\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2015 (2015), No. 288, pp. 1--24.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2015 Texas State University.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2015/288\hfil H\"older continuity] {H\"older continuity of bounded weak solutions to generalized parabolic $p$-Laplacian equations II: singular case} \author[S. Hwang, G. M. Lieberman \hfil EJDE-2015/288\hfilneg] {Sukjung Hwang, Gary M. Lieberman} \address{Sukjung Hwang \newline Center for Mathematical Analysis and Computation, Yonsei University, Seoul 03722, Korea} \email{sukjung\_hwang@yonsei.ac.kr} \address{Gary M. Lieberman Department of Mathematics, Iowa State University, Ames, IA 50011, USA} \email{lieb@iastate.edu} \thanks{Submitted July 17, 2015. Published November 19, 2015.} \subjclass[2010]{35B45, 35K67} \keywords{Quasilinear parabolic equation; singular equation; \hfill\break\indent generalized structure; a priori estimate; H\"older continuity} \begin{abstract} Here we generalize quasilinear parabolic $p$-Laplacian type equations to obtain the prototype equation \[ u_t - \operatorname{div} \Big(\frac{g(|Du|)}{|Du|} Du\Big) = 0, \] where $g$ is a nonnegative, increasing, and continuous function trapped in between two power functions $|Du|^{g_0 -1}$ and $|Du|^{g_1 -1}$ with $10$. For the most part, we are only concerned here with the case that $g_1\le2$, but some of our results do not need this additional restriction, so we shall always state it explicitly when it is used. Our results generalized those of Ladyzhenskaya and Ural'tseva \cite{LaUr62} and Chen and DiBenedetto \cite{ChDB88,ChDB92}, who proved H\"older continuity under the structure conditions \begin{equation} \label{EAp} \mathcal{A}(x,t,u,\xi)\cdot\xi \ge C_0|\xi|^p,\ |\mathcal{A}(x,u,\xi)| \le C_1|\xi|^{p-1} \end{equation} with $p=2$ and $p<2$, respectively. The structure \eqref{EAp} is contained in this model as the special case $g(s)=s^{p-1}$, in which case we may take $g_0=g_1=p$, but we consider a class of structure functions $g$ much wider than that of just power functions. In this way, we obtain a uniform proof of H\"older continuity (with appropriate uniformity of constants) for all $p\in(1,2]$ at once under the structure condition \eqref{EAp} as well as a proof of H\"older continuity under more general structure conditions. In \cite{HL1}, we have discussed our approach for a generalization of the case $p\ge2$, so we concern ourselves here with the points relevant to the generalization of the case $p\le2$. It is known that solutions of this problem generally become zero in finite time when $p<2$ (see \cite[Sections VII.2 and VII.3]{DB93} for a more complete discussion of this phenomenon) but not when $p=2$ (because of the Harnack inequality, first proved by Moser \cite{Mos64}), so our proof needs to take this behavior into account. In addition, \cite[Section 4]{DBGiVe10} gives a H\"older exponent which degenerates as $p$ approaches $2$; the proof must be further modified for $p$ close to $2$ if the H\"older exponent is to remain positive near $p=2$. Our method manages the whole range $1 1$, \[ \beta^{g_0} G(\sigma) \leq G(\beta \sigma) \leq \beta^{g_1} G(\sigma). \] \item[(c)] For $0< \beta < 1$, \[ \beta^{g_1} G(\sigma) \leq G(\beta \sigma) \leq \beta^{g_0} G(\sigma). \] \item[(d)] $ \sigma_1 g(\sigma_2) \leq \sigma_1 g(\sigma_1) + \sigma_2 g(\sigma_2)$. \item[(e)] \emph{(Young's inequality)} For any $\epsilon \in (0,1)$, \[ \sigma_1 g(\sigma_2) \leq \epsilon^{1-g_1} g_1 G(\sigma_1) + \epsilon g_1 G(\sigma_2). \] \end{itemize} \end{lemma} \begin{lemma}\label{S1:H-ineq} For any $\sigma >0$, let \[ h(\sigma) = \frac{1}{\sigma} \int_{0}^{\sigma} g(s) \,ds, \quad H(\sigma) = \int_{0}^{\sigma} h(s) \,ds. \] Then we have \begin{gather*} g_0 h(\sigma) \leq g(\sigma) \leq g_1 h(\sigma), \\ g_0 H(\sigma) \leq G(\sigma) \leq g_1 H(\sigma), \\ (g_0 - 1) h(\sigma) \leq \sigma h'(\sigma) \leq (g_1 - 1) h(\sigma), \\ \frac{1}{g_1} \sigma h(\sigma) \leq H(\sigma) \leq \frac{1}{g_0} \sigma h(\sigma), \\ \beta^{g_0} H(\sigma) \leq H(\beta \sigma) \leq \beta^{g_1} H(\sigma) \end{gather*} for any $\beta>1$. \end{lemma} Our next result concerns some inequalities about integration of a function over various intervals. We shall use these inequalities in the proof of the Main Lemma. This lemma is probably well-known, but we are unaware of any reference for it. \begin{lemma} \label{S1:fintegral} Let $f$ be a continuous, decreasing, positive function defined on $(0,\infty)$. Then, for all $\delta$ and $\sigma\in(0,1)$, we have \begin{equation} \label{S1:fintegral:E2} \int_0^1 f(\delta+s)\,ds \le \frac 1\sigma\int_0^{\sigma}f(\delta+s)\,ds. \end{equation} If, in addition, for all $\beta>1$ and $\sigma>0$, we have \begin{equation} \label{S1:fintegral:E} \beta f(\beta\sigma) \ge f(\sigma), \quad f(\beta\sigma) \le f(\sigma), \end{equation} then, for all $\delta\in (0,1)$, we have \begin{equation} \label{S1:fintegral:E1} \int_0^\delta f(\delta+s)\, ds \le \frac 2{2+\ln (1/\delta)} \ int_0^1 f(\delta+s)\, ds. \end{equation} \end{lemma} \begin{proof} To prove \eqref{S1:fintegral:E2}, we define the function \[ F(\sigma) = \sigma \int_0^1 f(\delta+s)\, ds - \int_0^{\sigma} f(\delta+s)\, ds. \] Since \[ F'(\sigma) = \int_0^1f(\delta+s)\, ds - f(\delta+\sigma), \] and $f$ is decreasing, it follows that $F'$ is increasing so $F$ is convex. Moreover \[ F(0)=F(1)=0, \] so $F(\sigma) \le0$ for all $\sigma\in(0,1)$. Simple algebra then yields \eqref{S1:fintegral:E2}. To prove \eqref{S1:fintegral:E1}, we first use a change of variables to see that, for any $j\ge1$, we have \[ \int_{j\delta}^{2j\delta} f(\delta+s)\, ds =\int_0^{j\delta} f((j+1)\delta +\sigma)\, d\sigma = j\int_0^{\delta} f((j+1)\delta+js))\, ds. \] Since $(j+1)\delta+js\le (j+1)(\delta+s)$ and $f$ is decreasing, we have \[ \int_{j\delta}^{2j\delta} f(\delta+s)\, ds \ge j \int_0^\delta f((j+1)(\delta+s))\,ds \] and then \eqref{S1:fintegral:E} gives \[ \int_{j\delta}^{2j\delta} f(\delta+s)\, ds \ge \frac j{j+1}\int_0^{\delta} f(\delta+s)\, ds. \] We now let $J$ be the unique positive integer such that $2^{-J}<\delta \le 2^{1-J}$ and we take $j=2^i$ with $i=0,\dots,J-1$. Since $j/(j+1)\ge 1/2$, it follows that \[ \int_0^\delta f(\delta+s)\, ds \le 2\int_{2^i\delta}^{2^{i+1}\delta} f(\delta+s)\,ds. \] Since \[ \int_0^{2^J\delta}f(\delta+s)\, ds = \int_0^\delta f(\delta+s)\, ds + \sum_{i=0}^{J-1} \int_{2^i\delta}^{2^{i+1}\delta} f(\delta+s)\,ds, \] we infer that \[ \int_0^{2^J\delta} f(\delta+s)\, ds \ge [1+\frac 12J]\int_0^\delta f(\delta+s)\, ds. \] The proof is completed by noting that $J>\ln(1/\delta)$ and that \[ \int_0^1 f(\delta+s)\, ds \ge\int_0^{2^J\delta} f(\delta+s)\, ds. \] \end{proof} Note that condition \eqref{S1:fintegral:E} is satisfied if $f(\sigma)=\sigma^{-p}$ with $0\le p \le 1$, in which case this lemma can be proved by computing the integrals directly. \section{Basic results and the proof of H\"older continuity} \label{S2} In this section, we prove the H\"older continuity of solutions of \eqref{I:gen} for singular equations (that is, equations with $g_1\le 2$) and for degenerate equations (that is, equations with $g_0\ge 2$). Our proof is based on some estimates for nonnegative supersolutions of the equation, and these estimates will be proved in the next section. Our Main Lemma states that a nonnegative supersolution $u$ of a singular equation is strictly positive in a subcylinder if $u$ is near to the maximum value in more than a half of cylinder. \begin{lemma}[Main Lemma] \label{S2:MainLemma} Let $\omega$ and $R$ be positive constants. Then there are positive constants $\delta$ and $\mu$, both less than one and determined only by the data such that, if $u$ is a nonnegative solution of \eqref{I:gen} in \[ Q =Q_{\delta\omega,2R}\big(\frac 34\big) \] with $g_1\le2$, and \begin{equation}\label{S2:MainLemma-hyp} \big| Q\cap \{u \le \frac{\omega}{2}\}\big| \le \frac{1}{2}|Q|, \end{equation} then \begin{equation} \label{S2:MainLemma:estimate} \operatorname{ess\,inf}_{\mathcal Q} u \geq \mu \omega, \end{equation} with $\mathcal Q= Q_{\mu\omega,R/2}$. \end{lemma} We shall prove this lemma in the next section. Here we show first how to infer a decay estimate for the oscillation of any bounded solution of \eqref{I:gen}. \begin{lemma} \label{S2:Lmainboth} Let $C_0$, $C_1$, $g_0$, $g_1$, $\rho$, and $\omega$ be positive constants with $C_0\le C_1$ and $10$. As noted in \cite{HL1}, $B'\Omega$ need not be the same as $B\Omega$. For $(x_0,t_0)\in B'\Omega$ and $\omega>0$, we write $\operatorname{dist}_B(x_0,t_0)$ for the supremum of the set of all numbers $r$ such that $Q^{+,x_0,t_0}_{\omega,r}\subset \Omega$ and $K_r^{x_0,t_0}\times\{0\}\subset \partial_P\Omega$. \begin{theorem} \label{S2:thminitial} Let $u$ be a bounded weak solution of \eqref{I:gen} in $\Omega$, and suppose $1< g_0\le g_1\le2$. Suppose also that the restriction of $u$ to $B'\Omega$ is continuous at some $(x_0,t_0)\in B'\Omega$. Then $u$ is locally continuous up to $(x_0,t_0)$. Specifically, if there is a continuous increasing function $\tilde\omega$ defined on $[0,\operatorname{dist}_B(x_0,t_0))$ with $\tilde\omega(0)=0$, \begin{equation} \label{S2:Lthmininitial-o} \frac 56 \tilde\omega(2r)\le \tilde\omega(r) \end{equation} for all $r\in(0,\operatorname{dist}_B(x_0,t_0)/2)$, and with \[ |u(x_0,t_0)-u(x_1,t_0)| \le \tilde\omega(|x_0-x_1|) \] for all $x_1$ with $|x_0-x_1| < \operatorname{dist}_B(x_,t_0)$, then there exist constants $\gamma$ and $\alpha\in(0,1)$ depending only upon the data such that, for any $(x,t)\in\Omega$ with $t\ge t_0$, we have \begin{equation} \label{S2:thminitial:E1} \begin{split} \left|u(x_0,t_0) - u(x,t)\right| &\leq \gamma U \Big(\frac{|x_0 - x|+ |t_0 - t|_{G}}{\operatorname{dist}_B(x_0,t_0)}\Big)^{\alpha} \\ &\quad{}+3\tilde\omega\left(\gamma|x_0-x|_\infty +\gamma \operatorname{dist}_B(x_0,t_0)^{1-\alpha}|t_0-t|_G^\alpha\right). \end{split} \end{equation} \end{theorem} \begin{proof} We start by taking $\omega_0=U$ and $\rho_0=\operatorname{dist}_B(x_0,t_0)$. If $(x,t)\notin Q^{+,x_0,t_0}_{\omega_0,\rho_0}$, then the result is immediate for any $\alpha$ as long as $\gamma \ge1$. With $\lambda$ as in Lemma~\ref{S2:Lmainin}, we set $\sigma=5/6$, and \[ \varepsilon=\min\{\lambda, \frac 12\sigma^{(2-g_0)/g_0}\}. \] If $(x,t)\in Q^{+,x_0,t_0}_{\omega_0,\rho_0}$, then we define $\rho_n=\lambda^n\rho_0$. We also define $\omega'_n$ for $n>0$ inductively as $\omega'_{n+1}=\max\{\frac 56\omega'_n, 3\omega^*(\rho_n)\}$, and we set \[ Q_n= Q^{+,x_0,t_0}_{\omega'_n,\rho_n}. \] It follows from Lemma~\ref{S2:Lmainin} that $\operatorname{ess\,osc}_{Q_n} u \le \omega'_n$, but this estimate must be improved. To this end, we set \[ \omega_n= \max\big\{ \big(\frac56\big)^n\omega_0,3\tilde\omega(\rho_{n-1})\big\}, \] and infer from the proof of \cite[Theorem 2.6]{HL1} that $\omega'_n\le\omega_n$ for $n>0$. Hence \[ \operatorname{ess\,osc}_{Q_n} u \le \omega_n. \] As before, we assume that $x\neq x_0$ and $t\neq t_0$, so there are nonnegative integers $n$ and $m$ such that \[ \rho_{n+1} \le |x_0-x|_\infty < \rho_n, \] and \[ \omega_{m+1}^2G\Big( \frac {\omega_{m+1}}{\rho_{m+1}}\Big)^{-1} \le |t_0-t|< \omega_m^2G\big( \frac {\omega_m}{\rho_{m}}\big)^{-1} . \] With $\alpha_1= \log_{1/2}(5/6)$, it follows that \[ \big(\frac56\big)^n \le \Big(\frac {2|x_0-x|_\infty}{\rho_0}\Big)^{\alpha_1}, \quad \tilde\omega(\rho_n) \le \tilde\omega(\frac 1\lambda|x_0-x|_\infty). \] Moreover, if we set $\beta= \varepsilon \sigma^{(2-g_0)/g_0}$ and $\widehat{\omega}_m= \beta^m\omega_0$, it follows that $\widehat{\omega}_{m+1} \le \omega'_{m+1}$, so (as in the proof of Theorem 2.4) \[ |t_0-t|_G \ge \beta^{m+1}\rho_0. \] For $\alpha_2= \log_\beta\sigma$, we infer again that \[ \widehat{\omega}_{m} \le \Big( \frac {|t_0-t|_G}{\rho_0}\Big)^{\alpha_2}\omega_0. \] In addition, for $\alpha_3=\log_\beta\lambda$ (which is in the interval $(0,1]$), we infer that \[ \rho_{m} \le \Big( \frac {|t_0-t|_G}{\beta\rho_0}\Big)^{\alpha_3}\rho_0. \] Therefore, \[ \bar\omega(\rho_m) \le \bar\omega\Big( \rho_0^{1-\alpha_3}|t_0-t|_G^{\alpha_3}\Big). \] And the proof is complete by combining all these inequalities and taking $\alpha=\min\{\alpha_1,\alpha_2,\alpha_3\}$. \end{proof} As in \cite[Theorem 2.5]{HL1}, condition \eqref{S2:Lthmininitial-o} involves no loss of generality in that any modulus of continuity for the restriction of $u$ to $B'\Omega$ is controlled by one satisfying this condition. \section{Proof of the main lemma}\label{S3} Throughout this section, $u$ is a bounded nonnegative weak solution of \eqref{I:gen} with \eqref{I:gen-str}. The proof of Lemma~\ref{S2:MainLemma} is composed of four steps under the assumption that $u$ is large at least half of a cylinder $Q_{\omega, 2R}$. First, Proposition~\ref{S3:prop1} gives spatial cube at some fixed time level on which $u$ is away from its minimum (zero value) on arbitrary fraction of the spatial cube. From the spatial cube, positive information spread in both later time and over the space variables with time limitations (Proposition~\ref{S3:prop2} and Proposition~\ref{S3:prop3a}). Controlling the positive quantity $\theta >0$ in $T_{k,\rho}(\theta)$ is key to overcoming those time restrictions. Once we have a subcylinder centered at $(0, 0)$ in $Q_{\omega, 4R}$ with arbitrary fraction of the subcylinder, we finally apply modified De Giorgi iteration (Proposition~\ref{S3:prop4}) to obtain strictly positive infimum of $u$ in a smaller cylinder around $(0,0)$. \subsection{Basic results} Our first proposition shows that if a nonnegative function is large on part of a cylinder, then it is large on part of a fixed cylinder. Except for some minor variation in notation, our result is \cite[Lemma 7.1, Chapter III]{DB93}; we refer the reader to \cite[Proposition 3.1]{HL1} for a proof using the present notation. \begin{proposition} \label{S3:prop1} Let $k$, $\rho$, and $T$ be positive constants. If $u$ is a measurable nonnegative function defined on $Q=K_\rho\times(-T,0)$ and if there is a constant $\nu_1\in [0,1)$ such that \[ |Q\cap \{u\le k\}| \le (1-\nu_1)|Q|, \] then there is a number \[ \tau_1\in \Big(- T, -\, \frac {\nu_1}{2-\nu_1}T\Big) \] for which \[ \left| \{ x\in K_\rho: u(x,\tau_1)\le k\}\right| \le \big(1-\frac {\nu_1}2\big) |K_\rho|. \] \end{proposition} Our next proposition is similar to \cite[Lemma IV.10.2]{DB93}. \begin{proposition}\label{S3:prop2} Let $\nu$, $k$, $\rho $, and $\theta$ be given positive constants with $\nu<1$. If $g_1\le 1$, then, for any $\epsilon \in (0,1)$, there exists a constant $\delta =\delta (\nu,\epsilon, \theta, \text{data})$ such that, if $u$ is a nonnegative supersolution of \eqref{I:gen} in $K_\rho\times(-\tau,0)$ with \begin{equation}\label{S3:prop2-hyp} \left|\{x\in K_{\rho}: u(x,-\tau) < k \}\right|< (1-\nu) \left|K_{\rho}\right| \end{equation} for some \begin{equation} \tau \leq \theta(\delta k)^2 G\Big(\frac{\delta k}{\rho}\Big)^{-1}, \label{S3:prop2-tauge} \end{equation} then \[ \left|\{ x\in K_{\rho}: u(x, -t) < \delta k \}\right| < \left(1- (1-\epsilon)\nu\right)|K_{\rho}| \] for any $-t \in (-\tau, 0]$. \end{proposition} \begin{proof} The proof is almost identical to that of \cite[Proposition 3.2]{HL1}. With $\Psi$ defined as \[ \Psi= \ln^+ \Big( \frac {k}{(1+\delta)k-(u-k)_-}\Big), \] we note that $\delta k |\Psi'| \leq 1$. It follows that \begin{align*} |\Psi'|^2 G\big(\frac{|D\zeta|}{\Psi'}\big) &\leq \left( \delta k |\Psi'|\right)^{2-g_1} \left( \delta k \right)^{-2} G\left( \delta k |D\zeta| \right) \\ &\leq \sigma^{-g_1} \left(\delta k\right)^{-2} G\left( \frac{\delta k}{\rho} \right). \end{align*} Arguing as in the proof of \cite[Proposition 3.2]{HL1} (and noting that $2^{g_1} \ge1$) then yields \begin{align*} & \int_{-\tau}^{-s}\int_{K_{\rho}} h(\Psi^2)|\Psi||\Psi'|^{2} G\big(\frac{|D\zeta|}{|\Psi'|}\big) \,dx\,dt \\ &\leq 2^{g_1}\theta h\left(j^2 (\ln 2)^2 \right) ( j \ln 2 ) \sigma^{-g_1} |K_{\rho}| \\ &\leq 2^{g_1}\theta \frac{H(j^2 (\ln 2)^2)}{j \ln 2} \sigma^{-g_1} |K_{\rho}| \end{align*} for any $s\in (0,\tau)$. This inequality is the same as \cite[(3.4)]{HL1}. Since the remainder of the proof of \cite[Proposition 3.2]{HL1} is valid for the full range $1\varepsilon\} \text { is convex for all }\varepsilon\in(0,1), \\ \zeta_2(-T)=0, \\ \zeta_2=1 \text { on } (-T_1,0), \\ 0 \le \zeta_2' \le \big(\frac 2k\big)^2G\big(\frac k \rho\big) \text { on } (-T,0). \label{S3:zetat} \end{gather} \end{subequations} Let us note that it's easy to arrange that $\zeta_2'\ge0$ and that \[ \frac 1{\zeta_2'}\ge \big(\frac k2\big)^2G\big(\frac k{2\rho}\big)^{-1} -\big(\frac k2\big)^2G\big(\frac {k}{\rho}\big)^{-1}. \] Since Lemma~\ref{S1:ineq}(b) implies that \[ G\big(\frac {k}{\rho}\big) \ge 2^{g_0}G\big(\frac {k}{2\rho}\big) \ge 2G\big(\frac {k}{2\rho}\big), \] we infer the second inequality of \eqref{S3:zetat}. Also, we introduce the notation $D^-$ to denote the derivative \[ D^-f(t)= \limsup_{h\to0^+} \frac {f(t)-f(t-h)}h. \] With these preliminaries, we can now state our integral inequality. Our proof of this inequality is essentially the same as that for \cite[Lemma IV.6.1]{DB93}; the new ingredient is a more careful estimate of the integral involving $\zeta_t$ (which we denote by $I_4$). In this way, we obtain an estimate which does not depend on $p-2$ being bounded away from zero, which was the case in \cite[(6.9) Chapter IV]{DB93}. \begin{lemma} \label{S3:integralinequality} If $g_1\le2$ and if $u$ is a weak supersolution of \eqref{I:gen} in $K_{2\rho}\times (-T,0)$ satisfying \eqref{E:51large}, then there are positive constants $\gamma$ and $\gamma_0$, determined only by $\nu$, $\nu_0$, and the data such that \begin{equation} \label{S3:Eintegralinequality} D^-\Big( \int_{K_{2\rho}} \Phi_\kappa(u(x,t))\zeta^q(x,t)\, dx\Big) + \gamma_0\int_{K_{2\rho}} \Psi_\kappa^{g_0}(u(x,t))\zeta^q(x,t)\, dx \le \gamma |K_{2\rho}| \end{equation} for all $t\in (-T,0)$, where \begin{equation} \label{S3:integralinequalityq} q= g_0/(g_0-1). \end{equation} \end{lemma} \begin{proof} With \[ u^*=\frac {(1+\delta)\kappa-(u-\kappa)_-}{2\rho}, \] we use the test function \[ \frac {\zeta^q((1+\delta)\kappa-(u-\kappa)_-)}{G(u^*)} \] in the weak form of the differential inequality satisfied by $u$ to infer that, for all sufficiently small positive $h$, we have \[ I_1 +I_2\le I_3+I_4 \] with \begin{gather*} I_1 =\int_{K_{2\rho}} \Phi_\kappa(u(x,t))\zeta^q(x,t)\,dx -\int_{K_{2\rho}} \Phi_\kappa(u(x,t-h))\zeta^q(x,t-h)\,dx, \\ \begin{aligned} I_2 &=\int_{t-h}^h\int_{K_{2\rho}} \zeta^q(x,\tau) D(u-\kappa)_-(x,\tau) A \frac 1{G(u^*(x,\tau))}\\ &\quad\times \Big[1- \frac{u^*(x,\tau) g(u^*(x,\tau))}{G(u^*(x,\tau))}\Big] \, dx\,d\tau, \end{aligned}\\ I_3 =q\int_{t-h}^t \int_{K_{2\rho}} D\zeta(x,\tau) A \zeta^{q-1}(x,\tau) \frac {(1+\delta)\kappa-(u-\kappa)_-} {G\left(u^*(x,\tau)\right)} \,dx\, d\tau, \\ I_4=q\int_{t-h}^t \int_{K_{2\rho}} \Phi_\kappa(u(x,\tau)) \zeta^{q-1}(x,\tau)\zeta_t(x,\tau)\,dx\,d\tau, \end{gather*} and $A$ evaluated at $(x,\tau,u(x,\tau),Du(x,\tau))$ in $I_2$ and $I_3$. We now use \eqref {I:gen-str1} and the first inequality in \eqref{I:DeltaNabla} to see that \[ I_2 \ge C_0(g_0-1) \int_{t-h}^t \int_{K_{2\rho}} \zeta^q(x,\tau) \frac {G(|D(u-\kappa)_-(x,\tau)|)}{G(u^*(x,\tau))}\,dx\,d\tau. \] Also, \eqref {I:gen-str2} and Lemma~\ref{S1:ineq}(e) (with $\sigma_1= (qC_1/C_0)|D\zeta(x,\tau)|\rho u^*(x,\tau)$, $\sigma_2= |D(u-\kappa)_-(x,\tau)|$, and $\epsilon = \zeta(x,\tau)(g_0-1)/(2g_1)$) imply that \[ qD\zeta(x,\tau) \cdot A (x,\tau,u,Du) \zeta^{q-1}(x,\tau) \frac {(1+\delta)\kappa-(u-\kappa)_-}{G\left(u^*(x,\tau)\right)} \le J_1+J_2 \] with \begin{gather*} J_1 = g_1^{g_1}(\frac 2{g_0-1})^{g_1-1}\zeta^{q-g_1} \frac {G(q(C_1/C_0)|D\zeta|\rho u^*)}{G(u^*)},\\ J_2 = \frac 12 C_0(g_0-1)\zeta^q \frac {G(|D(u-\kappa )_-|)}{G(u^*)}. \end{gather*} From our conditions on $\zeta$ and because $q\ge2\ge g_1$, we conclude that there is a constant $\gamma_1$, determined only by data, such that $J_1 \le \gamma_1$, so \[ I_3 \le \gamma _1h|K_{2\rho}|+ \frac 12I_2. \] Next, we estimate $\Phi_\kappa$. Since $\kappa\le k/2$ and $\delta\in (0,1)$, it follows that, for all $s\in (0,(u-\kappa)_-)$, we have $(1+\delta)\kappa-s \le 2k$ and hence \[ G\Big( \frac {(1+\delta)\kappa-s}{2\rho}\Big) \ge \Big(\frac {(1+\delta)\kappa-s}{2k}\Big)^2 G\big(\frac k \rho\big). \] It follows that \[ \Phi_\kappa(u) \le 4k^2G\big(\frac k \rho\big)^{-1} \int_0^{(u-\kappa)_-} [(1+\delta)\kappa-s]^{-1}\, ds =4k^2G\big(\frac k \rho\big)^{-1} \Psi_\kappa(u), \] and therefore \[ I_4\le 16q\int_{t-h}^t\int_{K_{2\rho}} \Psi_\kappa(u(x,\tau))\zeta^{q-1}(x,\tau)\, dx\, d\tau. \] Combining all these inequalities and setting \begin{gather*} I_{21} = \int_{t-h}^t \int_{K_{2\rho}} \zeta^2(x,\tau) \frac {G(|D(u-\kappa)_-(x,\tau)|)}{G(u^*(x,\tau))}\,dx\,d\tau, \\ I_{41} = \int_{t-h}^t\int_{K_{2\rho}}\Psi_\kappa(u(x,\tau)) \zeta^{q-1}(x,\tau)\, dx\, d\tau \end{gather*} yields \begin{equation} \label{S3:Lintegralinequality1} I_1+\frac 12C_0(g_0-1) I_{21} \le \gamma_1 h|K_{2\rho}| + 16qI_{41}. \end{equation} Our next step is to compare \[ I_{22} = \int_{t-h}^t\int_{K_{2\rho}}\zeta^q(x,\tau) \Psi_\kappa^{g_0}(u(x,\tau))\,dx\,d\tau \] with $I_{21}$. To this end, we first use Lemma~\ref{Lpoin} with $\varphi=\zeta_1^q$, $v=(u-\kappa)_-$, and $p=g_0$ to conclude that there is a constant $\gamma_2$ determined only by the data and $\nu_0$ such that, for almost all $\tau \in (t-h,t)$, we have \begin{equation} \label{S3:I21} \int_{K_{2\rho}} \zeta^q(x,\tau)\Psi_\kappa^{g_0}(u(x,\tau))\, dx \le \gamma_2\rho^{g_0}\int_{K_{2\rho}}\zeta^q(x,\tau) |D\Psi_\kappa(u(x,\tau))|^{g_0}\, dx. \end{equation} (Of course, we have multiplied \eqref{Lpoin:E} by $\zeta_2^q(\tau)$ here.) Now we use the explicit expression for $\Psi_\kappa$ to infer that \[ \rho|D\Psi_\kappa(u(x,\tau))| = \frac {|D(u-\kappa)_-(x,\tau)|}{2u^*(x,\tau)} \le\frac {|D(u-\kappa)_-(x,\tau)|}{u^*(x,\tau)}. \] Whenever $|D(u-\kappa)_-(x,\tau)| \le u^*(x,\tau)$, we conclude that \[ \rho^{g_0}|D\Psi_\kappa(u(x,\tau))|^{g_0}\le 1 \] and, wherever $|D(u-\kappa)_-(x,\tau)| > u^*(x,\tau)$, we infer from Lemma~\ref{S1:ineq} that \[ \rho^{g_0}|D\Psi_\kappa(u(x,\tau))|^{g_0} \le \frac {G(|D(u-\kappa)_-(x,\tau)|)}{G(u^*(x,\tau))}. \] It follows that, for any $(x,\tau)$, we have \[ \rho^{g_0}|D\Psi_\kappa(u(x,\tau))|^{g_0} \le 1+\frac {G(|D(u-\kappa)_-(x,\tau)|)}{G(u^*(x,\tau))}. \] Inserting this inequality into \eqref{S3:I21} and integrating the resultant inequality with respect to $\tau$ yields \[ I_{22} \le \gamma_2(I_{21}+h|K_{2\rho}|). \] By invoking \eqref{S3:Lintegralinequality1}, we conclude that \[ I_1+ \frac {1}{2\gamma_2}C_0(g_0-1)I_{22} \le \Big(\gamma_1+ \frac {1}{2\gamma_2}C_0(g_0-1)\Big)h|K_{2\rho}| +16qI_{41}. \] We now note that \[ \Psi_\kappa(u)\zeta^{q-1}= \left(\Psi_\kappa^{g_0}(u) \zeta^q\right)^{1/g_0}, \] so Young's inequality shows that \[ \Psi_\kappa(u)\zeta^{q-1} \le \varepsilon \Psi_\kappa^{g_0}(u) \zeta^q + \varepsilon^{-q} \] for any $\varepsilon\in(0,1)$. By choosing $\varepsilon$ sufficiently small, we see that there are constants $\gamma_0$ and $\gamma$ such that \[ I_1+\gamma_0 I_2 \le \gamma h|K_{2\rho}|. \] To complete the proof, we divide this inequality by $h$ and take the limit superior as $h\to0^+$. \end{proof} Our next step is to estimate the integral of $\zeta^q$ over suitable $N$-dimensional sets with $q$ defined by \eqref{S3:integralinequalityq}. Specifically, for each positive integer $n$ and a number $\delta\in(0,1)$ to be further specified, we define the set \[ K_{\rho,n}(t) = \{x\in K_{2\rho}: u(x,t)<\delta^nk\} \] and we introduce the quantities \[ A_n(t) = \frac 1{|K_{2\rho}|} \int_{K_{\rho,n}(t)} \zeta^q(x,t)\, dx, \quad Y_n = \sup_{-T\nu$, then \begin{equation} \label{S3:L7.6equation} A_{n+1}(t) \le \sigma Y_n. \end{equation} \end{lemma} \begin{proof} In this case, we define \[ t_*=\sup\Big\{\tau\in(-T,t): D^- \Big( \int_{K_{2\rho}} \zeta^q(x,\tau)\Phi_{\delta^nk} (u(x,\tau))\, dx \Big) \ge0 \Big\} \] (and note that this set is nonempty). From the definition of $t_*$, we have that \begin{equation} \label{S3:L7.6:Phiint} \int_{K_{2\rho}} \zeta^q(x,t) \Phi_{\delta^nk} u(x,t)\, dx \le \int_{K_{2\rho}} \zeta^q(x,t_{*}) \Phi_{\delta^nk} u(x,t_{*})\, dx. \end{equation} It follows from Lemma~\ref{S3:integralinequality} and the definition of $t_*$ that \[ \int_{K_{2\rho}} \zeta^q(x,t_{*}) \Psi^{g_0}_{\delta^nk} u(x,t_{*})\, dx\le C|K_{2\rho}|, \] with $C= \gamma/\gamma_0$. Now we set \[ K_*(s)=\{x\in K_{2\rho}: (u-\delta^nk )_-(x,t_*) >s\delta^nk\} \] for $s\in (0,1)$, and \[ I_1(s)= \int_{K_*(s)} \zeta^q(x,t_*)\, dx. \] As in the proof of Lemma~\ref{S3:L7.5}, we have that \[ \Phi_{s\delta^nk}(u(x,t_*)) \ge \ln\frac {1+\delta}{1+\delta-s}, \] so \begin{equation} \label{S3:L7.6:1} I_1(s)\le C\Big(\ln\frac {1+\delta}{1+\delta-s}\Big)^{-g_0}|K_{2\rho}|. \end{equation} Moreover, if $x\in K_*(s)$, then \[ u(x,t_*) <(1-s)\delta^nk \le \delta^nk, \] and hence $K_*(s)\subset K_{\rho,n}$, so \begin{equation} \label{S3:L7.6:2} I_1(s)\le Y_n|K_{2\rho}|. \end{equation} We now define \[ s_*= \Big[1-\exp\Big(- \big( \frac {2C}{\nu}\big)^{1/g_0}\Big)\Big](1+\delta_*), \] with $\delta_*\in(0,1)$ chosen so that $s_*<1$. Since $Y_n>\nu$, a simple calculation shows that \begin{equation} \label{S3:L7.6:CYn} C\Big(\ln \frac {1+\delta}{1+\delta -s}\Big)^{-g_0}\le \frac 12Y_n \end{equation} for $s>s_*$ provided $\delta\le \delta_*$. Next, we set \[ I_2 =\int_{K_{2\rho}} \zeta^q(x,t_*)\Phi_{\delta^nk}(u(x,t_*))\, dx \] and use Fubini's theorem to conclude that \begin{align*} I_2 &=\int_{K_{2\rho}} \zeta^q(x,t_*) \Big( \int_0^{\delta^nk} \frac { \chi_{\{(\delta^nk-u)+>s \}} ((1+\delta)\delta^nk-s) }{G\big( \frac {(1+\delta)\delta^nk-s}{2\rho}\big)}\, ds \Big)\, dx \\ &= \int_0^{\delta^nk} \frac {(1+\delta)\delta^nk-s} {G\big( \frac {(1+\delta)\delta^nk-s}{2\rho}\big)} \Big( \int_{K_{2\rho}} \zeta^q(x,t_*) \chi_{\{(\delta^nk-u)+>s \}}\, dx\Big)\, ds \end{align*} Using the change of variables $\tau= s/(\delta^{n}k)$, we see that \begin{align*} I_2 &=\int_0^1 \frac {(1+\delta)-\tau} {G\big( \frac {\delta^nk(1+\delta-\tau)}{2\rho}\big)} \Big( \int_{K_{2\rho}} \zeta^q(x,t_*) \chi_{\{(\delta^nk-u)+>\delta^nk\tau \}}\, dx\Big)\, d\tau \\ &= \int_0^1 \frac {(1+\delta)-\tau} {G\big( \frac {\delta^nk(1+\delta-\tau)}{2\rho}\big)}I_1(\tau)\, d\tau. \end{align*} Combining this equation with \eqref{S3:L7.6:1}, \eqref{S3:L7.6:2}, and \eqref{S3:L7.6:CYn} then yields \[ I_2\le Y_n|K_{2\rho}|\Big[\int_0^{s_*} \frac {(1+\delta)-\tau} {G\big( \frac {\delta^nk(1+\delta-\tau)}{2\rho}\big)}\, d\tau + \frac 12\int_{s_*}^1 \frac {(1+\delta)-\tau} {G\big( \frac {\delta^nk(1+\delta-\tau)}{2\rho}\big)}\, d\tau\Big]. \] We now define the function \[ f(\tau)= \frac {\tau}{G\big( \frac {\delta^nk\tau}{2\rho}\big)} \] and we set $\sigma_*= 1-s_*$. Using the change of variables $s=1-\tau$ then yields \[ I_2 \le Y_n|K_{2\rho}|\mathcal K, \] with \[ \mathcal K= \int_{\sigma_*}^1 f(\delta+s)\, ds + \frac 12\int_0^{\sigma_*}f(\delta+s)\, ds. \] Since \[ \mathcal K = \int_0^1f(\delta+s)\, ds -\frac 12 \int_0^{\sigma_*} f(\delta+s)\, ds, \] it follows from \eqref{S1:fintegral:E1} that \[ \mathcal K \le \big(1-\frac {\sigma_*}2\big) \int_0^1f(\delta+s)\, ds, \] and therefore \begin{equation} \label{S3:L7.6:Phi1} I_2 \le Y_n|K_{2\rho}|\left(1-\frac {\sigma_*}2\right) \int_0^1f(\delta+s)\, ds. \end{equation} Our next step is to obtain a lower bound for $I_2$. Taking into account \eqref{S3:L7.6:Phiint}, we have \[ I_2 \ge \int_{K_{\rho,n+1}(t)} \zeta^q(x,t)\Phi_{\delta^nk}(u(x,t))\, dx. \] Next, for $z<\delta^{n+1}k$, we have \begin{align*} \Phi_{\delta^nk}(z) &= \int_0^{(z-\delta^nk)_-}\frac {(1+\delta)\delta^nk -s} {G\big( \frac {(1+\delta)\delta^nk-s}{2\rho}\big)}\, ds \\ &\ge \int_0^{\delta^nk(1-\delta)}\frac {(1+\delta)\delta^nk -s} {G\big( \frac {(1+\delta)\delta^nk-s}{2\rho}\big)}\, ds \\ &= \int_0^{1-\delta} f(\delta+s)\, ds \\ &\ge \Big(1- \frac 2{2+\ln\delta}\Big) \int_0^1 f(\delta+s)\,ds \end{align*} by \eqref{S1:fintegral:E2}, so \[ \Phi_{\delta^nk}(u(x,t)) \ge \Big(1- \frac 2{2+\ln(1/\delta)}\Big) \int_0^\delta f(\delta+s)\, ds \] for all $x\in K_{\rho,n+1}(t)$ and hence \[ I_2 \ge \Big(1- \frac 2{2+\ln(1/\delta)}\Big) \Big(\int_{K_{\rho,n+1}(t)} \zeta^q(x,t)\, dx\Big) \Big(\int_0^\delta f(\delta+s)\, ds\Big). \] In conjunction with \eqref{S3:L7.6:Phi1}, this inequality implies that \[ A_{n+1}(t) \le \frac {1-(\sigma_*/2)}{1- ( 2/(2+\ln(1/\delta))}Y_n. \] By taking $\delta_2$ sufficiently small, we can make sure that \[ \sigma=\frac {1-(\sigma_*/2)}{1- ( 2/(2+\ln(1/\delta_2))} \] is in the interval $(0,1)$. If we take $\delta_1=\min\{\delta_*,\delta_2\}$, we then infer \eqref{S3:L7.6equation} for $\delta\le\delta_1$. \end{proof} As shown in \cite{DB93}, if $t_*$ and $t$ are equal in this proof, we can infer \eqref{S3:L7.5:Yn1} very simply. We are now ready to prove Proposition~\ref{S3:prop3a}. \begin{proof}[Proof of Proposition \ref{S3:prop3a}] Since $Y_{n+1} \le Y_n$, it follows from Lemmata~\ref{S3:L7.5} and \ref{S3:L7.6} that, for all positive integers $n$, we have \[ A_{n+1}(t) \le \max \{\nu,\sigma Y_n\} \] for all $t\in (-T,0)$ and hence $Y_{n+1}\le \max\{\nu,\sigma Y_n\}$. Induction implies that \[ Y_n \le \max\{\nu,\sigma^{n-1} Y_1\} \] for all $n$. In addition $Y_1\le1$, so there is a positive integer $n_0$, determined by $\nu$, $a_0$, and the data such that $Y_{n_0}\le\nu$. Next, we recall that $\zeta=1$ on $K_\rho\times(-T_1,0)$, and hence, for all $t\in(-T_1,0)$, we have \begin{align*} \big| \{ x\in K_\rho:u(x,t)\le \delta^{n_0}k\}\big| &= \int_{\{x\in K_\rho: u(x,t)\le \delta^{n_0}k\} }\zeta^q(x,t)\, dx \\ &\le \int_{\{x\in K_{2\rho}: u(x,t)\le \delta^{n_0}k\}} \zeta^q(x,t)\, dx \le Y_{n_0}. \end{align*} The proof is complete by using the inequality $Y_{n_0} \le \nu$ and taking $\delta^*=\delta^{n_0}$. \end{proof} \subsection{Proof of main lemma} \begin{proof} With $\delta$ to be chosen, we use Proposition~ \ref{S3:prop1} with $k=\omega/2$, $\rho=2R$, $\nu_1=\frac 12$, and \[ T= 3\big(\frac {\delta \omega}2\big)^2G\big( \frac {\delta \omega}{2R}\big)^{-1} \] to infer that there is a $\tau_1\in (-T, -T/3)$ such that \[ \left|\{ x\in K_{2R}: u(x,\tau_1)\le \frac \omega2\}\right| \le \frac 34|K_{2R}|. \] Next, we set $\nu=\frac 14$, $\rho=2R$, $k=\omega/2$, $\tau=\tau_1$, and $\theta=3$. Since $\tau_1\le T$ and \[ T = 3(\delta k)^2 G\big( \frac {2\delta k}\rho\big)^{-1} \le 3(\delta k)^2G\big( \frac {\delta k}{\rho}\big)^{-1}, \] it follows that \eqref{S3:prop2-hyp} is satisfied, so Proposition~ \ref{S3:prop2} implies that \begin{equation} \label{S2:MainLemmaE1} \big| \{x\in K_{2R}: u(x,t)\le \frac {\delta \omega}2\}\big| \le \frac 78|K_{2R}| \end{equation} for all $t\in (\tau_1,0)$ provided we take $\delta$ to be the constant from that proposition. (In particular, $\delta$ is determined only by the data.) Since $\tau_1\ge T/3$, it follows that \[ \tau_1 \ge \big(\frac {\delta \omega}2\big)^2 G\big( \frac {\delta \omega}{2R}\big)^{-1}, \] and hence \eqref{S2:MainLemmaE1} holds for all \[ t\in\Big( -\big(\frac {\delta \omega}2\big)^2 G\big( \frac {\delta \omega}{2R}\big)^{-1},0\Big). \] Now we use Proposition~\ref{S3:prop3a}, with $\omega=\delta \omega$ and $\nu$ to be chosen, to infer that there is a constant $\delta^*\in(0,1)$, determined only by the data and $\nu$ such that \begin{equation} \label{S2:MainLemmaE2} \big|\{ x\in K_R: u(x,t) \le \frac {\delta^*\delta \omega}2\} \big| \le \nu |K_{2R}| \end{equation} for all \[ t\in \Big( - \big( \frac {\delta \omega}2\big)^2 G\big( \frac {\delta \omega}R\big)^{-1}, 0\Big). \] Since $G\big( \frac {\delta \omega/2}R\big) \le G\big(\frac {\delta \omega}R\big)$ and $\delta^*\le1$, it follows that \[ \big(\frac {\delta \omega}2\big)^2 G\big(\frac {\delta \omega}R\big)^{-1} \ge \big(\frac {\delta \omega}2\big)^2 G\big(\frac {\delta \omega/2}R\big)^{-1} \ge \Big( \frac {\delta^*\delta \omega/2}R\Big)^2 G\Big( \frac {\delta^*\delta \omega/2}R\Big)^{-1}. \] We now take $\nu_0$ to be the constant corresponding to $\theta=1$ in Proposition~\ref{S3:prop4}, and we set $\nu=2^{-N}\nu_0$, which determines $\delta^*$. Then \eqref{S3:prop4:hyp} is satisfied for $k=\delta^*\delta \omega/2$ and $\rho=R/2$. Proposition~\ref{S3:prop4} then yields \eqref{S2:MainLemma:estimate} with $\mu =\delta^*\delta/4$. \end{proof} \section{Auxiliary theorems}\label{S4} We now present the basic results used in the previous sections of the paper. Since the results are either well-known or were proved in \cite{HL1}, we just state the results here. \subsection{Local energy estimate} The local energy estimate is a fundamental inequality playing an important role in the proofs of several results, especially Proposition~\ref{S3:prop1}, Proposition~\ref{S3:prop2}, and Proposition~\ref{S3:prop3a}. We refer the reader to \cite[Proposition 4.1]{HL1} for a proof. \begin{proposition} Let $G$ satisfy structure conditions \eqref{I:gen-str} in a cylinder $Q_{\rho}:=K_{\rho}\times [t_0,t_1]$, and let $\zeta$ be a cutoff function on the cylinder $Q_\rho$, vanishing on the parabolic boundary of $Q_\rho$ with $0\le\zeta\le1$. Define constants $r$, $s$, and $q$ by \begin{equation}\label{S4:LocalE-rsq} r=1- \frac{1}{g_1}, \quad s=\frac{g_0}{g_1}, \quad \text{and} \quad q= 2g_1. \end{equation} If $u$ is a locally bounded weak supersolution of \eqref{I:gen}, then there exist constants $c_0$, $c_1$, and $c_2$ depending on data such that \begin{equation}\label{S4:LocalE} \begin{split} &\int_{K_{\rho}\times \{t_1\}} G^{r-1} \Big(\frac{\zeta (u-k)_{-}}{\rho}\Big) (u-k)_{-}^{s+2} \zeta^{q} \,dx \\ & + c_0\iint_{Q_{\rho}} G\left(|D(u-k)_{-}|\right) G^{r-1} \Big(\frac{\zeta (u-k)_{-}}{\rho}\Big) (u-k)_{-}^{s} \zeta^{q} \,dx\,dt \\ &\leq c_1\iint_{Q_{\rho}} G^{r-1} \Big(\frac{\zeta (u-k)_{-}}{\rho}\Big) (u-k)_{-}^{s+2} \zeta^{q-1}| \zeta_{t}| \,dx\,dt \\ &\quad + c_2 \iint_{Q_{\rho}} G\left(|D\zeta| \zeta(u-k)_{-}\right) G^{r-1} \Big(\frac{\zeta (u-k)_{-}}{\rho}\Big) (u-k)_{-}^{s} \,dx\,dt \end{split} \end{equation} for any constant $k$. \end{proposition} We refer the reader to \cite[Proposition 4.1]{HL1} for the corresponding result about nonpositive subsolutions. \subsection{Logarithmic energy estimate} With the functions $h$ and $H$ defined in Lemma~\ref{S1:H-ineq}, the logarithmic energy estimate was proved as \cite[Proposition 4.2]{HL1}, which also contains the corresponding result for nonpositive weak subsolutions. \begin{proposition} Assume that $G$ satisfies \eqref{I:gen-str} in a cylinder $K_{R}\times [t_0, t_1]$. Let $q\ge g_1$ and $\delta\in (0,1)$ be constants, and let $\zeta$ be a cut-off function which is independent of the time variable. Let $u$ be a nonnegative weak supersolution of \eqref{I:gen}and let $k$ be a positive constant. Then \begin{equation}\label{S4:LogE}\begin{split} &\int_{K_{R}\times \{t_1\}} H(\Psi^2) \zeta^{q} \,dx + C_0 (4 g_0 - 2) \int_{t_0}^{t_1}\int_{K_{R}} G(|Du|) h(\Psi^2) (\Psi')^{2} \zeta^{q}\,dx\,dt \\ &\leq \int_{K_{R}\times \{t_0\}} H(\Psi^2) \zeta^{q} \,dx +C^* \int_{t_0}^{t_1}\int_{K_{R}} h(\Psi^2) \Psi (\Psi')^2 G\Big(\frac{|D\zeta|}{|\Psi'|}\Big) \zeta^{q-g_1}\,dx\,dt \end{split}\end{equation} where \[ C^*=\frac{C_0}{g_1} \left(\frac{2 q g_1 C_1}{C_0}\right)^{g_1},\quad \Psi(u) = \ln^{+} \left[\frac{k}{(1+ \delta) k - (u-k)_{-}}\right]. \] \end{proposition} \subsection{A Poincar{\'e} type inequality} For our proofs, we shall need the following result which is \cite[Proposition I.2.1]{DB93}. \begin{lemma} \label{Lpoin} Let $\Omega$ be a bounded convex subset of $\mathbb R^N$ and let $\varphi$ be a nonnegative continuous function on $\overline {\Omega}$ such that $\varphi\le1$ in $\Omega$ and such that the sets $\{x\in\Omega: \varphi(x)>k\}$ are convex for all $k\in(0,1)$. Then, for any $p\ge1$, there is a constant $C$ determined only by $N$ and $p$ such that \begin{equation} \label{Lpoin:E} \begin{aligned} &\Big( \int_\Omega \varphi|v|^p\, dx\Big)^{1/p}\\ &\le C\frac {(\operatorname{diam}\Omega)^N}{|\{x\in \Omega:v(x)=0,\ \varphi(x)=1\}|^{(N-1)/N}} \Big( \int_\Omega \varphi|Dv|^p\,dx \Big)^{1/p} \end{aligned} \end{equation} for all $v\in W^{1,p}$. \end{lemma} Note that if the set $\{x\in \Omega:v(x)=0,\ \varphi(x)=1\}$ has measure zero, then \eqref{Lpoin:E} is true because the right hand side is infinite. \subsection{Embedding theorem} Our next result is a variation on the Sobolev imbedding theorem, which is just \cite[Theorem 4.4]{HL1}. \begin{theorem}\label{S4:Embedding-theorem} For a nonnegative function $v\in W_{0}^{1,1}(Q)$ where $Q=K\times[t_0, t_1]$, $K\subset \mathbb{R}^{N}$, we have \begin{equation}\label{EB00} \begin{split} \iint_{Q} v \,dx\,dt &\leq C(N) |{Q} \cap \{v>0\}|^{\frac{1}{N+1}} \\ &\times \Big[\operatorname{ess\,sup}_{t_0 \leq t \leq t_1} \int_{K} v \,dx \Big]^{\frac{1}{N+1}} \Big[\iint_{Q} |Dv| \,dx\,dt \Big]^{\frac{N}{N+1}}. \end{split} \end{equation} \end{theorem} \subsection{Iteration} Finally, we recall \cite[Lemma I.4.1]{DB93}. \begin{lemma}\label{S4:Iteration-lemma} Let $\{Y_n\}$, $n=0,1,2,\ldots$, be a sequence of positive numbers, satisfying the recursive inequalities \begin{equation} \label{S4:Iteration-lemma:E} Y_{n+1} \leq C b^{n} Y_{n}^{1+\alpha} \end{equation} where $C,b >1$ and $\alpha >0$ are given numbers. If \[ Y_0 \leq C^{-\frac{1}{\alpha}}b^{-\frac{1}{\alpha^2}}, \] then $\{Y_n\}$ converges to zero as $n\rightarrow \infty$. \end{lemma} \subsection*{Acknowledgments} This work is based on the first author's thesis at Iowa State University. The first author was partially supported by EPSRC grant EP/J017450/1 and NRF grant 2015 R1A5A 1009350. \begin{thebibliography}{00} \bibitem{Chen84} Chen, Y. Z.; \emph{H\"older estimates for solutions of uniformly degenerate quasilinear parabolic equations}, A Chinese summary appears in Chinese Ann. Math. Ser. A \textbf{5} (1984), no. 5, 663, Chinese Ann. Math. Ser. B, \textbf{5} (1984), no. 4, 661--678. \bibitem{ChDB88} Chen, Y. Z.; DiBenedetto, E.; \emph{On the local behavior of solutions of singular parabolic equations}, Arch. Rational Mech. Anal., \textbf{103} (1988), no. 4, 319--345. \bibitem{ChDB92} Chen, Y. Z.; DiBenedetto, E.; \emph{H\"older estimates of solutions of singular parabolic equations with measurable coefficients}, Arch. Rational Mech. Anal., \textbf{118} (1992), no. 3, 257--271. \bibitem{DG57} De Giorgi, E.; \emph{Sulla differenziabilit\`a e l'analiticit\`a delle estremali degli integrali multipli regolari} (in Italian), Mem. Accad. Sci. Torino. Cl. Sci. Fis. Mat. Nat. (3), \textbf{3} (1957), 25--43. \bibitem{DB86j} DiBenedetto, E.; \emph{A boundary modulus of continuity for a class of singular parabolic equations}, J. Differential Equations, \textbf{6} (1986), no. 3, 418--447. \bibitem{DB86} DiBenedetto, E.; \emph{On the local behaviour of solutions of degenerate parabolic equations with measurable coefficients}, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), \textbf{13} (1986), no. 3, 487--535. \bibitem{DB93} DiBenedetto, E.; \emph{Degenerate parabolic equations}, Universitext, Springer-Verlag, New York (1993). \bibitem{DBGiVe10} DiBenedetto, E.; Gianazza, U.; Vespri, V.; \emph{A new approach to the expansion of positivity set of non-negative solutions to certain singular parabolic partial differential equations}, Proc. Amer. Math. Soc., \textbf{138} (2010), no. 10, 3521--3529. \bibitem{DBGiVe12} DiBenedetto, E.; Gianazza, U.; Vespri, V.; \emph{Harnack's inequality for degenerate and singular parabolic equations}, Springer Monographs in Mathematics, Springer, New York (2012), \bibitem{GiSuVe10} Gianazza, U.; Surnachev, M.; Vespri, V.; \emph{A new proof of the H\"older continuity of solutions to $p$-Laplace type parabolic equations}, Adv. Calc. Var., \textbf{3} (2010), no. 3, 263--278. \bibitem{Hwang} Hwang, S.; \emph{H\"older regularity of solutions of generalised $p$-Laplacian type parabolic equations}, Ph. D. thesis, Iowa State University, (2012), Paper 12667. http: //lib.dr.iastate.edu/etd/12667. \bibitem{HL1} Hwang. S.; Lieberman, G. M.; \emph{H\"older continuity of a bounded weak solution of generalized parabolic $p$-Laplacian equation I: degenerate case}, Electron. J. Diff. Equ., \textbf{2015} (2015), no. 287, 1--32. \bibitem{KrRu61} Krasnosel'ski\u\i, M. A.; Ruticki\u\i, Ja. B.; \emph{Convex functions and Orlicz spaces}, Translated from the first Russian edition by Leo F. Boron, P. Noordhoff Ltd., Groningen (1961). \bibitem{LaUr61} Lady\v{z}enskaja, O. A.; Ural'ceva, N. N.; \emph{Quasilinear elliptic equations and variational problems in several independent variables} (in Russian), Uspehi Mat. Nauk, \textbf{16}, (1961), no. 1 (97), 19--90. Russian Math. Surveys, \textbf{16}, (1961), 17--92. \bibitem{LaUr62} Lady\v{z}enskaja, O. A.; Ural'ceva, N. N.; \emph{A boundary value problem for linear and quasilinear parabolic equations}, (in Russian), Izv. Akad. Nauk SSSR Ser. Mat., \textbf{26} (1962), 5--52. Amer. Math. Soc. Transl. (2), \textbf{47} (1965), 217--267. \bibitem{LaSoUr67} Lady\v{z}enskaja, O. A.; Solonnikov, V. A.; Ural'ceva, N. N.; \emph{Linear and quasilinear equations of parabolic type} (in Russian), Translated from the Russian by S. Smith. Translations of Mathematical Monographs, Vol. 23, American Mathematical Society, Providence, R.I., (1967). \bibitem{Lie91} Lieberman, G. M.; \emph{The natural generalization of the natural conditions of Ladyzhenskaya and Ural'tseva for elliptic equations}, Comm. Partial Differential Equations, \textbf{16} (1991), no. 2-3, 311--361. \bibitem{Lie96} Lieberman, G. M.; \emph{Second order parabolic differential equations}, publisher={World Scientific Publishing Co. Inc.}, River Edge, NJ (1996). \bibitem{Mos64} Moser, J.; \emph{A Harnack inequality for parabolic differential equations}, Comm. Pure Appl. Math., \textbf{17} (1964), 101--137. \bibitem{RaRe91} Rao, M. M.; Ren, Z. D.; \emph{Theory of Orlicz spaces}, Monographs and Textbooks in Pure and Applied Mathematics, \textbf{146}, Marcel Dekker Inc., New York, (1991). \bibitem{Urb08} Urbano, J. M.; \emph{The method of intrinsic scaling}, Lecture Notes in Mathematics, \textbf{1930}, A systematic approach to regularity for degenerate and singular PDEs, Springer-Verlag, Berlin (2008). \end{thebibliography} \end{document}