1.Introduction
Recently, there has been a lot of interest in the systematic study of problems involving non-local operators due to their frequency in practical real-world applications, such as finance, optimization, soft thin films, stratified materials, and phase transitions. We refer the reader to see [32]. The elliptic theory for linear and quasilinear nonlocal operators has seen extensive research over the past few decades, particularly in the works of Caffarelli and collaborators [4, 5, 14]. Additionally, research on nonlocal nonlinear problems has been extensively explored in [30], we also refer to [9, 10, 11, 15, 22, 24, 25, 26, 31] on related existence results for the problems of elliptic and parabolic type involving non-local fractional Laplacian (p-Laplacian) operators.
In this paper, suppose that Ω is a bounded open domain of ℝn and T is a real positive number. We deal with the following initial boundary value problem:
(1.1)
\left\{{\matrix{{{{\partial u} \over {\partial t}} + (- \Delta)_p^su = f(x,t,u)} \hfill & {{\rm{in}}\,{Q_T} = \Omega \times (0,T),} \hfill \cr {u = 0} \hfill & {{\rm{in}}\,({{\mathbb R}^n}\backslash \Omega) \times (0,T),} \hfill \cr {u(x,0) = u(x)} \hfill & {{\rm{in}}\,\Omega,} \hfill \cr}} \right.
where 0 < s < 1 and 2 < p are real numbers, u: Ω × (0, T) → ℝm, m ∈ {0, 1, 2, . . .} is a vector-valued function and the function f satisfies the following hypothesis:
- (H1)
f : Ω × (0, T) × ℝm → ℝm is a Carathéodory function satisfying
\matrix{{\left| {f(x,t,r)} \right| \le {\alpha_0}(1 + |r{|^{q - 1}}),} \cr {{F_t}(x,t,r) \ge {\alpha_1}(- 1 - |r{|^q}),} \cr}
for all (x, t, r) ∈ Ω × (0, T) × ℝm, where α0, α1 are positive constants,
F(x,t,r) = \mathop \smallint \nolimits_0^r f(x,t,l)dl
and
{F_t} = {d \over {dt}}F
.
The fractional p-Laplacian operator
(- \Delta)_p^su
is defined as follows:
(- \Delta)_p^su(x,t) = P.V \int_{{\mathbb R}^n} {\matrix{{{{|u(x,t) - u(y,t){|^{p - 2}}(u(x,t) - u(y,t))} \over {|x - y{|^{n + ps}}}}dy,} & {x \in {{\mathbb R}^n},} \cr}}
where P.V stands for “in the principal value sense” and is a frequently used abbreviation. For more information on this operator, see [13].
Concerning the fractional Laplacian (p = 2), a famous model for anomalous diffusion is the following equation:
{{\partial u} \over {\partial t}} + {(- \Delta)^s}u = 0
, which comes asymptotically from basic random walk models (see [33, 34]). Also in [17], de Pablo et al. proposed the nonlinear anomalous diffusion equation
{{\partial u} \over {\partial t}} + {(- \Delta)^s}({u^m}) = 0
, the fractional porous medium equation with 0 < s < 1 and m > 0. We also refer to [34] for more details on this type of equation.
On the other hand, in the case p ≠ 2 and f = 0, Vázquez in [35] proved the existence and uniqueness of strong nonnegative solutions for (1.1). If u0 ∈ L2(Ω), the existence results of energy solution were studied in [29].
When it comes to the problem (1.1), the existence results are treated in several works, for example, the different issues of the existence and the regularity of energy-weak solutions to the problem same to (1.1) were investigated by Giacomoni et al. in [21]. In [1], the authors have studied the problem (1.1) with f depending only on x and t and proved the existence results with suitable regularity if (f, u0) ∈ L1 (ΩT) × L1(Ω) and has a nonnegative entropy solution if f0, u0 are nonnegative. The same author in [2] proved the asymptotic behavior result of entropy solutions when the right-hand side does not depend on time.
The idea of this work, motivated by all of the results above, is to study the existence of weak solutions to the problem (1.1) by using the Galerkin method combined with the theory of Young measures. To the best of our knowledge, the parabolic problem (1.1) has never been studied by the theory of Young measure. We suggest to the readers to consult [6, 7, 19] which treat some elliptic and parabolic systems by such a theory. In [8], the authors proved the existence of weak solutions to the elliptic case of (1.1) employing the Young measures theory and the Galerkin method.
This article is organized into four sections. In Section 2 we give some background information on fractional Sobolev spaces and a review of the Young measures theory. Later, under some assumptions, we obtain the existence of weak solutions using the Galerkin approximation and the Young measures. The final part is devoted to illustrating the feasibility of the hypotheses with an example.
2.Preliminaries and notations
In this section, we first recall some necessary results which will be used in the next section. Let 1 < p < ∞, s ∈ (0, 1), we define
p_s^*
the fractional critical exponent by:
p_s^* = \left\{{\matrix{\infty \hfill & {{\rm{if}}\,ps \ge n,} \hfill \cr {np/(n - ps)} \hfill & {{\rm{if}}\,ps < n.} \hfill \cr}} \right.
Let Ω ⊂ ℝn be an open set, QΩ = (ℝn × ℝn) \(𝒞Ω × 𝒞Ω), Qτ = Ω × (0, τ) for all τ ∈ (0, T ] and 𝒞Ω = ℝn\Ω. It is clear that Ω × Ω is strictly contained in QΩ. W is a linear space of Lebesgue measurable functions from ℝn to ℝm such that the restriction to Ω of any function u in W belongs to Lp(Ω; ℝm) and
\int\!\!\!\int_{{Q_\Omega}} {{{|u(x) - u(y){|^p}} \over {|x - y{|^{n + ps}}}}dydx < \infty.}
The space W is equipped with the norm
{\left\| u \right\|_W} = {\left\| u \right\|_{{L^p}(\Omega ;{{\mathbb R}^m})}} + {\left({\int\!\!\!\int_{{Q_\Omega}} {{{|u(x) - u(y){|^p}} \over {|x - y{|^{n + ps}}}}dydx}} \right)^{1/p}}.
Let us consider the closed linear subspace
{W_0} = \left\{{u \in W:u = 0\,{\rm{a}}.{\rm{e}}.\,{\rm{in}}\,{\cal C}\Omega} \right\}.
In W0, we may also use the norm
{\left\| u \right\|_{{W_0}}} = {\left({\int\!\!\!\int_{{Q_\Omega}} {{{|u(x) - u(y){|^p}} \over {|x - y{|^{n + ps}}}}dydx}} \right)^{1/p}}.
It is known that (W0, ∥ · ∥W0) is a uniformly convex reflexive Banach space (see [36]). The following Poincare’s inequality from [12] will be used below: there exists Cr > 0 such that
(2.1)
\matrix{{{{\left\| \phi \right\|}_{{L^r}\left({\Omega,{{\mathbb R}^m}} \right)}} \le {C_r}{{\left\| \phi \right\|}_{{W_0}}}} & {{\rm{for}}\,{\rm{all}}} & {\phi \in {W_0}} & {{\rm{and}}} & {r \in [1,p_s^*]} \cr}.
In the sequel, let
p < {n \over s}
and Ci, i = 1, 2, . . . be positive constants that vary from line to line, and are independent of the terms involved in any limit process. We note the following functional space Lp(0, T; W0), which is a separable and reflexive Banach space endowed with the norm
{\left\| u \right\|_{{L^p}\left({0,T;{W_0}} \right)}} = {\left({\mathop \smallint \nolimits_0^T \left\| u \right\|_{{W_0}}^pdt} \right)^{1/p}}.
The space
{\cal C}_0^\infty \,(\Omega ;{{\mathbb R}^m})
of infinitely differentiable functions with compact support on Ω is dense in W0.
The following embedding W0 ↪ Lr (Ω; ℝm) is compact for all
r \in [1,p_s^*)
, and continuous for all
r \in [1,p_s^*]
.
In the following, 𝒞0 (ℝm) stands for the space of continuous functions on ℝm with compact support with regards to the ∥·∥∞-norm. The space of signed Radon measures with finite mass is noted ℳ (ℝm). The corresponding duality is given by
\{\mu,\rho \rangle = \int_{{{\mathbb R}^m}} {\rho (\lambda)d\mu (\lambda)}.
Let {zj}j≥1 be a bounded sequence in L∞ (Ω; ℝm). Then there exist a subsequence {zk} ⊂ {zj} and a Borel probability measure µx on ℝm for almost every x ∈ Ω, such that for a.e. ρ ∈ 𝒞 (ℝm) we have
\rho ({z_k}) \rightharpoonup *\bar \rho
weakly in L∞(Ω), where
\bar \rho (x) = \{{\mu_x},\rho \rangle = \int_{{{\mathbb R}^m}} {\rho (\lambda)d{\mu_x}(\lambda)}
for a.e. x ∈ Ω.
Let Ω ⊂ ℝn be Lebesgue measurable (not necessarily bounded) and zj from Ω to ℝm, for j ∈ ℕ, be a sequence of Lebesgue measurable functions. Then there exist a subsequence zk and a family {µx}x∈Ω of non-negative Radon measures on ℝm, such that
- (i)
∥µx∥ℳ(ℝm) := ∫ℝm dµx(λ) ≤ 1 for almost every x ∈ Ω.
- (ii)
\rho ({z_k}) \rightharpoonup *\bar \rho
weakly in L∞(Ω) for all 𝒞0 (ℝm), where
\bar \rho (x) = \left\langle {{\mu_x},\rho} \right\rangle
.
- (iii)
If for all M > 0
(2.2)
\mathop {\lim}\limits_{N \to \infty} \mathop {\sup}\limits_{k \in {\mathbb N}} |\{x \in \Omega \cap {B_M}(0):\left| {{z_k}(x)} \right| \ge N\} | = 0,
then ∥µx∥ = 1 for a.e. x ∈ Ω, and for any measurable Ω′ ⊂ Ω we have
\rho ({z_k}) \rightharpoonup \bar \rho = \left\langle {{\mu_x},\rho} \right\rangle
weakly in L1 (Ω′) for continuous function ρ provided the sequence ρ (zk) is weakly precompact in L1 (Ω′).
3.Local existence of weak solutions
In this section, we define a weak solution to the problem (1.1) and prove the main result (Theorem 3.2 below). We start with the following definition:
Definition 3.1
A function u ∈ Lp(0, T; W0) is called a weak solution of (1.1), if
{{\partial u} \over {\partial t}} \in {L^2}({Q_T};{{\mathbb R}^m})
and
\matrix{{\int_{{Q_T}} {{{\partial u} \over {\partial t}}\phi dxdt}} \hfill \cr {+ \mathop \smallint \nolimits_0^T \int\!\!\!\int_{{Q_\Omega}} {{{|u(x,t) - u(y,t){|^{p - 2}}(u(x,t) - u(y,t))} \over {|x - y{|^{n + ps}}}}(\phi (x,t) - \phi (y,t))dxdydt}} \hfill \cr \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,{= \int_{{Q_T}} {f(x,t,u)\phi dxdt},} \hfill \cr}
holds for all
\phi \in {C^1}\,(0,T;C_0^\infty (\Omega))
.
Theorem 3.2
If u0 ∈ W0,
2 < q < {{(2 + p)p_s^* - 2p} \over {p_s^*}} < p_s^*
and (H1) is satisfied, then there exists a constant T0 > 0 such that problem (1.1) has at least one weak solution as T < T0.
Proof
The proof is divided into three assertions.
Assertion 1: Galerkin approximation
Similar to that in [27], we take a sequence
{\{{w_j}\}_{j \ge 1}} \subset C_0^\infty \,(\Omega ;\,{{\mathbb R}^m})
, such that
C_0^\infty \,(\Omega ;\,{{\mathbb R}^m}) \subset {\overline {\bigcup\nolimits_{k \ge 1} {{U_k}}}^{{C_1}(\overline \Omega)}}
, where {wj}j≥1 is an orthonormal basis in L2 (Ω; ℝm) and Uk = span {w1, . . . , wk}.
Lemma 3.3
For the function u0 ∈ W0, there exists a subsequence ξk ∈ Uk such that ξk → u0 in W0 as k → ∞.
Proof
Since u0 ∈ W0, we can find a sequence {vk} in
C_0^\infty \,(\Omega ;\,{\rm{@}}{{\rm{R}}^m})
such that vk → u0 in W0. Since
\{{v_k}\} \subset C_0^\infty \,(\Omega ;\,{{\mathbb R}^m}) \subset {\bigcup\nolimits_{M \ge 1} {\overline {{U_M}}}^{{C^1}(\overline \Omega ;{{\mathbb R}^m})}}
, there exists a sequence
\{v_k^i\} \subset \bigcup\nolimits_{M \ge 1} {{U_M}}
such that
v_k^i \to {v_k}
in
{C^1}\left({\bar \Omega ;{{\mathbb R}^m}} \right)
as i tends to ∞. For
{1 \over {{2^k}}}
, there exists ik ≥ 1 such that
{\left\| {v_k^{{i_k}} - {v_k}} \right\|_{{C^1}(\overline \Omega)}} - \le {1 \over {{2^k}}}
. Therefore
{\left\| {v_k^{{i_k}} - {u_0}} \right\|_{{W_0}}} \le {C_1}{\left\| {v_k^{{i_k}} - {v_k}} \right\|_{{C^1}\left({\overline \Omega} \right)}} + {\left\| {{v_k} - {u_0}} \right\|_{{W_0}}}.
Hence
v_k^{{i_k}} \to {u_0}
in W0 as k tends to ∞. We denote
{u_k} = v_k^{{i_k}}
. Since uk ∈ UM≥1 UM, there exists UMk such that uk ∈ UMk, without loss of generality, we assume that UM1 ⊂ UM2 as M1 ≤ M2. We suppose that M1 > 1 and define ξk as follows:
\left\{{\matrix{{{\xi_k}(x) = 0,} \hfill & {{\rm{for}}\,k = 1, \ldots,{M_1} - 1,} \hfill \cr {{\xi_k}(x) = {u_1},} \hfill & {{\rm{for}}\,k = {M_1}, \ldots,{M_2} - 1,} \hfill \cr {{\xi_k}(x) = {u_2},} \hfill & {{\rm{for}}\,k = {M_2}, \ldots,{M_3} - 1,} \hfill \cr \vdots \hfill & \vdots \hfill \cr}} \right.
Then {ξk} is the desired sequence such that ξk → u0 in W0 as k → ∞.
We define the function Rk : [0, T) × ℝk → ℝk where k is fixed:
\matrix{{{{(R(t,\varsigma))}_i}} \hfill & {= \int\!\!\!\int_{{Q_\Omega}} {{{{{\left| {\sum\nolimits_{j = 1}^k {{{({\varsigma_j}(t))}_j}{w_j}(x)} - \sum\nolimits_{j = 1}^k {{{({\varsigma_j}(t))}_j}{w_j}(y)}} \right|}^{p - 2}}} \over {x - y{|^{n + ps}}}}}} \hfill \cr {} \hfill & {\times \left({\sum\limits_{j = 1}^k {{{({\varsigma_j}(t))}_j}{w_j}(x)} - \sum\limits_{j = 1}^k {{{({\varsigma_j}(t))}_j}{w_j}(y)}} \right)({w_i}(x) - {w_i}(y))dxdy,}\hfill }
for ς ∈ ℝk and i = 1, . . . , k. The function R(t, ς) is continuous in t and ς.
Now, we shall construct the approximating solutions for (1.1) as follows:
{u_k}(x,t) = \sum\limits_{j = 1}^k {{{({b_j}(t))}_j}{w_j}(x)},
where unknown functions (b(t))j are determined by the following system of ODE:
(3.1)
\left\{{\matrix{{b^{'}(t) + {R_k}(t,b(t)) = {S_k}(t,b(t)),} \hfill & {0 < t < T,} \hfill \cr {b(0) = {\psi_k}(0),} \hfill & {} \hfill \cr}} \right.
where
{({S_k}(t,b))_i} = \int_\Omega {\matrix{{f(x,t,\sum\limits_{j = 1}^k {{b_j}{w_j}}){w_i}dx,} & {{{({\psi_k}(0))}_i} = \int_\Omega {{\xi_k}(x){w_i}dx}} \cr}},
and
\matrix{{{\xi_k}(x) \to {u_0}} & {{\rm{in}}\,{W_0}} & {{\rm{as}}\,k} \cr} \to \infty \,{\rm{where}}\,{\xi_k}(x) \in {U_k}.
Multiplying (3.1) by b(t), we get
(3.2)
b^{'}b + {R_k}(t,b)b = {S_k}(t,b)b.
According to (H1), the following inequalities hold
(3.3)
\matrix{{{S_k}(t,b)b} \hfill & {\le {\alpha_0}\int_\Omega {\left( {{{\left| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right|}^q} + \left| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right|} \right)} dx} \hfill \cr {} \hfill & {\le {\alpha_0}\int_\Omega {{{\left| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right|}^q}dx} + {\alpha_0}{C_2}\int_\Omega {{{\left| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right|}^2}dx} .} \hfill \cr}
Since
2 < q < p_s^*
, using the interpolation inequality (see [3, Theorem 2.11]) and (2.1), we get
(3.4)
\matrix{{\int_\Omega {{{\left| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right|}^q}dx}} \hfill & {\le \left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^{\theta q}\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^{p_s^*}}(\Omega ;{{\mathbb R}^m})}^{(1 - \theta)q}} \hfill \cr {} \hfill & {\le {C_{p_s^*}}\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^{\theta q}\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{W_0}}^{(1 - \theta)q},} \hfill \cr}
where θ ∈ (0, 1) satisfies
{1 \over q} = {\theta \over 2} + {{1 - \theta} \over {p_s^*}} \cdot
We observe that
(1 - \theta)q = {{p_s^*\left({q - 2} \right)} \over {p_s^* - 2}} < p_s^*
and
\lambda : = {{p\theta q} \over {p - (1 - \theta)q}} = {{2p(p_s^* - q)} \over {p_s^*(p - q + 2) - 2p}} > 2.
For any ϵ ∈ (0, 1), the Young inequality implies
(3.5)
\matrix{{\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^{\theta q}\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{W_0}}^{(1 - \theta)q}} \hfill \cr {\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, \le \epsilon \left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{W_0}}^p + C(\epsilon)\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^\lambda.} \hfill \cr}
Then, (3.4) is transformed into the following inequality
(3.6)
{\int_\Omega {\left| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right|}^q}dx \le {C_{p_s^*}}\epsilon \left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{W_0}}^p + C(\epsilon)\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^\lambda.
Plugging inequalities (3.3), (3.4) and (3.6) into (3.2), we deduce that
\matrix{{{1 \over 2}{{d|b(t){|^2}} \over {dt}} + \left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{W_0}}^p \le {C_{p_s^*}}{\alpha_0}\epsilon \left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{W_0}}^p} \hfill \cr {\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, + {\alpha_0}C(\epsilon)\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^\lambda + {\alpha_0}\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^2.} \hfill \cr}
By choosing
\epsilon = {1 \over {2{\alpha_0}{C_{p_s^*}}}}
, we get
(3.7)
\matrix{{{1 \over 2}{{d|b(t){|^2}} \over {dt}} + {1 \over 2}\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{W_0}}^p \le {\alpha_0}C(\epsilon)\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^\lambda} \hfill \cr {\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, + {\alpha_0}\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^2.} \hfill \cr}
It follows that
{{d|b(t){|^2}} \over {dt}} \le 2{C_3}\left({\left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^\lambda + \left\| {\sum\limits_{j = 1}^k {{b_j}{w_j}}} \right\|_{{L^2}(\Omega ;{{\mathbb R}^m})}^2} \right).
Denote z(t) = |b(t)|2, then
(3.8)
{{dz(t)} \over {dt}} \le 2{C_3}\left({z{{(t)}^{{\lambda \over 2}}} + z(t)} \right).
Integrating (3.8) from 0 to t, and using the property
z(0) = |b(0){|^2} = \int_\Omega {\xi_k^2(x)dx \le {C_4}},
we can conclude that
\matrix{{z(t) \le {\rm{exp}}(2{C_3}t){{\left({C_4^{1 - {\lambda \over 2}} - {\rm{exp}}({C_3}(\lambda - 2)t)} \right)}^{{2 \over {2 - \lambda}}}},} & {{\rm{as}}\,t < {{{\rm{ln(}}C_4^{1 - {\lambda \over 2}}{\rm{)}}} \over {{C_3}(\lambda - 2)}}.} \cr}
For
0 < T < {T_0} = {{\ln (C_4^{1 - {\lambda \over 2}})} \over {{C_3}(\lambda - 2)}}
, we obtain that |b(t)| ≤ C(T) ∀t ∈ [0, T ], where
C(T) = \exp (2{C_3}T){(C_4^{1 - {\lambda \over 2}} - \exp ({C_3}(\lambda - 2)T))^{{2 \over {2 - \lambda}}}}
Put
\matrix{{{{\cal J}_k} = \mathop {\max}\limits_{(t,b) \in [0,T] \times B(b(0),2C(T))} \left| {{S_k} - {R_k}(t,b)} \right|} & {{\rm{and}}} & {{\beta_k} = \min \left\{{T,{{2C(T)} \over {{{\cal J}_k}}}} \right\}} \cr},
where B(b(0), 2C(T)) is the ball of center b(0) and radius 2C(T). By [16, Peano theorem], we know that problem (3.1) has a C1 solution on [0, βk]. Let b (βk) be an initial value, then we can repeat the above process and get a C1 solution on [βk, 2βk]. Without loss of generality, we assume that
\matrix{{T = \left[ {{T \over {{\beta_k}}}} \right]{\beta_k} + \left({{T \over {{\beta_k}}}} \right){\beta_k},} & {0 < \left({{T \over {{\beta_k}}}} \right) < 1,} \cr}
where
\left[ {{T \over {{\beta_k}}}} \right]
is the integer part of
{T \over {{\beta_k}}}
and
\left({{T \over {{\beta_k}}}} \right)
is the decimal part of
{T \over {{\beta_k}}}
. We can divide [0, T ] into [(i − 1)βk, iβk], i = 1, . . . , N and [Nβk, T ] where
N = \left[ {{T \over {{\beta_k}}}} \right]
, then there exist C1 solution
b_k^i(t)
in [(i − 1)βk, iβk], i = 1, . . . , N and
b_k^{N + 1}(t)
in [Nβk, T ]. Therefore, we get a solution bk(t) ∈ C1([0, T ]) defined by
{b_k}(t) = \left\{{\matrix{{b_k^1(t),} \hfill & {{\rm{if}}\,t \in \left[ {0,{\beta_k}} \right],} \hfill \cr {b_k^2(t),} \hfill & {{\rm{if}}\,t \in \left({{\beta_k},2{\beta_k}} \right],} \hfill \cr \vdots \hfill & \vdots \hfill \cr {b_k^N(t),} \hfill & {{\rm{if}}\,t \in \left({(N - 1){\beta_k},N{\beta_k}} \right],} \hfill \cr {b_k^{N + 1}(t),} \hfill & {{\rm{if}}\,t \in \left({N{\beta_k},T} \right].} \hfill \cr}} \right.
As a result, we get the desired Galerkin approximation solution.
Assertion 2: A priori estimates
By (3.1), we have
(3.9)
\matrix{{\int_\Omega {{{\partial {u_k}} \over {\partial t}}{w_i}dx}} \hfill \cr {+ \int\!\!\!\int_{{Q_\Omega}} {{{|{u_k}(x,t) - {u_k}(y,t){|^{p - 2}}({u_k}(x,t) - {u_k}(y,t))} \over {|x - y{|^{n + ps}}}}({w_i}(x) - {w_i}(y))dxdy}} \hfill \cr {\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, = \int_\Omega {f(x,t,{u_k}){w_i}dx,}} \hfill \cr}
where 1 ≤ i ≤ k and t ∈ [0, T ] (T < T0).
Multiplying (3.9) by (b(t))i (resp. by
{d \over {dt}}{(b(t))_i}
) and summing with respect to i from 1 to k, we arrive at (integrating with respect to t from 0 to τ (τ ∈ (0, T ]))
(3.10)
\matrix{\,\,\,\,\,\,\,\,\,\,{\int_{{Q_\tau}} {{{\partial {u_k}} \over {\partial t}}{u_k}dxdt} + \int_0^\tau {\left\| {{u_k}(x,t)} \right\|_{{W_0}}^pdt} = \int_{{Q_\tau}} {f(x,t,{u_k}){u_x}dxdt},} \hfill \cr \,\,\,\,\,\,\,\,\,\,{\int_\Omega {{{\left| {{{\partial {u_k}} \over {\partial t}}} \right|}^2}dx}} \hfill \cr {+ \int\!\!\!\int_{{Q_\Omega}} {{{|{u_k}(x,t) - {u_k}(y,t){|^{p - 2}}({u_k}(x,t) - {u_k}(y,t))} \over {|x - y{|^{n + ps}}}}\left({{{\partial {u_k}(x,t)} \over {\partial t}} - {{\partial {u_k}(y,t)} \over {\partial t}}} \right)dxdy}} \hfill \cr {\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, = \int_\Omega {f(x,t,{u_k}){{\partial {u_k}} \over {\partial t}}dx}.} \hfill \cr}
According to (3.7), we have
{1 \over 2}{d \over {dt}}\int_\Omega {{u_k}{{(x,t)}^2}dx + {1 \over 2}\left\| {{u_k}(x,t)} \right\|_{{W_0}}^p} \le {C_5}\left({{{\left({\int_\Omega {{{\left| {{u_k}} \right|}^2}dx}} \right)}^{\lambda /2}} + \int_\Omega {{{\left| {{u_k}} \right|}^2}dx}} \right).
Similar to the estimation of b(t), we have
(3.11)
\matrix{{{{\int_\Omega {\left| {{u_k}(x,t)} \right|}}^2}dx \le C(T)\,} & {\forall t \in [0,T]} & {(T < {T_0})} \cr}.
Moreover
(3.12)
{\left\| {{u_k}} \right\|_{{L^p}\left({0,T;{W_0}} \right)}} \le {C_6}.
Hence, we get
(3.13)
{\left\| {{u_k}} \right\|_{{L^\infty}(0,T;{L^2}(\Omega ;{{\mathbb R}^m}))}} \le {C_7}.
According to (3.10) and (H1), we get
(3.14)
\matrix{{\int_\Omega {{{\left| {{{\partial {u_k}} \over {\partial t}}} \right|}^2}dx}} \hfill \cr {+ \int\!\!\!\int_{{Q_\Omega}} {{{|{u_k}\left({x,t} \right) - {u_k}\left({y,t} \right){|^{p - 2}}\left({{u_k}\left({x,t} \right) - {u_k}\left({y,t} \right)} \right)} \over {|x - y{|^{n + ps}}}}\left({{{\partial {u_k}\left({x,t} \right)} \over {\partial t}} - {{\partial {u_k}\left({y,t} \right)} \over {\partial t}}} \right)dxdy}} \hfill \cr {\,\,\,\,\,\,\,\,\,\,\,\,\,\, - {d \over {dt}}\int_\Omega {F\left({x,t,{u_k}} \right)dx} \le \int_\Omega {{F_t}(x,t,{u_k})dx} \le {\alpha_1}\int_\Omega {|{u_k}{|^q}dx + {\alpha_1}.}} \hfill \cr}
From the fact
\matrix{{{1 \over p}{d \over {dt}}\left\| {{u_k}\left({x,t} \right)} \right\|_{{W_0}}^p} \hfill \cr {= \int\!\!\!\int_{{Q_\Omega}} {{{|{u_k}\left({x,t} \right) - {u_k}\left({y,t} \right){|^{p - 2}}\left({{u_k}\left({x,t} \right) - {u_k}\left({y,t} \right)} \right)} \over {|x - y{|^{n + ps}}}}\left({{{\partial {u_k}\left({x,t} \right)} \over {\partial t}} - {{\partial {u_k}\left({y,t} \right)} \over {\partial t}}} \right)dxdy,}} \hfill \cr}
applied to (3.14), we deduce
(3.15)
\matrix{{\int_\Omega {{{\left| {{{\partial {u_k}} \over {\partial t}}} \right|}^2}dx} + {d \over {dt}}\left({{1 \over p}\left\| {{u_k}(x,t)} \right\|_{{W_0}}^p - \int_\Omega {F(x,t,{u_k})dx}} \right)} \hfill \cr {\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, \le {\alpha_1}\left({\int_\Omega {|{u_k}{|^q}dx + 1}} \right).} \hfill \cr}
By using the same technique in (3.5) and using (3.11) to the term in the right-hand side of (3.15), we get
(3.16)
\matrix{{\int_\Omega {{{\left| {{{\partial {u_k}} \over {\partial t}}} \right|}^2}dx}} \hfill & {+ {d \over {dt}}\left({{1 \over p}\left\| {{u_k}(x,t)} \right\|_{{W_0}}^p - \int_\Omega {F(x,t,{u_k})dx}} \right)} \hfill \cr {} \hfill & {\le {\alpha_1}\epsilon {C_{p_s^*}}\left\| {{u_k}\left({x,t} \right)} \right\|_{{W_0}}^p + {\alpha_1}C(\epsilon){{\left({\int_\Omega {{u_k}{|^2}dx}} \right)}^{\lambda /2}} + {\alpha_1}} \hfill \cr {} \hfill & {\le {C_8}\left({\left\| {{u_k}\left({x,t} \right)} \right\|_{{W_0}}^p + 1} \right).} \hfill \cr}
Integrating (3.16) with respect to t from 0 to τ (τ ∈ (0, T ]) and using the strong convergence in uk(x, 0) → u0(x) in W0, we get
(3.17)
\matrix{{\int_{{Q_\tau}} {{{\left| {{{\partial {u_k}} \over {\partial t}}} \right|}^2}dxdt + {1 \over p}\left\| {{u_k}\left({x,\tau} \right)} \right\|_{{W_0}}^p}} \hfill & {\le {C_9}\left({\int_0^\tau {\left\| {{u_k}(x,t)} \right\|_{{W_0}}^pdt + 1}} \right)} \hfill \cr {} \hfill & {+ \int_\Omega {F\left({x,\tau,{u_k}} \right)dx.}} \hfill \cr}
By assumption (H1) and interpolation inequality used in (3.5), we get
(3.18)
\int_\Omega {F(x,\tau,{u_k})dx \le {\alpha_1}\epsilon {C_{p_s^*}}\left\| {{u_k}(x,\tau)} \right\|_{{W_0}}^p + {\alpha_1}C(\epsilon){{\left({\int_\Omega {{{\left| {{u_k}} \right|}^2}dx}} \right)}^{\lambda /2}}}.
Plugging (3.18) in (3.17), we arrive at
\matrix{{\int_{{Q_\tau}} {{{\left| {{{\partial {u_k}} \over {\partial t}}} \right|}^2}dxdt + {1 \over p}\left\| {{u_k}\left({x,\tau} \right)} \right\|_{{W_0}}^p \le {C_9}\left({\int_0^\tau {\left\| {{u_k}(x,t)} \right\|_{{W_0}}^pdt + 1}} \right)}} \hfill \cr {\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, + {\alpha_1}\epsilon {C_{p_s^*}}\left\| {{u_k}(x,\tau)} \right\|_{{W_0}}^p + {\alpha_1}C(\epsilon){{\left({\int_\Omega {|{u_k}(x,\tau){|^2}dx}} \right)}^{\lambda /2}}.} \hfill \cr}
By choosing
\epsilon = {1 \over {2{\alpha_1}p{C_{p_s^*}}}}
, we get
\int_{{Q_\tau}} {{{\left| {{{\partial {u_k}} \over {\partial t}}} \right|}^2}dxdt + {1 \over {2p}}\left\| {{u_k}(x,\tau)} \right\|_{{W_0}}^p} \le {C_{10}}\left({\int_0^\tau {\left\| {{u_k}(x,\tau)} \right\|_{{W_0}}^pdt + 1}} \right).
The Gronwall inequality implies that
\int_0^\tau {\left\| {{u_k}(x,t)} \right\|_{{W_0}}^pdt \le {C_{11}}}
for each τ ∈ [0, T ]. Therefore
\int_{{Q_\tau}} {{{\left| {{{\partial {u_k}} \over {\partial t}}} \right|}^2}dxdt + {1 \over {2p}}\left\| {{u_k}(x,\tau)} \right\|_{{W_0}}^p} \le {C_{12}}.
We finally get
(3.19)
{\left\| {\frac{{\partial {u_k}}}{{\partial t}}} \right\|_{{L^2}\left( {{Q_T}} \right)}} + {\left\| {{u_k}} \right\|_{{L^\infty }\left( {0,T;{W_0}} \right)}} \le {C_{13}}.
The assumption (H1) implies that
(3.20)
{\left\| {f\left( {x,\;t,\;{u_k}} \right)} \right\|_{{L^{q'}}\left( {{Q_T}} \right)}} \le {C_{14}}.
Assertion 3: Passage to the limit
By virtue of (3.12), (3.13), (3.19), and (3.20), we get the existence of a subsequence of (uk) still denoted by (uk) such that
(3.21)
\left\{ {\begin{array}{*{20}{l}}{{u_k} \rightharpoonup *\;u\;\;{\rm{in}}\;{L^\infty }\left( {0,\;T;{L^2}\left( {\Omega ;{\mathbb{R}^m}} \right)} \right) \cap {L^\infty }\left( {0,\;T;{W_0}} \right)\;,}\\{{u_k} \rightharpoonup \;u\;\;{\rm{in}}\;{L^p}\left( {0,\;T;{W_0}} \right)\;,}\\{\frac{{\partial {u_k}}}{{\partial t}} \rightharpoonup \frac{{\partial u}}{{\partial t}}\;\;{\rm{in}}\;{L^2}\left( {{Q_T};{\mathbb{R}^m}} \right)\;,}\\{f\left( {x,\;t,\;{u_k}} \right) \rightharpoonup \;\chi \;{\rm{in}}\;{L^{q'}}\left( {{Q_T},\;{\mathbb{R}^m}} \right)\;.}\end{array}} \right.
[28, Theorem 5.1] and (3.21) imply that uk → u in Lp(0, T, L2(Ω; ℝm)) and a.e. on QT (for a subsequence), and [28, Lemma 1.3] implies that f(x, t, u) = χ. We can conclude from the continuity in (H1),
f\left( {x,\;t,\;{u_k}} \right){u_k} \to f\left( {x,\;t,\;u} \right)u\;\;\;{\rm{a}}{\rm{.e}}{\rm{.}}\;\;{\rm{in}}\;{Q_T}.
Using the Vitali Theorem, we get
\mathop {\lim }\limits_{k \to \infty } \int_{{Q_T}} {f\left( {x,\;t,\;{u_k}} \right){u_k}dxdt} = \int_{{Q_T}} {f\left( {x,\;t,\;u} \right)udxdt} .
By
\int_\Omega {{u_k}{{(x,\;T)}^2}dx \le {C_{15}}}
, we get the existence of a subsequence of (uk) still denoted by (uk) and a function û in L2 (Ω; ℝm) such that uk(x, T) → û in L2 (Ω; ℝm). Then, for any b(t) ∈ C1([0, T ]) and
\phi \in C_0^\infty \left( \Omega \right)
,
\int_Q {\frac{{\partial {u_k}}}{{\partial t}}b\phi dxdt} = \int_\Omega {{u_k}\left( {x,\;T} \right)b\left( T \right)\phi dx} - \int_\Omega {{u_k}\left( {x,\;0} \right)b\left( 0 \right)\phi dx} - \int_Q {{u_k}\frac{{\partial b}}{{\partial t}}\phi dxdt} .
Tending k to ∞, we get
\int_\Omega {\left( {\hat u - u\left( {x,\;T} \right)} \right)b\left( T \right)\phi dx} - \int_\Omega {\left( {{u_0}\left( x \right) - u\left( {x,\;0} \right)} \right)b\left( 0 \right)\phi dx} = 0.
Choosing b(T) = 1, b(0) = 0 or b(T) = 0, b(0) = 1, we have û = u(x, T) and u0(x) = u(x, 0).
As stated in the introduction, Young measure is the tool we use to prove the existence of a weak solution. To identify the weak limit, we consider the following lemma:
Lemma 3.4
Suppose that (3.12) holds. Then, the Young measure µ(x,y,t) generated by
\frac{{{u_k}\left( {x,t} \right) - {u_k}\left( {y,t} \right)}}{{{{\left| {x - y} \right|}^{\frac{n}{p} + s}}}} \in {L^p}\left( {{Q_\Omega } \times \left( {0,\;T} \right);{\mathbb{R}^m}} \right)
has the following properties:
- (a)
∥µ(x,y,t)∥ℳℝm= 1 for a.e. (x, y, t) ∈ QΩ × (0, T), i.e. µ(x,y,t) is a probability measure.
- (b)
\left\langle {{\mu _{\left( {x,y,t} \right)}},id} \right\rangle = \int_{{\mathbb{R}^m}} {\lambda d{\mu _{\left( {x,y,t} \right)}}\left( \lambda \right)}
is the weak L1-limit of
\frac{{{u_k}\left( {x,t} \right) - {u_k}\left( {y,t} \right)}}{{{{\left| {x - y} \right|}^{\frac{n}{p} + s}}}}
.
- (c)
\left\langle {{\mu _{\left( {x,y,t} \right)}},id} \right\rangle = \frac{{u\left( {x,t} \right) - u\left( {y,t} \right)}}{{{{\left| {x - y} \right|}^{\frac{n}{p} + s}}}}
for a.e. (x, y, t) ∈ QΩ × (0, T).
Proof
- (a)
For simplicity reasons, we consider
(3.22)
{v_k}\left( {x,\;y,\;t} \right) = \frac{{{u_k}\left( {x,t} \right) - {u_k}\left( {y,t} \right)}}{{{{\left| {x - y} \right|}^{\frac{n}{p} + s}}}} \in {L^p}\left( {{Q_\Omega } \times \left( {0,\;T} \right);{\mathbb{R}^m}} \right).
We know that for any M > 0, (Ω ∩ BM)2 ⊆ Ω × Ω ⫅̸ QΩ, where BM is the ball centered in 0 with radius M. Let N ∈ ℝ be such that
{Q_N} \equiv \left\{ {\left( {x,\;y,\;t} \right) \in \Omega \cap {B_M} \times \Omega \cap {B_M} \times \left( {0,\;T} \right)\;:\;\left| {{v_k}\left( {x,\;y,\;t} \right)} \right| \ge N} \right\}.
Using (3.12), we get
\begin{array}{*{35}{l}} {{\left\| {{v}_{k}} \right\|}_{{{L}^{p}}\left( {{Q}_{\Omega }}\times \left( 0,T \right);{{\mathbb{R}}^{m}} \right)}} & ={{\left( \int_{0}^{T}{\iint_{{{Q}_{\Omega }}}{\frac{{{\left| {{u}_{k}}\left( x,t \right)-{{u}_{k}}\left( y,t \right) \right|}^{p}}}{{{\left| x-y \right|}^{n+ps}}}}dxdydt} \right)}^{1/p}} \\ {} & ={{\left\| {{u}_{k}} \right\|}_{{{L}^{p}}\left( 0,T;{{W}_{0}} \right)}}\le M. \\ \end{array}
Consequently, there exists C16 ≥ 0 such that
(3.23)
{{C}_{16}}\ge \iint_{{{Q}_{\Omega }}\times \left( 0,T \right)}{{{\left| {{v}_{k}}\left( x,~y,~t \right) \right|}^{p}}dxdy}\ge \iint_{{{Q}_{N}}}{{{\left| {{v}_{k}}\left( x,~y,~t \right) \right|}^{p}}dxdy\ge {{N}^{p}}\left| {{Q}_{N}} \right|},
where |QN | is the Lebesgue measure of QN. According to (3.23), the sequence (vk) satisfies (2.2). Hence, a Young measure noted by µ(x, y, t) is generated by vk such that ∥µ(x, y, t)∥ℳ(ℝm) = 1 for a.e. (x, y, t) ∈ QΩ × (0, T).
- (b)
By (3.12), there exists a subsequence still denoted by (vk) that converges in Lp (QΩ × (0, T); ℝm). Since Lp (QΩ × (0, T); ℝm) is reflexive, then vk is weakly convergent in L1 (QΩ × (0, T); ℝm). By the third assertion in Lemma 2.4, we replace the function ρ by the identity function, to obtain
{v_k}\; \rightharpoonup \;\left\langle {{\mu _{\left( {x,y,t} \right)}},id} \right\rangle = \int_{{\mathbb{R}^m}} {\lambda d{\mu _{\left( {x,y,t} \right)}}\left( \lambda \right)\;\;\;{\rm{weakly}}\;{\rm{in}}\;} {L^1}\left( {{Q_\Omega } \times \left( {0,\;T} \right);{\mathbb{R}^m}} \right).
- (c)
According to (3.12), vk is bounded in Lp (QΩ × (0, T); ℝm), then there exists a subsequence such that vk ⇀ v in Lp (QΩ × (0, T); ℝm). Owing to the previous arguments, we get from the uniqueness of limits that
\left\langle {{\mu _{\left( {x,y,t} \right)}},id} \right\rangle = v\left( {x,\;y,\;t} \right) = \frac{{u\left( {x,t} \right) - u\left( {y,t} \right)}}{{{{\left| {x - y} \right|}^{\frac{n}{p} + s}}}}\;\;\;{\rm{for}}\;{\rm{a}}{\rm{.e}}{\rm{.}}\;\;\left( {x,\;y,\;t} \right) \in {Q_\Omega } \times \left( {0,\;T} \right).
Now, let {vk} be the sequence given in (3.22), i.e.
{v_k}\left( {x,\;y,\;t} \right) = \frac{{{u_k}\left( {x,t} \right) - {u_k}\left( {y,t} \right)}}{{{{\left| {x - y} \right|}^{\frac{{n + ps}}{p}}}}}.
The weak convergence given in Lemma 3.4 shows that
(3.24)
\begin{array}{*{20}{l}}{{{\left| {{v_k}\left( {x,\;y,\;t} \right)} \right|}^{p - 2}}{v_k}\left( {x,\;y,\;t} \right)}&{ \rightharpoonup \;\int_{{\mathbb{R}^m}} {{{\left| \lambda \right|}^{p - 2}}\lambda d{\mu _{\left( {x,y,t} \right)}}\left( \lambda \right)} }\\{}&{ = {{\left| {v\left( {x,\;y,\;t} \right)} \right|}^{p - 2}}v\left( {x,\;y,\;t} \right)}\\{}&{ = \frac{{{{\left| {u\left( {x,t} \right) - u\left( {y,t} \right)} \right|}^{p - 2}}\left( {u\left( {x,t} \right) - u\left( {y,t} \right)} \right)}}{{{{\left| {x - y} \right|}^{\frac{{n + ps}}{{p'}}}}}}}\end{array}
weakly in L1 (QΩ × (0, T); ℝm). Since the space Lp is reflexive and |vk(x, y, t)|p−2vk(x, y, t) is bounded in Lp′(QΩ × (0, T); ℝm), the sequence |vk(x, y, t)|p−2vk(x, y, t) converges in Lp′ (QΩ × (0, T); ℝm). Hence its weak Lp′-limit is also |v(x, y, t)|p−2v(x, y, t). Thus, for any φ ∈ Lp(0, T; W0) we have
\frac{{\varphi \left( {x,t} \right) - \varphi \left( {y,t} \right)}}{{{{\left| {x - y} \right|}^{\frac{{n + ps}}{p}}}}} \in {L^p}\left( {{Q_\Omega } \times \left( {0,\;T} \right);{\mathbb{R}^m}} \right).
According to the weak limit in (3.24), we get
\begin{array}{*{35}{l}} \underset{k\to \infty }{\mathop{\lim }}\,\int_{0}^{T}{\iint_{{{Q}_{\Omega }}}{\frac{{{\left| {{u}_{k}}\left( x,t \right)-{{u}_{k}}\left( y,t \right) \right|}^{p-2}}\left( {{u}_{k}}\left( x,t \right)-{{u}_{k}}\left( y,t \right) \right)}{{{\left| x-y \right|}^{n+ps}}}\left( \varphi \left( x,~t \right)-\varphi \left( y,~t \right) \right)dxdydt}} \\ =\int_{0}^{T}{\iint_{{{Q}_{\Omega }}}{\frac{{{\left| u\left( x,t \right)-u\left( y,t \right) \right|}^{p-2}}\left( u\left( x,t \right)-u\left( y,t \right) \right)}{{{\left| x-y \right|}^{n+ps}}}\left( \varphi \left( x,~t \right)-\varphi \left( y,~t \right) \right)dxdydt}} \\ \end{array}
for every φ ∈ Lp(0, T; W0).
From (3.9), for ϕ ∈ C1 (0, T; UM), M ≤ k, we have
\begin{array}{*{35}{l}} \int_{{{Q}_{T}}}{\frac{\partial {{u}_{k}}}{\partial t}\phi dxdt} \\ +\ \int_{0}^{T}{\iint_{{{Q}_{\Omega }}}{\frac{{{\left| {{u}_{k}}\left( x,t \right)-{{u}_{k}}\left( y,t \right) \right|}^{p-2}}\left( {{u}_{k}}\left( x,t \right)-{{u}_{k}}\left( y,t \right) \right)}{{{\left| x-y \right|}^{n+ps}}}\left( \phi \left( x,~t \right)-\phi \left( y,~t \right) \right)dxdydt}} \\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ =\int_{{{Q}_{T}}}{f\left( x,~t,~{{u}_{k}} \right)\phi dxdt}. \\ \end{array}
For k tending to ∞, it follows from the above results, that
(3.25)
\begin{array}{*{35}{l}} \int_{{{Q}_{T}}}{\frac{\partial u}{\partial t}\phi dxdt} \\ +\ \int_{0}^{T}{\iint_{{{Q}_{\Omega }}}{\frac{{{\left| u\left( x,t \right)-u\left( y,t \right) \right|}^{p-2}}\left( u\left( x,t \right)-u\left( y,t \right) \right)}{{{\left| x-y \right|}^{n+ps}}}\left( \phi \left( x,~t \right)-\phi \left( y,~t \right) \right)dxdydt}} \\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ =\int_{{{Q}_{T}}}{f\left( x,~t,~u \right)\phi dxdt}. \\ \end{array}
for all
\phi \in {C^1}\left( {0,\;T;\bigcup\limits_{M \ge 1} {{U_M}} } \right)
. Letting M goes to infnity, consequently, (3.25) holds for all
\phi \in {C^1}\left( {0,\;T;C_0^\infty \left( \Omega \right)} \right)
.
4.An example
We consider the following problem
\left\{ {\begin{array}{*{20}{l}}{\frac{{\partial u}}{{\partial t}} + ( - \Delta )_p^su = a\left( {x,\;t} \right)|u{|^{q - 2}}u}&{{\rm{in}}\;{Q_T} = \Omega \times \left( {0,\;T} \right)\;,}\\{u = 0}&{{\rm{in}}\;{\mathcal{C}}\Omega \times \left( {0,\;T} \right)\;,}\\{u\left( {x,\;0} \right) = {u_0}\left( x \right)}&{{\rm{in}}\;\;\Omega ,}\end{array}} \right.
comparing it with problem (1.1) where f(x, t, u) = a(x, t) |u|q−2u,
F\left( {x,\;t,\;u} \right) = \frac{{a\left( {x,t} \right)}}{q}{\left| u \right|^q}
, and Ft(x, t, u) ⩾ C(−|r|q − 1). If
2 < q < p_s^*
, then by Theorem 3.2, there exists a constant T0 > 0 such tha the problem (1.1) has a weak solutions as T < T0.