Brought to you by:
Paper

Stabilization of the solution of a doubly nonlinear parabolic equation

and

© 2013 Russian Academy of Sciences (DoM), London Mathematical Society, Turpion Ltd.
, , Citation È. R. Andriyanova and F. Kh. Mukminov 2013 Sb. Math. 204 1239 DOI 10.1070/SM2013v204n09ABEH004338

1064-5616/204/9/1239

Abstract

The method of Galerkin approximations is employed to prove the existence of a strong global (in time) solution of a doubly nonlinear parabolic equation in an unbounded domain. The second integral identity is established for Galerkin approximations, and passing to the limit in it an estimate for the decay rate of the norm of the solution from below is obtained. The estimates characterizing the decay rate of the solution as $x\rightarrow \infty$ obtained here are used to derive an upper bound for the decay rate of the solution with respect to time; the resulting estimate is pretty close to the lower one.

Bibliography: 17 titles.

Export citation and abstract BibTeX RIS

§ 1. Introduction

Let $\Omega$ be a domain in ${R}_{n}=\lbrace x=(x_1,x_2,\ldots,x_n)\rbrace$, $n \geqslant 2$. We consider the first mixed problem

Equation (1.1)

Equation (1.2)

for a doubly nonlinear parabolic equation in the cylindrical domain $D=\lbrace t>0\rbrace \times \Omega$.

Here,

The existence and uniqueness of the solution of this problem was examined by Lions [1], Raviart [2], Grange and Mignot [3], Bamberger [4], Alt and Luckhaus [5], Bernis [6] and others. For the most part, these problems were dealt with on a bounded domain $\Omega$ and a bounded time interval $[0,T]$ with arbitrary $T>0$. The strong solution of this problem in a bounded domain $\Omega$ was found by Raviart by replacing the evolution derivative with a difference relation. The existence of a weak solution in an unbounded domain $\Omega$ was proved by Bernis, who applied a limiting procedure to the solutions built by Grange and Mignot on bounded domains. However, working with a weak solution involves a difficulty consisting, for example, in examining the decay of the solution as $t\rightarrow \infty$. Bamberger proved the uniqueness of a strong positive solution of the problem.

In this paper, Galerkin approximations are employed in the conventional way to construct a strong solution of the problem across the whole time interval $[0,\infty)$ (the domain $\Omega$ may be unbounded). In [7], this type of solution was constructed on a bounded interval $[0,T]$ for any $T>0$. The ideas of [7] were used in [8] to construct a solution of an anisotropic parabolic equation for $\alpha >2$ on a bounded time interval.

Derivation of the estimates we need for the Galerkin approximations is facilitated by the fact that they are smooth functions; then passing to the limit we can extend these estimates to the solution of problem (1.1), (1.2). In particular, for the equation

Equation (1.3)

on a bounded domain $\Omega$, with $p>\alpha$, it will be shown that

Equation (1.4)

Conceptually, these results continue the studies started by Gushchin [9] on the behaviour of solutions of linear parabolic equations for large $t$. The left-hand side of estimate (1.4) with $\alpha =2$ was obtained by Alikakos and Rostamian in [10] for the Cauchy problem. Tedeev [11] obtained analogous estimates for the solution of a high-order quasi-linear equation. Sharp, two-sided estimates for the decay rate of the norm of the solution to a linear and a quasi-linear parabolic equation in an unbounded domain were established by Kozhevnikova [12] and Karimov and Kozhevnikova [13], and for an anisotropic parabolic equation with $\alpha =2$, by Kozhevnikova and Mukminov [14]. Nearly sharp estimates for the solution of an anisotropic parabolic equation with $\alpha >2$ were obtained in [8] (a solution on a bounded time interval).

§ 2. Statements of the main results

We define the space $\mathring{W}_{\alpha,p}^1(\Omega)$ to be the completion of $C_0^{\infty }(\Omega)$ in the norm

where

We also let $V(D^T)$, where $D^T=D_0^T$ and $D_a^b=(a,b)\times \Omega$, denote the completion of $C_0^{\infty }(D^T)$ in the norm

Definition. A generalized solution of problem (1.1), (1.2) is a function $u$ in the space $V(D^T)$ for all $T>0$ and satisfying the identity

Equation (2.1)

for all $\varphi \in C_0^{\infty }(D_{-1}^T)$.

Here and below, $(f,\varphi)_{Q}$ denotes the value of a distribution $f$ at an element $\varphi \in C_0^\infty (Q)$, where $Q$ is a domain in $ \mathbb {R}_{n}$ or in $\mathbb {R}\times \mathbb {R}_{n}$. However, in formula (2.1), it is assumed that $f\in (V(D^T))^{\prime }$.

We shall write a function $u(t,x)$ as $u$ or $u(t)$ when there is no ambiguity.

Throughout we assume that $u_0\in \mathring{W}_{\alpha,p}^1(\Omega)$.

Theorem 1. Let $f,f_t\in (V(D^T))^{\prime }$ for all $T>0$. Then there exists a generalized solution $u(t,x)$ of problem (1.1), (1.2) such that

Equation (2.2)

Equation (2.3)

Equation (2.4)

Equation (2.5)

for all $T>0$. Here, $\alpha ^{\prime }=\alpha /(\alpha -1)$.

Bamberger [4] defined a strong solution by the condition $(|u|^{\alpha -2}u)_t\in L_1(D^T)$ and proved, in particular, that if

where $\Omega$ is a bounded domain, then the problem has at most one strong solution. He also showed that this solution $u(t)$ vanishes for $t\geqslant t_0$, provided that $p<\alpha <q$, $1/q=1/p-1/n$.

It is easily checked that the solution in Theorem 1 is strong.

The paper [10] puts forward a series of assertions (not all of them are fully justified) about the behaviour of the solution of the Cauchy problem for the equation

Equation (2.6)

The following upper and lower bounds for the solution of the Cauchy problem were obtained under the assumption that $u_0\in L_1({R}_{n})\cap L_2({R}_{n})$ and $\nabla u_0\in L_p({R}_{n})$.

The upper estimates read as follows:

The lower estimates are as follows:

In [15] the following estimate for the solution of the Cauchy problem for equation (2.6) for any $q\in [q_0,\infty)$ is given, under the assumption that $u_0\in L_{q_0}({R}_{n})$:

The paper [11] looks at the problem

Equation (2.7)

Equation (2.8)

where

(In fact [11] deals with somewhat more general equations.) A solution $u$ in the space $ L_p(0,T; \mathring{W}_p^m(\Omega)) \cap C([0,T];L_2(\Omega))\cap L_q(D^T)$, $T>0$, is considered under the assumption that $u_0\in \mathring{W}_p^m(\Omega)\cap L_2(\Omega)\cap L_q(\Omega)$. In particular, the following bounds were proved for a bounded domain $\Omega$:

Above we have only given basic results on the behaviour of the solutions of equations (2.6), (2.7). For more detailed results the reader is referred to the book [16].

The remaining assertions of the present paper are established under the assumption that $p>\alpha$. In Theorems 25, a domain $\Omega$ may be bounded or unbounded. For an unbounded $\Omega$, it is assumed that the initial function has bounded support

Equation (2.9)

where $\Omega ^r=\lbrace x\in \Omega \mid |x|<r\rbrace$.

Theorem 2. Let $p>\alpha$. Assume that condition (2.9) is satisfied and that the initial function $u_0$ is bounded. Then there exist a positive number $C=C(u_0,\alpha,p)$ and a bounded solution $u(t,x)$ of problem (1.3), (1.2) such that, for all $t\geqslant 0$,

Equation (2.10)

Below we shall need the characteristic function of the layer between two concentric spheres

when $a=0$ or $b=\infty$ the limit indices will be omitted: $\chi ^{b}=\chi _{0}^{b}$, $\chi _{a}=\chi _{a}^{\infty }$.

The proof of the lower bound when $\Omega$ is unbounded depends heavily on the following fact.

Theorem 3. Let $p>\alpha$, let (2.9) be satisfied, and let $u(t,x)$ be a generalized solution of problem (1.3), (1.2) with bounded initial function $u_0$. Assume that $u(t,x)$ satisfies properties (2.2)–(2.5). Then $u(t,x)$ is also bounded and there exist positive numbers $A$ and $\kappa$ such that

Equation (2.11)

for all $t>0$, $r\geqslant 2R_0$.

To prove Theorem 3 we shall employ the maximum principle (see § 5).

For $r\sim t^\varepsilon$, estimate (2.11) becomes trivial. Hence, we need a different estimate in order to bound the decay rate of the solution from above. We shall assume that the origin lies on the boundary of $\Omega$. Also let $\gamma _r=\lbrace {x}\in \Omega \mid |x|=r\rbrace$, $r>0$.

Consider the following geometrical characteristic of the unbounded domain $\Omega$. We set

Equation (2.12)

Assume that the domain $\Omega$ satisfies the condition

Equation (2.13)

For example, this condition is satisfied for a domain lying inside some cone.

Theorem 4. Let $p>\alpha$ and let (2.9) and (2.13) be satisfied. Then there exist positive numbers $\kappa (p)$, $\mathscr{M}(p)$ and a generalized solution $u(t,{x})$ of problem (1.3), (1.2) such that

Equation (2.14)

for all $t\geqslant 0$, $r\geqslant 2R_0$.

Let $r(t)$ be an arbitrary positive function satisfying the inequality

Equation (2.15)

for all $t>0$. Clearly, such a function always exists.

Theorem 5. Let $p > \alpha$. Assume that (2.9) and (2.15) are satisfied. Then there exist a positive number $M$ and a solution $u(t,x)$ of problem (1.3), (1.2) in a domain $\Omega$ lying inside some cone (of aperture $<2\pi$) such that

Equation (2.16)

If the domain $\Omega$ satisfies the condition

Equation (2.17)

then we can set $r(t)=t^{\varepsilon _1}$, $t> 0$, and (2.16) assumes the form

Taking $\varepsilon _1$ sufficiently small, $\varepsilon$ can be made arbitrarily close to zero. The exponent in the last estimate is close to the exponent $-1/(p-\alpha)$ in the lower bound (2.10).

As an example, consider the domain of revolution

$x^{\prime }\!=\!(x_2,\ldots,x_n)$, with a positive function $g(x_1)\!<\!\infty$. The only requirement imposed on $g$ is that the set $\Omega (g)$ be a domain and that

For these domains, one can easily prove the relation

from which condition (2.17) follows.

§ 3. Proof of the existence theorem

Under the above assumptions on $f$ we have

In particular, $f(0)\in (\mathring{W}_{\alpha,p}^1(\Omega))^{\prime }$.

3.1. The case $\alpha \geqslant 2$.

Consider a sequence $\omega _k \in C_0^{\infty }(\Omega)$ of linearly independent functions whose linear hull is dense in $\mathring{W}_{\alpha,p}^1(\Omega)$. We set $I_m=\bigcup _{k=1}^m \operatorname{supp}\omega _k$. Galerkin approximations to the solution will be sought in the form

where the functions $c_{mk}(t)$ are determined from the equations

Equation (3.1)

The numbers $b_m>0$ will be chosen later. We claim that equations (3.1) can be solved for the derivatives $c_{mk}^{\prime }$. Clearly, they are as follows:

For each $t$, the coefficient matrix

is a Gram matrix of the system of linearly independent vectors $\omega _k$, $k=1,2,\dots,m$, and hence is invertible. From equations (3.1) with initial conditions $c_{mk}(0)$, chosen so that $u_m(0,x)\rightarrow u_0(x)$ in $\mathring{W}_{\alpha,p}^1(\Omega)$, we find the functions $c_{mk}(t)$. At first, these functions are determined on a short time interval. Then, the fact that the Galerkin approximations are bounded (to be proved later) lets us define them on an infinite time interval. The numbers $b_m$ will be chosen so as to have $\Vert u_m(0)\Vert _2^2/b_m\rightarrow 0$ as $m\rightarrow \infty$.

Now we proceed to estimate the Galerkin approximations. Multiplying equations (3.1) by $c_{mj}(t)$ and summing, this gives

Integrating with respect to $t$ shows that

Equation (3.2)

The last integral on the right is bounded by a constant independent of $m$, thanks to the above convergences. Further, for $t\in (0,T)$ we have

Equation (3.3)

Hence, from (3.2) and Gronwall's lemma we see that the sequence $u_m$ is bounded in the spaces $C([0,T];L_{\alpha }(\Omega))$ and $ V(D^T)$ for all $T>0$ by the constant $C(T,f,\Vert u_0\Vert _\alpha)$, which remains unchanged if the sequence $b_m$ is replaced by a larger one $b_m^{\prime}>b_m$.

Now we multiply equations (3.1) by $c_{mj}^{\prime }(t)$ and sum:

Integrating with respect to $t$ shows that

Equation (3.4)

Integrating the last term by parts, we obtain

Note that

Further, since $u_m$ is bounded in the space $V(D^T)$, we have

Now we find from (3.4) that, for any $T>0$, the sequence $|u_m|^{(\alpha -2)/2}u_m^{\prime }$ is bounded in $L_2(D^T)$ and that the sequence $\nabla u_m$ is bounded in $C([0,T]; L_p(\Omega))$. Using the diagonal process, we can now choose a subsequence $u_{m_k}$ which is weakly convergent in these spaces. For brevity of notation we shall omit the subscript $k$:

The convergence holds for any $T=1,2,\dots$, and the limit functions agree in their common domain of definition. Consequently, we have convergence for any $T>0$.

Below it will be shown that $\widetilde{u}=(|u|^{(\alpha -2)/2}u)^{\prime }$, $\chi =A(u)$ and that the function $u$ is a generalized solution of problem (1.1), (1.2). We will proceed in three steps, which we will also need later in the subsequent analysis.

Step 1. The sequence $u_m(t)$ is bounded in the space $\mathring{W}^1_{\alpha,p}(\Omega)$ on any finite interval:

We fix a countable dense subset $\lbrace t_s\rbrace$ of $[0,\infty]$. We can assume that $t_0=0$. Given any bounded domain $\Omega ^r\subset \Omega$ with smooth boundary, the embedding $\mathring{W}_1^1(\Omega ^r)\subset L_1(\Omega ^r)$ is compact. Hence, using the diagonal process we can choose a subsequence $u_{m_k}(t_s)\rightarrow h_s$ converging strongly in $L_1(\Omega ^r)$ for all natural $s$. Choosing, if necessary, a further subsequence (and dropping the subscripts), we assume that $u_m(t_s,x)\rightarrow h_s(x)$ almost everywhere in $\Omega ^r$ for any $t_s$. In particular, for $t_0=0$ we have $u_m(0,x)\rightarrow u_0(x)$ almost everywhere in $\Omega$.

For the next step we require the following result.

Lemma 1. Let the sequence $v_m(t)\!\in \! C([0,T];L_2(\Omega))$ satisfy the following properties:

  • 1)  
    $v_m(t_s,x)$ converges almost everywhere in $\Omega ^r$ for each $t_s$ and some $r>0$;
  • 2)  
    the sequence $v_m^{\prime }(t)$ is bounded in $L_2(D^T)$ for each $T>0$.

Then it contains a subsequence $v_{m_k}$ converging to the function $v$ in the space $C([0,T];L_1(\Omega ^r))$, and $v_{m_k}\rightarrow v$ almost everywhere in $(0,T)\times \Omega ^r$.

Proof. The sequence $v_m(t)$ is equicontinuous in $t$ in the space $L_2(\Omega)$:

Equation (3.5)

Next, the sequence $v_m(t)$ is bounded in the space $C([0,T];L_2(\Omega))$. Consequently, it contains a subsequence $v_{m_k}(t)$ converging weakly in $L_2(\Omega)$ for the same $t_s$ as above. Along with the almost everywhere convergence in $\Omega ^r$, this implies strong convergence in $L_1(\Omega ^r)$ for any $t_s$ (see [1], Ch. 1, § 12.2). We again write $v_m(t)$ for $v_{m_k}(t)$.

For a bounded domain $\Omega ^r$, it follows easily from (3.5) that $v_m(t)$ is a uniformly Cauchy sequence in the norm $L_1(\Omega ^r)$:

Choosing a finite grid $t_{s_k}$ of small spacing and then increasing $n$ and $m$, we ensure that the right-hand side is uniformly small in $t$.

Thus, we see that $v_{m_k}\rightarrow v$ in $C([0,T];L_1(\Omega ^r))$ for any $T=1,2,\dots$ . The convergence also takes place in $L_1((0,T)\times \Omega ^r)$. Hence, one may select a subsequence converging in $(0,T)\times \Omega ^r$ almost everywhere. The lemma is proved.

Step 2. We apply Lemma 1 to the sequence $v_m=|u_m|^{(\alpha -2)/2}u_m$. Since $r>0$ is arbitrary and $T=1,2,\dots$, we can apply the diagonal process to select a subsequence $v_{m_k}$ converging in $D$ almost everywhere. Hence, the sequence $u_{m_k}$ converges almost everywhere in $D$ to $u$ (see [1], Ch. 1, § 1.4, Lemma 1.3). We quote this result.

Lemma 2. Let $Q$ be a domain in ${R}_{n}$ or in ${R}_{n+1}$. Assume that a sequence $g_m$ converges to $g$ almost everywhere in $Q$ and is bounded in $L_q(Q)$. Then $g_m\rightarrow g$ weakly in $L_q(Q)$.

Clearly, this result, which is stated in [1] for bounded domains, also holds for unbounded domains.

From Lemma 2 it also follows that $v_{m_k}\rightarrow v=|u|^{(\alpha -2)/2}u$ weakly in $L_2(D^T)$ for any $T>0$. Given a fixed $T$, the convergence $v_{m_k}(T)\rightarrow v(T)$ in $L_1(\Omega ^r)$ implies that there is a subsequence convergent almost everywhere in $\Omega ^r$ for all $r>0$. Hence,

Equation (3.6)

The sequence $u_m(T)$ is bounded in the space $\mathring{W}_{\alpha,p}^1(\Omega)$, and hence we can choose a subsequence such that

Equation (3.7)

also, if $f=0$, then the constant $C$ depends only on $\Vert u_0\Vert _\alpha$. The constant $C$ is monotone in $T$, so this proves (2.2).

Further,

Passing to the limit, we get

It follows that $\widetilde{u}=v^{\prime }=(|u|^{(\alpha -2)/2}u)^{\prime }$.

We claim that the sequence $|u_m|^{\alpha -2}u_m^{\prime }$ is bounded in $L_{\alpha ^{\prime }}(D^T)$. Indeed,

Hence we can assume that $ |u_m|^{\alpha -2}u_m^{\prime }\rightarrow |u|^{\alpha -2}u^{\prime }$ weakly in $L_{\alpha ^{\prime }}(D^T)$. As a result, $|u|^{\alpha -2}u\in C([0,\infty);L_{\alpha ^{\prime }}(\Omega))$. We also have

Equation (3.8)

In fact, in Step 1 we noted that $u_m(0,x)\rightarrow u_0(x)$ converges almost everywhere in $\Omega$. Now, $v_m(0)\rightarrow v(0)=|u(0)|^{(\alpha -2)/2}u(0)$ in $L_1(\Omega ^r)$, $r>0$, by Lemma 1, thereby proving (3.8).

Step 3. We will now prove the equality $\chi =A(u)$. For this we need some integral relations. We multiply equation (3.1) by the smooth function $d_j(t)$, integrate over $t$, and let $m\rightarrow \infty$, denoting $d_j(t)\omega _j(x)$ by $\varphi$ in the resulting expression:

Equation (3.9)

We note that

as $u_m$ is bounded in $C([0,T];L_{\alpha }(\Omega))$ and $b_m\rightarrow \infty$. Clearly, any function from $V(D^T)$ can be approximated by linear combinations

Hence, (3.9) also holds for functions from the space $ V(D^T)$. So, in view of (3.8), $u$ will be a generalized solution of problem (1.1), (1.2) if we prove that $\chi =A(u)$.

Note that $v,v^{\prime }\in L_2(D^T)$ implies that $v\in C([0,T];L_2(\Omega))$ and that $\Vert u(t)\Vert _\alpha \in C([0,T])$, $\alpha >1$. Substituting the function $\varphi =u$ into (3.9), this gives

Equation (3.10)

Our next argument depends on the operator $A$ being monotone. It is readily verified that

Equation (3.11)

From equations (3.1) it is easily seen that

Hence,

Consequently, since $\liminf \Vert u_m(T)\Vert _{\alpha }\geqslant \Vert u(T)\Vert _{\alpha }$, by (3.6)

Equation (3.12)

Applying (3.10) shows that

Setting $h=u-\lambda \omega$, $\lambda >0$, $\omega \in V(D^T)$, we have

Letting $\lambda \rightarrow 0$, we see that $(\chi -A(u),\omega)\geqslant 0$ for any $\omega$. Hence, $\chi =A(u)$.

3.2. The case $\alpha <2$.

Galerkin approximations to the solution will be sought in the previous form, but now the functions $c_{mk}(t)$ are determined from the equations

Equation (3.13)

Here, the functions $v_m=u_m^2+\varepsilon _m$ are introduced for regularization; the numbers $\varepsilon _m>0$ will be chosen later. We claim that equations (3.13) can be solved with respect to the derivatives $c_{mk}^{\prime }$. Clearly, they are as follows

For each $t$, the coefficient matrix

is a Gram matrix of the system of linearly independent vectors $\omega _k$, $k=1,2,\dots$, and hence is invertible. From equations (3.13) with initial conditions $c_{mk}(0)$, chosen so that $u_m(0,x)\rightarrow u_0(x)$ in the space $\mathring{W}_{\alpha,p}^1$, we find the functions $c_{mk}(t)$.

Now we estimate the Galerkin approximations. Multiplying equations (3.13) by $c_{mj}(t)$ and summing, we find that

We have $u_mu_m^{\prime }=v_m^{\prime }/2$ and $\operatorname{supp} u_m \subset I_m$, and hence, integrating with respect to $t$ shows that

Equation (3.14)

Since the sets $I_m$ are bounded, we may choose $\varepsilon _m$ to satisfy

Equation (3.15)

Now the last integral on the right of (3.14) is bounded by a constant independent of $m$, thanks to the above convergences. As before, inequality (3.3) holds. Hence, using (3.14) and Gronwall's lemma it follows that the integrals $\displaystyle \int _{I_m}v_m(t)^{\alpha /2}\,dx$ are uniformly bounded in $t$ and $m$, and hence so is the sequence $u_m$ in the spaces $C([0,T];L_{\alpha }(\Omega))$ and $ V(D^T)$ for all $T>0$.

We now multiply equations (3.13) by $c_{mj}^{\prime }(t)$ and sum:

Integrating over $t$ gives

Equation (3.16)

As before, $ |(f,u_m^{\prime })_{D^T}|\leqslant C(T)$. In addition, $(\alpha -1)u_m^2+\varepsilon _m\geqslant (\alpha -1)v_m$. Setting

it follows from (3.16) that the sequence $v_m^{(\alpha -2)/4}u_m^{\prime }=(g_m(u_m))^{\prime }$ is bounded in $L_2(D^T)$ and the sequence $\nabla u_m$ is bounded in $C([0,T];L_p(\Omega))$. In view of the above results, we can choose a subsequence $u_{m_k}$ that is weakly convergent in the spaces specified below. For brevity, the subscript $k$ will be dropped:

Repeating Steps 1–3, as adapted to the new setting, we will be able to prove that $\widetilde{u}=|u|^{(\alpha -2)/2}u^{\prime }$.

Proceeding as in Step 1, we can assume (dropping the subscripts) that the sequence $u_m(t_s,x)$ converges almost everywhere in $\Omega ^r$ for each $t_s$.

At Step 2, Lemma 1 is applied to the sequence $v_m=g_m(u_m)$ to produce a subsequence converging almost everywhere in $(0,T)\times \Omega ^r$. Since $r>0$, $T>0$ are arbitrary, the diagonal process can be applied to extract a subsequence $g_m(u_{m_k})$ converging almost everywhere in $D$. The sequence of inverse functions $g_m^{-1}(v)$ converges pointwise. Now, Lemma 2 shows that the sequence $u_{m_k}$ converges almost everywhere to $u$ in $D$. Hence, we have proved that

Moreover, $\widetilde{u}=|u|^{(\alpha -2)/2}u^{\prime }$.

As in the case $\alpha \geqslant 2$, (3.6) and (3.7) are satisfied for a fixed $T$.

For $\alpha <2$, the sequence $u_m^{\prime }$ is bounded in $L_{\alpha }(D^T)$. Indeed, using (3.16),

Hence, we can assume that $u_m^{\prime }\rightarrow u^{\prime }$ weakly in $L_{\alpha }(D^T)$, $T>0$, and so $u\in C([0,\infty);L_{\alpha }(\Omega))$. This proves (2.4).

At Step 3, our aim is to show that $\chi =A(u)$. For this purpose we shall need some integral relations. We multiply equations (3.13) by a smooth function $d_j(t)$, integrate with respect to $t\in (0,T)$, and then integrate the first term by parts. Now, writing $\varphi$ for $d_j(t)\omega _j(x)$, we have

We note that

since $(v_m^{(\alpha -1)/2})^{\alpha ^{\prime }}=v_m^{\alpha /2}$ is a bounded sequence in $C([0,T];L_1(\Omega ^r))$. Hence, we can choose a subsequence to ensure that $v_m^{\alpha /2-1}u_m\rightarrow |u|^{\alpha -2}u$ weakly in $L_{\alpha ^{\prime },{\text{loc }}}(D^T)$ and $v_m^{\alpha /2-1}(T)u_m(T)\rightarrow |u|^{\alpha -2}u(T)$ weakly in $L_{\alpha ^{\prime },{\text{loc}}}(\Omega)$. That the limit functions have exactly this form is justified by the fact that, as we said above, the subsequence $u_m$ converges almost everywhere in $D^T$ and almost everywhere in $\Omega$ for $t=T$ (see (3.6)). Now making $m\rightarrow \infty$, in view of (3.8) it follows that

Equation (3.17)

As in the case $\alpha \geqslant 2$, we show that this equality is satisfied for all functions $\varphi \in C_0^\infty (D_{-1}^\infty)$. Relation (3.17) also means that $u$ is a generalized solution of problem (1.1), (1.2), provided that $\chi =A(u)$.

Substituting $\varphi =u$ into (3.17), we see that

Equation (3.18)

Our next argument depends on $A$ being monotone. Using equations (3.14), it is easily seen that

Equation (3.19)

Employing inequality (3.15), by (3.19) we have

Now (3.12) follows because $\liminf \Vert v_m^{\alpha /4}(T)\Vert _{2}\geqslant \Vert u(T)\Vert _{\alpha }$ and $\varepsilon _m\rightarrow 0$. Finally, a similar argument as in the case $\alpha \geqslant 2$ shows that $\chi =A(u)$.

The proof of Theorem 1 is complete.

§ 4. Proof of Theorem 2

Assume now that $\alpha < p$ and that a domain $\Omega$ is bounded. Our aim is to estimate the decay rate of the solution to problem (1.3), (1.2) from below as $t\rightarrow \infty$.

4.1. The case $\alpha \geqslant 2$.

We define

dropping the subscripts where possible. After differentiating with respect to $t$, for $f=0$ formula (3.2) assumes the form

Equation (4.1)

Differentiating (3.4), we obtain

Equation (4.2)

The following estimates hold:

Applying the Cauchy-Schwarz inequality for the inner product in $\mathbb {R}_2$, this gives

Hence,

Equation (4.3)

Using (4.1), we rewrite this as follows:

As a result, $\gamma {E^{\prime }}/{E}\geqslant {H^{\prime }}/{H}$, or, on integrating,

Consequently,

or

It follows that

Thus,

Equation (4.4)

In view of (3.7), for fixed $t>0$ and $\alpha \leqslant p$ if the domain is bounded one may choose a subsequence $u_{m_k}(t)$ which converges strongly in the space $L_\alpha (\Omega)$. The functions

lie in the linear hull of the functions $\omega _1,\omega _2,\dots,\omega _m$. In a finite-dimensional space all the norms are equivalent, and hence

We choose numbers $b_m$ so as to have $ \widetilde{c_m}\leqslant b_m/m$. Now

Passing to the limit in (4.4) as $m\rightarrow \infty$, we obtain

Equation (4.5)

4.2. The case $\alpha <2$.

In this case we set

Note that (3.15) implies that $E(t)\geqslant \varepsilon _m^{\alpha /4}$.

Differentiating (3.14) with respect to $t$ and rewriting it for $f=0$, this establishes

Equation (4.6)

From (3.16) it follows that

Equation (4.7)

For any $\mu >0$, clearly

In view of (3.15),

Now (4.3) follows by minimizing the right-hand side over $\mu$.

The rest of the proof is similar to the case $\alpha \geqslant 2$.

We claim that the estimate (4.5), in the case of a bounded domain $\Omega$, is sharp. To prove this, we need the following Friedrichs-Steklov-type inequality

We note in passing that the following Friedrichs-Steklov inequality holds for an unbounded domain located inside a cone:

Equation (4.8)

For $p>\alpha$, we have

Equation (4.9)

Differentiating (3.10) or (3.18) with respect to $T$ and employing (4.9) for $u$, we see that

Solving this differential inequality, we obtain the estimate

which proves that (4.5) is sharp.

4.3. An unbounded domain.

Our aim here is to derive (4.5) for the solution of problem (1.3), (1.2) in the case when $\Omega$ is unbounded.

We let $u^{l}$, $l=1,2,\dots$, denote the solutions in $[0,\infty)\times \Omega ^{l}$ with fixed initial function $u_0$. We can assume that these solutions are extended by zero outside $[0,\infty)\times \Omega ^{l}$. Using (3.7), we have the estimate

Using the properties of the solutions $u^{l}$ listed in Theorem 1, we can choose a subsequence $l_i$ such that $u^{l_i}(t)\rightarrow u(t)$ weakly in $V(D^T)$ as $i\rightarrow \infty$ for all $T>0$, and then, applying Steps 1–3, prove that the limit function $u(t)$ is a generalized solution of problem (1.3), (1.2).

Further, given a fixed $t>0$ we can assume that $u^{l_i}(t)\rightarrow u(t)$ weakly in $\mathring{W}_{\alpha,p}^1(\Omega)$ as $i\rightarrow \infty$. As the embedding $\mathring{W}_{\alpha,p}^1(\Omega ^{r})\subset L_\alpha (\Omega ^{r})$ is compact, $u^{l_i}(t)\rightarrow u(t)$ strongly in $L_\alpha (\Omega ^{r})$ as $i\rightarrow \infty$ for any $r>0$.

By (2.11), for any $\varepsilon$ there exists $r$ such that

Estimate (4.5) holds for $u^l$. Now

Using the strong convergence in $L^{\alpha }(\Omega ^{r})$, we pass to the limit as $l\rightarrow \infty$ and then as $r\rightarrow \infty \ (\varepsilon \rightarrow 0)$. Thus, we have now proved (2.10) for an unbounded domain $\Omega$ as well.

The proof of Theorem 2 is complete.

§ 5. Proof of Theorem 3

For the sake of completeness we recall the maximum principle to be used in the proof of Theorem 3. Note that the inequality $\alpha <p$ is not used in the proof of the maximum principle.

Proposition. Let $u_0(x)\leqslant B$, $x\in \Omega$, $0\leqslant B<\infty$. Then a generalized solution of problem (1.1), (1.2) with properties (2.2)–(2.5) is bounded:

Equation (5.1)

Proof. The truncation $u^{(B)}(t,x)=\max (u(t,x)-B,0)$ of a function $u\in V(D^T)$ is known to lie in the same space (see, for example, [17], Ch. 2, § 4, Lemmas 4.1–4.3).

Assume first that $\alpha \in (1,2)$. That the function $\varphi =u^{(B)}\xi$ can be substituted into (3.17) is justified by the passage to the limit; here $\xi (x)$ is a Lipschitz function with bounded support. We have

Equation (5.2)

We choose $\xi =\xi (|x|)$, where

In this case, $|\nabla \xi |\leqslant {1}/(R-r)$. Note that $u^{(B)}(0)=u_0^{(B)}\equiv 0$. We have $f\leqslant 0$, and hence

Further,

where $0<\theta (x)<1$. Thus, (5.2) implies that

Equation (5.3)

Without loss of generality we may assume that $B>0$. Since $u\in C([0,T];L_{\alpha }(\Omega))$, we have $\operatorname{mes}\lbrace \Omega \cap \lbrace u(t)>B\rbrace \rbrace <C$, $t\in [0,T]$. Hence, $\Vert u^{(B)}(t)\Vert _{p}\leqslant c\Vert \nabla u^{(B))}(t)\Vert _{p}$, $t\in [0,T]$ (see, for example, [17], Ch. 2, § 1, inequality (2.12)). So, the right-hand side of (5.3) tends to zero as $R\rightarrow \infty$. Therefore,

for almost all $x\in \Omega ^r$. Hence, $u^{(B)}(T,x)=0$ almost everywhere in $\Omega ^r$. This proves (5.1) since $r>0$ and $T>0$ are arbitrary.

When $\alpha \geqslant 2$, we have to substitute $\varphi =u^{(B)}\xi$ into (3.9) and proceed according to the above scheme. For example, we transform the first integral in (3.9) (putting $\varphi =u^{(B)}\xi$):

here $h(s)$, $s>0$, is a monotonic function: $h^{\prime }(s)=(\alpha -1)(s+B)^{\alpha -2}s$. The rest of the argument proceeds as above.

Proof of Theorem 3. Assume first that $\alpha \in (1,2)$. We set $\varphi =u\xi ^p$, where the function $\xi (x)$ will be chosen later. Substituting $\varphi$ into (3.17) with $f=0$, we have

Equation (5.4)

By choosing the functions $\xi$ so that the supports of the functions $u_0$ and $\xi$ are disjoint, we can assume that the right-hand side of (5.4) is zero. Next, using Young's inequality,

Equation (5.5)

For $\alpha \geqslant 2$, the same result is obtained by substituting $\varphi =u\xi ^p$ into (3.9).

Let $\xi =\xi (|x|)$, $\xi (t)=0$ for $t<r$, $\xi (t)=(t-r)/{\rho }$ for $t\in (r,r+\rho]$, and $\xi (t)=1$ for $t>r+\rho$. Also let $c$ be such that $|u(t,x)|\leqslant c$. Then

where we have set $\Omega _r^{b}=\lbrace x\in \Omega \mid r<|x|<b\rbrace$, and the superscript $b=\infty$ will be omitted. It follows that

Setting

we have

Equation (5.6)

In view of (3.10), the function $H_r(t)$ is bounded; that is, $H_r(t)\leqslant A$ for all $r>0$, $t>0$. Hence, by (5.6),

Equation (5.7)

Proceeding by induction and applying Stirling's formula, this gives

Let $r=R_0+k\rho$. Then

We choose $k$ so that $e\leqslant \rho ^p k/(e\delta T)\leqslant 2^pe$. Then

As a result,

Consequently,

which implies (2.11). If $k$ cannot be chosen as required, the inequality $\smash[b]{\dfrac{(r-R_0)^p}{e\delta T}<e}$ holds, and estimate (2.11) is trivial.

The proof of Theorem 3 is complete.

§ 6. An upper estimate for the norm of the solution

Our aim in this section is to prove Theorems 4 and 5.

6.1. Proof of Theorem 4.

Let $\theta (\rho)$, $\rho >0$, be an absolutely continuous function equal to 1 for $\rho \geqslant r$, to 0 for $\rho \leqslant R_0$, linear for $\rho \in [R_0,2R_0]$ and satisfying the equation

Equation (6.1)

(the constant $\delta$ will be specified below). Solving this equation, we find, in particular, that

Equation (6.2)

For any function $v(x)\in C_0^{\infty }(\Omega)$ it follows from the definition of $\nu (\rho)$ that

Multiplying by $\theta ^p(\rho)$ and integrating over $\rho \in [a,b]$, we derive the inequality

which now holds for any function ${v}\in \mathring{W}_{\alpha,p}^1(\Omega)$. For $a=2R_0$ and $b=r$, we rewrite the last inequality as follows

Equation (6.3)

Substituting $\xi (x)=\theta (|x|)$ into (5.5), we have

Equation (6.4)

Using (6.1), (6.2), one easily reduces (6.4) to the form

Equation (6.5)

Applying (4.8) with $b=2R_0$ and using (3.18), it follows that

Equation (6.6)

Using (6.3) shows that

Equation (6.7)

Taking $\delta =C^{-1/p}$ and combining (6.5)–(6.7), we obtain

thereby proving inequality (2.14).

The proof of Theorem 4 is complete.

6.2. Proof of Theorem 5.

We choose a positive number $r\geqslant 2R_0$. Setting

in view of (2.14), we have

By (4.8),

Setting $\mu (r)=C_1r^{1+n(p-\alpha)/(p\alpha)}$ with an appropriate factor $C_1$, we have

Equation (6.8)

Let $t_r$ denote a point in the interval $[0,\infty)$ such that $E(t_r)=\Vert u(t_r)\Vert _\alpha ^{\alpha }=\varepsilon (r)$. If $E(t)>\varepsilon (r)$ for any $t\geqslant 0$, we let $t_r=\infty$. Since the function $E(t)$ is monotone nonincreasing for $t\in [0,t_r)$, we have $E(t)>\varepsilon (r)$. Now (6.8) implies that

Equation (6.9)

Differentiating (3.18) with respect to $T$, we see that

Next, in view of (6.9),

Equation (6.10)

Solving this differential inequality, we obtain

The substitution $\varepsilon (r)$ into the last inequality shows that, for $t\in (0,t_r)$,

Equation (6.11)

Note that, for $t\in [t_r,\infty)$, the inequality $E(t)\leqslant \varepsilon (r)$ holds and estimate (6.11) is also valid.

In (6.11) we set $r=r(t)$ and use (2.15):

This proves inequality (2.16).

The proof of Theorem 5 is complete.

Please wait… references are loading.
10.1070/SM2013v204n09ABEH004338