$ satisfies \eqref{i.5}, \eqref{i.6}, then \begin{equation}\label{in.1} u_t-F_r(t,x,u_x,\lambda u_{xx})u_{xx}-f_2(t,u,u_x)=0,\quad \lambda\in [0,1], \end{equation} is proper. Moreover in \cite{KK}, in particular, it was shown that if $F_r=F_r(p,r)$ is locally strictly elliptic and $F(t,x,u,p,0)=f_2(u,p)$ satisfies \eqref{i.5}, \eqref{i.6}, then there exists a unique continuous viscosity solution to the Dirichlet problem \begin{equation}\label{in.2} \begin{gathered} u_t-F_r(u_x,\lambda u_{xx})u_{xx}-f_2(u,u_x)=0,\quad \lambda\in [0,1], \\ u(t,-l)=u(t,l)=0, \quad u(0,x)=u_0(x), \end{gathered} \end{equation} provided (\ref{in.2}) has a sub- and supersolution satisfying initial-boundary data. Comparing this result with the results of \cite{TT}, we conclude that a viscosity solution of the mentioned above problem becomes classical, if additionally $f_2$ satisfies \eqref{i.7} and the coefficients have sufficient smoothness. The situation becomes more complicated, when $F_r$ and $f_2$ depends also on $t$ and $x$. First of all we have to assume that the elliptic operator is uniformly proper. It means that $f_2$ is strictly decreasing in $u$ \begin{equation}\label{n.1} f_2(t,x,u_1,p)-f_2(t,x,u_2,p)\geq \gamma_0(u_2-u_1) \end{equation} for $u_2\geq u_1$, $x\in[-l,l]$, $p\in {\bf{R}}$, for fixed $t\in [0,T]$, where $\gamma_0$ is a positive constant. The second assumption is a structure condition on the continuity of the elliptic operator in $x$ (see \cite{CLI}). Assumptions that we use in order to improve \eqref{i.5}, \eqref{i.6} were inspired by these two assumptions under which the existence of a viscosity solution can be proved. We proceed now to the statement of main results of the paper. Assume that $F(t,x,u,p,r)$ is defined for $(t,x)\in \overline{Q}_T$, $u\in [-M,M]$ and arbitrary $(p,r)$ and is bounded on every compact set in $\overline{Q}_T\times [-M,M]\times {\bf{R}}^2$. Suppose that \begin{equation}\label{n.3} |f_2(t,x,u,p)-f_2(t,y,u,p)|\leq K_1(t,x,y,u,p)|x-y| \end{equation} for $t\in[0,T]$, $x,y\in[-l,l]$, $00$. Denote by ${\bf{V}}$ the following set $$ {\bf{V}}=\{(t,x,y)\in \overline{Q}_T, 00$ for problem \eqref{1.1}, \eqref{1.2}, \eqref{1.5}. Analogous results we obtain for problem \eqref{1.1}, \eqref{1.3}, \eqref{1.5} (see Corollary \ref{b}). In the case of problem \eqref{1.1}, \eqref{1.4}, \eqref{1.5} we also need assumption \eqref{i.7} to obtain the nonexistence of the gradient blow-up on the boundary as well as in the interior of $Q_T$ (see Corollary \ref{c}). Note (see Remark \ref{rmk1}) that if $f_2(t,x,0,p)=0$, then condition \eqref{i.7} is a simple consequence of the strict monotonicity of $f_2$ in $u$ (see condition \eqref{n.4}). Consider now the case when $f_2(t,x,u,p)=X(t,x)U(u)H(p)$ (special case). Suppose that \begin{equation}\label{3.1} |f_2(t,x,u,p)-f_2(t,y,u,p)|\leq K_2|U(u)||H(p)||x-y| \end{equation} for $t\in[0,T]$, $x,y\in[-l,l]$, $0u_1$, $p\in [-q_1, -q_0]\cup [q_0, q_1]$, where $\gamma_1,\gamma_0>0$ and without loss of generality we assume that $U(u(t,x))$ is a strictly decreasing function. Put $\gamma_2=\min_{t,x\in [0,T]\times [-l,l]}|X(t,x)|>0$. Consider problem \eqref{1.1}, \eqref{1.2}, \eqref{1.5}. \begin{theorem}\label{d}. Let $u(t,x)$ be a classical solution of problem \eqref{1.1}, \eqref{1.2}, \eqref{1.5}. Suppose that conditions \eqref{i.1}, \eqref{i.4}, \eqref{i.8} - \eqref{i.10}, \eqref{3.1}, \eqref{3.2} are fulfilled. Then in $\overline{Q}_T$ the inequality $$ |u_x(t,x)|\leq C_4 $$ holds, where the constant $C_4$ depends on $\mathop{\rm osc}(u)$, $M$, $\psi$, $K_2$, $\gamma_1$ and $\gamma_2$. \end{theorem} From Theorem \ref{d} it follows that the gradient of a bounded solution of \eqref{i.2} cannot blow-up in the interior of $Q_T$ for any $T>0$ if $f_2=X(t,x)U(u)H(p)$ is Lipschitz continuous in $x$ and strictly decreasing in $u$. Concerning the special case see also Remark \ref{rmk5}. Comparing Theorem \ref{a} with Theorem \ref{d} one can easily see that in the case where $f_2$ is an arbitrary function of variables $t$, $x$, $u$, $p$, besides the Lipschitz continuity in $x$ and strict monotonicity in $u$, we need to impose some additional structure conditions regarding the behavior of $f_2$ in $p$ (see also Remark \ref{rmk3}) . In Section 1 we obtain the a priori estimate of the gradient of a bounded solution in the general case $f_2=f_2(t,x,u,p)$. In Section 2 we obtain the a priori estimate of the gradient of a bounded solution in the case where $f_2=X(t,x)U(u)H(p)$. We remark that based on these a priori estimates one can prove the existence theorems for the initial-boundary value problems for \eqref{1.1}, using the well-known fixed point theorem (see \cite{LIB}). The proofs are exactly the same as in \cite{TT}. \section{Gradient estimates in the general case} In this section we obtain global a priori estimates of the gradient of classical solutions for boundary value problems for \eqref{1.1}, in the case where $f_2(t,x,u,p)$ is an arbitrary bounded function of variables $t$, $x$, $u$, $p$. Recall that a classical solution is a function belonging to $C^{1,2}_{t,x}(Q_T)\cap C^{0,1}_{t,x}(\bar{Q}_T)$ in the case of problem \eqref{1.1}, \eqref{1.2}, \eqref{1.5} or \eqref{1.1}, \eqref{1.3}, \eqref{1.5} and to $C^{1,2}_{t,x}(Q_T)\cap C^{0}(\bar{Q}_T)$ for problem \eqref{1.1}, \eqref{1.4}, \eqref{1.5}. We use here Kruzhkov's idea of introducing a new spatial variable \cite{SNK,KRYJ} and the technique developed in \cite{TT}. \begin{proof}[Proof of Theorem \ref{a}] Consider equation \eqref{1.1} in the form \eqref{i.2} at two different points $(t,x)$ and $(t,y)$: \begin{gather}\label{2.1} u_t=F_r(t,x,u,u_x,\lambda u_{xx}) u_{xx} + F(t,x,u,u_x,0), \quad \lambda\in [0,1], \quad u=u(t,x), \\ \label{2.2} u_t=F_r(t,y,u,u_y,\mu u_{yy}) u_{yy} + F(t,y,u,u_y,0), \quad \mu\in [0,1], \quad u=u(t,y). \end{gather} Introduce the function $v(t,x,y)=u(t,x)-u(t,y)$. In $\Omega=\{(t,x,y):00 \quad \text{for} \quad \tau\in [0,\tau_0]. \end{equation} Represent the solution of \eqref{2.5}, \eqref{2.6} in parametrical form $$ h(q)=\int^{q_1}_q \frac{\rho d\rho}{\psi(\rho)}, \quad \tau(q)=\int^{q_1}_q \frac{d\rho}{\psi(\rho)}. $$ The parameter $q$ varies in the interval $[q_0,q_1]$, where $K^*< q_00, \\ \tilde{w}_x(t_1,x_1,y_1)=e^{-t}[u_{x}(t_1,x_1) - h'(x_1-y_1)]=0, \\ \tilde{w}_y(t_1,x_1,y_1)=e^{-t}[-u_{y}(t_1,y_1) + h'(x_1-y_1)]=0 \end{gather*} and as a consequence \begin{equation}\label{31.2} u(t_1,x_1)>u(t_1,y_1), \quad u_{x}(t_1,x_1)=u_{y}(t_1,y_1)=h'(x_1-y_1)>0. \end{equation} Represent the right-hand side of inequality \eqref{31.1} in the following way \begin{equation}\label{n1.1} \begin{aligned} &e^{-t}[f_2(t,y,u(t,y),u_y(t,y)) - f_2(t,x,u(t,x),u_x(t,x))]\\ &=e^{-t}[f_2(t,y,u(t,y),u_y(t,y)) -f_2(t,x,u(t,y),u_y(t,y))\\ &\quad + f_2(t,x,u(t,y),u_y(t,y)) - f_2(t,x,u(t,x),u_x(t,x))], \end{aligned} \end{equation} where we subtract and add the term $f_2(t,x,u(t,y),u_y(t,y))$. So at the maximum point $(t_1,x_1,y_1)$, using \eqref{n.3}, \eqref{n.4}, (\ref{31.2}), we obtain \begin{equation}\label{n1.2} \begin{aligned} &\widetilde{L}_1(\tilde{w})\\ &\geq e^{-t_1}[f_2(t_1,y_1,u(t_1,y_1),u_y(t_1,y_1)) - f_2(t_1,x_1,u(t_1,y_1),u_y(t_1,y_1))\\ &\quad + f_2(t_1,x_1,u(t_1,y_1),u_y(t_1,y_1)) - f_2(t_1,x_1,u(t_1,x_1),u_x(t_1,x_1))]\\ &\geq e^{-t_1}\Big[-K_1\Big(t_1,x_1,y_1,u(t_1,y_1),h'(x_1-y_1)\Big)(x_1-y_1)\\ &\quad + \gamma\Big(t_1,x_1,y_1,u(t_1,x_1),u(t_1,y_1),h'(x_1-y_1)\Big) (u(t_1,x_1)-u(t_1,y_1))\Big]. \end{aligned} \end{equation} Consider now the difference $u(t_1,x_1)-u(t_1,y_1)$. Due to the fact that $$ \tilde{w}(t_1,x_1,y_1)=e^{-t}[u(t_1,x_1) - u(t_1,y_1) - h(x_1-y_1)]>0, $$ we have $$ u(t_1,x_1) - u(t_1,y_1)> h(x_1-y_1)=h(x_1-y_1)-h(0)=h'(\xi)(x_1-y_1) $$ for some $\xi\in [0, \tau_0]$. Thus one can rewrite the inequality (\ref{n1.2}) in the following way \begin{equation}\label{n1.3} \begin{aligned} \widetilde{L}_1(\tilde{w}) &\geq e^{-t_1}\Big[-K_1\Big(t_1,x_1,y_1,u(t_1,y_1),h'(x_1-y_1)\Big)\\ &\quad + \gamma\Big(t_1,x_1,y_1,u(t_1,x_1),u(t_1,y_1),h'(x_1-y_1)\Big) h'(\xi)\Big](x_1-y_1). \end{aligned} \end{equation} Using now \eqref{n.7} we conclude that \begin{equation}\label{n1.4} \widetilde{L}_1(\tilde{w})\geq e^{-t_1}\Big[-Ch'^{\alpha}(x_1-y_1)+ h'(\xi)\Big]\gamma_0(x_1-y_1). \end{equation} To obtain the contradiction with $\widetilde{L}_1(\tilde{w})\big|_{(t_1,x_1,y_1)}<0$ we have to show that (recall that $x_1>y_1$) $$ -Ch'^{\alpha}(x_1-y_1)+ h'(\xi)\geq 0. $$ Using the fact that $q_0\leq h'\leq q_1$, we arrive to the inequality $$ -Cq^{\alpha}_1+q_0\geq 0. $$ Thus if \begin{equation}\label{n1.5} q_0\geq Cq_1^{\alpha}, \end{equation} then $\widetilde{L}_1(\tilde{w})\big|_{(t_1,x_1,y_1)}\geq 0$. Obviously, when $\alpha<1$, there exist $q_0$ and $q_1>q_0$ such that inequality \eqref{n1.5} takes place. Solving the system $$ q_0\geq Cq_1^{\alpha}, \quad q_1>q_0, $$ one can easily obtain that for $q_0>C^{\frac{1}{1-\alpha}}$ inequality \eqref{n1.5} takes place for $q_00$ \begin{gather*} -\tilde{w}_y(t,x,-l)=e^{-t}(u_{y}(t,-l) - h'(x+l))=-e^{-t}h'(x+l)<0, \\ \tilde{w}_x(t,l,y)= e^{-t}(u_{x}(t,l) - h'(l-y))=-e^{-t}h'(l-y)<0. \end{gather*} Thus the function $\tilde{w}(t,x,y)$ cannot attain its positive maximum neither on $Q_1$ nor on $Q_2$ since $-\partial/\partial y$ and $\partial/\partial x$ are here outward normal derivatives with respect to $P$. Consequently, $\tilde{w}\big|_{\Gamma}\leq 0$ and hence $\tilde{w}(t,x,y)\leq 0$ in $\overline{P}$. The case when $\tau_0\geq 2l$ can be treated similarly. The only difference is the absence of the boundary $x-y=\tau_0$. We put $\widetilde{Q}_1=\{(t,x):0u(\tilde{t}_1,\tilde{x}_1), \quad u_x(\tilde{t}_1,\tilde{x}_1)=u_{y}(\tilde{t}_1,\tilde{y}_1)= -h'(\tilde{x}_1-\tilde{y}_1)<0. $$ Using inequalities \eqref{n.3} - \eqref{n.7} we obtain in the same way that $\widetilde{L}_1(\tilde{w}_1)\geq 0$. From this contradiction it follows that $\tilde{w}_1$ cannot attain its positive maximum in $\overline{P}\setminus \Gamma$. Consider $\tilde{w}_1$ on $\Gamma$. One can easily see that all considerations concerning the estimate of the function $\tilde{w}$ on the boundary $\Gamma$ can be done without any changes in estimate of $\tilde{w}_1$. Thus we have that \begin{equation}\label{5.1} u(t,y) - u(t,x)\leq h(x-y) \quad {\rm in} \quad \overline{P}. \end{equation} Combining (\ref{5.1}) with (\ref{41.3}) we get $$ |u(t,x) - u(t,y)|\leq h(x-y) \quad \text{in } \overline{P}. $$ In view of the symmetry of the variables $x$, $y$ in the same manner we examine the case $y>x$. As a result we have that for $$ 0\leq t\leq T, \quad |x|\leq l, \quad |y|\leq l, \quad 0<|x-y|\leq \tau_0 $$ the inequality $$ \big|\frac{u(t,x) - u(t,y)}{x-y}\big| \leq \frac{h(|x-y|) -h(0)}{|x-y|} $$ holds and as a consequence we have $$ |u_x(t,x)|\leq h'(0)=q_1=C_1. $$ Theorem \ref{a} is proved. \end{proof} Let us pass to problem \eqref{1.1}, \eqref{1.3}, \eqref{1.5}. \begin{corollary}\label{b} Let $u(t,x)$ be a classical solution of \eqref{1.1}, $\eqref{1.3}$, $\eqref{1.5}$ and all conditions of Theorem $\ref{a}$ are fulfilled. Then in $\overline{Q}_T$ the inequality $$ |u_x(t,x)|\leq C_2 $$ holds, where the constant $C_2$ depends only on $\mathop{\rm osc}(u)$, $N_1$, $N_2$, $\psi$, $C$, $\alpha$, where $N_i=\sup|\sigma_i|$ (the supremum is taken over the set $[0,T]\times [-M,M]$). \end{corollary} \begin{proof} The proof of this corollary differs from the proof of Theorem \ref{a} only in the selection of $q_0$ and in analyzing of the behavior of $\tilde {w}(t,x,y)$ on bounds $Q_1$ ($\widetilde{Q}_1$) and $Q_2$ ($\widetilde{Q}_2$). We select the quantity $q_0$ so that \begin{equation}\label{6.1} q_0>max \left \{C^{\frac{1}{1-\alpha}},K, N_1, N_2 \right \}. \end{equation} and follow the proof of \cite[Lemma 2, and Corollary 1.2]{TT}. Corollary \ref{b} is proved. \end{proof} Consider now problem \eqref{1.1}, \eqref{1.4}, \eqref{1.5}. \begin{corollary}\label{c}. Let $u(t,x)$ be a classical solution of \eqref{1.1}, \eqref{1.4}, \eqref{1.5} and all conditions of Theorem \ref{a} are fulfilled. Suppose in addition that condition \eqref{i.7} is fulfilled and $u_0(\pm l)=0$. Then in $\overline{Q}_T$ the following inequality $$ |u_x(t,x)|\leq C_3 $$ holds, where the constant $C_3$ depends only on $\mathop{\rm osc}(u)$ and $\psi$, $C$, $\alpha$. \end{corollary} \begin{proof} The proof of this corollary differs from the proof of Theorem $\ref{a}$ only in analyzing of the behavior of $w(t,x,y)$ on $Q_1$ ($\tilde{Q}_1)$ and $Q_2$ ($\tilde{Q}_2)$ (see the proof of \cite[Lemma 3 and Corollary 1.3]{TT}). Corollary \ref{c} is proved. \end{proof} \begin{remark} \label{rmk1} \rm One can easily see that from Theorem \ref{a} and Corollary \ref{b} it immediately follows that conditions \eqref{n.3}-\eqref{n.7} are sufficient for the nonexistence of the gradient blow-up of a bounded solution in the interior of $Q_T$ for problems \eqref{1.1}, \eqref{1.2}, \eqref{1.5} and \eqref{1.1}, \eqref{1.3}, \eqref{1.5}. Concerning problem \eqref{1.1}, \eqref{1.4}, \eqref{1.5}, one has to impose condition \eqref{i.7} supplementary to \eqref{n.3}-\eqref{n.7} in order to obtain the nonexistence of the gradient blow-up of a bounded solution in $\overline{Q}_T$. Obviously, if we suppose that $f_2(t,x,0,p)=0$, then condition \eqref{i.7} is a simple consequence of the strict monotonicity of $f_2$ in $u$ (see condition \eqref{n.4}) Thus if $f_2(t,x,0,p)=0$, then the nonexistence of the gradient blow-up in $\overline{Q}_T$ for problem \eqref{1.1}, \eqref{1.4}, \eqref{1.5} follows from \eqref{n.3}-\eqref{n.7} and we do not need condition \eqref{i.7}. \end{remark} \begin{remark} \label{rmk2} \rm Note that if $f_1$ satisfies Bernstein-Nagumo condition \eqref{i.3} then for every $q_0$ there always exists $q_1>q_0$ such that \eqref{i.9} takes place. Thus if $\alpha\leq 0$ in \eqref{n.7}, then inequality \eqref{n1.5} does not depend on $q_1$ and we can always construct $h(\tau)$ that satisfies \eqref{2.6} (due to the divergence of the integral in \eqref{i.9} in this case). \end{remark} \begin{remark} \label{rmk3} \rm Put $f_1=0$. In this case $\psi(p)=0$ and $h''=0$. From \eqref{2.6} it follows that $h=\frac{osc\, u}{\tau_0}\tau$, $h'=\frac{osc\, u}{\tau_0}$. Thus we have that $q_0=q_1=\frac{osc\, u}{\tau_0}$. Consequently \eqref{n1.5} takes the form \begin{equation}\label{r3.1} q_0\geq Cq_0^{\alpha} \end{equation} that is fulfilled for $q_0\geq C^{\frac{1}{1-\alpha}}$ and $\alpha<1$. Obviously, when $\alpha=1$, then in order to obtain the gradient a priori estimate one has to suppose that $C\leq 1$. One can easily check that even in the case where $f_1=0$ \eqref{r3.1} holds for $q_0\leq C^{\frac{1}{1-\alpha}}$ if $\alpha>1$. Thus we can prove Theorem \ref{a} with $\alpha>1$ only for $K\leq C^{\frac{1}{1-\alpha}}$. Note that condition \eqref{n.7} can be generalized in the following way $$ \max_{{\bf{V}}}\frac{K_1(t,x,y,u_1,p)}{\gamma(t,x,u_1,u_2,p)}\leq \Psi(|p|), $$ where $\Psi(\rho)$ is a nondecreasing positive function. As a consequence condition \eqref{n1.5} appears as $$ q_0\geq \Psi(q_1). $$ \end{remark} \begin{remark} \label{rmk4} \rm Let us give some simple examples of functions that satisfy \eqref{n.3}-\eqref{n.7} and at the same time do not satisfy \eqref{i.5}, \eqref{i.6}. Easy calculations show that, for example, the functions $$ f_2(t,x,u,p)=x|p|^{\mu}-u|p|^{\nu}, \quad \forall \quad \mu,\nu \quad \hbox{such that } \mu-\nu\leq 1, $$ $$ f_2(t,x,u,p)=-ue^{(x+c)p^{\alpha}}, \quad \alpha<1, \; c>l, $$ where $\alpha$ is such that $p^{\alpha}$ is defined, satisfy \eqref{n.3}-\eqref{n.7} and do not satisfy \eqref{i.5}, \eqref{i.6}. Moreover, in the case when $f_2=-ue^{(x+c)p^{\alpha}}$, $\alpha<1$, $c>l$, problem \eqref{1.4}, \eqref{1.5} as well as problems \eqref{1.2}, \eqref{1.5} and \eqref{1.3}, \eqref{1.5}, for example, for the equation \begin{equation}\label{sup.1} u_t=a(t,x,u,u_x)u_{xx}+f_2, \quad a>0, \end{equation} have a global classical solution for any Lipschitz continuous initial data. When $f_2(t,x,u,p)=x|p|^{\mu}-u|p|^{\nu}$, $\mu-\nu\leq 1$, problems \eqref{sup.1}, \eqref{1.2}, \eqref{1.5} and \eqref{sup.1}, \eqref{1.3}, \eqref{1.5} have a global classical solution for any Lipschitz continuous initial data. \end{remark} \section{Gradient estimates in the special case} In this section we obtain global a priori estimates of the gradient of classical solutions for boundary-value problems for equation \eqref{1.1} where $f_2=X(t,x)U(u)H(p)$. One can easily see that in this case conditions \eqref{n.3}, \eqref{n.4} take the form $$ |f_2(t,x,u,p)-f_2(t,y,u,p)|\leq K_2|U(u)||H(p)||x-y| $$ for $t\in[0,T]$, $x,y\in[-l,l]$, $0u_1$, $p\in [-q_1, -q_0]\cup [q_0, q_1]$, where $\gamma_1,\gamma_0>0$. Recall that without loss of generality we assume that $U(u(t,x))$ is a strictly decreasing function. The last assumption means that $X(t,x)H(p)>0$. Put $\gamma_2=\min_{t,x\in [0,T]\times [-l,l]}|X(t,x)|>0$. Obviously if $f_2(t,x,u,p)=X(t,x)U(u)H(p)$, then condition \eqref{n.7} is fulfilled with $C=\frac{K_2|U(-M)|}{\gamma_1\gamma_2}$, $\alpha=0$ and as a consequence can be dropped. \begin{proof}[Proof of Theorem \ref{d}] The proof differs from the proof of the previous theorem only in the choice of the quantity $q_0$. In the general case $q_0$ depends on $q_1$ and $\alpha$ (see \eqref{n1.5}). We will show now that if $f_2(t,x,u,p)=X(t,x)U(u)H(p)$, then $q_0$ is independent of $q_1$ and $\alpha$. Following the proof of Theorem \ref{a} we arrive to (\ref{n1.3}) that appears in the form \begin{equation}\label{3.30} \begin{aligned} \widetilde{L}_1(\tilde{w}) &\geq e^{-t_1}[-K_2|U(u(t_1,y_1))||H(h'(x_1-y_1))|(x_1-y_1)\\ &\quad +\gamma_1 X(t_1,x_1)H(h'(x_1-y_1))(u(t_1,x_1))-u(t_1,y_1))]. \end{aligned} \end{equation} Due to \eqref{3.2} we have that $\gamma_1 X(t,x)H(p)>0$ and consequently \begin{equation}\label{3.300} \gamma_1 X(t,x)H(p)=\gamma_1 |X(t,x)||H(p)|. \end{equation} Thus from (\ref{3.30}), (\ref{3.300}) we obtain that (recall that $U(u(t,x))$ is a strictly decreasing function and $|u|\leq M$) \begin{equation}\label{3.31} \begin{aligned} \widetilde{L}_1(\tilde{w}) &\geq e^{-t_1}|H(h'(x_1-y_1))|[-K_2|U(-M)|(x_1-y_1)\\ &\quad + \gamma_1|X(t_1,x_1)|(u(t_1,x_1)-u(t_1,y_1))]. \end{aligned} \end{equation} Due to the fact that $u(t_1,x_1))-u(t_1,y_1)\geq h'(\xi)(x_1-y_1)$, from (\ref{3.31}) it follows (recall that $h'(\xi)\geq q_0$) \begin{equation}\label{3.4} \begin{aligned} \widetilde{L}_1(\tilde{w}) &\geq e^{-t_1}|H(h'(x_1-y_1))|(x_1-y_1)[-K_2|U(-M)|+\gamma_1 |X(t_1,x_1)|h'(\xi)]\\ &\geq e^{-t_1}|H(h'(x_1-y_1))|(x_1-y_1)[-K_2|U(-M)|+\gamma_1 |X(t_1,x_1)|q_0]. \end{aligned} \end{equation} To obtain the contradiction with $\widetilde{L}_1(\tilde{w})\big|_{(t_1,x_1,y_1)}<0$ we have to show that \begin{equation}\label{3.5} -K_2|U(-M)|+\gamma_1 |X(t_1,x_1)|q_0\geq 0. \end{equation} Obviously this is the case when \begin{equation}\label{3.6} q_0\geq \frac{K_2|U(-M)|}{\gamma_1\gamma_2}. \end{equation} From this contradiction we conclude that $\tilde{w}$ cannot attain its positive maximum in $\bar{P}\setminus \Gamma$. Similarly one can prove that $\tilde{w}_1=e^{-t}(u(t,y)-u(t,x)-h(x-y))$ cannot attain its positive maximum in $\bar{P}\setminus \Gamma$. Further without any changes we follow the proof of Theorem \ref{a}. Note here that in (\ref{2.61}) one has to suppose that $q_0\geq \max\{K,\frac{K_2|U(-M)|}{\gamma_1\gamma_2}\}$. Theorem \ref{d} is proved. \end{proof} \begin{remark} \label{rmk5} \rm In the case of problems \eqref{1.1}, \eqref{1.3}, \eqref{1.5} and \eqref{1.1}, \eqref{1.4}, \eqref{1.5} one can easily formulate results that are similar to those of Corollaries \ref{b} and \ref{c}. The proofs of these results is an easy compilation of the proofs of Theorems \ref{a}, \ref{d}, Corollary \ref{b} and of Theorems \ref{a}, \ref{d}, Corollary \ref{c} respectively. \end{remark} \begin{remark} \label{rmk6} \rm From the mentioned above it follows that the strict monotonicity of $f_2(t,x,u,p)=X(t,x)U(u)H(p)$ in $u$, coupled with the Lipschitz continuity in $x$, guarantee the nonexistence of the gradient blow-up of a bounded solution in the interior of $Q_T$ for problems \eqref{1.1}, \eqref{1.2}, \eqref{1.5} and \eqref{1.1}, \eqref{1.3}, \eqref{1.5}. If additionally $f_2(t,x,0,p)=0$ then the strict monotonicity of $f_2(t,x,u,p)$ in $u$, coupled with the Lipschitz continuity in $x$, guarantee the nonexistence of the gradient blow-up of a bounded solution in $\overline{Q}_T$ for problem \eqref{1.1}, \eqref{1.4}, \eqref{1.5}. Concerning problem \eqref{1.1}, \eqref{1.4}, \eqref{1.5} in the case where $f_2(t,x,0,p)\not= 0$, one has to impose condition \eqref{i.7} supplementary to \eqref{3.1}, \eqref{3.2} in order to obtain the nonexistence of the gradient blow-up of a bounded solution in $\overline{Q}_T$. \end{remark} \begin{remark} \label{rmk7} \rm Let us give some simple examples of functions that satisfy \eqref{3.1}, \eqref{3.2} and at the same time do not satisfy \eqref{i.5}, \eqref{i.6}. Easy calculations show that, for example, the functions $$ f_2(t,x,u,p)=g(x)up^{2m}, \quad f_2(t,x,u,p)=g(x)u^3|p|^{\nu}, \quad f_2(t,x,u,p)=g(x)u e^p, $$ where $g(x)<0$ is an arbitrary Lipschitz continuous function, $m>1$ is an integer number, $\nu>2$ is a real number, satisfy \eqref{3.1}, \eqref{3.2} and do not satisfy \eqref{i.5}, \eqref{i.6}. Moreover, in the case where $f_2$ has one of the above representations, problems \eqref{sup.1}, \eqref{1.2}, \eqref{1.5}, \eqref{sup.1}, \eqref{1.3}, \eqref{1.5} and \eqref{sup.1}, \eqref{1.4}, \eqref{1.5} have a global classical solution for any Lipschitz continuous initial data. \end{remark} \begin{thebibliography}{00} \bibitem{ANFI} S. Angenent, M. Fila, \emph{Interior gradient blow-up in a semilinear parabolic equation}, Diff. and Integral Equations, 9 (1996), N.5, 865-877. \bibitem{ASISH} K. Asai, N. Ishimura, \emph{On the interior derivative blow-up for the curvature evolution of capillary surfaces}, Proc. Amer. Math. Soc., 126 (1998), N3, 835 - 840. \bibitem{souplet} J. P. Bartier, Ph. Souplet, \emph{Gradient bounds for solutions of semilinear parabolic equations without Bernstein's quadratic condition}, Comp. Rend. Math., 338(7) (2004), 533 - 538. \bibitem{BERN} S. N. Bernstein, \emph{Sur les equations du calcul des variations}, Ann. Sci. Ecole Norm. Sup., 29 (1912), 431 - 485. \bibitem{SNB} S. N. Bernstein, \emph{On the equations of calculus of variation}, Sobr. soch. M.: Izd-vo AN SSSR (1960) T.3, p. 191-242 [Russian]. \bibitem{ACONST} A. Constantin, J. Escher, \emph{Global solutions for quasilinear parabolic problems, J. Evol. Equ.}, 2 (2002), n. 1, 97 - 111. \bibitem{CLI} M. G. Crandall, H. Ishii and P.L.Lions, \emph{User's guide to viscosity solutions of second order partial differential equations}, Bull. Amer. Math. Soc. 27(1992), 1-67. \bibitem{KK} B. Kawohl, N. Kutev, \emph{Comparison principle and Lipschitz regularity for viscosity solutions of some classes of nonlinear partial differential equations}, Funk. Ekvac., 43 (2000), 241 - 253. \bibitem{DL} T. Dlotko, \emph{Examples of parabolic problems with blowing-up derivatives}, J. Math. Anl. Appl. 154 (1991) 226 - 237. \bibitem{FIL} M. Fila, G. Lieberman, \emph{Derivative blow-up and beyond for quasilinear parabolic equations}, Diff. Integ. Equations, 7 (1994), No 3/4, 811-822. \bibitem{AFF} A. F. Filippov, \emph{On condition for the existence of a solution of a quasilinear parabolic equation}, Dokl. Akad. Nauk SSSR 141 (1961), 568-570(Russian). English trasl. Soviet Math. Dokl. 2(1961), 1517-1519. \bibitem{Y} Y. Giga, \emph{Interior blow-up for quasilinear parabolic equations}, Discrete and Continuous Dynamical Systems 1 (1995), 449-461. \bibitem{SNK} S. N. Kruzhkov, \emph{Quasilinear parabolic equations and systems with two independent variables}, Trudy Sem. Petrovsk. 5(1979), 217 - 272 [Russian. English translation in ``Topics in Modern Math.", Consultant Bureau, New York, 1985]. \bibitem{KRYJ} S.N. Kruzhkov, \emph{Nonlinear parabolic equations in two independent variables}, Transact. Moscow Math. Soc. 16 (1968), 355 - 373. \bibitem{KRYL} N. V. Krylov, \emph{Nonlinear elliptic and parabolic equations of second order}, Mathematics and its Applications (Soviet Series), 7. D. Reidel Publishing Co., Dordrecht, 1987, 462. \bibitem{KUT} N. Kutev, \emph{Global solvability and boundary gradient blow up for one dimensional parabolic equations}, Progress in PDE: Elliptic and Parabolic problems (C.Bandle, et al., eds.), Longman, 1992, 176 - 181. \bibitem{LSV} O. A. Ladyzhenskaja, V. A. Solonnikov, N. N. Uralceva, \emph{Linear and Quasilinear Equations of Parabolic Type}, Amer. Math. Soc. Transl. (2), vol. 23, 1968. \bibitem{LIB} G. M. Lieberman, \emph{Second Order Parabolic Equations}, World Scientific Publishing Co., Inc., River Edge, N J, 1996. \bibitem{Nag} M. Nagumo, \emph{Uber die Differentialgleichung $y''=f(t,y,y')$}, Proc. Phys. Math. Soc. Japan 19 (1937) 861-866. \bibitem{SER} J. Serrin, \emph{The problem of Dirichlet for quasilinear elliptic differential equations with many independent variables}, Philos. Trans. Roy. Soc. London Ser. A 264, 413-496 (1969). \bibitem{SOUPL} Ph. Souplet, \emph{Gradient blow-up for multidimensional nonlinear parabolic equation with general boundary conditions}, Dif. Int. Equat. 15 (2002), N2, 237 - 256. \bibitem{TT} Al. S. Tersenov, Ar. S. Tersenov, \emph{On the Bernstein-Nagumo's condition in the theory of nonlinear parabolic equations}, J. Reine Angew. Math. 572 (2004), 197 - 217. \bibitem{VZ} M.P. Vishnevskii, T. I. Zelenyak, M. M. Lavrentiev, \emph{Behavior of solutions to parabolic equations for large time value}, Sib. Mat. J. 36 (1995), N.3, 435 - 453. \end{thebibliography} \end{document}