1. Introduction
The objective of the paper is to study the nonnegative weak solutions of nonlinear parabolic equation with the type
ut=div(∣∇um∣p−2∇um)−a(x)umq1|∇um|p1+f0(um)∫ΩK(y)|um(y,t)|βdy+g(x), in S=Ω×(0,∞),
|
(1.1)
|
u(x,t)=0, (x,t)∈∂Ω×(0,∞),
|
(1.3)
|
where Ω⊂RN is a bounded open domain with smooth boundary ∂Ω, ∫ΩK(y)|u(y,t)|βdy represents a nonlocal function dependent on spatial domain Ω, a(x)≥0 is a bounded function,
K(x) and g(x) are bounded functions too, and ∇ is the spatial gradient operator. We assume that p>1,m>1,
p1≤2, p>2p1,
N≥1,
0≤um0(x)∈Lq−1+1m(Ω),q>1, |f0(s)|≤c|s|1m, s∈R1=(−∞,∞).
|
(1.4)
|
As usual, the here and after, the constants c may be different from one to another. The equation with the type of (1.1) has been suggested as the mathematical model for a variety of problems in mechanics, physics and biology, which can be found in [10,11,15,17] et al. Equation (1.1) has been widely researched, whether it is linear or nonlinear, is uniformly parabolic or degenerate parabolic. In what follows, we only give a very roughly review.
If a(x)=g(x)=f0≡0, the existence of nonnegative solution of the problem (1.1)-(1.3), defined in weak sense, is well established (see [10], [6] et al.).
If g(x)=f≡0, some special cases of equation (1.1) had been researched by Bertsh [3], Zhou [36] and Zhang [34] et al. For examples, the existence and the properties of the viscosity solution to the following equation are obtained in [3,36]
where γ is a positive constant. The existence and the properties of the viscosity solution to the following equation are obtained in [34]
ut=△u−b(x)|u|q−1|∇u|2,
|
(1.6)
|
where b(x) is a known function. The most important characteristic of the equation (1.5) or (1.6) lies in that, generally, the uniqueness of the solutions is not true, one can refer to [4,9,29,34,36] for the details. Thus, for the equation with the type of (1.1), one mainly concerns with the existence of the viscosity solution and the related properties such as the large time behavior, one can refer to [8,20,33,35] et al. for some progresses in the direction.
But if p1=0, it is well-known the uniqueness of the solutions is true. Aassila [1] studied equation (1.1) when p=2,m=1 and proved the existence of solution by Schauder fixed point theorem, studied the convergence of the solution towards a steady state by using the point of view in dynamical systems. Cholewa and Dlotko [7], Teman [28] considered the following problem
ut−div(∣∇u∣p−2∇u)+|u|αu=f0(u)+g(x),
|
(1.7)
|
and proved the existence of global attractor in L2 which is in fact a bounded set in W1,p0⋂Lα+2. Chen [20] studied the long time behavior of solutions for following equation
ut−div(∣∇u∣p−2∇u)+a(x)|u|αu=f0(u)∫ΩK(y)|u(y,t)|βdy+g(x),
|
(1.8)
|
and obtained the existence and Lp estimate of the global attractor.
While the papers, first by Nakao-Chen [25] and later by Chen-Wang [6], had studies the global existence and the gradient estimate for the quasilinear parabolic equation of m-Laplacian type with a nonlinear convection term, the typical equations included in [6,25] are with the form as
ut=div(ur|∇u|p−2∇u)+∇A(u).
|
(1.9)
|
In our paper, we will study the global solution of equation (1.1) with the initial value (1.2) and homogeneous boundary value (1.3) by the usual regularized method. The main techniques are inspired by [6,25]. However, due to the local and the nonlocal nonlinearity of the equation we considered, even to prove the initial value condition, we have to put some restrictions in the exponents of m,p,p1,q1. In particular, as we have said, instead of the nonlinear convection term ∇A(u) in equation (1.9), equation (1.1) contains the damping term −a(x)umq1|∇ump1|, the uniqueness of the solutions generally is not true. We can only prove the uniqueness of the solutions under the condition p1=0. If p1≠0 we only can prove the uniqueness of the viscosity solutions. At the same time, comparing with [5], since equation (1.1) is more complicated, how to get the estimate in the gradient term of the solution, and how to prove the continuity of the solution etc, become more difficult. A clear promotion lies in that we put not any restrictions in the derivative f′0(s) of the function f0(s), while it must satisfy that |f′0(s)|≤c|s|r−1 in [5]. Other related works on equation (1.1), one can refer to the references [2,14,16,18,19,22,24,27,30,31,32] et al.
Now we quote the following definition.
Definition 1.1.
A nonnegative function u(x,t) is called a weak solution of (1.1)-(1.3) if u satisfies
(i)
u∈L∞loc(0,∞;L∞(Ω)),
|
(1.10)
|
ut∈L2loc(0,∞;L2(Ω)), um∈L∞loc(0,∞;W1,p0(Ω)),
|
(1.11)
|
(ii)
∬S[uφt−∣∇um∣p−2∇um⋅∇φ−a(x)umq1|∇um|p1φ]dxdt+∬S[f0(um)∫ΩK(y)|um(y,t)|βdy+g(x)]φdxdt=0, ∀φ∈C10(S);
|
(1.12)
|
(iii)
limt→0∫Ω∣u(x,t)−u0(x)∣dx=0.
|
(1.13)
|
We are to get the solution of problem (1.1)-(1.3) by considering the regularized equation
ut=div((|∇um|2+1k)p−22∇um)−a(x)umq1|∇um|p1+f0(um)∫ΩK(y)|um(y,t)|βdy+g(x),
|
(1.14)
|
with the initial value (1.2) and the homogeneous boundary value (1.3). Here 0≤u0k(x) is a suitable smooth function such that u0k(x)∈L∞(Ω), limk→∞‖um0k‖q−1+1m=‖um0‖q−1+1m.
Definition 1.2.
If uk is the solution of the initial boundary value problem of (1.14)-(1.2)-(1.3), limk→∞uk=u, a.e in S, u is a weak solution of (1.1)-(1.3), then u is said to be a viscosity solution.
We need some important lemmas in order to get our results.
Lemma 1.1.
If 1≤l<N,
1+β≤q, 1≤r≤q≤(1+β)Nl/(N−l),
u1+β∈W1,l(Ω), then
‖u‖q≤c1/(1+β)‖u‖1−θr‖u1+β‖θ/(1+β)1,l,
|
(1.15)
|
where
θ=(β+1)(r−1−q−1)/(N−1−l−1+(β+1)r−1).
This lemma is a general version of Gagliardo-Nirenberg inequality, it is first proved by M. Nakao [23].
Lemma 1.2.
Let y(t) be a nonnegative function on
(0,T]. If it satisfies
y′(t)+Atλθ−1y1+θ(t)≤Bt−ky(t)+Ct−δ,0<t≤T,
|
(1.16)
|
where A,θ>0, λθ≥1, B,C≥0,k≤1,
then
y(t)≤A−1θ(2λ+2BT1−k)1θt−λ+2C(λ+BT1−k)−1t1−δ,0<t≤T.
|
(1.17)
|
This lemma can be found in [26].
Lemma 1.3.
Suppose L1≥1, r,R,M>0,
λ1>0. For n=2,3,⋯, let
Ln=RLn−1−M, θn=NR(1−Ln−1L−1n)(N(R−1)+r)−1,
|
βn=(Ln+M)θ−1n−Ln, λn=(1+λn−1(βn−M))β−1n.
|
Then
limn→∞λn=L1λ1r+Nl1+MN.
|
(1.18)
|
This lemma also was first proved in [25], then used in [6].
In our paper, we assume that p>1+1m, so equation (1.1) is a doubly degenerate parabolic equation. By considering the solution uk of the regularized problem (1.14)-(1.2)-(1.3) and using Moser iteration technique, we get uk's local bounded properties and the local bounded properties of the Lp-norm of the gradient ∇uk. By the compactness theorem, we get the existence of the viscosity solution of the diffusion equation itself. In details, we will prove the following theorems.
Theorem 1.1.
It is supposed that K, g are suitable smooth bounded functions, a(x)∈C(¯Ω) and exists a0>0, such that a(x)≥a0 in Ω, f0 satisfies (1.4). If p>1+1m, u0(x)≥0,
um0(x)∈Lq−1+1m(Ω),3>q>2−1m,
|
(1.19)
|
p1≤2, 2p1<p, β<max{p−1−1m,q−1+1m},
|
(1.20)
|
ϵ=max{mNq1Nm(p−1)−N+mq+p1(m(p−1)+m−2)m(p−1)−1,(β+m)NNm(p−1)−N+mq}<1,
|
(1.21)
|
then the problem (1.1)-(1.3) has a weak viscosity solution u, satisfying
um∈L∞loc(0,∞;Lq+1−1m(Ω))⋂L∞loc(0,∞;W1,p0(Ω)),
|
(1.22)
|
and
‖um(t)‖∞≤c(1+t−λ)(1+t)−1/(p−1−1m),t>0,
|
(1.23)
|
where λ=N(pq+(p−1−1m)N)−1. Moreover, if p>2,
then
‖∇um‖p≤c(1+t−δ1)(1+t)−σ,t>0,
|
(1.24)
|
where
δ1=max{1+m−1m(p−1)−1,δ−1},δ=max{m+1m,2β},
|
and
σ=p[m(2q1+1)−1]+mp1[m(p−1)−1](p−p1).
|
Remark 1.1.
The condition (1.21) is only used to prove (1.13). We conjecture that this condition can be weaken.
Theorem 1.2.
Let u be a nonnegative weak solution of problem (1.1)-(1.3). If g(x)≤0, f′0(s)≥0, if
p>1+1m, p1+q1>(p−1) then
suppu(.,s)⊂suppu(.,t),
|
(1.25)
|
for all s,t with 0<s<t.
2. The L∞ estimation of the solution
Instead of considering the regularized problem (1.14)-(1.2)-(1.3) directly as one deals with the case m=1, we have to consider the following approximate problem. For small s>0, we consider
ut=div((|∇um|2+1k)p−22∇um)−a(x)umq1|∇um|p1+f0(um)∫ΩK(y)|um(y,t)|βdy+g(x),
|
(2.1)
|
u(x,0)=u0k(x)+s,x∈Ω,
|
(2.2)
|
where 0≤u0k(x) is a suitable smooth function such that u0k(x)∈L∞(Ω), limk→∞‖um0k‖q−1+1m=‖um0‖q−1+1m.
Similar as the chapter 8 of [13], in which the existence of the initial boundary value problem of the quaslinear equation in the divergent form is obtained, by Leray-Schauder fixed point theory, using the condition p1≤2, we know that problem (2.1)-(2.3) has a nonnegative classical solution uks, we omit the details here.
Let s→0. In a similar way as [33], we are able to prove that
|∇umks|p−2∇umksxi⇀∗ |∇umk|p−2∇umkxi,weakly star
in L∞loc(0,∞;Lpp−1(Ω)),
|
and uk is the solution of equation (2.1) with the following initial boundary values
Lemma 2.1.
Assume that
(H1) a(x)∈C(¯Ω) and exists a0>0, such that a(x)≥a0 in Ω;
(H2) f0(s)∈C(R1),
|f0(s)|≤K0|s|1m, for some K0>0.
(H3) g(x),K(x)∈L∞.
In addition, β+1m<q1, 3>q≥2−1m,
then umk∈L∞loc(0,∞;Lq−1+1m(Ω)) and
‖umk‖q−1+1m≤c(1+t)−1p−1−1m,t≥0.
|
(2.6)
|
Proof. In the proof what follows, we only denote uk as u for simplicity. We only give the proof of the case q>2−1m, if q=2−1m, one can get the conclusion just a minor version. Let An=(q−2)n3−q,Bn=(3−q)n2−q, and
fn(s)={sq−1, if s≥1n,Ans2+Bns, if 0≤s<1n.
|
Suppose that n>k, multiply (2.1) with fn(um) and integrate it on Ω. Since f′(s)>0, then we have
∫Ωfn(um)div(|∇um|2+1k)p−22∇um)dx=−∫Ω(|∇um|2+1k)p−22|∇um|2f′n(um)dx
|
≤−∫Ω|∇um|pf′n(um)dx=−∫Ω∣∇∫um0(f′n(s))1pds∣pdx,
|
(2.7)
|
−∫Ωa(x)fn(um)umq1|∇um|p1dx≤0.
|
(2.8)
|
Suppose that ∣f0(s)∣≤K0sr. Then
|∫Ω⋂{um≤1n}f0(um)fn(um)∫ΩK(y)|um(y,t)|βdydx|
|
≤c(K)∫Ω⋂{um≤1n}umr(Anu2m+Bnum)dx∫Ω|u|mβdy
|
≤c(K)n1−q−r∫Ω|u|mβdy≤c(K)n1−q−r‖um‖βq−1+1m.
|
If r=1m,
|∫Ω⋂{um>1n}f0(um)fn(um)∫ΩK(y)|um(y,t)|βdydx|
|
≤c(K)∫Ωum(r+q−1)dx∫Ω|u|mβdy≤c‖um‖q−1+1m+βq−1+1m,
|
we have
|∫Ωf0(um)fn(um)∫ΩK(y)|um(y,t)|βdydx|
|
≤c‖um‖βq−1+1m[n1−q−1m+‖um‖q−1+1mq−1+1m].
|
(2.9)
|
|∫Ω⋂{um>1n}fn(um)g(x)dx|≤c(g)∫Ωum(q−1)dx≤c(g)‖um‖q−1q−1+1m.
|
(2.10)
|
From the above calculations, we have
∫Ωfn(um)utdx+∫Ω∣∇∫um0(f′n(s))1pds∣pdx≤c‖um‖q−1+1m+βq−1+1m+O(1nq−1),
|
(2.11)
|
by Poincare inequality, we have
∫Ωfn(um)utdx+c∫Ω∣∫um0(f′n(s))1pds∣pdx≤c‖um‖q−1+1m+βq−1+1m+O(1nq−1).
|
(2.12)
|
Let n→∞ in (2.12). We can deduce that
ddt∫Ωum(q−1)+1dx+c∫Ωum[q−1+1m+p−1−1m]dx≤c‖um‖q−1+1m+βq−1+1m.
|
(2.13)
|
By Jessen inequality, from (2.13) we get
ddt‖um‖q−1+1mq−1+1m+c‖um‖q−1+1m+p−1−1mq−1+1m≤c‖um‖q−1+1m+βq−1+1m.
|
If
by young inequality,
ddt‖um‖q−1+1mq−1+1m+c‖um‖q−1+1m+p−1−1mq−1+1m≤c,
|
then
‖um‖q+1−1m≤c(1+t)−1p−1−1m.
|
We get the desired result.
Lemma 2.2.
If p>1+1m, uk is the solution of problem (2.1)-(2.4)-(2.5), then
‖umk‖∞≤ct−λ, 0<t≤1,
|
(2.14)
|
‖umk‖∞≤c(1+t)−1p−1−1m,t≥1,
|
(2.15)
|
where λ=N(p−1−1m)N+qp.
Proof. Multiply (2.1) with um(l−1), and integrate it on Ω, then
∫Ωum(l−1)utdx=∫Ωdiv(|∇um|+1k)p−22∇um)um(l−1)dx−∫Ωa(x)umq1|∇um|p1um(l−1)dx
|
+∫Ωf0(um)um(l−1)∫ΩK(y)|um(y,t)|βdydx+∫Ωg(x)um(l−1)dx
|
=−(l−1)∫Ω(|∇um|+1k)p−22|∇um|2um(l−2)dx−∫Ωa(x)umq1|∇um|p1um(l−1)dx
|
+∫ΩK(y)|um(y,t)|βdy∫Ωf0(um)um(l−1)dx+∫Ωg(x)um(l−1)dx
|
≤−(l−1)∫Ω(|∇um|+1k)p−22|∇um|2um(l−2)dx
|
+c(K)∫Ω|um(y,t)|βdy∫Ωum(l−1)+1dx+c(g)∫Ωum(l−1)dx,
|
which deduces that
ddt‖um‖l−1+1ml−1+1m+c(l−1+1m)2−p∫Ω∣∇ump+l−1+1m−1−1mp∣pdx≤c‖um‖l−1+1ml−1+1m‖um‖q−1+1m+βq−1+1m+c‖um‖l−1l−1+1m
|
≤‖um‖l−1+1ml−1+1m+c‖um‖l−1l−1+1m,(by (2.6)).
|
Set L=l−1+1m. Then
ddt‖um‖LL+cL2−p∫Ω∣∇umL+p−1−1mp∣pdx≤c‖um‖L+βL+c‖um‖L−1mL,
|
(2.16)
|
where c is a constant independent of l.
Now, if we choose L1=q−1+1m,Ln=rLn−1−(p−1−1m),
θn=rN(1−Ln−1L−1n)(p+N(r−1))−1,
μn=(Ln+p−1−1m)θ−1n−Ln,
r>1+(p−1−1m)q−1, n=2,3,⋯, by Lemma 1.3, we have
‖um‖Ln≤cp/(Ln+p−1−1m)‖um‖1−θnLn−1‖∇um(Ln+p−1−1m)/p‖pθn/(p−1−1m+Ln)p.
|
(2.17)
|
If we choose L=Ln in (2.16), by (2.17), we have
ddt‖um‖LnLn+c−p/θnL2−pn‖um‖Ln+μnLn‖um‖p−1−1m−μnLn−1≤c‖um‖Ln+βLn+c‖um‖Ln−1mLn. 0<t≤1.
|
(2.18)
|
We will prove that there exist two bounded sequences {ξn},{λn} such that
‖um‖Ln≤ξnt−λn, 0<t≤1.
|
(2.19)
|
Without loss of the generality, we may assume that ‖um‖Ln≥1. Otherwise, choosing ξn≡1,
(2.17) is true naturally. Thus, by (2.16), we have
ddt‖um‖LnLn+c−p/θnL2−pn‖um‖Ln+μnLn‖um‖p−1−1m−μnLn−1≤c‖um‖Ln+βLn. 0<t≤1.
|
If n=1, by Lemma 2.1, λ1=0,ξ1=supt≥0‖um(t)‖q−1+1m makes (2.19) sure. If (2.19) is true for n−1, from (2.18),
ddt‖um‖LnLn+c−p/θnL2−pn‖um‖Ln+μnLnξp−1−1m−μnn−1t−(p−1−1m−μn)λn−1≤c‖um‖Ln+βLn. 0<t≤1.
|
(2.20)
|
we can choose
λn=(λn−1(μn−p+1+1m)+1)μ−1n, ξn=ξn−1(cp/θnLp−1nλn)1/μn, n=2,3,⋯,
|
ddt‖um‖LnLn+c‖um‖Ln+λnLn≤c‖um‖Ln+βLn. 0<t≤1.
|
(2.21)
|
Suppose that
and notice that as n→∞,
λn→λ=N(p−1−1m)N+pq.
ddt‖um‖LnLn+c‖um‖Ln+λnLn≤0. 0<t≤1.
|
(2.23)
|
By Lemma 1.2 and (2.23), we know (2.19) is true.
Moreover, it is easy to see that {ξn} is bounded. Thus, by Lemma 1.2, (2.14) is true.
To prove (2.15), we set τ=log(1+t),t≥1,
w(τ)=(1+t)1p−1−1mum(t). By (2.16), we have
ddτ‖w(τ)‖LL+cL2−p‖∇wL+p−1−1mp‖pp≤Lp−1−1m‖w(τ)‖LL+c‖w(τ)‖L+βL, τ≥log2.
|
(2.21)
|
By the lemma 3.1 in [24], we can get (2.15), we omit details here.
3. The L∞ estimation of the gradient
Lemma 3.1.
If p>max{2,1+1m}, uk is the solution of problem (2.1)-(2.4)-(2.5), then
‖∇umk‖p≤ct−(1+m−1m(p−1)−1)+ct1−δ, 0<t≤1,
|
(3.1)
|
‖∇umk‖p≤c(1+t)−p(m(2q1+1)−1)+mp1(m(p−1)−1)(p−p1),t≥1.
|
(3.2)
|
Here δ=max{m−1m,2β}.
Proof. Multiply (2.1) with umt, and integrate it on Ω, then
m∫Ωum−1(ut)2dx=∫Ωdiv((|∇um|2+1k)p−22∇um)umtdx−∫Ωa(x)umq1|∇um|p1umtdx
|
+∫Ωf0(um)umtdx∫ΩK(y)|um(y,t)|βdy+∫Ωg(x)umtdx.
|
(3.3)
|
∫Ωdiv((|∇um|2+1k)p−22∇um)umtdx=−∫Ω(|∇um|+1k)p−22∇um∇umtdx
|
=−12∫Ω(|∇um|2+1k)p−22|∇um|2tdx
|
=−12∫Ωddt∫|∇um|20(s+1k)p−22dsdx=−12ddtΓk(|∇um|2),
|
(3.4)
|
where we define that
Γk(|∇um|2)=∫Ω∫|∇um|20(s+1k)p−22dsdx.
|
At the same time,
|−a(x)umq1|∇um|p1umtdx|≤m2∫Ωum−1(ut)2dx+c∫Ω|um|2q1+m−1m|∇um|2p1dx.
|
(3.5)
|
By Lemma 2.1, using Young inequality and H¨older inequality,
|∫Ωf0(um)umtdx∫ΩK(y)|um(y,t)|βdy|
|
≤c(ε∫Ωum−1(ut)2dx+c∫Ωum+1dx)‖um‖βq−1+1m
|
≤cε∫Ωum−1(ut)2dx+c∫Ωum+1dx
|
|∫Ωg(x)umtdx|≤ε∫Ωum−1(ut)2dx+c∫Ωum−1dx.
|
By (3.3)-(3.5), we have
∫Ωum−1(ut)2dx+1mddtΓk(|∇um|2)≤c∫Ω|um|2q1+m−1m|∇um|2p1dx+c∫Ωum+1dx+c∫Ωum−1dx.
|
(3.6)
|
Multiply (2.1) with um, and integrate it on Ω, then
1m+1∫Ωddtum+1dx=∫Ωdiv((|∇um|2+1k)p−22∇um)umdx−∫Ωa(x)umq1|∇um|p1umdx
|
+∫Ωf0(um)um∫ΩK(y)|um(y,t)|βdydx+∫Ωg(x)umdx
|
=−∫Ω(|∇um|2+1k)p−22|∇um|2dx−∫Ωa(x)umq1|∇um|p1umdx
|
+∫Ωf0(um)um∫ΩK(y)|um(y,t)|βdydx+∫Ωg(x)umdx
|
and
Γk(|∇um|2)≤∫Ω(|∇um|2+1k)p−22|∇um|2dx
|
=−1m+1∫Ωddtum+1dx−∫Ωa(x)umq1|∇um|p1umdx+∫Ωf0(um)um∫ΩK(y)|um(y,t)|βdydx+∫Ωg(x)umdx
|
≤1m+1‖um+12‖2‖um−12ut‖2+c(K)∫Ω|um(y,t)|βdy∫Ωum+1dx+c(g)∫Ωumdx,
|
so
1mddtΓk(|∇um|2)+(m+1)2‖um+12‖−22Γ2k(|∇um|2)
|
≤c∫Ω|um|2q1+m−1m|∇um|2p1dx+c∫Ωum+1dx+c∫Ωum−1dx
|
+c‖um+12‖−22(∫Ω|um(y,t)|βdy∫Ωum+1dx+∫Ωumdx)2
|
≤c∫Ω|um|2q1+m−1m|∇um|2p1dx+c∫Ωum+1dx+c∫Ωum−1dx
|
+c(∫Ω|um(y,t)|βdy)2∫Ωum+1dx+c‖um+12‖2mm+1−22.
|
(3.7)
|
Setting 2γ=2q1+1−1m, for ∀a∈[0,2γ], if we notice that p>2p1, then we have
∫Ω|um|2a|∇um|2p1dx≤‖um(t)‖a∞(∫Ω|um|(2γ−a)pp−2p1dx)p−2p1p‖∇um‖2p1p.
|
(3.8)
|
If 2γ≥(p−2p1)(N+1)/N, let a=(2γ−(p−2p1)(1+qN))+. By Lemma 1.3,
(∫Ω|um|(2γ−a)pp−2p1dx)p−2p1p≤c‖um(t)‖(2γ−a)(1−θ)s‖∇um‖p−2p1p,
|
(3.9)
|
where θ=(s−1−(1−2p1p)(2γ−a)−1)/(N−1−p−1+s−1), and s=(2γ−p+2p1−a)N/(p−2p1) when 2γ≥(p−2p1)(1+q/N), s=q when (p−2p1)(1+N−1)≤2γ≤(p−2p1)(1+q/N). By Lemma 2.1 and Lemma 2.2, from (3.8), we have
∫Ω|um|2a|∇um|2p1dx≤ct−λa‖∇um‖pp≤ct−λaΓk(|∇um|2). 0<t≤1.
|
(3.10)
|
At the same time, if we choose q=2 in Lemma 2.1, we have
‖um‖1+1m=(∫Ωum+1dx)mm+1≤c(1+t)−(p−1−mm+1)−1≤c,
|
and
∫Ωum−1dx≤ctm−1mλ, ‖um+12‖22=∫Ωum+1dx≤c.
|
(3.11)
|
By (3.7) and Lemma 2.2, we have
Γ′k(t)+ctm+1m(p−1)−1Γ2k(t)≤ct−λaΓk(t)+c(t−λm−1m+t−2βλ), 0<t≤1.
|
(3.12)
|
If 2γ<(p−2p1)(N+1)/N and p−2p1≤2a≤2γ,
∫Ω|um|2a|∇um|2p1dx≤c‖∇um‖2a(1−θ)1‖∇um‖2aθ+2p1p≤c‖∇um‖pp≤cΓk(|∇um|2). 0<t≤1.
|
(3.13)
|
If 2\gamma < (p-2p_1)(N+1)/N and p-2p_1\geq 2a\geq 0,
\int_{\Omega}|u^{m}|^{2a}|\nabla u^{m}|^{2}dx\leq c(1+\|\nabla
u^{m}\|_{p}^{p})\leq c(1+\Gamma_{k}(|\nabla u^{m}|^{2})).\ 0<t\leq
1.
|
(3.14)
|
(3.13) and (3.14) imply that (3.12) is still true when 2\gamma < (p-2p_1)(N+1)/N. Using Lemma 1.2,
\Gamma_{k}(t)\leq ct^{-(1+\frac{m-1}{m(p-1)-1})}+ct^{1-\delta}, \
0<t\leq 1,
|
where \delta =\max\{\frac{m+1}{m}, 2\beta\}. Then (3.1) is true. Now, we will prove (3.2). For t\geq 1, by (2.15)
\int_{\Omega}|u^{m}|^{2a}|\nabla u^{m}|^{2p_1}dx\leq c\|\nabla
u^{m}\|_{p}^{2}\|u^{m}(t)\|_{2\gamma p/p-2p_1}^{2\gamma}\leq
c(1+t)^{-2\gamma/(p-1-\frac{1}{m})}\|\nabla u^{m}\|_{p}^{2p_1}.\
t\geq 1.
|
(3.15)
|
\Gamma_{k}(|\nabla u^{m}|^{2})=\int_{0}^{|\nabla
u^{m}|^{2}}(s^{2}+\frac{1}{k})^{\frac{p-2}{2}}ds\leq c \|\nabla
u^{m}\|_{p}^{p}=c(\|\nabla u^{m}\|_{p}^{2p_1})^{\frac{p}{2p_1}}, \
t\geq 1.
|
(3.16)
|
\|u^{\frac{m+1}{2}}\|^{2}_{2}=\left(\int_{\Omega}u^{m+1}dx\right)^{2}\leq
c(1+t)^{-(p-1-\frac{1}{m})^{-1}}, \ t\geq 1.
|
(3.17)
|
by (3.7), using (3.15)-(3.17)
\Gamma_{k}'(t)+c(1+t)^{-(p-1-\frac{1}{m})^{-1}}\Gamma_{k}^{2}(t)\leq
c(1+t)^{2\gamma/(p-1-\frac{1}{m})}(\Gamma_{k}(t))^{\frac{2p_1}{p}}
|
+c\int_{\Omega}u^{m+1}dx+c\int_{\Omega}u^{m-1}dx
+c(\int_{\Omega}|u^m(y, t)|^{\beta}dy)^{2}\int_{\Omega}u^{m+1}dx+c\|u^{\frac{m+1}{2}}\|_{2}^{2(m-1)},
|
by Young inequality,
\Gamma_{k}'(t)+c(1+t)^{-(p-1-\frac{1}{m})^{-1}}\Gamma_{k}^{2}(t)\leq
c(1+t)^{\frac{-m(2\gamma
p+p_1)}{(m(p-1)-1)(p-p_1)}}+c(1+t)^{-\frac{m(m+1)}{m(p-1)-1}}
|
=c(1+t)^{-\frac{p(m(2q_1+1)-1)+mp_1}{(m(p-1)-1)(p-p_1)}}+c(1+t)^{-\frac{m(m+1)}{m(p-1)-1}},
|
which means (3.2) is true.
Lemma 3.2.
If p> 1+\frac{1}{m}, u_{k} is the solution of problem (2.1) -(2.4)-(2.5), then
\int_{t}^{T}\int_{\Omega}u^{m-1}_{k}(u_{kt})^{2}dxds\leq
ct^{-(1+\frac{m-1}{m(p-1)-1})}+ct^{-(\lambda
\gamma+\frac{m-1}{m(p-1)-1})}+ct^{-\frac{1+m}{m}\lambda}, \ 0<t\leq
T.
|
(3.18)
|
Proof. From (3.6), (3.10) and (2.14), we have
\int_{\Omega}u^{m-1}(u_{t})^{2}dx+\frac{1}{m}\frac{d}{dt}\Gamma_{k}(|\nabla
u^{m}|^{2})\leq c\int_{\Omega}|u^{m}|^{2q_1+\frac{m-1}{m}}|\nabla
u^{m}|^{2p_1}dx+c\int_{\Omega}u^{m+1}dx+c\int_{\Omega}u^{m-1}dx
|
\int_{t}^{T}\int_{\Omega}u^{m-1}(u_{t})^{2}dxds\leq
\Gamma_{k}(t)+c\int_{t}^{T}\int_{\Omega}|u^{m}|^{2q_1+\frac{m-1}{m}}|\nabla
u^{m}|^{2p_1}dxds+c\int_{t}^{T}\int_{\Omega}u^{m+1}dx
|
\leq\Gamma_{k}(t)+c\int_{t}^{T}s^{-\lambda(2q_1+\frac{m-1}{m})}\Gamma_{k}(s)ds+c\int_{t}^{T}\int_{\Omega}u^{m+1}dx
|
\leq ct^{-(1+\frac{m-1}{m(p-1)-1})}+ct^{-(\lambda
\gamma+\frac{m-1}{m(p-1)-1})}+ct^{-\frac{1+m}{m}\lambda}.
|
(3.19)
|
4. The proof of Theorem 1.1
The proof of Theorem 1.1 from Lemma 2.1, Lemma 2.2, Lemma 3.1 and Lemma 3.2, using the compactness theory (cf [21]), there is a sequence (still denoted it as \{u_{k}\}) of \{u_{k}\} such that when k\rightarrow \infty, u_{k}\rightarrow u, \ a.e. \ {\rm in}\ S and so
\lim\limits_{k\rightarrow
\infty}f_0(u_{k}^{m})\int_{\Omega}K(y)|u_{k}^{m}(y, t)|^{\beta}dy=
f_0(u^{m})\int_{\Omega}K(y)|u^{m}(y, t)|^{\beta}dy.
|
Moreover, we have
u_{k}\rightharpoonup \ u, \text{weakly* star in}\
L^{\infty}_{\rm loc}(0, \infty; L^{m(q-1)+1}(\Omega)),
|
(4.1)
|
u_{kt}\rightharpoonup u_{t}, \text{ weakly in} \ L^{2}(0, \infty;
L^{2}(\Omega)), \nabla u^{m}_{k}\rightharpoonup \nabla u^{m}, \
\text{weakly in} \ L^{p}_{{\rm loc}}(0, \infty;
L^{p}(\Omega))
|
(4.2)
|
|\nabla u^{m}_{k}|^{p-2} u^{m}_{kx_{i}}\rightharpoonup\
\chi_{i}, \text{weakly* in}\ L_{\text{loc}}^{\infty}(0, \infty;
L^{\frac{p}{p-1}}(\Omega)),
|
(4.3)
|
a(x)u_{k}^{mq_1}|\nabla u^{m}_{k}|^{p_1}\rightharpoonup\ \nu,
\text{weakly* in}\ L_{\text{loc}}^{\infty}(0, \infty;
L^{\frac{p}{p_1}}(\Omega)),
|
(4.4)
|
where \chi=\{\chi_{i}: 1\leq i\leq N\} and every \chi_{i} is a function in L_{loc}^{\infty}(0, T; L^{\frac{p}{p-1}}(\Omega)), \nu \in L_{loc}^{\infty}(0, \infty; L^{\frac{p}{p_1}}(\Omega)). (4.1) and (4.2) are clearly true.
In what follows, we only need to prove that
\chi=|\nabla u^{m}|^{p-2}\nabla u^{m}, \ \ {\rm in}\ L_{{\rm
loc}}^{\infty}(0, \infty; L^{\frac{p}{p-1}}(\Omega)).
|
(4.5)
|
and
\nu=a(x)u^{mq_1}|\nabla u^{m}|^{p_1}, \ \ {\rm in} \ L_{{\rm
loc}}^{\infty}(0, \infty; L^{\frac{p}{p_1}}(\Omega)).
|
(4.6)
|
It is easy to know that
\iint_{S}\left(u\varphi_{t}-\chi\cdot \nabla\varphi-\nu \varphi+
f_0(u^{m})\int_{\Omega}K(y)|u^{m}(y, t)|^{\beta}dy\varphi+g(x)\varphi\right)dxdt=0,
\ \forall \varphi \in C_{0}^{\infty}(S),
|
(4.7)
|
so, if we can prove that
\iint_{S}\mid \nabla u^{m}\mid^{p-2}\nabla u^{m}\cdot \nabla\varphi
dxdt=\iint_{S}\chi\cdot \nabla\varphi dxdt, \ \forall \varphi \in
C_0^{1}(S);
|
(4.8)
|
\iint_{S}a(x)u^{mq_1}|\nabla u^{m}|^{p_1}\varphi
dxdt=\iint_{S}\nu \varphi dxdt, \ \ \forall \varphi \in
C_0^{1}(S);
|
(4.9)
|
then (4.5), (4.6) and (1.12) are true.
First, for any \psi\in C^{\infty}_{0}(S), 0\leq\psi\leq 1;
v^{m}\in L^{p}_{\text{loc}}(0, T; W^{1, p}_{0}(\Omega)), we have
\iint_{S}\psi(\mid \nabla u_{k}^{m}\mid^{p-2}\nabla u_{k}^{m}-\mid
\nabla v^{m}\mid^{p-2}\nabla
v^{m})\cdot\nabla(u_{k}^{m}-v^{m})dxdt\geq 0,
|
(4.10)
|
If we multiply with u^{m}_{k}\psi on two sides of (2.1), then we have
\iint_{S}\psi\left(|\nabla
u^{m}_{k}|^{2}+\frac{1}{k}\right)^{\frac{p-2}{2}}|\nabla
u^{m}_{k}|^{2}dxdt=\frac{1}{m+1}\iint_{S}\psi_{t}u^{m+1}_{k}dxdt-\iint_{S}u^{m}_{k}\left(|\nabla
u^{m}_{k}|^{2}+\frac{1}{k}\right)^{\frac{p-2}{2}}\nabla
u^{m}_{k}\cdot\nabla\psi dxdt
|
-\iint_{S}a(x)u_{k}^{m(q_1+1)}|\nabla u^{m}_{k}|^{p_1}\psi
dxdt+\iint_{S}[
f_0(u_{k}^{m})\int_{\Omega}K(y)|u_{k}^{m}(y, t)|^{\beta}dy+g(x)]u^{m}_{k}\psi
dxdt.
|
(4.11)
|
Noticing that when 1 < p < 2,
|\nabla u^{m}_{k}|^{2}\geq (|\nabla
u_{k}^{m}|^{2}+\frac{1}{k})^{\frac{p}{2}}-(\frac{1}{k})^{\frac{p}{2}},
|
(|\nabla u_{k}^{m}|^{2}+\frac{1}{k})^{\frac{p-2}{2}}|\nabla
u^{m}_{k}|\leq (|\nabla u_{k}^{m}|^{2}+\frac{1}{k})^{\frac{p-1}{2}},
|
and when p\geq 2,
(|\nabla u_{k}^{m}|^{2}+\frac{1}{k})^{\frac{p-2}{2}}|\nabla
u_{k}^{m}|^{2}\geq|\nabla u_{k}^{m}|^{p},
|
(|\nabla u_{k}^{m}|^{2}+\frac{1}{k})^{\frac{p-2}{2}}|\nabla
u_{k}^{m}|\leq(|\nabla u_{k}^{m}|^{p-1}+1),
|
by (4.10), (4.11), we have
\frac{1}{m+1}\iint_{S}\psi_{t}u^{m+1}_{k}dxdt-\iint_{S}u^{m}_{k}\left(|\nabla
u^{m}_{k}|^{2}+\frac{1}{k}\right)^{\frac{p-2}{2}}\nabla
u^{m}_{k}\cdot\nabla\psi dxdt
|
-\iint_{S}a(x)u_{k}^{m(q_1+1)}|\nabla u^{m}_{k}|^{p_1}\psi dxdt
+(\frac{1}{k})^{\frac{p-2}{2}}\text{mes}\Omega
|
+\iint_{S}[
f_0(u_{k}^{m})\int_{\Omega}K(y)|u_{k}^{m}(y, t)|^{\beta}dy+g(x)]u^{m}_{k}\psi
dxdt
|
-\iint_{S}\psi|\nabla u_{k}^{m}|^{p-2}\nabla u_{k}^{m}\cdot\nabla
v^{m}dxdt-\iint_{S}\psi|\nabla v^{m}|^{p-2}\nabla v^{m}\cdot
\nabla(u_{k}^{m}-v^{m})dxdt\geq 0.
|
(4.12)
|
Since
\left(|\nabla
u^{m}_{k}|^{2}+\frac{1}{k}\right)^{\frac{p-2}{2}}\nabla
u^{m}_{k}=|\nabla u_{k}^{m}|^{p-2}\nabla
u_{k}^{m}+\frac{p-2}{2k}\int_{0}^{1}(|\nabla
u_{k}^{m}|^{2}+\frac{s}{k})^{\frac{p-4}{2}}ds\nabla u^{m}_{k},
|
and
\lim\limits_{k\rightarrow\infty}\iint_{S}\int_{0}^{1}(|\nabla
u_{k}^{m}|^{2}+\frac{s}{k})^{\frac{p-4}{2}}ds\nabla
u^{m}_{k}\cdot\nabla\psi u^{m}_{k}dxdt=0,
|
if we let k\rightarrow\infty in (4.12), we have
\frac{1}{m+1}\iint_{S}\psi_{t}u^{m+1}dxdt -\iint_{S}u^m\nu\psi dxdt-\iint_{S}u^m\chi\nabla\psi dxdt
|
-\iint_{S}\psi \chi\cdot\nabla v^{m}dxdt-\iint_{S}\psi|\nabla
v^{m}|^{p-2}\nabla v^{m}\cdot \nabla(u^{m}-v^{m})dxdt
|
+\iint_{S}[
f_0(u^{m})\int_{\Omega}K(y)|u^{m}(y, t)|^{\beta}dy+g(x)]u^{m}\psi
dxdt \geq 0.
|
(4.13)
|
Now, we choose \varphi=\psi u^{m} in (4.7),
\frac{1}{m+1}\iint_{S}\psi_{t}u^{m+1}dxdt-\iint_{S}u^m\nu\psi dxdt
-\iint_{S}\chi\cdot\nabla \psi u^m dxdt
|
+\iint_{S}[f_0(u^{m})\int_{\Omega}K(y)|u^{m}(y, t)|^{\beta}dy+g(x)]\psi
u^{m}dxdt =\iint_{S}\psi \chi\cdot\nabla u^m dxdt.
|
From this formula and (4.13), we have
\iint_{S}\psi(\chi-|\nabla v^{m}|^{p-2}\nabla v^{m})\cdot
\nabla(u^{m}-v^{m})dxdt\geq 0.
|
(4.14)
|
Let v^{m}=u^{m}-\lambda\varphi, \lambda\geq 0, \varphi\in C^{\infty}_{0}(S). Then
\iint_{S}\psi(\chi_{i}-|\nabla(u^{m}-\lambda\varphi)|^{p-2}(u^{m}-\lambda\varphi)_{x_{i}})dxdt\geq
0.
|
Let \lambda\rightarrow 0. We obtain
\iint_{S}\psi(\chi_{i}-|\nabla u^{m}|^{p-2}u^{m}_{x_{i}})dxdt\geq0,
\forall \varphi\in C^{\infty}_{0}(S).
|
Moreover, if we choose \lambda\leq 0, we are able to get
\iint_{S}\psi(\chi_{i}-|\nabla u^{m}|^{p-2}u^{m}_{x_{i}})dxdt\leq0,
\forall \varphi\in C^{\infty}_{0}(S).
|
Now, if we choose \psi such that \text{supp}\varphi\subset\text{supp}\psi, and on \text{supp}\varphi, \psi=1, then we can get (4.8).
By a process of limitation, we can choose the test function \varphi in (4.8) as u^m, then we have
\lim\limits_{k\rightarrow 0}\iint_{S}|\nabla u_{k}^{m}|^{p}dxdt=\iint_{S}\chi\cdot \nabla u^m dxdt=\iint_{S}|\nabla u^{m}|^{p}dxdt.
|
(4.15)
|
Due to (1.20), 2p_{1} < p, then by H\ddot{o}lder inequality, we have
\lim\limits_{k\rightarrow 0}\iint_{S}|\nabla u_{k}^{m}|^{p_1}dxdt=\iint_{S}\chi\cdot \nabla u^m dxdt=\iint_{S}|\nabla u^{m}|^{p_1}dxdt.
|
(4.16)
|
By a refinement of Fatou's lemma, the theorem 1.4.1 in [12], we are easy to prove (4.9), and so (1.12) is true.
Secondly, we are to prove (1.13).
For small r>0, denote \Omega_{r}=\{x\in\Omega:
\text{dist}(x, \partial \Omega)\leq r\}. For any \eta>0, let
\text{sgn}_{\eta}(s)=\left\{\begin{array}{ccc} 1, \ \ &\ \ {\rm if}\
\
s>\eta, \\
\frac{s}{\eta}, \ \ &\ \ {\rm if}\ \
|s|\leq\eta, \\
-1, \ \ &\ \ {\rm if}\ \ s<-\eta.
\end{array}
\right.
|
For any given small r>0, large enough k, l, we declare that
\int_{\Omega_{2r}}|u_{k}(x, t)-u_{l}(x, t)|dx\leq
\int_{\Omega_{r}}|u_{k}(x, 0)-u_{l}(x, 0)|dx+c_{r}(t),
|
(4.17)
|
where c_{r}(t) is independent of k, l, and \lim_{t\rightarrow 0}c_{r}(t)=0. By (2.1)
\int_{0}^{t}\int_{\Omega_{r}}\varphi (u_{kt}-u_{lt})dxd\tau+\int_{0}^{t}\int_{\Omega_{r}}\nabla \varphi
[(|\nabla u_{k}^{m}|^{2}+\frac{1}{k})^{\frac{p-2}{2}}\nabla
u_{k}^{m}-(|\nabla u_{l}^{m}|^{2}+\frac{1}{l})^{\frac{p-2}{2}}\nabla
u_{l}^{m}]dxd\tau
|
+\int_{0}^{t}\int_{\Omega_{r}}a(x)(u^{mq_1}_{k}|\nabla u_{k}^{m}|^{p_1}-u^{mq_1}_{l}|\nabla u_{l}^{m}|^{p_1})\varphi
dxd\tau
|
+\int_{0}^{t}\int_{\Omega_{r}}[f_0(u^{m}_{k})\int_{\Omega}K(y)|u^{m}_{k}(y, t)|^{\beta}dy
-f_0(u^{m}_{l})\int_{\Omega}K(y)|u^{m}_{l}(y, t)|^{\beta}dy]\varphi
dxd\tau
=0,
|
(4.18)
|
for \forall \varphi \in L^{p}(0, T; W_{0}^{1, p}(\Omega)). Suppose that \xi(x)\in C_{0}^{1}(\Omega_{r}) such that
0\leq\xi\leq 1; \ \ \xi\mid_{\Omega_{2r}}=1,
|
and choose \varphi=\xi\text{sgn}_{\eta}(u_{k}^{m}-u_{l}^{m}) in (4.18), then
\int_{0}^{t}\int_{\Omega_{r}} \xi\text{sgn}_{\eta}(u_{k}^{m}-u_{l}^{m})
(u_{kt}-u_{lt})dxd\tau
|
+\int_{0}^{t}\int_{\Omega_{r}}
[(|\nabla u_{k}^{m}|^{2}+\frac{1}{k})^{\frac{p-2}{2}}\nabla
u_{k}^{m}-(x|\nabla u_{l}^{m}|^{2}+\frac{1}{l})^{\frac{p-2}{2}}\nabla
u_{l}^{m}]\nabla\xi \text{sgn}_{\eta}(u_{k}^{m}-u_{l}^{m}) dxd\tau
|
+\int_{0}^{t}\int_{\Omega_{r}}
[(|\nabla u_{k}^{m}|^{2}+\frac{1}{k})^{\frac{p-2}{2}}\nabla
u_{k}^{m}-(x|\nabla u_{l}^{m}|^{2}+\frac{1}{l})^{\frac{p-2}{2}}\nabla
u_{l}^{m}]\nabla(u_{k}^{m}-u_{l}^{m})\xi \text{sgn}'_{\eta}(u_{k}^{m}-u_{l}^{m}) dxd\tau
|
+\int_{0}^{t}\int_{\Omega_{r}}a(x)(u^{mq_1}_{k}|\nabla
u_{k}^{m}|^{p_1}-u^{mq_1}_{l}|\nabla
u_{l}^{m}|^{p_1})\xi\text{sgn}_{\eta}(u_{k}^{m}-u_{l}^{m})
dxd\tau
|
+\int_{0}^{t}\int_{\Omega_{r}}[f_0(u^{m}_{k})\int_{\Omega}K(y)|u^{m}_{k}(y, t)|^{\beta}dy
-f_0(u^{m}_{l})\int_{\Omega}K(y)|u^{m}_{l}(y, t)|^{\beta}dy]\xi\text{sgn}_{\eta}(u_{k}^{m}-u_{l}^{m})
dxd\tau
\leq0.
|
(4.19)
|
If we notice that the third term in the left hand side on (4.19) is nonnegative when \eta\rightarrow 0, then we have
\lim\limits_{\eta\rightarrow 0}\int_{0}^{t}\int_{\Omega_{r}}
\xi\text{sgn}_{\eta}(u_{k}^{m}-u_{l}^{m})
(u_{kt}-u_{lt})dxd\tau
|
+\lim\limits_{\eta\rightarrow 0}\int_{0}^{t}\int_{\Omega_{r}}[(|\nabla
u_{k}^{m}|^{2}+\frac{1}{k})^{\frac{p-2}{2}}\nabla u_{k}^{m}-(|\nabla
u_{l}^{m}|^{2}+\frac{1}{l})^{\frac{p-2}{2}}\nabla
u_{l}^{m}]\nabla\xi \text{sgn}_{\eta}(u_{k}^{m}-u_{l}^{m}) dxd\tau
|
+\lim\limits_{\eta\rightarrow
0}\int_{0}^{t}\int_{\Omega_{r}}a(x)(u^{mq_1}_{k}|\nabla
u_{k}^{m}|^{p_1}-u^{mq_1}_{l}|\nabla
u_{l}^{m}|^{p_1})\xi\text{sgn}_{\eta}(u_{k}^{m}-u_{l}^{m})
dxd\tau
|
+\lim\limits_{\eta\rightarrow
0}\int_{0}^{t}\int_{\Omega_{r}}[f_0(u^{m}_{k})\int_{\Omega}K(y)|u^{m}_{k}(y, t)|^{\beta}dy
-f_0(u^{m}_{l})\int_{\Omega}K(y)|u^{m}_{l}(y, t)|^{\beta}dy]\xi\text{sgn}_{\eta}(u_{k}^{m}-u_{l}^{m})
dxd\tau
=0.
|
(4.20)
|
At the same time,
\lim\limits_{\eta\rightarrow 0}\int_{0}^{t}\int_{\Omega_{r}}
\xi\text{sgn}_{\eta}(u_{k}^{m}-u_{l}^{m})
(u_{kt}-u_{lt})dxd\tau=\int_{0}^{t}\int_{\Omega_{r}}\xi\text{sgn}(u_{k}^{m}-u_{l}^{m})
(u_{kt}-u_{lt})dxd\tau
|
=\int_{0}^{t}\int_{\Omega_{r}}\xi\text{sgn}(u_{k}-u_{l})
(u_{kt}-u_{lt})dxd\tau
|
\lim\limits_{\eta\rightarrow 0}\int_{0}^{t}\int_{\Omega_{r}}
\xi\text{sgn}_{\eta}(u_{k}-u_{l})
(u_{kt}-u_{lt})dxd\tau=\lim\limits_{\eta\rightarrow
0}\int_{0}^{t}\int_{\Omega_{r}}
\xi(\int_{0}^{u_{k}-u_{l}}\text{sgn}_{\eta}(s)ds)_{\tau}dxd\tau
|
=\lim\limits_{\eta\rightarrow 0}\int_{0}^{t}\int_{\Omega_{r}}
\xi\int_{0}^{u_{k}-u_{l}}\text{sgn}_{\eta}(s)ds\mid_{0}^{t}dx
=\int_{\Omega_{r}}\xi|u_{k}-u_{l}|dx-\int_{\Omega_{r}}\xi|u_{0k}-u_{0l}|dx.
|
(4.21)
|
By (4.20) (4.21), we have
\int_{\Omega_{2r}}\xi|u_{k}-u_{l}|dx\leq\int_{\Omega_{r}}|u_{0k}-u_{0l}|dx
+c\int_{0}^{t}\int_{\Omega_{r}}[(|\nabla
u_{k}^{m}|^{2}+\frac{1}{k})^{\frac{p-1}{2}}+(|\nabla
u_{l}^{m}|^{2}+\frac{1}{l})^{\frac{p-1}{2}}]dxd\tau
|
+\int_{0}^{t}\int_{\Omega_{r}}a(x)|u^{mq_1}_{k}|\nabla
u_{k}^{m}|^{p_1}-u^{mq_1}_{l}|\nabla u_{l}^{m}|^{p_1}|
dxd\tau
|
\int_{0}^{t}\int_{\Omega_{r}}|f_0(u^{m}_{k})\int_{\Omega}K(y)|u^{m}_{k}(y, t)|^{\beta}dy
-f_0(u^{m}_{l})\int_{\Omega}K(y)|u^{m}_{l}(y, t)|^{\beta}dy|dxd\tau.
|
(4.22)
|
By Lemma 2.2 and Lemma 3.1, if 0 < t\leq 1,
\int_{0}^{t}\int_{\Omega_{r}}a(x)|u^{mq_1}_{k}|\nabla
u_{k}^{m}|^{p_1}-u^{mq_1}_{1}|\nabla u_{1}^{m}|^{p_1}|
dxd\tau\leq c\int_{0}^{t}\int_{\Omega_{r}} t^{-\epsilon}dxd\tau,
|
which means (4.17) is true. Here
\epsilon=\max\{\frac{mNq_1}{Nm(p-1)-N+mq}+\frac{p_1(m(p-1)+m-2)}{m(p-1)-1}, \frac{(\beta+m)N}{Nm(p-1)-N+mq}\}<1
|
Now, for any given small r, if k, l are large enough, by (4.17), we have
\int_{\Omega_{2r}}|u(x, t)-u_{0}(x)|dx\leq
\int_{\Omega_{r}}|u(x, t)-u_{k}(x, t)|dx+\int_{\Omega_{2r}}|u_{0k}(x)-u_{0l}(x)|dx
|
+\int_{\Omega_{2r}}|u_{l}(x, t)-u_{0l}(x)|dx+\int_{\Omega_{2r}}|u_{0l}(x)-u_{0}(x)|dx
|
letting t\rightarrow 0, we get (1.13).
5. The uniqueness of the solutions and the proof of Theorem 1.2
As we have said in the introduction, the uniqueness of the solutions of problem (1.1)-(1.3) is not true generally. But it is not difficult to prove the following theorems.
Theorem 5.1.
Let u_1, u_2 be the two solutions of the problem (1.1)-(1.3) with the different initial values u_{01}(x), u_{02}(x) respectively. If (\frac{1}{m}+\beta-2)\setminus q_{1} < 1
and
then
\int_{\Omega}|u_{1}(x, t)-u_{2}(x, t)|dx\leq
\int_{\Omega}|u_{01}(x)-u_{02}(x)|dx, \ \forall t\geq 0.
|
(5.2)
|
Proof. Let u_1(t), u_2(t) be two solutions of equation (1.1). Let v_1=u_{1}^{m}(t), v_2= u_{2}^{m}(t). Denote w(t)=v_1^{\frac{1}{m}}(t)-u_2^{\frac{1}{m}}(t), v(t)=v_{1}(t)-v_{2}(t). Then w(t), v_1(t), v_2(t) satisfy that
w'(t)-[{\rm div}(\|\nabla v_{1}\|^{p-2}\nabla v_{1})-{\rm
div}(\|\nabla v_{2}\|^{p-2}\nabla
v_{2})+a(x)(v_{1}^{q_1}-v_{2}^{q_1})
|
=f_{0}(v_{1})\int_{\Omega}K(y)|v_{1}|^{\beta}dy-f_{0}(v_{2})\int_{\Omega}K(y)|v_{2}|^{\beta}dy.
|
(5.3)
|
For any positive integer n, let g_{n}(s) be an odd function and
g_{n}(s)=\left\{\begin{array}{cc} 1, \ \ &\ \ {\rm if}\ \
s>\frac{1}{n}, \\
n^{2}s^{2}e^{1-n^{2}s^{2}}, \ \ &\ \ {\rm if}\ \ s\leq\frac{1}{n}.
\end{array}
\right.
|
Clearly, when |s|\geq n^{-1}, g'_{n}(s)=0; when |s|\leq n^{-1}, 0\leq g'_{n}(s)=6s^{-1}.
Multiplying (5.3) with g_{n}(v_{1}-v_2) and integrating on \Omega, we have
\int_{\Omega}g_{n}(v)w'(t)dx+\int_{\Omega}[|\nabla
|v_{1}|^{p-2}|\nabla v_{1}-|\nabla |v_{2}|^{p-2}|\nabla
v_{2}]\nabla(v_{1}-v_2)g_{n}'(v)dx+\int_{\Omega}a(x)(v_{1}^{q_1}-v_{2}^{q_1})g_{n}(v)dx
|
=\int_{\Omega}g_{n}(v)
[f_{0}(v_{1})\int_{\Omega}K(y)|v_{1}|^{\beta}dy-f_{0}(v_{2})\int_{\Omega}K(y)|v_{2}|^{\beta}dy]dx.
|
(5.4)
|
Moreover,
\lim\limits_{n\rightarrow
\infty}\int_{\Omega}g_{n}(v)w'(t)dx=\frac{d}{dt}\|w(t)\|_{1},
|
\int_{\Omega}[|\nabla |v_{1}|^{p-2}|\nabla v_{1}-|\nabla
|v_{2}|^{p-2}|\nabla v_{2}]\nabla(v_{1}-v_2)g_{n}'(v)dx\geq 0,
|
\int_{\Omega}a(x)(v_{1}^{q_1}-v_{2}^{q_1})g_{n}(v)dx\geq 0,
|
|\int_{\Omega}g_{n}(v)
[f_{0}(v_{1})\int_{\Omega}K(y)|v_{1}|^{\beta}dy-f_{0}(v_2)\int_{\Omega}K(y)|v_{2}|^{\beta}dy]dx|
|
\leq|\int_{\Omega}K(y)|v_{1}|^{\beta}dy\int_{\Omega}[f_{0}(v_{1})-f_{0}(v_{2})]dx|
+c|\int_{\Omega}f_{0}(v_{2})dx||\int_{\Omega}\int_{v_{2}}^{v_{1}}s^{\beta-1}dsdy|
|
\leq
c\|w(t)\|_{1}\|v_{1}\|^{\beta}_{\beta}+c\|v_{2}\|_{1}\int_{\Omega}|\xi|^{\beta-1}|v(t)|dx,
|
where \xi\in [v_{1}, v_{2}].
So
\frac{d}{dt}\|w(t)\|_{1}\leq
c\|w(t)\|_{1}\|v_{1}\|^{\beta}_{\beta}+c\|v_{2}\|_{1}(\|v_{1}\|^{\beta}_{\beta}+\|v_{2}\|^{\beta}_{\beta}),
|
(5.5)
|
By using (1.23)-(1.24) of Theorem 1.1 to (5.5), letting n\rightarrow \infty. By Gronwall's inequality, for any given T>0, we can deduce that
\|w(t)\|_{1}\equiv 0, 0\leq t\leq T.
|
(5.6)
|
Another aim of the section is to prove the uniqueness of the viscosity solution of problem (1.1)-(1.3)
Theorem 5.2.
Suppose that a(x) and K(x) are bounded functions. If u(x, t)\in L^{\infty}(S), |\nabla u|\leq c in addition, 2\geq p_{1}\geq 1, then the viscosity solution of (1.1)-(1.3) is unique.
Proof. Let u, v be the two viscosity solutions of (1.1)-(1.3). Then there are two sequences \{u_{k}\} and \{v_{l}\}, which are the solutions of problem (1.14)-(1.2)-(1.3), such that
\lim\limits_{k\rightarrow \infty} u_{k}=u, \ \ \lim\limits_{l\rightarrow \infty}
v_{l}=v, \ {\rm a.e. in } \ S.
|
(5.7)
|
Clearly, since u(x, t), v(x, t)\in L^{\infty}(S), we may assume
\|u_{k}\|_{\infty}\leq c, \ \|v_{l}\|_{\infty}\leq c.
|
(5.8)
|
Let
w =u_k-v_l, \ \ w_{1}=u^{m}_k-v^{m}_l.
|
Then
w_{t}=\left(a_{ij}(x, t)w_{1x_{j}}\right)_{x_{i}}+b(x, t, w, \nabla
w), (x, t)\in \Omega\times (0, \infty)
|
(5.9)
|
w(x, 0)=u_{0k}(x)-v_{0l}(x), \ x\in \Omega
|
(5.10)
|
w(x, t)=0, \ (x, t)\in \partial \Omega\times (0, \infty),
|
(5.11)
|
where
a_{ij}(x, t)=\int_{0}^{1}{\left| s\nabla u_{k}^{m}+(1-s)\nabla
v_{l}^{m}\right|}^{p-2}ds\cdot\delta_{ij}
|
+\int_{0}^{1}(p-2)\left|s\nabla
u_{k}^{m}+(1-s)\nabla v_{l}^{m}\right|
^{p-4}(su_{kx_{i}}^{m}+(1-s)v_{lx_{i}}^{m})(su_{kx_{j}}^{m}+(1-s)v_{lx_{j}}^{m})ds,
|
and since p_1\geq 1, using the convexity of the function s^{p_1}, by (5.8), we have
b(x, t, w, \nabla w)= a(x)[u^{mq_1}_k|\nabla
u^{m}_k|^{p_1}-v^{mq_1}_l|\nabla v^{m}_l|^{p_1}]
|
+f_0(u^{m}_k)\int_{\Omega}K(y)|u^{m}_k(y, t)|^{\beta}dy-f_0(v^{m}_l)\int_{\Omega}K(y)|v^{m}_l(y, t)|^{\beta}dy,
|
|b(x, t, w, \nabla w)|\leq c|\nabla (u^{m}_k-v^{m}_l)|^{p_1}\leq
c|\nabla w|^{p_{1}}\leq c|\nabla w|^{2}+c.
|
By the chapter 8 of [13], we know that
\|u_{k}(x, t)-v_{l}(x, t)\|_{\infty}\leq c\|u_{0k}-v_{0l}\|_{\infty}.
|
Let k, l\rightarrow \infty, we know that the uniqueness of the viscosity solution (1.1)-(1.3) is true.
Suppose that the viscosity solution of problem (1.1)-(1.3) is unique in what follows. Then, by considering the regularized problem (1.14)-(1.2)-(1.3), we easily get the following Theorem 5.3, and Theorem 1.2 is a simple corollary of Theorem 5.3.
Theorem 5.3.
Let u be a weak solution of problem (1.1)-(1.3). If
v satisfies
v_{t}\geq\text{div}(\mid \nabla v^{m}\mid^{p-2}\nabla
v^{m})-a(x)v^{mq_1}|\nabla v^{m}|^{p_1}
|
+f_0(v^{m})\int_{\Omega}K(y)|v^{m}(y, t)|^{\beta}dy+g(x)\ \text{ in
}S=\Omega\times (0, \infty),
|
(5.12)
|
v(x, 0)\geq u_{0}(x), \ \ x\in \Omega,
|
(5.13)
|
v(x, t)=0, \ \ (x, t)\in \partial \Omega\times (0, \infty),
|
(5.14)
|
then
u(x, t)\geq v(x, t), \ \forall (x, t)\in S.
|
(5.15)
|
Now, let
v(x, t)=u_{kr}(x, t)=ru_{k}(x, r^{m(p-1)-1}t), \;\ r\in (0, 1).
|
Then
v_{t}(x, t) =\text{div}(\mid Dv^{m}\mid ^{p-2}Dv^{m})
-a(x)r^{m(p-1-q_{1}-p_{1})}v^{mq_{1}}\mid Dv^{m}\mid^{p_{1}}
|
+r^{m[p-1-\beta]}f_{0}(r^{-m}v^{m})\int_{\Omega}K(y)|v^{m}|^{\beta}dy+r^{m(p-1)}g(x), (x, t)\in
\Omega\times(0, \infty)
|
(5.16)
|
v(x, 0) =ru_{k}(x, 0), x\in \Omega,
|
(5.17)
|
v(x, t)=0, \ (x, t)\in \partial \Omega\times(0, \infty).
|
(5.18)
|
Noticing that g(x)\leq 0, f_{0}(r^{-m}v^{m})\geq f_{0}(v^{m}), and
p_{1}+q_{1}<p-1, p-1-\beta<0, \ 0<r<1,
|
which implies that
r^{m(p-1-q_{1}-p_{1})}<1, \ r^{m[p-1-\beta]}>1,
|
v_{t}(x, t)\geq\text{div}(\mid Dv^{m}\mid ^{p-2}Dv^{m})
-a(x)v^{q_{1}m}\mid Dv^{m}\mid^{p_{1}}
+f_{0}(v^{m})\int_{\Omega}K(y)|v^{m}|^{\beta}dy+g(x),
|
using the argument similar to that in the proof Lemma 3.5 of [35], we can prove
It follows that
\frac{u_{k}(x, r^{m(p-1)-1}t)-u_{k}(x, t)}{(r^{m(p-1)-1}-1)t}
|
\geq \frac{r-1}{(1-r^{m(p-1)-1})t}u_{k}(x, r^{m(p-1)-1}t).
|
Letting r\rightarrow 1, we get
u_{kt}\geq -\frac{u_{k}}{(m(p-1)-1)t}.
|
(5.19)
|
By (5.19), we can easily get Theorem 1.2.
Acknowledgement
The paper is supported by Natural Science Foundation of Fujian province in China (No: 2015J01592), supported by Science Foundation of Xiamen University of Technology.
Conflict of Interest
All authors declare no conflicts of interest in this paper.