In this paper, we address the existence, uniqueness, decay estimates, and the large-time behavior of the Radon measure-valued solutions for a class of nonlinear strongly degenerate parabolic equations involving a source term under Neumann boundary conditions with bounded Radon measure as initial data.
{ut=Δψ(u)+h(t)f(x,t) in Ω×(0,T),∂ψ(u)∂η=g(u) on ∂Ω×(0,T),u(x,0)=u0(x) in Ω,
where T>0, Ω⊂RN(N≥2) is an open bounded domain with smooth boundary ∂Ω, η is an outward normal vector on ∂Ω. The initial value data u0 is a nonnegative bounded Radon measure on Ω, the function f is a solution of the linear inhomogeneous heat equation under Neumann boundary conditions with measure data, and the functions ψ, g and h satisfy the suitable assumptions.
Citation: Quincy Stévène Nkombo, Fengquan Li, Christian Tathy. Stability properties of Radon measure-valued solutions for a class of nonlinear parabolic equations under Neumann boundary conditions[J]. AIMS Mathematics, 2021, 6(11): 12182-12224. doi: 10.3934/math.2021707
[1] | Andrey Muravnik . Wiener Tauberian theorem and half-space problems for parabolic and elliptic equations. AIMS Mathematics, 2024, 9(4): 8174-8191. doi: 10.3934/math.2024397 |
[2] | Zhanwei Gou, Jincheng Shi . Blow-up phenomena and global existence for nonlinear parabolic problems under nonlinear boundary conditions. AIMS Mathematics, 2023, 8(5): 11822-11836. doi: 10.3934/math.2023598 |
[3] | Apassara Suechoei, Parinya Sa Ngiamsunthorn . Extremal solutions of φ−Caputo fractional evolution equations involving integral kernels. AIMS Mathematics, 2021, 6(5): 4734-4757. doi: 10.3934/math.2021278 |
[4] | Yanyan Cui, Chaojun Wang . Dirichlet and Neumann boundary value problems for bi-polyanalytic functions on the bicylinder. AIMS Mathematics, 2025, 10(3): 4792-4818. doi: 10.3934/math.2025220 |
[5] | Hao-Yue Liu, Wei Zhang . Neumann gradient estimate for nonlinear heat equation under integral Ricci curvature bounds. AIMS Mathematics, 2024, 9(2): 3881-3894. doi: 10.3934/math.2024191 |
[6] | Nataliya Gol’dman . Nonlinear boundary value problems for a parabolic equation with an unknown source function. AIMS Mathematics, 2019, 4(5): 1508-1522. doi: 10.3934/math.2019.5.1508 |
[7] | Yonghui Zou, Xin Xu, An Gao . Local well-posedness to the thermal boundary layer equations in Sobolev space. AIMS Mathematics, 2023, 8(4): 9933-9964. doi: 10.3934/math.2023503 |
[8] | Hongmei Li . Convergence of smooth solutions to parabolic equations with an oblique derivative boundary condition. AIMS Mathematics, 2024, 9(2): 2824-2853. doi: 10.3934/math.2024140 |
[9] | Heping Jiang . Complex dynamics induced by harvesting rate and delay in a diffusive Leslie-Gower predator-prey model. AIMS Mathematics, 2023, 8(9): 20718-20730. doi: 10.3934/math.20231056 |
[10] | Fan Wan, Xiping Liu, Mei Jia . Ulam-Hyers stability for conformable fractional integro-differential impulsive equations with the antiperiodic boundary conditions. AIMS Mathematics, 2022, 7(4): 6066-6083. doi: 10.3934/math.2022338 |
In this paper, we address the existence, uniqueness, decay estimates, and the large-time behavior of the Radon measure-valued solutions for a class of nonlinear strongly degenerate parabolic equations involving a source term under Neumann boundary conditions with bounded Radon measure as initial data.
{ut=Δψ(u)+h(t)f(x,t) in Ω×(0,T),∂ψ(u)∂η=g(u) on ∂Ω×(0,T),u(x,0)=u0(x) in Ω,
where T>0, Ω⊂RN(N≥2) is an open bounded domain with smooth boundary ∂Ω, η is an outward normal vector on ∂Ω. The initial value data u0 is a nonnegative bounded Radon measure on Ω, the function f is a solution of the linear inhomogeneous heat equation under Neumann boundary conditions with measure data, and the functions ψ, g and h satisfy the suitable assumptions.
In this paper, we study the existence, uniqueness, decay estimates, and the large-time behavior of the solutions for a class of the nonlinear strongly degenerate parabolic equations involving the linear inhomogeneous heat equation solution as a source term under Neumann boundary conditions with bounded Radon measure as initial data. This problem is described as follows:
{ut=Δψ(u)+h(t)f(x,t) in Q:=Ω×(0,T),∂ψ(u)∂η=g(u) on S:=∂Ω×(0,T),u(x,0)=u0(x) in Ω, | (P) |
where T>0, Ω⊂RN(N≥2) is an open bounded domain with smooth boundary ∂Ω, η is an unit outward normal vector. The initial value data u0 is a nonnegative bounded Radon measure on Ω. The functions ψ and g fulfill the following assumptions
{(i)ψ∈L∞(R+)∩C2(R+), ψ(0)=0, ψ′>0 in R+,(ii)ψ′,ψ″∈L∞(R+) and ψ′(s)→0ass→+∞,(iii)ψ(s)→γass→+∞,(iv)∣ψ″∣ψ′≤κinR∗+,for someκ∈R∗+, | (I) |
and
{(i)g∈L∞(R+)∩C1(R+), g′<0 in R+ and g>0 in R+,(ii)g′∈L∞(R+)andg(s)→0ass→+∞, | (A) |
where R+≡[0,+∞), R∗+≡(0,+∞) and γ∈R∗+. By ψ′ and ψ″ we denote the first and second derivatives of the function ψ. The assumption (I)-(iii) stem from (I)-(i), hence we extend the function ψ in [0,+∞] defining ψ(+∞)=γ.
The typical example of the functions ψ and g are given
ψ(s)=γ[1−e1−(1+s)m]andg(s)=e1−(1+s)m. | (1.1) |
where 0<m≤1.
The function h satisfies the following hypothesis
h∈C1(R+)∩L1(R+),h(0)=0,h′>0inR+. | (J) |
The function f is a solution of the linear inhomogeneous heat equation under Neumann boundary conditions with measure data
{ft=Δf+μ in Q:=Ω×(0,T),∂f∂η=g(f) on S:=∂Ω×(0,T),f(x,0)=u0(x) in Ω, | (H) |
where μ is a nonnegative bounded Radon measure on Q and g fulfills the assumption (A).
Throughout this paper, we consider solutions of the problem (P) as maps from (0, T) to the cone of nonnegative finite Radon measure on Ω, which satisfy (P) in the following sense: For a suitable class of test functions ξ there holds
∫T0⟨ur(⋅,t),ξt(⋅,t)⟩Ωdt+∫T0h(t)⟨f(⋅,t),ξ(⋅,t)⟩Ωdt+⟨u0,ξ(⋅,0)⟩+ |
+∫T0⟨g(ur(⋅,t)),ξ⟩∂Ωdt=∫T0⟨∇ψ(ur)(⋅,t),∇ξ(⋅,t)⟩Ωdt | (1.2) |
(see Definition 2.1). Here the measure u(⋅,t) is defined for almost every t∈(0,T), ur∈L1(Q).
The type of the problem (P) has been intensively studied by many authors for instance (see [5,18,19,20,27,28,30]) few to mention. For the general form of the problem (P), we consider the following problem studied in [18],
{ut=div(∇ϕ(x,t,u)+h(x,t,u))+F(x,t,u) in ΩT,(∇ϕ(x,t,u)+h(x,t,u))⋅η=r(x,t,u) on ∑T,u(x,0)=u0 in (∂Ω∖∑)T∪¯Ω×{0}, | (A.1) |
where ΩT=Ω×(0,T), ∑T=∑×(0,T), (∂Ω∖∑)T=(∂Ω∖∑)×(0,T) with ∑ is a relative open subset of ∂Ω, ¯∑ and ∂Ω∖∑ are C2 surface with boundary which meet in C2 manifold dimension N−2 and 0≤u0∈L∞((∂Ω∖∑)T∪¯Ω×{0}). The author in [18], proved the local existence, uniqueness and the blow-up at the finite time of the degenerate parabolic equations (A.1). Furthermore, the existence and regularity of the solutions to the quasilinear parabolic systems under nonlinear boundary conditions is discussed in detail by the studies [28,29]
{ut+A(t,u)u=F(t,u) in Ω×(s,T),β(t,u)u=r(t,u) in ∂Ω×(s,T],u(s)=u0 on Ω, | (A.2) |
where s<t≤T and u0∈Wτ,p(Ω,RN)(τ∈[0,∞)) and the definition of the operators A(t,u)u and β(t,u)u are in [28]. Similarly, studies in [19,20] showed the existence and regularity of the degenerate parabolic equations with nonlinear boundary conditions and u0∈L2(Ω) as an initial datum. Thus, we point out that the difference between the previous works (A.1), (A.2) and our work is on the following points; firstly, the initial value u0∈M+(Ω) (the nonnegative bounded Radon measure on Ω), secondly, the assumptions of the functions ψ, g given by (I) and (A). Finally, the source term f is a solution to the linear inhomogeneous heat equation under Neumann boundary conditions with measure data.
Furthermore, the study of the degenerate parabolic problem with forcing term has been intensively investigated by many authors (see [31,32,33]). In particular, [31] deals with existence solutions in the sense distributions of the nonlinear inhomogeneous porous medium type equations
ut−divA(x,t,u,Du)=μinQ:=Ω×(0,T) | (A.3) |
where μ is a nonnegative Radon measure on Q with μ(Q)<∞ and μ|RN+1∖Q=0. In last decade, some authors studied the existence, uniqueness and qualitative properties of the Radon measure-valued solutions to the nonlinear parabolic equations under zero Dirichlet or zero Neumann boundary conditions with bounded Radon measure as initial data (e.g. [1,6,7,9,10,11,12,13,15,25] and references therein). Specially, [6] discuss the existence, uniqueness and the regularity of the Radon measure-valued solutions for a class of nonlinear degenerate parabolic equations
{ut=Δθ(u) in Q,θ(u)=0 on S,u0(x,0)=u0 on Ω, | (A.4) |
where u0∈M+(Ω) and the function θ fulfills the assumptions expressed in [6]. The difference between the abovementioned studies and the problem (P) is the presence of the nonzero-Neumann boundary conditions and the source term which is a solution of the linear inhomogeneous heat equations under Neumann boundary conditions with measure data.
In general, the study of the partial differential equations through numerical methods is investigated by several authors (e.g. [47,48,49,50]). In particular, there are some authors who deal with the computation of the measure-valued solutions of the incompressible or compressible Euler equations (see [47,48]). Mostly, the authors employ the numerical experiment corresponding to initial data of the partial differential equations and prove that the resulting approximation converge to a weak solution. For instance, in [50], the authors study numerical experiment to prove that the convergence of the solution to the nonlinear degenerate parabolic equations is measure-valued. Similarly, [49] employs the numerical method to show that the resulting approximation of a non-coercive elliptic equations with measure data converges to a weak solution. Hence, the numerical experiments represent the straightforward application of the theoretical study of the type of the problem (P).
To address the large-time behavior of the Radon measure-valued solutions of the problem (P), we construct the steady-state problem as a nonlinear strongly degenerate elliptic equations given as follows
{−Δψ(U)+U=u0 in Ω,∂ψ(U)∂η=g(U) on ∂Ω, | (E) |
where u0∈M+(Ω) and the function ψ and g satisfy the hypotheses (I) and (A) respectively.
We consider solutions of the problem (E) as maps from Ω to the cone of nonnegative bounded Radon measure on Ω which satisfies (E) in the following sense: For a suitable class of test function φ, there holds
∫Ω∇ψ(Ur)∇φdx+∫ΩUφdx=∫Ωφdu0(x)+∫∂Ωg(Ur)φdH(x) |
(see Definition 2.6), where Ur∈L1(Ω) denotes the density of the absolutely continuous part of U with respect to the Lebesgue measure.
The nonlinear elliptic equations under Neumann boundary conditions with absorption term and a source term has been intensively studied by several authors [26,34,38,39,40]. In these studies, the authors dealt with the existence, uniqueness and regularity of the solutions. Furthermore, in [34], the following problem is considered
{LU+B(U)=f2 in Ω,∂U∂η+C(U)=g2 on ∂Ω, | (A.5) |
where B(U)∈L1(Ω), C(U)∈L1(∂Ω), f2∈L1(Ω), g2∈L1(∂Ω) and the expression of the differential operator L in [34, Section 2]. The authors proved the existence, uniqueness and regularity of the solutions U∈W1,1(Ω) to the problem (A.5) (see [34, Section 4, Theorem 22 and Corollary 21). The difference between the previous studies mentioned above and (A.5) is that we study the nonlinear strongly degenerate elliptic equations and the solutions obtained are Radon measure-valued. However, the existence, uniqueness, and regularity of the Radon measure-valued solutions of the quasilinear degenerate elliptic equations under zero Dirichlet boundary conditions are discussed in detail [13] by considering the following problem
{−div(A(x,U)∇U)+U(x)=¯μ in Ω,U(x)=0 on ∂Ω, | (A.6) |
where ¯μ∈M(Ω) and A(x,U) satisfies the hypothesis in [13]. In this case, the difference between the problem (E) and (A.4) is a boundary conditions with the assumptions on ψ.
In this paper, we study a class of nonlinear parabolic problems involving a forcing term and initial data is a nonnegative Radon measure. In the recent years, there are different papers that investigate these kind of problems in the setting in which the solution is a Radon measure for positive time. This type of study was done for parabolic and hyperbolic equations. One of the main tool is to search a solution by an approximation of the initial data and then try to pass to the limit in a very weak topology. The innovative part of this work is mainly the study of the large time behavior of the solutions. In my opinion, it is essential to highlight that the explicit examples of equations study in this work have not already been dealt with in literature and the novelties of the techniques that they introduced in the work. Finally, the study of the asymptotic behavior is a novelty.
The main difficulty to study the problem (P) is due to the presence of the forcing term which depends on the property solutions of the inhomogeneous heat equation (H).
The main motivation of this study comes from the desire to deal with parabolic equations in which the forcing term can be either Radon measure or Lp(Q)(1≤p<∞) functions. Whence, the idea to consider the linear inhomogeneous heat equation solution with measure data as a forcing term.
To deal with the existence and the uniqueness of the weak solutions to the problem (P), we use the definition of the Radon measure-valued solutions of the parabolic equations and the natural approximation method. In particular, to show the uniqueness of the problem (P), we will distinguish two cases for the forcing term f, either the function is in L2((0,T),H1(Ω)) or the Radon measure on Q. Notice that when the linear inhomogeneous heat equation (H) does not admit an unique solution, the problem (P) has no unique solution as well.
Furthermore, we prove the necessary and sufficient condition between measure data and capacity in order to deal with the existence of the weak solutions to the problem (P).
To establish the decay estimates of the Radon measure-valued solutions to the problem (P), we construct the suitable function and we use it as a test function in the approximation of the problem (P). Then we easily infer the decay estimates after the use of some measure properties.
To address the large-time behavior of the Radon measure-valued solutions of the problem (P), we first show that the problem (E) has a Radon measure-valued solutions in Ω.
To the best of our knowledge no existing result of decay estimates and large-time behavior of Radon measure-valued solutions obtained as limit of the approximation of the problem (P) are known in the literature. Hence, this interesting case will be discussed in this paper. This paper is organized as follows: In the next section, we state the main results, while in Section 3, we present important preliminaries. In Section 4, we study the existence and uniqueness of the heat equation (H). Finally, we prove the main results in the Sections 5-8.
To study the weak solution of the problem (P), we refer to the following definition.
Definition 2.1. For any u0∈M+(Ω) and μ∈M+(Q), a measure u is called a weak solution of problem (P), if u∈M+(Q) such that
(i) u∈L∞((0,T),M+(Ω)),
(ii) ψ(ur)∈L2((0,T),H1(Ω)),
(iii) g(ur)∈L1(S),
(iv) for every ξ∈C1((0,T),C1(Ω)), ξ(⋅,T)=0 in Ω, u satisfies the identity
∫T0⟨u(⋅,t),ξt(⋅,t)⟩Ωdt+∫T0h(t)⟨f(⋅,t),ξ(⋅,t)⟩Ωdt+⟨u0,ξ(⋅,0)⟩Ω+ |
+∫T0⟨g(ur(⋅,t)),ξ⟩∂Ωdt=∫T0⟨∇ψ(ur),∇ξ⟩Ωdxdt | (2.1) |
where ur is the nonnegative density of the absolutely continuous part of Radon-measure with respect to the Lebesgue measure such that ur∈L∞((0,T),L1(Ω)) and the function f is the solution of the problem (H).
Throughout this paper, we assume that Ω is a strong C1,1 open subset of RN. Also, we assume that there exists a finite open cover (Bj) such that the set Ω∩Bj epigraph of a C1,1 function ζ:RN−1→R that is
Ω∩Bj={x∈Bj/xN>ζ(¯x)}and∂Ω∩Bj={x∈Bj/xN=ζ(¯x)} |
where x=(¯x,xN), the local coordinates with ¯x=(x1,x2,…,xN−1). We denote ϑ={¯x,x∈Ω∩Bj}⊆RN−1, the projection of Ω∩Bj onto the (N−1) first components, and ϑς={¯x,x∈supp(ς)∩Ω}.
If a function ϕ is defined on S, we denote ϕS the function defined on (Bj∩Q)×[0,T] by ξS(x,t)=ξ(¯x,ζ(¯x),t). Notice that the restriction of ξS to [0,T)×ϑ.
The next definition of the trace is corresponding to the problem (P) adapts to the context of [36, Theorem 2.1].
Definition 2.2 Let F∈[L2(Q)]N+1 be such that divF is a bounded Radon measure on Q. Then there exists a linear functional Tη on W12,2(S)∩C(S) which represents the normal traces F⋅ν on S in the sense that the following Gauss-green formula holds:
(i) For all ξ∈C∞c(¯Q),
⟨Tν,ξ⟩=∫QξdivF+∫Q∇ξ⋅F |
where ⟨Tν,ξ⟩ depends only on ξS.
(ii) If (Bj,ς,f) is an above subsequence localization near boundary, then for all ξ∈C∞c([0,T)ׯΩ) there holds
⟨Tν,ξ⟩=−lims→01s{∫Ts∫ϑ∫ζ(¯x)+sζ(¯x)F⋅(−∇ζ(¯x)10)ξσdxNd¯xdt}−lims→01s∫s0∫ΩF⋅(001)ξσdxdt | (2.2) |
where the divergence of the fields,
F(x,t)=(u(x,t)∇ψ(ur(x,t))) |
is a bounded Radon measure on Q.
The following result states the existence of the trace of the boundary condition to the problem (P).
Lemma 2.1 Let Ω is a strong C1,1 open subset of RN. Then there exists an unique trace Tη:W1,1(Ω)→L1(∂Ω) such that
⟨Tη,ξ⟩=∫Sg(ur)ξdH(x)dt | (2.3) |
where the function g(ur)∈L1(S) and ξ∈C∞c([0,T)ׯΩ).
To prove the uniqueness of the solution to the problem (P), we define the notion very weak solution of the problem (P) as follows.
Definition 2.3. For any μ∈M+(Q) and u0∈M+d,2(Ω), a measure u is called a very weak solution to problem (P) if u∈L∞((0,T),M+(Ω)) such that
∫T0⟨u(⋅,t),ξt(⋅,t)⟩Ωdt+∫Qψ(ur)Δξdxdt+∫Qh(t)f(x,t)ξdxdt+∫Sg(u)ξdHdt+⟨u0,ξ(0)⟩Ω=0 | (2.4) |
for every ξ∈C2,1(¯Q), which vanishes on ∂Ω×[0,T], for t=T.
To prove the uniqueness of the problem (P) when f lies in M+(Q), we consider the following every weak solution gives below:
Definition 2.4 Let u0∈M+d,2(Ω) and μ∈M+(Q) such that
u0=f0−divG0,f0∈L1(Ω)andG0∈[L2(Ω)]N. |
A function u is called a very weak solutions obtained as limit of approximation, if
un∗⇀uinM+(Q) | (2.5) |
where {un}⊆L∞(Q)∩L2((0,T),H1(Ω)) is the sequences of weak solutions to problem (Pn) satisfies
{u0n=f0n−F0n∈C∞c(Ω),F0n→divG0in(H1(Ω))∗,f0n→f0inL1(Ω). | (2.6) |
We denote (H1(Ω))∗ the dual space of H1(Ω) and the embedding H1(Ω)⊂L2(Ω)⊂(H1(Ω))∗ holds.
Definition 2.5 Let u0∈M+d,2(Ω) and μ∈M+d,2(Q) such that
u0=f0−divG0,f0∈L1(Ω)andG0∈[L2(Ω)]N. |
μ=f1−divG+φt,f1∈L1(Q),G∈[L2(Q)]Nandφ∈L2((0,T),H1(Ω)). |
A measure f is called a very weak solutions obtained as limit of approximation, if
fn∗⇀finM+(Q) | (2.7) |
where {un} and {fn}⊆L∞(Q)∩L2((0,T),H1(Ω)) are the sequences of weak solutions to problem (Pn) and (Hn) respectively satisfy
{μn=f1n−Fn+gnt∈C∞c(Q),u0n=f0n−F0n∈C∞c(Ω),f1n→f1inL1(Q),Fn→divGinL2((0,T),(H1(Ω))∗),φn→φinL2((0,T),H1(Ω)),F0n→divG0in(H1(Ω))∗,f0n→f0inL1(Ω). | (2.8) |
Then, the function u is very weak solutions of the problem (P) obtained as limit of approximation if the function f is a very weak solutions of the problem (H) obtained as limit of approximation.
Notice that
un∗⇀uinM+(Q),μn∗⇀μinM+(Q)andu0n∗⇀u0inM+(Ω). |
M+d,2(Ω) denotes the set of nonnegative measures on Ω which are diffuse with respect to the Newtonian capacity and the definition of the diffuse measure with respect to the parabolic capacity M+d,2(Q) will be recalled in the Section 3.
Before dealing with the existence of the problem (P), we first prove the existence and uniqueness of the solutions to the problem (H) given by the following result.
Theorem 2.1. Assume that u0∈M+(Ω) and μ∈M+(Q) hold.
(i) Then, there exists a nonnegative Radon measure-valued solution to the problem (H) in the space L∞((0,T),M+(Ω)) such that
f(x,t)=∫ΩGN(x−y,t)du0(y)+∫t0∫ΩGN(x−y,t−σ)dμ(y,σ)+∫t0∫∂ΩGN(x−y,t−σ)g(f(y,σ))dH(y)dσ | (2.9) |
for almost every t∈(0,T). Furthermore, the Radon measure-valued solution f satisfies the following estimate
∥f(⋅,t)∥M+(Ω)≤eCt(∥μ∥M+(Q)+∥u0∥M+(Ω)) | (2.10) |
for any C=C(T) a positive constant.
(ii) Suppose that u0∈M+d,2(Ω), μ∈M+d,2(Q) and g(f)=¯K almost everywhere on S (¯K is a positive constant) are satisfied. Then, the nonnegative weak Radon measure-valued solution to the problem (H) obtained as limit of the approximation is unique in L∞((0,T),M+(Ω)).
We denote by GN(x−y,t−s) as the Green function of the heat equation under homogeneous Neumann boundary conditions. By [4], the Green function satisfies the following properties
GN(x−y,t−s)≥0,x,y∈Ω,0≤s<t<T, | (2.11) |
∫ΩGN(x−y,t−s)dx=1,y∈Ω,0≤s<t<T. | (2.12) |
There exist two positive constants τ1 and τ2 such that
|GN(x−y,t−s)−1∣Ω∣|≤τ1e−τ2(t−s),x,y∈Ω,1+s<t. | (2.13) |
limt→s∫ΩGN(x−y,t−s)ϕ(y)dy=ϕ(x) | (2.14) |
for any ϕ∈Cc(Ω) and ∣Ω∣ is a Lebesgue measure of the set Ω.
Remark 2.1 (i) For any test function ξ∈C1((0,T),C1(Ω)) such that ξ(⋅,T)=0 in Ω and ∂ξ∂η=0 on S, the inner product ⟨f(⋅,t),ξ(⋅,t)⟩Ω in (2.1) is given by the following expression
⟨f(⋅,t),ξ(⋅,t)⟩Ω=∫Ω∫ΩG(x−y,t)ξ(y,0)du0(y)dx+ |
+∫Ω∫t0∫ΩGN(x−y,t−σ)(fξσ−2∇f∇ξ−fΔξ)dydσdx+ |
+∫Ω∫t0∫ΩGN(x−y,t−σ)ξ(y,σ)dμ(y,σ)dx+ |
+∫Ω∫t0∫∂ΩGN(x−y,t−σ)ξ(y,σ)g(f(y,σ))dH(y)dσdx | (2.15) |
where ξσ is a first derivative order of ξ with respect σ.
(ii) By the regularity properties of the Green function GN(x−y,t−σ) in [42], the solution of the problem (H) given by (2.9), f∈L2((0,T),H1(Ω)).
(iii) By virtue of the assumptions (J), (2.11) and (2.12), the term h(t)f(x,t) is well-defined at t=0. Indeed, the function t↦∫ΩGN(x−y,t−σ)h(σ)dμ(y,σ), t↦∫∂ΩGN(x−y,t−σ)h(σ)g(f(y,σ))dH(y) and t↦∫ΩGN(x−y,t−σ)f(y,σ)h′(σ)dy are continuous in R+. Then there holds
limt→0+h(t)f(x,t)=limt→0+∫t0∫ΩGN(x−y,t−σ)f(y,σ)h′(σ)dydσ+ |
+limt→0+∫t0∫∂ΩGN(x−y,t−σ)h(σ)g(f)dH(y)dσ+limt→0+∫t0∫ΩGN(x−y,t−σ)h(σ)dμ(y,σ)=0. |
Hence we extend the function h(t)f(x,t) in [0, T] defining h(0)f(x,0)=0. Furthermore, the presence of the function h is to well-defined the forcing term of the nonlinear parabolic problem (P).
In order to study the existence and uniqueness of the solutions to the problem (P), we give the necessary and sufficient condition on the measures μ and u0 for the existence of the weak solutions to the problem (P) with respect to the parabolic and Newtonian capacity respectively. This result is given by the following theorem.
Theorem 2.2. Suppose that the hypotheses (I), (A), μ∈M+(Q) and u0∈M+(Ω) hold. For any function h satisfying (J), there exists t∈(0,T) such that ∫t0h(σ)dσ=1 and u is a weak solution to the problem (P). Then μ and u0 are absolutely continuous measures with respect to the parabolic capacity.
Notice that Newtonian and parabolic capacity are equivalent, then μ and u0 are absolutely continuous measures with respect to C2-capacity as well.
In the next theorem, we present the result of the existence Radon measure-valued solutions to the problem (P).
Theorem 2.3 Suppose that the assumptions (I), (J), (A) μ∈M+(Q) and u0∈M+(Ω) are satisfied. Then there exists a weak solution u to problem (P) obtained as a limiting point of the sequence {un} of solutions to problems (Pn) such that for every t∈(0,T)∖H∗, there holds
∥u(⋅,t)∥M+(Ω)≤C(∥μ∥M+(Q)+∥u0∥M+(Ω)). | (2.16) |
The result of the uniqueness of the problem (P) is given by the following theorem:
Theorem 2.4 Assume that the hypotheses (I), (J) and (A), μ∈M+d,2(Q) and u0∈M+d,2(Ω) hold. Then there exists a unique very weak solution obtained as the limit of approximation u of the problem (P), if g(ur)=L almost everywhere in S, whenever L is a positive constant.
To establish the decay estimate of the solution to the problem (P), we recall two particular problems of the problem (P). Now we consider the following problem.
{vt=Δϑ(v) in Q,∂ϑ(v)∂η=g1(v) on S,v(x,0)=u0 in Ω, | (P0) |
The functions ψ and g satisfy the assumption (I) and (A) respectively and have the same properties with the functions ϑ and g1 given as follows
ϑ(s)=γ[1−1(1+s)m](m>0)andg1(s)=1(1+s)m | (2.17) |
where m>0 and s>0. Therefore, by Theorem 2.3, the problem (P0) possesses a solution in the space L∞((0,T),M+(Ω)), such that
∥v(⋅,t)∥M+(Ω)≤C∥u0∥M+(Ω) |
for almost every t∈(0,T).
Similarly, we consider the following problem
{wt=Δψ(w)+h(t)f(x,t) in Q,∂ψ(w)∂η=g(w) on S,w(x,0)=0 in Ω, | (P1) |
By Theorem 2.3, the problem (P1) admits a solution in L∞((0,T),M+(Ω)), such that
∥w(⋅,t)∥M+(Ω)≤C∥μ∥M+(Ω) |
for almost every t∈(0,T).
Now we state the decay estimates in the next theorem:
Theorem 2.5 Suppose that (I), (J), (A), μ∈M+(Q) and u0∈M+(Ω) are satisfied. The measure u is the weak solution to the problem (P). According to Theorem 2.3, v is the weak solution to the problem (P0) and w is the weak solution to the problem (P1). Then for every t∈(0,T)∖H∗ with ∣H∗∣=0, there holds
∥u(⋅,t)−v(⋅,t)∥M+(Ω)≤C(T−t)α(∥μ∥M+(Q)+∥u0∥M+(Ω)), | (2.18) |
∥u(⋅,t)−w(⋅,t)∥M+(Ω)≤C∥u0∥M+(Ω)(T−t)α, | (2.19) |
and
∥u(⋅,t)∥M+(Ω)≤Ctα(∥u0∥M+(Ω)+∥μ∥M+(Q)) | (2.20) |
for any positive constant C and α>1.
To deal with the large-time behavior of the Radon measure-valued solutions to the problem (P), we first extend (0,T) to (0,+∞), then we assume that the hypothesis
lim supt→+∞∥u(⋅,t)∥M+(Ω)≤C | (2.21) |
where C is a positive constant.
To analyze the large-time behavior of the Radon measure-valued solutions, we first study the existence of the Radon measure-valued solutions corresponding to the steady state problem (E) by considering the following definition.
Definition 2.6 Assume that the hypotheses (I), (A) and u0∈M+(Ω) are satisfied. A measure U is a solution of the problem (E), if U∈M+(Ω) such that
(i) ψ(Ur)∈W1,1(Ω),
(ii) g(Ur)∈L1(∂Ω),
(iii) for every φ∈C1(Ω), the following assertion
∫Ω∇ψ(Ur(x))∇φ(x)dx+∫ΩU(x)φ(x)dx=∫Ωφ(x)du0(x)+∫∂Ωg(U(x))φ(x)dH(x) | (2.22) |
holds true.
The existence result of the problem (E) is given by the following theorem:
Theorem 2.6 Suppose that the hypotheses (I), (A) and u0∈M+(Ω) are satisfied. Then there exists a weak solution U∈M+(Ω) of the problem (E) obtained as a limiting point of the sequence {Un} of solutions to the approximation problem (En) such that
∥U∥M+(Ω)≤C∥u0∥M+(Ω) | (2.23) |
where C>0 is a constant.
The result of the large-time behavior of the Radon measure-valued solutions of the problem (P) is given by the following theorem
Theorem 2.7. Suppose that the assumption (I), (A), (J), u0∈M+(Ω) and μ∈M+(Q). U is a Radon measure-valued solutions of the steady-state problem (E) in sense of Theorem 2.6 and u is a Radon measure-valued solutions in the sense of Theorem 2.3 such that (2.21) holds. Then there holds
u(⋅,t)→UinM+(Ω)ast→∞ | (2.24) |
In the following section, we define the truncation function for k>0 and s∈R,
Tk(s)=min{∣s∣,k}sign(s). |
To prove the main results from the previous section, we need to recall the preliminaries about capacity and measure collected in [9,10,11,12,13,14,15,16]. Likewise, we recall some important notations as follows:
For any Borel set E⊂Ω, the C2-capacity of E in Ω is defined as
C2(E)=inf{∫Ω(∣u∣2+∣∇u∣2)dx/u∈ZEΩ} |
where ZEΩ denotes the set of u which belongs to H1(Ω) such that 0≤u≤1 almost everywhere in Ω, and u=1 almost everywhere in a neighborhood E.
Let W={u∈L2((0,T),H1(Ω))andut∈L2((0,T),(H1(Ω))∗)} endowed with its natural norm ∥u∥W=∥u∥L2((0,T),H1(Ω))+∥ut∥L2((0,T),(H1(Ω))∗) a Banach space. For any open set U⊂Q, we define the parabolic capacity as
Cap(U)=inf{∥u∥W/u∈VUQ} |
where VUQ denotes the set of u belongs to W such that 0≤u≤1 almost everywhere in Q, and u=1 almost everywhere in a neighborhood U.
Let M(B) be the space of bounded Radon measures on B, and M+(B)⊂M(B) the cone of nonnegative bounded Radon measures on B. For any μ∈M(B) a bounded Radon measure on B, we set
∥μ∥M(Ω):=∣μ∣(B) |
where ∣μ∣ stands for the total variation of μ.
The duality map ⟨⋅,⋅⟩B between the space M(B) and Cc(B) is defined by
⟨μ,φ⟩B=∫Bφdμ. |
M+s(B) denotes the set of nonnegative measures singular with respect to the Lebesgue measure, namely
M+s(B):={μ∈M+(Ω)/∃ a Borel setF⊆Bsuch that∣F∣=0,μ=μ⌞B} |
we will consider either ∣.∣ the Lebesgue measure on RN or RN+1. Similarly, M+ac(B) the set of nonnegative measures absolutely continuous with respect to the Lebesgue measure, namely
M+ac(B):={μ∈M+(Ω)/μ(F)=0, for every Borel setF⊆Bsuch that∣F∣=0} |
Let M+c,2(B) be the set of nonnegative measures on B which are concentrated with respect to the Newtonian capacity
M+c,2(B):={μ∈M+(B)/∃ a Borel setF⊆B,such thatμ=μ⌞FandC(F)=0} |
M+d,2(B) denotes the set of nonnegative measures on B which are diffuse with respect to the Newtonian capacity
M+d,2(B):={μ∈M+(B)/μ(F)=0, for every Borel setF⊆Bsuch thatC(F)=0}. |
It is known that a measure ¯μd,2∈M+d,2(Ω) (resp. μd,2∈M+d,2(Q)) if there exist f0∈L1(Ω) and G0∈[L2(B)]N (resp. if μd,2∈M+d,2(Q), there exist f∈L1(Q), g∈L2((0,T),H1(Ω)) and G∈[L2(Q)]N) such that
¯μd,2=f0−divG0inD′(Ω)(resp.μd,2=f−divG+gtinD′(Q)). | (3.1) |
For any λ∈M+(B), if there exists a unique couple λd,2∈M+d,2(B), λc,2∈M+c,2(B) such that
λ=λd,2+λc,2. | (3.2) |
On the other hand, there exists a unique couple λac∈M+ac(B),λs∈M+s(B) such that
λ=λac+λs | (3.3) |
where either B=Ω, C(F)=C2(E) or B=Q, C(F)=Cap(U).
By L∞((0,T),M+(Ω)), the set of nonnegative Radon measures u∈M+(¯Q) such that for every t∈(0,T), there exists a measure u(⋅,t)∈M+(Ω) such that
(i) for every ξ∈C(¯Q) the map
t↦⟨u(⋅,t),ξ(⋅,t)⟩Ωis Lebesgue measurable |
and
⟨u(⋅,t),ξ(⋅,t)⟩Ω=∫T0⟨u(⋅,t),ξ(⋅,t)⟩Ωdt |
(ii) there exists a constant C>0 such that
esssupt∈(0,T)∥u(⋅,t)∥M+(Ω)≤C |
with the norm denotes
∥u∥L∞((0,T),M+(Ω))=esssupt∈(0,T)∥u(⋅,t)∥M+(Ω). | (3.4) |
In the literature, many authors dealt with the existence, uniqueness, blow-up at finite and infinite time, decay estimates, stability properties and asymptotic behavior of the solutions to the heat equation under Neumann boundary conditions with a source term and initial data, such as (see [2,3,4,5,42,43] and references therein). Moreover, most of the authors employed the maximum principle theorem through the monotonicity technique and semi-group method to show the existence, blow-up, stability properties and asymptotic behavior of these solutions. Meanwhile, in this section we prove the existence and uniqueness of the linear inhomogeneous heat equation (H) by using the fundamental solution of the heat equation (see [2,3,4,42]). Also, we use the definition of the Radon measure-valued solutions in [9] and some properties of the Radon measure provided in [24,44]. Moreover, we consider for every n∈N, the approximation of problem (H) such that
{fnt=Δfn+μn in Q,∂fn∂η=g(fn) on S,fn(x,0)=u0n(x) in Ω, | (Hn) |
Since u0∈M+(Ω), the approximation of the Radon measure u0 is given by [9, Lemma 4.1] such that {u0n}⊆C∞c(Ω) satisfies the following assumptions
{u0n∗⇀u0inM+Ω),u0n→u0ra.e inΩ,∥u0n∥L1(Ω)≤∥u0∥M+(Ω). | (4.1) |
Moreover μ∈M+(Q), the approximation of the Radon measure μ is given by [15] such that {μn}⊆C∞c(Q) fulfills the following hypotheses
{μn∗⇀μinM+(Q),μn→μra.e inQ,∥μn∥L1(Q)≤∥μ∥M+(Q), | (4.2) |
for every n∈N. By [21,22,43], the approximation problem (Hn) has a unique solution fn in C1((0,T),L2(Ω))∩L2((0,T),H1(Ω))∩L∞(Q).
In the next proposition, we establish the relationship between the approximation solution fn and any test function in (Pn).
Proposition 4.1. Suppose that ξ∈C1c(Q) such that ∂ξ∂η=0 on S, the test function in (Hn) and fn the approximation solution of the problem (Hn). Then, the following expression holds
fn(x,t)ξ(x,t)=∫ΩGN(x−y,t)ξ(y,0)u0n(y)dy+∫t0∫ΩGN(x−y,t−σ){fnξσ−2∇fn∇ξ−fnΔξ}dydσ+ |
+∫t0∫ΩGN(x−y,t−σ)ξ(y,σ)μn(y,σ)dydσ+∫t0∫∂ΩGN(x−y,t−σ)ξ(y,σ)g(f(y,σ))dH(y)dσ | (4.3) |
where ξσ is first-order derivative order of ξ with respect to σ.
Remark 4.2. Assume that the test function ξ=ρ∈C2c(Ω), then we obtain
fn(x,t)ρ(x)=∫ΩGN(x−y,t)ρ(y)u0n(y)dy−∫t0∫ΩGN(x−y,t−σ){2∇fn∇ρ+fnΔρ}dydσ+ |
+∫t0∫ΩGN(x−y,t−σ)ρ(y)μn(y,σ)dydσ+∫t0∫ΩGN(x−y,t−σ)ρ(y)g(f(y,σ))dH(y)dσ. | (4.4) |
On the other hand, we suppose that the test function ξ=θ(t)∈C1(0,T) then (4.3) reads
fn(x,t)θ(t)=∫ΩGN(x−y,t)θ(0)u0n(y)dy+∫t0∫ΩGN(x−y,t−σ)fn(y,σ)θ′(σ)dydσ+ |
+∫t0∫ΩGN(x−y,t−σ)θ(σ)μn(y,σ)dydσ+∫t0∫ΩGN(x−y,t−σ)θ(σ)g(f(y,σ))dH(y)dσ. | (4.5) |
Proof of Proposition 4.1. Assume that ξ∈C1((0,T),C2c(Ω)) such that ∂ξ∂η=0 on S, a test function in (Hn), then the following equation
{(fnξ)t=Δ(fnξ)+fnξt−2∇fn∇ξ−fnΔξ+μnξ in Ω×(0,T),∂(fnξ)∂η=g(fn)ξ on ∂Ω×(0,T),fn(x,0)ξ(x,0)=u0nξ(x,0) in Ω, | (Hξ) |
is well-defined. By [35, Chapter 20, Section 20.2], the problem (Hξ) admits a unique solution fnξ expressed in (4.3).
Proof of Theorem 2.1 (i) We argue this proof into two steps:
Step 1. We show that {fn(⋅,t)} is a Cauchy sequence in L1(Ω) a.e in (0,T). To attain this, we use the expression (4.3) to prove the Cauchy sequence. Indeed, for any m,n∈N there holds
fn(x,t)−fm(x,t)=∫ΩGN(x−y,t)[u0n(y)−u0m(y)]dy+ |
+∫t0∫ΩGN(x−ξ,t−σ)[μn(ξ,σ)−μm(y,σ))]dyds+ |
+∫t0∫∂ΩGN(x−y,t−σ)[g(fn(ξ,σ))−g(fm(ξ,s))]dH(y)ds. | (4.6) |
From the assumption (2.12), the Eq (4.6) yields
∫Ω∣fn(x,t)−fm(x,t)∣dx≤∫Ω∣u0n(y)−u0m(y)∣dy+∫t0∫Ω∣μn(y)−μm(y)∣dydσ |
+∫t0∫∂ΩGN(x−ξ,t−s)dH(ξ)(∫Ω∣g(fn(x,s))−g(fm(x,s))∣dx)ds. | (4.7) |
Furthermore, by using the mean value theorem, we find that there exists a function θ(x,s) which is continuous in ¯QT1 such that α1<θ(x,s)<α2, g(fn(x,s))−g(fm(x,s))=g′(θ(x,s))(fn(x,s)−fm(x,s)), where g′(θ(x,s))∈L∞(R+)(see assumption (I)-(i)) and 0<α1<α2 are constants, therefore we obtain
∫Ω∣fn(x,t)−fm(x,t)∣dx≤∫Ω∣u0n(y)−u0m(y)∣dy+∫T0∫Ω∣μn(y)−μm(y)∣dy+ |
+C(T1)∫t0∫Ω∣fn(x,σ)−fm(x,σ)∣dxdσ, | (4.8) |
whenever C(T1)=sup(ξ,σ)∈¯QT1∫∂ΩGN(x−ξ,T1)g′(θ(x,s))dH(ξ)>0. By the property (2.13) of the Green function GN(x−ξ,t−σ) of the heat equation with nonhomogeneous Neumann boundary and the fact that g′(θ(x,s))∈L∞(R+), then C(T1) is a constant depending on T1. From the Gronwall's inquality, the inequality (4.8) yields
∫Ω∣fn(x,t)−fm(x,t)∣dx≤C(T,T1)∫Ω∣u0n(y)−u0m(y)∣dy+ |
+C(T,T1)∫Ω∣μn(y,σ)−μm(y,σ)∣dydσ | (4.9) |
for a.e 0≤t<T1<T and C(T,T1)=1+C(T1)TeC(T1)T>0 a constant.
Since the sequences {u0n} and {u0m} are satisfying the assumption (4.1) and {μn} and {μm} are verifying the assumption (4.2), then by passing to the limit as n and m go to infinity, there holds
limn,m→+∞∫Ω∣fn(x,t)−fm(x,t)∣dx≤C(T,T1)lim supn→+∞∫Ω∣u0n(y)−u0r(y)∣dy+ |
+C(T,T1)lim supm→+∞∫Ω∣u0m(y)−u0r(y)∣dy+ |
+C(T,T1)lim supn→+∞∫Ω∣μn(y,σ)−μr(y,σ)∣dydσ+ |
+C(T,T1)lim supm→+∞∫Ω∣μm(y,σ)−μr(y,σ)∣dydσ≤0. | (4.10) |
Hence the sequence {fn(⋅,t)} is Cauchy in L1(Ω) for almost every t∈(0,T).
Step 2. We show that fn(⋅,t)∗⇀f(⋅,t) in M+(Ω) a.e in (0,T).
Since the function fn(x,t) is a solution of the approximation problem (Hn) and μn(x)≥0 in Q, u0n(x)≥0 in Ω, g>0 in R+, then we apply the maximum principal theorem in [22,43] and then the solution of the approximation problem (Hn) is nonnegative in ¯Q. Likewise, we assume that ξ(x,t)≡1, then we obtain
fn(x,t)=∫ΩGN(x−y,t−σ)u0n(y)dy+∫t0∫ΩGN(x−y,t−σ)μn(y,σ)dydσ+ |
+∫t0∫∂ΩGN(x−y,t−σ)g(f(y,σ))dH(y)dσ. | (4.11) |
By the assumptions (A), (2.12), (2.13), (4.1) and (4.2), we infer that
∫Ωfn(x,t)dx≤∥u0∥M+(Ω)+∥μ∥M+(Q)+C∫t0∫Ωfn(x,σ)dxdσ. | (4.12) |
By Gronwall's inequality, we deduce that
∥fn(⋅,t)∥L1(Ω)≤eCt(∥u0∥M+(Ω)+∥μ∥M+(Q)), | (4.13) |
for almost every t∈(0,T).
By Step 1, the sequence {fn(⋅,t)} is Cauchy in L1(Ω), then we infer that fn(⋅,t)→f(⋅,t) a.e in (0,T). Hereby we argue as in [9, Proposition 5.3], one proves that f(⋅,t)∈M+(Ω) and the following convergence
fn(⋅,t)∗⇀f(⋅,t)inM+(Ω) | (4.14) |
for almost every t∈(0,T) holds.
From [44, Chapter 5, Section 5.2.1, Theorem 1], the estimate (4.13) yields
∥f(⋅,t)∥M+(Ω)≤lim infn→+∞∫Ωfn(⋅,t)dx≤eCt(∥u0∥M+(Ω)+∥μ∥M+(Q)). |
The estimate (2.10) is achieved.
(ii) Now we show the uniqueness solutions to the problem (H).
To attain this, we consider f1 and f2 two every weak solutions of the problem (H) in sense of Definition 2.5 with initial data u01 and u02 respectively.
Let {f1n},{f2n}⊆L∞(Q)∩L2((0,T),H1(Ω)) be two weak solutions given by the proof (i) of Theorem 2.1. Assume that {u01n},{u02n}{μ1n},{μ2n} are approximating Radon measures in sense of Definition 2.4 and f1n, f2n in (4.11) hold. Since we have assumed that g(f)=¯K almost everywhere in S, thus g(f1n)=g(f2n)=¯K on S. For any ξ∈C1c(Q) such that ∂ξ∂η=0 on S, there holds
∫T0[f1n(y,t)−f2n(y,t)]ξ(y,t)dt=∫Q[f1n(x,t)−f2n(x,t)]δy(x)ξ(x,t)dxdt |
=∫T0∫ΩGN(0,t)(u01n(y)−u02n(y))ξ(y,t)dydt+ |
+∫T0∫t0∫ΩGN(0,t−σ)(μ1n(y,σ)−μ2n(y,σ))ξ(y,t)dydσ=:I1+I2. | (4.15) |
Let us evaluate the limit of I1 and I2 when n→∞. To attain this, we begin with the expression I1:
I1=[∫Ω(u01n(y)−u01n(y))ξ(y,t)dy][∫T0GN(0,t)dt]. |
Taking ξ(y,t)=ρ(y)˜h(t) with ρ∈C2(Ω) such that ∂ρ∂η=0 on ∂Ω and ˜h∈Cc(0,T), then we have
I1=∫T0˜h(t)GN(0,t)dt∫Ω(u01n(y)−u02n(y))ρ(y)dy |
=∫T0˜h(t)G(0,t)dt[∫Ω(f01n(y)−f02n(y))ρ(y)dy−∫Ω(F01n(y)−F02n(y))ρ(y)dy]. |
Passing to the limit when n→∞, there holds
limn→∞I1=0. | (4.16) |
Now we consider the expression I2,
I2=[∫T0∫Ω(μ1n(y,t)−μ2n(y,t))ξ(y,t)dydt][∫t0G(0,t−σ)dσ]. |
According to Definition 2.4, it is worth observing that
I2=[∫t0G(0,t−σ)dσ][∫Q(f11n(y,t)−f12n(y,t))ξ(y,t)dydt]− |
−[∫t0G(0,t−σ)dσ][∫Q(F1n(y,t)−F2n(y,t))ξ(y,t)dydt]+ |
+[∫t0G(0,t−σ)dσ][∫Q(φ1n(y,t)−φ2n(y,t))ξt(y,t)dydt]+ |
+[∫t0G(0,t−σ)dσ][∫Ω(φ1n(y,0)−φ2n(y,0))ξ(y,0)dy]. |
We pass to the limit when n goes to infinity, therefore
limn→∞I2=0. | (4.17) |
By (4.16), (4.17) and Dominated Convergence theorem, we obtain
∫Q(f1(x,t)−f2(x,t))ξ(x,t)dxdt=0, | (4.18) |
which leads to
f1n∗⇀f1M+(Q)andf2n∗⇀f2M+(Q). |
Hence f1=f2 holds.
Remark 4.1 (i) Since f∈M+(Q), then it is worthy observing that fn in (4.11) is a sequence of the approximation Radon measure f satisfying the following properties
{fn∗⇀finM+(Q),fn→fa.e inQ,∥fn∥L1(Q)≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)),fis given in(2.9), | (4.19) |
for every n∈N and C>0 is a constant.
(ii) By (2.11)-(2.13) and the assumption (A), we deduce from the compactness theorem in [23] the approximation problem (Hn) possesses a weak solution f in L2((0,T),H1(Ω)) such that the properties
{fn=Tn(f),fn→fa.e inQ,∣fn∣≤∣f∣, | (4.20) |
hold true.
Proof of Lemma 2.1. To prove this result, we use Definition 2.2 and we recall the Gauss-green formula given by the functional
⟨Tν,ξ⟩=∫QξdivF+∫Q∇ξ⋅F | (5.1) |
Since there exists a linear continuous functional Tν on W12,2(S)∩C(S) which stands for F⋅ν, then we define a notion of the normal trace of the flux ∇ψ(ur)⋅ν such that
⟨Tη,ξ⟩=⟨Tν,ξ⟩+∫Ωξ(x,0)du0(x)+∫Qh(t)f(x,t)ξdxdt. | (5.2) |
The definition make sense because of the definition of the weak solution when we assume that the value of the initial data
lims→01s∫s0∫Ωu(x,t)ξ(x,t)dxdt=∫Ωξ(x,0)du0 | (5.3) |
holds, for any s−tsχ(0,s)(t)ϕ as a test function in (2.1). In particular ⟨Tν,ξσ⟩ depends only on ξS and from (2.2), we infer the formula
⟨Tν,ξ⟩=−lims→01s∫Ts∫ϑ∫ζ(¯x)+sζ(¯x)∇ψ(ur)(−∇ζ(¯x)1)ξςdxNd¯xdt | (5.4) |
for any ξ∈C∞c([0,T)ׯΩ). We denote {vδ} a boundary-layer sequence of C2(Ω)∩C(¯Ω) such that
limδ→0+vδ=1a.e in Ω,0≤vδ≤1,vδ=0on∂Ω. | (5.5) |
For more properties concerning the boundary-layer sequence {vδ} (see [37, Lemma 5.5 and Lemma 5.7]). If ξ∈(H1(Ω))N, then
limδ→0+∫Ωςξ∇vδ=−limδ→0+∫Ωdiv(ςξ)vδ=−∫Ωdiv(ςξ)=−∫∂Ωςξ⋅ηdH(x) | (5.6) |
The previous statement (5.6) explains that for any function-valued F:Ω→RN, then −F⋅∇vδ approaches the normal trace F⋅η. Let ξ∈C∞c([0,T)ׯΩ) and ξ=ξ(1−vδ) on S, it implies that ⟨Tν,ξς⟩=⟨Tν,ςξ(1−vδ)⟩. By Definition 2.2 and the equation (5.2), the Gauss-Green formula yields
⟨Tη,ςξ⟩=⟨Tν,ςξ(1−vδ)⟩+∫Ωξ(x,0)ς(1−vδ)du0+∫Qh(t)f(x,t)ξ(x,t)ς(1−vδ)dxdt |
=∫Qξς(1−vδ)divF+∫Q∇(ξς(1−vδ))⋅F+∫Ωξ(x,0)ς(1−vδ)du0+ |
+∫Qh(t)f(x,t)ξ(x,t)ς(1−vδ)dxdt. |
Since 0≤vδ≤1 and vδ→1 a.e in Ω as δ→0+, then Dominated Convergence Theorem ensures that
limδ→0+∫Qξς(1−vδ)divF=0and⟨Tη,ξς⟩=−limδ→0+∫Q∇ξ(ur)∇vδξςdxdt | (5.7) |
On the other hand, we consider ξς(1−vδ) as a test function in the problem (P) then the following expression holds
−∫Ωϕ(x,0)ς(1−vδ)du0+∫Ωu(x,T)ξ(x,T)ς(1−vδ)dx−∫Qu(x,t)ξt(x,t)ς(1−vδ)dxdt |
=−∫Q∇ψ(ur)∇(ξς)(1−vδ)dxdt+∫Q∇ψ(ur)∇vδξςdxdt+∫Sg(ur)ξςdH(x)dt+. |
+∫Qh(t)f(x,t)ξ(x,t)ς(1−vδ)dxdt. |
Since 0≤vδ≤1 and vδ→1 a.e in Ω as δ→0+, then Dominated Convergence Theorem yields
∫Sg(ur)ξςdH(x)dt=−limδ→0+∫Q∇ψ(ur)∇vδξςdxdt. | (5.8) |
By combining the assertions (5.7) with (5.8), the statement (2.3) is satisfied.
Proof of Theorem 2.2. Assume that for any compact set K=K0×[0,T]⊂Ω×(0,T)⊂RN×R+ (resp. for any compact set K0⊂Ω⊂RN) such that μ−(K)=0 (resp. u−0(K0)=0) and Cap(K)=0 (resp. C2(K0)=0). To show that μ and u0 are absolutely continuous measures with respect to the parabolic capacity, it is enough to prove that μ+(K)=0 (resp. u+0(K0)=0). To this purpose Cap(K)=0 (resp. C2(K0)=0), there exists a sequence {φn(t)}⊂C∞c(RN×R+) (resp. {φn(0)}⊂C∞c(RN)) such that 0≤φn(t)≤1 in Q (resp. 0≤φn(0)≤1 in Ω), φn(t)≡1 in K (resp. φn(0)≡1 in K0) and φn(t)→0 in W as n→∞ (resp. φn(0)→0 in H1(Ω) as n→∞). In particular ∥Δφn(t)∥L1(Q)→0 as n→∞.
Let us consider the nonnegative function φn(t)∈C∞c(RN×R+) such that φn(T)=0 in Ω and ∂φn(t)∂η=0 in S as a test function in the problem (P), then there holds
∫Ωφn(0)du0+∫Quφnt(t)dxdt=−∫Qψ(ur)Δφn(t)dxdt−∫Qh(t)f(x,t)φn(t)dxdt− |
−∫Sg(ur)φn(t)dH(x)dt. | (5.9) |
By (4.3)(Probably n is large enough), the following statement holds
∫Qf(x,t)h(t)φn(t)dxdt=∫Q∫t0∫ΩGN(x−y,t−σ)f(y,σ)(h(σ)φn(σ))σdydσdxdt |
−∫Q∫t0∫ΩGN(x−y,t−σ)[2∇f∇(h(σ)φn(σ))+fΔ(h(σ)φn(σ))]dydσdxdt+ |
+∫Q∫t0∫ΩGN(x−y,t−σ)h(σ)φn(σ)dμ(y,σ)dxdt+ |
+∫Q∫t0∫∂ΩGN(x−y,t−σ)h(σ)φn(σ)g(f(y,σ))dH(y)dσdxdt. | (5.10) |
Combining the Eq (5.9) with (5.10), we obtain
∫Q∫Ω∫t0GN(x−y,t−σ)h(σ)φn(σ)dμ(y,σ)dxdt+∫Ωφn(0)du0= |
=∫Q∫t0∫Ωh(σ)GN(x−y,t−σ)[2∇f∇φn(σ)+fΔφn(σ)]dydσdxdt− |
−∫Q∫t0∫ΩGN(x−y,t−σ)f(y,σ)(h(σ)φn(σ))σdydσdxdt− |
−∫Q∫t0∫∂ΩGN(x−y,t−σ)h(σ)φn(σ)g(f(y,σ))dH(y)dσdxdt− |
−∫Sφn(t)g(ur)dH(x)dt−∫Qψ(ur)Δφn(t)dxdt−∫Quφnt(t)dxdt. | (5.11) |
By (2.14), f∈L2((0,T),H1(Ω)) (see Remark 4.1-(ii)) and letting σ→t, ∫t0h(σ)dσ=1 and dropping down the nonnegative terms on the left hand-side of the previous equation. Therefore (5.11) yields
∫Qφn(t)dμ(x,t)+∫Ωφn(0)du0(x)≤∥u∥L2((0,T),H1(Ω))∥φn(t)∥W+ |
C(γ)∫Q∣Δφn(t)∣dxdt+∥f∥L2((0,T),H1(Ω))∥φn(t)∥W. | (5.12) |
Since the following assertions are valid, then
μ+(K)≤∫Qφn(t)dμ(x,t)+∫Qφn(t)dμ−(x,t) | (5.13) |
where μ+(K)=μ(K)+μ−(K) and
u+0(K0)≤∫Ωφn(0)du0(x)+∫Ωφn(0)du−0(x) | (5.14) |
with u+0(K0)=u0(K0)+u−0(K0). In view of (5.13) and (5.14), the inequality (5.12) reads as
μ+(K)+u+0(K0)≤∥u∥L2((0,T),H1(Ω))∥φn(t)∥W+C(γ)∫Q∣Δφn(t)∣dxdt+ |
+∥f∥L2((0,T),H1(Ω))∥φn(t)∥W+∫Qφn(t)dμ−(x,t)+∫Ωφn(0)du−0(x). | (5.15) |
Since μ−(K)=0 (resp. u−0(K0)=0), then for any ϵ>0 one has
∫Qφn(t)dμ−(x,t)<ϵ2(resp.∫Ωφn(0)du−0(x)<ϵ2). | (5.16) |
Then, the limit in (5.16) as n→∞, the following holds μ+(K)+u+0(K0)≤ϵ. Therefore, μ+(K)=0 for any compact set K⊂Q (resp. u+0(K0)=0 for any compact set K0⊂Ω).
To prove the existence and decay estimates of the solutions, we consider the following problem
{unt=Δψn(un)+h(t)fn(x,t) in Q,∂ψn(un)∂η=g(un) on S,un(x,0)=u0n in Ω, | (Pn) |
where the sequence {u0n}⊆C∞c(Ω) satisfies the assumption (4.1) and the sequence {fn}⊆C∞c(¯Q) fulfills the hypothesis (4.19). We set
ψn(s)=ψ(s)+1n | (5.17) |
By [8,18,21,22], the approximating problem (Pn) has a solution un in C((0,T),L1(Ω))∩L∞(Q). Then, the definition of the weak solution {un}⊆C∞(¯Q) of (Pn) satisfies the following expression
∫T0⟨un(⋅,t),ξt(⋅,t)⟩Ωdt+∫T0h(t)⟨fn(⋅,t),ξ(⋅,t)⟩Ωdt+⟨u0n,ξ(⋅,0)⟩Ω+ |
+∫T0⟨g(un),ξ⟩∂Ωdt=∫T0⟨∇ψn(un),∇ξ⟩Ωdxdt | (5.18) |
for every ξ in C1(¯Q) such that ξ(⋅,T)=0 in Ω and ∂ξ∂η=0 on S.
Now we establish some technical estimates which will be used in the proof of the existing solution.
Lemma 5.2 Assume that (I), (J), (A), μ∈M+(Q) and u0∈M+(Ω) are satisfied. Let un be the solution of the approximation problem (Pn), then
∥un(⋅,t)∥L1(Ω)≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)). | (5.19) |
∥∇ψn(un)∥L2(Q)+∥ψn(un)∥L2(Q)≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)). | (5.20) |
for almost every t∈(0,T) and C is a positive constant.
The sequence {[ψn(un)]t} is bounded in L2((0,T),(H1(Ω))∗)+L1(¯Q).
Proof of Lemma 5.2. To prove the estimate (5.19), we consider the approximation problem (Pn) such that
{uns=Δψn(un)+h(s)fn(x,s)inΩ×(τ,τ+t),∂ψn(un)∂η=g(un)on∂Ω×(τ,τ+t),un(x,τ)=u0n(x)inΩ×{τ}, | (5.21) |
where τ+t≤T and τ, t∈(0,T).
Let us consider ξ∈C1,2(¯Ω×[τ,τ+t]) such that ∂ξ∂η=0 on ∂Ω×(τ,τ+t) and ξ(⋅,τ+t)=0 in Ω as a test function in the above approximation problem (5.21), then we have
∫Ω×(τ,τ+t)unξsdxds+∫Ω×(τ,τ+t)ψn(un)Δξdxds+∫∂Ω×(τ,τ+t)g(un)ξdH(x)ds+ |
+∫Ω×(τ,τ+t)h(s)fn(x,s)ξ(x,s)dxds+∫Ωμn(x)ξ(x,τ)dx=0. | (5.22) |
By the mean value theorem and the assumption (I), the Eq (5.22) yields
∫Ω×(τ,τ+t)un(ξs+θnΔξ)dxds+∫∂Ω×(τ,τ+t)g(un)ξdH(x)ds+ |
+∫Ω×(τ,τ+t)h(s)fn(x,s)dxds+∫Ωμn(x)ξ(x,τ)dx=0, | (5.23) |
where θn(x,t)=∫10ψ′un(αun)dα.
On the other hand, we consider the following backward parabolic equations
{−ϕs−θϵΔϕ=1τinQτ=Ω×(τ,τ+t),∂ϕ∂η=0onSτ=∂Ω×(τ,τ+t),ϕ(⋅,τ+t)=0inΩ×{τ+t}, | (5.24) |
has an unique solution ϕ in C1,2(¯Qτ)∩C(Qτ) and 0<ϕ≤C for any τ,t∈(0,T) (see [18, Lemma 4.2]). Then for ξ=ϕ, there holds
1τ∫τ+tτ∫Ωun(x,s)dxds=∫Ωμn(x)ϕ(x,τ)dx+∫τ+tτ∫∂Ωg(un)ϕdH(x)ds+ |
+∫τ+tτ∫Ωh(s)fn(x,s)ϕdxds | (5.25) |
By the assumptions (A), (J), (4.19) and (4.1), there exists a positive constant C such that the expression below is satisfied
1τ∫τ+tτ∫Ωun(x,s)dxds≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)). | (5.26) |
By letting τ→0+, we obtain the estimate (5.19). Where C=C(h(T),∥g(un)∥L∞(R+),∣S∣)>0. To prove the estimate (5.20), we consider Tγ+1(ψn(un)) as a test function in the approximation problem (Pn), then we have
∫{(x,t)∈QT/ψn(un)≤γ+1}∣∇ψn(un)∣2dxdt=∫Ω(∫u0n(x)0Tγ+1(ψn(s))ds)dx− |
−∫Ω(∫un(x,T)0Tγ+1(ψn(s))ds)dx+∫T0∫∂Ωg(un)Tγ+1(ψn(un))dH(x)dt+ |
+∫T0∫ΩTγ+1(ψn(un))h(t)fn(x,t)dxdt | (5.27) |
where Tλ(s)=min{λ,s}. It follows that there exists a positive constant C such that
∫{(x,t)∈QT/ψn(un)≤γ+1}∣∇ψn(un)∣2dxdt |
≤(γ+1)∫Ωμn(x)dx+C(γ+1)∥g(un)∥L∞(R+)+h(T)∫Qfn(x,t)dxdt. |
For the suitable positive constant C=C(h(T),γ,∥g(un)∥L∞(R+),∥μ∥M+(Ω),∥u0∥M+(Ω),∣S∣)>0, the following estimate holds
∫{(x,t)∈QT/ψn(un)≤γ+1}∣∇ψn(un)∣2dxdt≤C. | (5.28) |
On the other hand, we assume that Gλ(s)=max{λ,s} and we choose Gγ+1(ψn(un)) as a test function in the approximation problem (Pn), then we have
∫{(x,t)∈QT/ψn(un)>γ+1}∣∇ψn(un)∣2dxdt=∫Ω(∫u0n(x)0Gγ+1(ψn(s)))ds)dx− |
−∫Ω(∫un(x,T)0Gγ+1(ψn(s))ds)dx+∫T0∫∂Ωg(un)Gγ+1(ψn(un))dH(x)dt+ |
+∫T0∫Ωh(t)fn(x,t)Gγ+1(ψn(un))dH(x)dt. | (5.29) |
It implies that
∫{(x,t)∈Q/ψn(un)>γ+1}∣∇ψn(un)∣2dxdt |
≤(γ+1)∫Ωμn(x)dx+(γ+1)M∥g(un)∥L∞(R+)+h(T)∫Qfn(x,t)dxdt |
It follows that
∫{(x,t)∈Q/ψn(un)>γ+1}∣∇ψn(un)∣2dxdt≤C. | (5.30) |
Combining the inequality (5.28) with (5.30), we deduce that
∫Q∣∇ψn(un)∣2dxdt≤C | (5.31) |
By the assumption (I), then ψn(un)∈L2(Q), whence the estimate (5.20) holds.
To end the proof of this Lemma, we consider that for every ξ∈C1c(Q) such that if we choose ϕ=ψ′n(un)ξ arbitrary as a test function in problem (Pn), then the following stands true
∫Qξt[ψn(un)]dxdt=−∫Qξψ′n(un)div(∇ψn(un))dxdt−∫Qh(t)fn(x,t)ψ′n(un)ξdxdt | (5.32) |
It follows that
∫Qξt[ψn(un)]dxdt=∫Qψ′n(un)∇ψn(un)∇ξdxdt−∫Qh(t)fn(x,t)ϕdxdt− |
−∫Sg(un)ψ′n(un)ξdH(x)dt+∫Qψ″n(un)ψ′n(un)∣∇ψn(un)∣2ξdxdt. | (5.33) |
Now we estimate each term in the right hand side of (5.33), we obtain
|∫Qψ′n(un)∇ψn(un)∇ξdxdt|≤∥ψ′n∥L∞(R+)∫Q∣∇ξ∣∣∇ψn(un)∣dxdt. | (5.34) |
From Hölder's inequality and (5.31), the inequality (5.34) reads as
|∫Qψ′n(un)∇ψn(un)∇ξdxdt|≤C∥∇ξ∥L2(Q). | (5.35) |
By the assumption (J) and (4.19), we deduce the estimate
|∫Qh(t)fn(x,t)ξdxdt|≤C∥ξ∥L∞(Q) | (5.36) |
where C=C(h(T),∥μ∥M+(Ω),∥u0∥M+(Ω))>0 is a constant.
By the assumptions (A) and (I), there exists a positive constant C=C(∥g(un)∥L∞(R+),∥ψ′n(un)∥L∞(R+))>0 such that
∫Sg(un)ψ′n(un)ξdH(x)dt≤C∥ξ∥L∞(S). | (5.37) |
Furthermore, one has
|∫Qψ″n(un)ψ′n(un)∣∇ψn(un)∣2ξdxdt|≤κ∥ξ∥L∞(Q)∫Q∣∇ψn(un)∣2dxdt. |
In view of (5.28), the expression below holds true
|∫Qψ″n(un)ψ′n(un)∣∇ψn(un)∣2ξdxdt|≤C∥ξ∥L∞(Q) | (5.38) |
where C=C(κ,∥u0∥M+(Ω),∥μ∥M+(Q))>0. By (5.35)-(5.38) and (5.33), we infer that the sequence {[ψn(un)]t} is bounded in L2((0,T),(H1(Ω))∗)+L1(¯Q).
Now we study the limit points of the sequences {un} and ψn(un) as n→∞.
Proposition 5.1 Suppose that the assumptions (I), (A) and (J) are satisfied. Let un be the solution of the approximation problem (Pn). Then there exists a subsequence {unk}⊆{un} and v∈L2((0,T),H1(Ω))∩L∞(Q) such that
ψnk(unk)∗⇀vinL∞(Q). | (5.39) |
ψnk(unk)⇀vinL2((0,T),H1(Ω)). | (5.40) |
[ψnk(unk)]t⇀vtinL2((0,T),(H1(Ω))∗). | (5.41) |
ψnk(unk)→va.e inQ, | (5.42) |
where vt∈L1(Q) and v≤γ.
Proof of Proposition 5.1. The convergences (5.39) and (5.40) are the consequence of assumption (I)-(i) and estimate (5.20) respectively. By Lemma 5.1, the sequence {[ψn(un)]t} is bounded in L2((0,T),(H1(Ω))∗)+L1(¯Q). By [45], there exists a subsequence {unk}⊆{un} and v∗∈L2((0,T),H1(Ω))∩L∞(¯Q) such that
ψnk(unk)→v∗a.e inQ. |
Furthermore, by [9, Proposition 5.1] and (5.41) holds true and we have
ψnk(unk)→va.e inQ |
with v=v∗ which leads to (5.42) be satisfied. In view of the assumptions (I)-(i) and (5.17), we get
∥ψnk(unk)−ψ(unk)∥L∞(Q)=1nk. |
Therefore the following convergence ψ(unk)∗⇀vinL∞(Q) holds true.
Remark 5.1 For any subsequence {unk}⊆{un} and v the function given in Proposition 5.1, the following assertions
ψ−1(v)∈L∞((0,T),L1(Ω)), unk→ψ−1(v)a.e in Q and unk→g−1(v)a.e in S hold.
Proposition 5.2 Assume that the hypotheses (I), (J), (A), μ∈M+(Q) and u0∈M+(Ω) are satisfied. Let {unk} be the subsequence and v the function mentioned in Proposition 5.1. Then there exist a subsequence {unk(⋅,t)}⊆{un(⋅,t)} and ua,u(⋅,t),ub(⋅,t)∈M+(Ω) such that
unk(⋅,t)∗⇀u(⋅,t):=ua(⋅,t)+ub(⋅,t)inM+(Ω), | (5.43) |
ψnk(unk)(⋅,t)∗⇀ψ(ub)(⋅,t)inM+(Ω), | (5.44) |
g(unk)(⋅,t)∗⇀g(ub)(⋅,t)inL∞(∂Ω). | (5.45) |
Moreover, there hold
ub(⋅,t)=ur(⋅,t)a.e inΩandua(⋅,t)=us(⋅,t)inM+(Ω) | (5.46) |
for almost every t∈(0,T). Furthermore u∈L∞((0,T),M+(Ω)) and for almost every t∈(0,T), there holds
∥u(⋅,t)∥M+(Ω)≤C(∥μ∥M+(Q)+∥u0∥M+(Ω)). | (5.47) |
Proof. By the assumption (I)-(i), ψnk(unk)∈L∞(Q) and using Hölder's inequality, we have
∫Q∣∇ψnk(unk)∣dxdt≤[∫Q∣∇ψnk(unk)∣2(1+ψnk(unk))2dxdt]12[∫Q(1+ψnk(unk))2dxdt]12 |
≤C[∫Q∣∇ψnk(unk)∣2dxdt]12. |
From the estimate (5.20), there exists a positive constant C=C(∥ψnk(unk)∥L∞(R+),∥μ∥M+(Q),∥u0∥M+(Ω))>0 such that
∫Q∣∇ψnk(unk)∣dxdt≤C. | (5.48) |
According to Lemma 5.1, the assumption (I) and (5.48), we infer that
∥ψnk(unk)∥BV(Q)=∥ψnk(unk)∥L1(Q)+∥∇ψnk(unk)∥L1(Q)+∥[ψnk(unk)]t∥L1(Q)≤C. | (5.49) |
By Fatou's Lebesgue Lemma, we obtain
∫T0lim infk→∞∫Ω{∣ψnk(unk)∣+∣∇ψnk(unk)∣+∣[ψnk(unk)]t∣}≤C. | (5.50) |
Then there exists zero Lebesgue measure set N1⊂(0,T) such that
lim infk→∞∫Ω{∣ψnk(unk)∣+∣∇ψnk(unk)∣+∣[ψnk(unk)]t∣}(x,t)≤C | (5.51) |
for every t∈(0,T)∖N1. In view of (5.51), the sequence {ψnk(uk)(⋅,t)}⊆BV(Ω) for every t∈(0,T)∖N1. By [44, Chapter IV, Section 1.1, Proposition 5], there exists a subsequence {ψnk(unk)(⋅,t)} and v(⋅,t)∈M+(Ω) a.e in (0,T) such that the convergence
ψnk(unk)(⋅,t)∗⇀v(⋅,t)inM+(Ω) | (5.52) |
holds true. Furthermore, from the assertions (5.19), (5.52) and the Prohorov Theorem (see [44, Chapter II, Section 2.6, Theorem 1] or [25, Proposition A.2] or [17, Proposition 1]), there exists a sequence {˜τnk} of the Young measures associated with the sequence {unk}⊆{un} converges narrowly over ¯Q×R to a Young measure ˜τ which the disintegration ¯μ(⋅,t) is the Dirac mass concentrated at the point ψ−1(v(⋅,t)) for a.e in Ω (see [17]). By [25, Proposition A.4], there exist sequences of measure sets Ak⊆Ω, Ak⊆Ak+1 and ∣Ak∣→0, such that
ukj(⋅,t)χΩ∖Ak⇀ub(⋅,t):=∫[0,+∞)λd¯μ(⋅,t)(λ)inL1(Ω), | (5.53) |
where ub∈L∞((0,T),L1(Ω)), ub≥0 is a barycenter of the limiting Young measure ¯μ(⋅,t) associated with the subsequence {unk(⋅,t)} and supp¯μ(⋅,t)⊆[0,+∞) for almost every t∈(0,T).
By (5.19) and the compactness result, the sequence {unkχΩ∖Aj} is uniformly bounded in L1(Ω). Therefore, there exists a Radon measure ua(⋅,t)∈M+(Ω) such that unk(⋅,t)∗⇀u(⋅,t) in M+(Ω). Finally, the sequence unk is of unk(⋅,t)=unk(⋅,t)χAk+unk(⋅,t)χΩ∖Ak∗⇀ua(⋅,t)+ub(⋅,t) in M+(Ω). Hence u(⋅,t):=ua(⋅,t)+ub(⋅,t) in M+(Ω) and the statement (5.43) is completed. By the assumption (I)-(iii), there holds
lims→+∞ψnk(s)s=0. | (5.54) |
By the assertion (5.54) and [45, Proposition 5.2] or [25], we obtain
ψnk(uk)(⋅,t)∗⇀ψ∗(⋅,t)inM+(Ω) | (5.55) |
where ψ∗(⋅,t)∈L1(Ω) and
ψ∗(⋅,t)=∫[0,+∞)ψ(λ)d¯μ(⋅,t)(λ). | (5.56) |
Furthermore, we also obtain the next result via (5.55)
ψ∗(⋅,t)=∫[0,+∞)ψ(λ)d¯μ(⋅,t)(λ)=ψ(∫[0,+∞)λd¯μ(⋅,t)(λ))=ψ(ub)(⋅,t). |
By combining the assertion (5.53) and the previous equality, we conclude that ψ(ub)(⋅,t)=v(⋅,t) a.e in (0,T), when the convergence (5.44) is satisfied.
By virtue of the convergence (5.53), the next convergence result
g(unk)→g(ψ−1(v)):=g(ub)a.e inS | (5.57) |
holds true. Since the function g(unk)∈L∞(R+) (see assumption (H)-(i)) and from Fatou's Lebesgue Lemma, then there exists a positive constant C such that
∫T0lim infk→+∞∫∂Ωg(unk)dxdt≤C. | (5.58) |
Therefore, there exists a zero Lebesgue measure set N2⊆(0,T) such that
lim infk→+∞∫∂Ωg(unk)(x,t)dx≤C | (5.59) |
for every t∈(0,T)∖N2. In view of (5.59) and (5.57), there exists a function z(⋅,t):=g(ub)(⋅,t)∈L∞(∂Ω) such that the convergence (5.45) is achieved.
To show (5.46), we consider the functions F,G:R+→R+ defined by setting
Fϵ(s)={0 if s≤1ϵ,(ϵs−1)22ϵ2 if 1ϵ≤s≤1ϵ+1,s−1ϵ−12 if s≥1ϵ+1, |
and Gϵ(s)=s−Fϵ(s) for every ϵ>0. It is worthy observing that F′ϵ(s)≥0 in R+ and 0≤F″ϵ(s)≤χ{s≥1ϵ}(s). According to the above results, there exists a subsequence {unk} in Lemma 5.1 and Proposition 5.1. For any nonnegative function ρ∈C2(¯Ω), we choose F′ϵ(unk)ρ(x) as a test function in the approximation problem (Pn), then we obtain the following identity
∫ΩFϵ(unk)(⋅,τ)ρ(x)dx≤∫ΩFϵ(u0nk)ρ(x)dx−∫τ0∫ΩF′ϵ(unk)∇ψnk(unk)∇ρ(x)dxdt+ |
+∫τ0∫∂Ωg(unk)F′ϵ(unk)ρ(x)dH(x)dt+∫τ0∫Ωh(t)fnk(x,t)F′ϵ(unk)ρ(x)dxdt | (5.60) |
where τ∈(0,T). Since the sequence {F′ϵ(unk)} is uniformly bounded in L∞(Q), then F′ϵ(unk)→0 as ϵ→0+ and Fϵ(unk)→0 as ϵ→0+. By Lemma 5.1 and Proposition 5.1, and by applying the Dominated Convergence Theorem, results to
limk→+∞∫τ0∫ΩF′ϵ(unk)∇ψnk(unk)∇ρ(x)dxdt=∫τ0∫ΩF′ϵ(ψ−1(v))∇v∇ρ(x)dxdt. | (5.61) |
Similary, we get
limk→+∞∫τ0∫∂Ωg(unk)F′ϵ(unk)ρ(x)dH(x)dydt=∫τ0∫∂Ωg(ψ−1(v))F′ϵ(ψ−1(v))ρ(x)dH(x)dt, | (5.62) |
By the statement (4.19) and Proposition 5.1, we have
limk→+∞∫τ0∫Ωh(t)fnk(x,t)F′ϵ(unk)ρ(x)dxdt=∫τ0∫Ωh(t)f(x,t)F′ϵ(ψ−1(v))ρ(x)dH(x)dt. | (5.63) |
Given the properties of the sequence {Fϵ(unk)} and passing to limit in (5.61), (5.62) and (5.63) when ϵ→0+, then the following holds
limϵ→0+limk→+∞∫τ0∫ΩF′ϵ(unk)∇ψnk(unk)∇ρ(x)dxdt=0. | (5.64) |
Similarly we obtain
limϵ→0+limk→+∞∫τ0∫∂Ωg(unk)F′ϵ(unk)ρ(x)dH(x)dt=0. | (5.65) |
And
limϵ→0+limk→+∞∫τ0∫Ωh(t)fnk(x,t)F′ϵ(unk)ρ(x)dxdt=0. | (5.66) |
On the other hand, we have
Fϵ(u0nk)=u0nk−Gϵ(u0nk)=u0rnk+u0snk−Gϵ(u0nk). |
Since u0rnk→u0r in L1(Ω), u0snk∗⇀u0s in M+(Ω) and the sequence {Gϵ(u0nk)} is uniformly bounded in L∞(Ω), then we deduce that
u0rnk−Gϵ(u0nk)→u0r−Gϵ(u0r):=Fϵ(u0r)inL1(Ω). | (5.67) |
According to the convergence statement (5.43), we have
Fϵ(unk)(⋅,t)=unk(⋅,t)−Gϵ(unk)(⋅,t)∗⇀ua(⋅,t)+ψ−1(v)−Gϵ(ψ−1(v))(⋅,t)inM+(Ω) | (5.68) |
where Fϵ(ψ−1(v))(⋅,t):=ψ−1(v)(⋅,t)−Gϵ(ψ−1(v))(⋅,t).
Furthermore, from the Eqs (5.43) and (5.66) we obtain the following
limϵ→0+limk→+∞∫ΩFϵ(unk)(⋅,t)ρ(x)dx=⟨ua(⋅,t),ρ⟩Ω+limϵ→0+∫ΩFϵ(ψ−1(v))(⋅,t)ρ(x)dx. | (5.69) |
It follows that
limϵ→0+limk→+∞∫ΩFϵ(unk)(⋅,t)ρ(x)dx=⟨ua(⋅,t),ρ⟩Ω. | (5.70) |
Likewise, from (5.67) one has
limϵ→0+limk→+∞∫ΩFϵ(u0nk)(⋅,t)ρ(x)dx=⟨u0s,ρ⟩Ω+limϵ→0+∫ΩFϵ(u0r)ρ(x)dx. |
It implies that
limϵ→0+limk→+∞∫ΩFϵ(u0nk)ρ(x)dx=⟨u0s,ρ⟩Ω. | (5.71) |
Combining the statements (5.64)-(5.66), (5.70), (5, 71) with (5.60) yields
⟨ua(⋅,t),ρ⟩Ω≤⟨u0s,ρ⟩Ω. |
Since ua(⋅,t) is a singular measure with respect to the Lebesgue measure ua(⋅,t)=[ua(⋅,t)]s=us(⋅,¯t) for a suitable ¯t∈(0,T)∖H∗, where H∗ is zero Lebesgue measure in (0,T). Hence the assertion (5.46) is obtained.
From [44, Chapter 5, Section 5.2.1, Theorem 1], the estimate (5.19) yields
∥u(⋅,t)∥M+(Ω)≤lim infk→+∞∫Ωunk(⋅,t)dx≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)). | (5.72) |
The estimate (5.47) is completed.
Proof of Theorem 2.2. By Proposition 5.1 and Proposition 5.2, we have ψ(ur)=v a.e in Q. Hence the problem (P) has a weak Radon measure-valued solution u in L∞((0,T),M+(Ω)).
Remark 5.1 By Theorem 2.2, the result holds
[u(⋅,t)]s≤u0sinM+(Ω) | (5.73) |
for almost every t∈(0,T). By (5.73), there exists zero Lebesgue measure set N3⊂(0,T) such that
[u(⋅,t)]c,2(E)≤[u0]c,2(E)inM+(Ω) | (5.74) |
for all Borel sets E⊂Ω, with C2(E)=0 and t∈(0,T)∖N3.
Proposition 5.3. Suppose that the assumptions (I) and (A) are fulfilled. Let {unk} be the subsequence and v the function given in Proposition 5.1. Then the following sets
S={(x,t)∈¯Q∖ψ(ur)(x,t)=γ}andN={(x,t)∈¯Q∖g(ur)(x,t)=0} |
have zero Lebesgue measure. Moreover S⊆N and B=S∪N has zero Lebesgue measure.
Proof of Proposition 5.3. By [9, Proposition 5.2], the set S has zero Lebesgue measure. Assume that
Aj={(x,t)∈¯Q∖v(x,t)≤1j}. |
Then, it is worth observing that
Aj+1⊇Aj,N=∞⋃j=1Aj,∣N∣=limj→+∞∣Aj∣ | (5.75) |
To prove that ∣N∣=0, it is enough to show that ∣Aj∣→0 as j→+∞.
Since the function g′<0 in R+ (see the assumption (A)-(i)), then we have
g(unk)≤2j⇔unk≥g−1(2j)((x,t)∈¯Q). | (5.76) |
It follows that
g−1(2j)∫{(x,t)∈¯Q∖v(x,t)≤1j}χ{g(unk)≤2j}dxdt≤∫Qunk(x,t)dxdt. | (5.77) |
By the estimate (5.19), we have
g−1(2j)∣Aj∣≤CT∥μ∥M+(Ω). | (5.78) |
Since g−1(2j)→+∞ as j→+∞, then (3.62) yields ∣Aj∣→0 as j→+∞.
Assume that (x0,t0)∈S, then ψ(ur(x0,t0))=γ for every γ∈(0,+∞). Since g(ur(x0,t0))=∂ψ(ur)∂η(x0,t0)=∂∂η(γ)=0. Therefore, (x0,t0))∈N, that is S⊆N holds true. The fact that S⊆N, then B=N. Consequently, B is zero Lebesgue measure set.
Proposition 6.1. Under assumptions (I), (A) and (J). Let u be a very weak Radon measure-valued solution to the problem (P) and for every ρ∈C2(¯Ω) such that ∂ρ∂η=0 on ∂Ω, there holds
esslimt→0+⟨u(⋅,t),ρ⟩Ω=⟨u0,ρ⟩Ω | (6.1) |
Proof of Proposition 6.1 Let us consider that for every τ>0, the smooth function ητ∈C1c(0,T), 0≤ητ≤1 such that
ητ(t)={0if0≤t≤t1−τ,1τ(t+τ−t1)ift1−τ≤t≤t1,1ift1≤t≤t2,1τ(−t+τ+t2)ift2≤t≤t2+τ,0ift2+τ≤t≤T. | (6.2) |
Let us choose ρj(x)ητ(t) as a test function in (P), there holds
∫T0∫Ω{−uρj(x)η′τ(t)−ψ(ur)ητ(t)Δρj(x)}dxdt= |
∫T0∫∂Ωg(ur)ρj(x)ητ(t)dH(x)dt+∫T0∫Ωh(t)fητ(t)ρj(x)dxdt. | (6.3) |
It is worth observing that the first term on the left hand side of the equality (6.3) gives
∫T0∫Ω−uρj(x)η′τ(t)dxdt=−1τ∫t1t1−τ∫Ωu(x,t)ρj(x)dxdt+ |
+1τ∫t2+τt2∫Ωu(x,t)ρj(x)dxdt. | (6.4) |
Let us consider a zero Lebesgue measure set Dj in (0,T) such that for any t1,t2∈(0,T)∖Dj, one has
limτ→0∫T0∫Ω−uρj(x)η′τ(t)dxdt=−∫Ωu(x,t1)ρj(x)dx+∫Ωu(x,t2)ρj(x)dx. | (6.5) |
We assume that a sequence {ρj(x)} of test functions in Ω such that
ρ∈C2(Ω),ρj(x)→ρ(x)withρ(x)∈C2(Ω) |
and
Δρj(x)→Δρ(x)uniformly in Ω. |
Then for every t∈(0,T)∖Dj, there holds
∫Ωu(x,t)ρj(x)dx−∫Ωu0ρj(x)dx=∫t0∫Ωψ(ur)Δρj(x)dxds+ |
+∫t0∫∂Ωg(ur)ρj(x)dH(x)dt+∫t0∫Ωh(t)fρj(x)dxdt. | (6.6) |
By Dominated Convergence Theorem, we obtain
∫Ωu(x,t)ρ(x)dx−∫Ωu0ρ(x)dx=∫t0∫Ωψ(ur)Δρ(x)dxds+ |
+∫t0∫∂Ωg(ur)ρ(x)dH(x)dt+∫t0∫Ωh(t)fρ(x)dxdt | (6.7) |
for every t∈(0,T)∖D with D=⋃j≥0Dj
Since ψ(ur)∈L∞(Q), for every ρ∈C2(Ω) and for every sequence {tj}⊆(0,T)∖D, tj→0+ as j→∞ such that
∫Ωu(x,tj)ρ(x)dx−∫Ωu0ρ(x)dx=∫tj0∫Ωψ(ur)Δρ(x)dxds+ |
+∫tj0∫∂Ωg(ur)ρ(x)dH(x)dt+∫tj0∫Ωh(t)fρ(x)dxdt | (6.8) |
holds true.
Since u∈L∞((0,T),M+(Ω)), then we have
supj∥u(⋅,tj)∥M+(Ω)≤C. | (6.9) |
So that there exists a subsequence {tjm}⊆{tj} and a Radon measure μ0∈M+(Ω) such that
u(⋅,tjm)∗⇀μ0inM+(Ω)asjm→∞. | (6.10) |
By the standard density arguments, one has
esslimjm→∞⟨u(⋅,tjm),ρ⟩Ω=⟨u0,ρ⟩Ω | (6.11) |
where μ0=u0, hence (6.1) is obtained.
Proof of Theorem 2.4 Let u1,u2 be two very weak solutions obtained as limit of approximation of (P) with initial data u01n and u02n respectively. Let {u1n}, {u2n}⊆L∞(Q)∩L2((0,T),H1(Ω)) be two approximating sequence solutions to the problem (Pn). We consider a test function ξ∈C2,1(Q) such that ξ(⋅,T)=0 in Ω and ∂ξ∂η=0 on ∂Ω×(0,T) in the approximation problem (Pn) in the sense of the Definition 2.3, then there holds
∫Q(u1n−u2n)ξtdxdt=−∫Q(ψn(u1n)−ψn(u2n))Δξdxdt− |
−∫Qh(t)(f1n−f2n)ξdxdt−∫S(g1n−g2n)ξdH(x)dt− |
∫Ω(u01n−u02n)ξ(x,0)dx, | (6.12) |
where {f1n}, {f2n}, {u01n}, and {u02n} are two approximating functions.
By the assumption g(ur)=L a.e in S, then for any sequences {u1n}, {u2n} one has g(u1n=g(u2n)=L on S. Consequently the third term on the right hand-side of the equation (6.12) vanishes.
For almost every (x,t)∈Q, we consider the function an(x,t) defined as
an(x,t)={ψn(u1n(x,t))−ψn(u2n(x,t))u1n(x,t)−u2n(x,t)ifu1n(x,t)≠u2n(x,t),ψ′n(u1n(x,t))ifu1n(x,t)=u2n(x,t). | (6.13) |
Obviously an∈L∞(Q) and for every n∈N there exists a positive constant Cn such that
essinf(x,t)∈Qan(x,t)≥Cn>0. | (6.14) |
This ensures that for every z∈C2c(Q), the problem
{ξnt+anΔξn+z=0inQ,∂ξn∂η=0onS,ξn(⋅,T)=0inΩ, | (6.15) |
has a unique solution ξn∈L∞((0,T),H2(Ω))∩L2((0,T),H1(Ω)) with ξnt∈L2(Q) (see [18,21]).
Moreover, it can be seen that
∣ξn(x,t)∣≤(T−t)∥z∥L∞(Q). | (6.16) |
Let us consider the function β such that for any t1+1<t2 and t1,t2∈(0,T)
β(t)={0if0≤t≤t1,t−t1ift1<t<t2,t2−t1ift≥t2. | (6.17) |
Choosing βΔξn as a test function in (6.15), then we obtain
∫Qξntβ(t)Δξndxdt+∫Qβ(t)an(x,t)[Δξn]2dxdt+∫Qzβ(t)Δξndxdt=0. | (6.18) |
It follows that
12∫Q∣∇ξn∣2dxdt+∫Qan(x,t)[Δξn]2dxdt≤C0(T,z) | (6.19) |
holds, for some constant C0(T,z) independent on n.
From (6.16) and (6.19), there exists a constant C1(T,z) such that
∥ξn∥L2((0,T),H1(Ω))+∥√anΔξn∥L2(Q)≤C1(T,z). | (6.20) |
On the other hand, multiplying (6.15) by Δξn and we obtain
−∫Q∇ξn∇ξnt+∫Qan[Δξn]2dxdt=−∫QξnΔzdxdt |
which leads to
12∫Ω∣∇ξn∣2(x,0)dx+∫Qan[Δξn]2dxdt≤C2(T,z), | (6.21) |
where C2(T,z)=C(∥ξn∥L∞(Q),∥z∥C2(¯Q))>0. Therefore, we get
∥ξn(⋅,0)∥H1(Ω)+∥√anΔξn∥L2(Q)≤C2(T,z). | (6.22) |
By standard density arguments, we can choose ξ=ξn as a test function in (6.15). It implies that (6.12) yields
∫Q(u1n−u2n)zdxdt=∫Qh(t)(f1n−f2n)ξn(x,t)dxdt+ |
+∫Ω(u01n−u02n)ξn(x,0)dx. | (6.23) |
Letting n to infinity in (6.23). Then it enough to observe from (6.20), there exists ξn∈L∞((0,T),H2(Ω))∩L2((0,T),H1(Ω)) which is obtained by extracting the subsequence of the {ξn} such that
ξn(x,t)∗⇀ξ(x,t)inL∞(Q). | (6.24) |
ξn(x,t)⇀ξ(x,t)inL2((0,T),H1(Ω)). | (6.25) |
Since ξnt∈L2(Q) and the compactness theorem states in [21], we deduce that
ξnt(x,t)→ξt(x,t)inL2((0,T),(H1(Ω))∗), | (6.26) |
ξn(x,t)→ξ(x,t)a.e inQ. | (6.27) |
By (6.16) and (6.22), there exists ξ(⋅,0)∈L∞(Ω)∩H1(Ω) such that the following statements
ξn(x,0)∗⇀ξ(x,0)inL∞(Ω), | (6.28) |
ξn(x,0)⇀ξ(x,0)inH1(Ω) | (6.29) |
holds true. By Theorem 2.1, the solutions of the problem (H) are unique in M+(Q). Therefore f1n∗⇀f in M+(Q) and f2n∗⇀f in M+(Q). Furthermore, the sequences {u01n} and {u02n} satisfy the assumption (2.6). By combining the above assumptions and Dominated Convergence Theorem, the Eq (6.23) reads
∫Q(u1−u2)z(x,t)dxdt=limn→+∞∫Q[h(t)(f1n−f2n)]ξ(x,t)dxdt+ |
+limn→∞∫Ω(f01n−f02n)ξ(x,0)dxdt−limn→∞∫Ω(F01n−F02n)ξ(x,0)dx=0 |
It follows that u1=u2 in M+(Q).
In this section, we prove the result of decay estimate solutions.
Proof of Theorem 2.5. We consider un and vn two solutions of the approximation problems (Pn) and (P0n) respectively. For any ξ∈C1((0,T),C1(Ω)) such that ξ(⋅,T)=0 in Ω and ∂ξ∂η=0 on S as a test function of the approximation problem (Pn)−(P0n), then there holds
∫Q(un−vn)ξt(x,t)dxdt=∫Q∇[ψ(un)−ϑ(vn)]∇ξdxdt−∫Qh(t)fn(x,t)ξdxdt− |
−∫S(g(un)−g1(vn))ξdH(x)dt. | (7.1) |
For every ϵ>0, we consider {zϵ} be a sequence of smooth functions such that ∥zϵ∥L1(0,T)≤C and zϵ(t)∗⇀δt in M+(0,T). Let us choose ξ(x,t)=sign(un(x,t)−vn(x,t))∫Ttzϵ(s)(T−s)αds(α>1) into the Eq (7.1), then (7.1) reads
[∫T0zϵ(t)(T−t)αdt][∫Ω∣un(⋅,t)−vn(⋅,t)∣dx] |
=−[∫Tth(t)(∫T0zϵ(s)(T−s)αds)dt][∫Ωfnsign(un(x,t)−vn(x,t))dx]− |
−[∫Tt(∫T0zϵ(s)(T−s)αds)dt][∫∂Ω(g(un)−g1(un))sign(un(x,t)−vn(x,t))dH(x)] | (7.2) |
Letting ϵ→0+ in the previous equation and using the properties of the Dirac mass at t, then we have the following expression
(T−t)α∫Ω∣un(⋅,t)−vn(⋅,t)∣dx≤C∫Qfn(x,t)(T−t)αdxdt | (7.3) |
for any t∈(0,T)∖H∗ with ∣H∗∣=0 and C=C(∣S∣,∥g(un)∥L∞(R+),∥g1(vn)∥L∞(R+),Tα)>0 is a constant. On the other hand, by (4.5) we have
fn(x,t)(T−t)α=Tα∫ΩGN(x−y,t)u0n(y)dy+∫t0∫∂ΩGN(x−y,t−σ)g(fn)(T−σ)αdydσ+ |
+∫t0∫ΩGN(x−y,t−σ){−αfn(T−σ)α−1+μn(T−σ)α}dydσ. |
By (2.11)-(2.13) and the properties of the Green function GN, we get the following result
∫Qfn(x,t)(T−t)αdxdt≤Tα+1∫Ωu0n(y)dy+α∫t0∫Qfn(y,σ)(T−σ)αdydσdt+ |
+∫t0∫Qμn(T−σ)αdydσdt+∫T0∫Sg(fn)(T−σ)αdydσdt. |
By the assumptions (A), (4.1) and (4.2), there exists a positive constant C=C(Tα+1,∥g(fn)∥L∞(R+),∣S∣)>0 such that
∫Qfn(x,t)(T−t)αdxdt≤C(∥u0∥M+(Ω)+∥μ∥M+(Ω))+α∫t0(∫Qfn(x,σ)(T−σ)αdxdt)dσ. | (7.4) |
By Gronwall's inequality, (7.4) yields
∫Qfn(x,t)(T−t)αdxdt≤CeαT(∥u0∥M+(Ω)+∥μ∥M+(Ω)) | (7.5) |
where C=C(Tα+1,∥g(fn)∥L∞(R+),∣S∣,eαT)>0 is a constant. Combining (7.3) with (7.5), we deduce that
(T−t)α∫Ω∣un(⋅,t)−vn(⋅,t)∣dx≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)). | (7.6) |
By [24, Chapter V, Section 5.2.1, Theorem 1], the semi-continuity of the total variation yields,
(T−t)α∥u(⋅,t)−v(⋅,t)∥M+(Ω)≤(T−t)αlim infn→∞∫Ω∣un(⋅,t)−vn(⋅,t)∣dx |
≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)). | (7.7) |
Hence (2.18) holds.
We consider un and wn two solutions of the approximation problems (Pn) and (P1n) respectively. For any ξ∈C1((0,T),C1(Ω)) such that ξ(⋅,T)=0 in Ω and ∂ξ∂η=0 on S as a test function of the approximation problem (Pn)−(P1n). Therefore, we have the following equation
∫Q(un−wn)ξt(x,t)dxdt=∫Q∇[ψ(un)−ψ(wn)]∇ξdxdt−∫Ωu0nξ(x,0)dxdt | (7.8) |
Taking ξ(x,t)=sign(un(x,t)−wn(x,t))∫Ttzϵ(s)(T−s)αds(α>1) into the equality (7.8), then we obtain
[∫T0zϵ(t)(T−t)αdt][∫Ω∣un(⋅,t)−wn(⋅,t)∣dx] |
=−[∫T0zϵ(s)(T−s)αdt][∫Ωu0nsign(un(x,0)−wn(x,0))]. |
Letting ϵ→0+ in the previous equation and using the properties of the Dirac mass at t, then we have
(T−t)α∫Ω∣un(⋅,t)−wn(⋅,t)∣dx≤Tα∫Ωu0ndx. | (7.9) |
By (4.1), the above inequality (7.9) yields
(T−t)α∫Ω∣un(⋅,t)−wn(⋅,t)∣dx≤C∥u0∥M+(Ω). | (7.10) |
By [24, Chapter V, Section 5.2.1, Theorem 1], the semi-continuity of the total variation yields,
(T−t)α∥u(⋅,t)−w(⋅,t)∥M+(Ω)≤(T−t)αlim infn→∞∫Ω∣un(⋅,t)−wn(⋅,t)∣dx |
≤C∥u0∥M+(Ω) |
where C=C(T,α)>0 a constant. Hence (2.19) is achieved. Now we consider the auxiliary function Wn such that
Wn(x,t)=tαun(x,t)sign(Wn) | (7.11) |
for every α>1. The derivation of the expression Wn with respect to the variable t gives
Wnt(x,t)=αtα−1un(x,t)sign(Wn)+tαunt(x,t)sign(Wn). | (7.12) |
Since unt=Δψ(un)+h(t)fn(x,t) and we multiply the Eq (7.12) by the function sign(Wn) and then we integrate the result over Ω×(0,t)(for any t∈(0,T)), then we obtain
∫Ω∣Wn∣(x,t)dx=α∫t0sα−1∫Ωun(x,s)dxds+∫t0∫∂Ωg(un)sαdH(x)ds+∫t0∫Ωsαh(s)fn(x,s)dxds. | (7.13) |
By replacing the expression of Wn in (7.13), we deduce that
tα∫Ωun(x,t)dx≤αTα∫Qun(x,t)dxdt+Tα∫Sg(un)dH(x)dt+h(T)Tα∫Qfn(x,t)dxdt. | (7.14) |
By assumptions (A), (J), (2.16) and (4.19), there exists a constant C=C(αTα+1,h(T)Tα,∥g(un)∥L∞(R+))>0 such that
tα∫Ωun(x,t)dx≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)). | (7.15) |
According to [24, Chapter V, Section 5.2.1, Theorem 1], we conclude from the estimate (7.15), the following estimate
tα∥u(⋅,t)∥M+(Ω)≤tαlim infn→∞∫Ωun(⋅,t)dx≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)). |
Hence the estimate (2.21) is completed.
To show the existence of the problem (E), we employ the natural approximation method. Therefore, the solution of the problem (P) is constructed by limiting point of a family {un} of solutions to the approximation problem. To this purpose, we consider the function ϕ∈C∞c(Ω) such that 0≤ϕ≤1 and ϕ=1 in K0 (for any compact set K0⊂Ω⊂RN), then we get
−Δ(ϕψ(U))+ϕU=ϕu0+ε(ϕ)inD′(Ω) |
where ε(ϕ)=−ψ(U)Δϕ−2∇ϕ∇ψ(U) and ε(ϕ)=0 in K0 with ε(ϕ)∈L1(Ω).
Now we consider the approximation of problem (E)
{−Δψ(Un)+Un=u0n in Ω,∂ψ(Un)∂η=g(Un) on ∂Ω, | (En) |
where u0n=(ϕu0+ε(ϕ))∗ρn and {ρn} a sequence of standard mollifiers. Furthermore, the sequence {u0n}⊆C∞(¯Ω) satisfies the assumption (4.1).
Then for every n∈N, there exists Un∈H1(Ω)∩L∞(Ω) solution of the approximation problem (En).
In the next Lemma, we state the technical estimates important for the proof of the existing solutions.
Lemma 8.1 Assume that (I), (A) and u0∈M+(Ω) are satisfied. The sequence {Un} be a weak solution of the approximation problem (En). Then, there holds
∥Un∥L1(Ω)≤C∥u0∥M+(Ω), | (8.1) |
∥∇ψ(Un)∥L2(Ω)+∥ψ(Un)∥L2(Ω)≤C, | (8.2) |
where C>0 is a constant. Moreover, for every 1≤p<NN−1 there holds
∥∇ψ(Un)∥Lp(Ω)+∥ψ(Un)∥Lp(Ω)≤C, | (8.3) |
where C=C(p)>0 is a constant.
Proof of Lemma 8.1 We consider φ∈C1(¯Ω) as a test function in the approximation problem (En), then we have
∫Ω∇ψ(Un)∇φdx+∫ΩUnφdx=∫Ωu0nφdx+∫∂Ωg(Un)φdH(x) | (8.4) |
Assume that Ω−={x∈Ω/Un(x)≤0in the sense ofL1(Ω)} and φ(x)=infx∈Ω{Un(x),0}. It is worth observing that φ(x)∈L1(Ω). To show that Un≥0 in Ω, it is enough to prove that φ(x)=0 in Ω. Indeed, we choose φ(x)=sign(Un(x)), then we get
∫Ω−∣Un(x)∣dx=∫Ω−u0n(x)sign(Un(x))dx+∫∂Ω−g(Un)sign(Un(x))dH(x)≤0 | (8.5) |
where u0n≥0 in Ω and g>0 in R+ (see the assumption (A)). Therefore φ(x)=0 a.e in Ω. Hence the solution of the approximation problem (En), Un(x)≥0 a.e x∈Ω.
Now we consider the regularizing sequence {Tϵ}⊆C1(R+) for every ϵ>0 such that
(i) 0≤Tϵ(s)≤1 in R+, Tϵ(s)=0, T′ϵ≥0 in R+,
(ii) Tϵ(s)→1 as ϵ→0+ for every s≠0.
We choose Tϵ(Un)∈H1(Ω)∩L∞(Ω) as a test function in the approximation problem (En) and by employing the assumptions (A) and (I), then we get
∫ΩT′ϵ(Un)ψ′(Un(x))∣∇Un(x)∣2dx+∫ΩUn(x)Tϵ(Un)dx≤C∥u0∥M+(Ω) | (8.6) |
where C=C(∥g(Un)∥L∞(R+),∣∂Ω∣)>0. Since T′ϵ(Uk)ψ′(Un(x))≥0 in R+ (see the hypothesis (I)), then (8.6) reads
∫ΩUn(x)Tϵ(Un)dx≤C∥u0∥M+(Ω) | (8.7) |
Again, by considering the limit when ϵ→0+, the estimate (8.1) holds true. Now we consider another regularizing sequence {Tϵ}⊆C1(R+) for every ϵ>0 such that Tϵ(s)=1 if 0≤s≤1ϵ, Tϵ(s)=ϵs if 1ϵ≤s≤2ϵ, Tϵ(s)=2 if s≥2ϵ. It is obvious to see that 1≤Tϵ(s)≤2 in R+. We take the function φ(s)=∫s0Tϵ(σ)dσ and we choose φ(ψ(Un)) as a test function in (En), then we obtain
∫Ω∣∇ψ(Un)∣2Tϵ(ψ(Un))dx+∫ΩUnφ(ψ(Un))dx=∫Ωφ(ψ(Un))u0ndx+∫∂Ωg(Un)φ(ψ(Un))dH(x). | (8.8) |
Since 1≤Tϵ(ψ(Un))≤2 and ψ(Un)≤φ(ψ(Un))≤2ψ(Un), therefore there exists a positive constant C such that
∫Ω∣∇ψ(Un)∣2dx≤C∥u0∥M+(Ω) | (8.9) |
where C=C(∥ψ(Un)∥L∞(R+),∥g(Un)∥L∞(R+),∣∂Ω∣)>0. By the assumption (I), the statement ψ(Un)∈L2(Ω) holds. Whence the estimate (8.2) is achieved.
Again, recalling the Hölder's inequality, we get
∫Ω∣∇ψ(Un)∣pdx≤[∫Ω∣∇ψ(Un)∣2(1+ψ(Un))2dx]1q[∫Ω(1+ψ(Un))q′dx]1q′ |
where q:=2p and q′:=22−p. Therefore, there exists a positive constant C=C(p,∥ψ(Un)∥L∞(R+))>0 such that
∫Ω∣∇ψ(Un)∣pdx≤C | (8.10) |
By the assumption (I), the statement ψ(Un)∈Lp(Ω) holds. Hence the estimate (8.3) is achieved.
Proof of Theorem 2.6. From the estimate (8.2) and assumption (A), we can extract from {ψ(Un)} a subsequence {ψ(Unk)} such that
ψ(Unk)→VinH1(Ω)andψ(Unk)→Va.e inΩ | (8.11) |
g(Unk)∗⇀VL∞(∂Ω)andg(Unk)→Va.e in∂Ω | (8.12) |
By (8.3), the sequence {ψ(Unk)}⊆BV(Ω) and applying [44, Chapter IV, Section 1.1, Proposition 5], there exists a subsequence {ψ(Unk)} and V1∈M+(Ω) such that the convergence
ψ(Unk)∗⇀V1inM+(Ω). | (8.13) |
By repeating the same method as in the Proposition 5.2, we deduce that
Unk∗⇀U:=ψ−1(V)+λ1inM+(Ω) | (8.14) |
where Ur=ψ−1(V) a.e in Ω, Us=λ1 in M+(Ω) and Ur=g−1(V) a.e in ∂Ω.
By [24, Chapter V, Section 5.2.1, Theorem 1], the estimate (8.1) yields
∥U∥M+(Ω)≤lim infn→∞∫ΩUn(x)dx≤C∥u0∥M+(Ω). |
Hence the estimate (2.23) is completed.
Remark 8.1 The sets
S0={x∈¯Ω∖ψ(Ur)(x)=γ}andN0={x∈¯Ω∖g(Ur)(x)=0} |
have zero Lebesgue measure. Moreover S0⊆N0 and supp(Us)⊆S0.
Proof of Theorem 2.7. We choose ξ(x,t)=sign(un(x,t)−Un(x))∫Ttzϵ(s)sαds(α>1) as a test function in the approximation problem (Pn)−(En), then we have
∫Ω∫T0∣un(x,t)−Un(x)∣zϵ(t)tαdt=∫Ω∫T0∣u0n(x)−Un(x)∣zϵ(t)tαdtdx+ |
+∫Ω∫T0[g(un)−g(Un)]sign(un(x,t)−Un(x))∫Ttzϵ(s)sαdsdtdH(x)+ |
+∫Ω∫T0un(x,t)−Un(x)]sign(un(x,t)−Un(x))∫Ttzϵ(s)sαdsdtdx+ |
+∫Ω∫T0h(t)fn(x,t)sign(un(x,t)−Un(x))∫Ttzϵ(s)sαdsdtdx. |
By the previous proof mentioned above, we deduce that
∫Ω∫T0∣un(x,t)−Un(x)∣zϵ(t)tαdt≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)) | (8.15) |
where C=C(αTα+1,h(T),Tα,∥g(un)∥L∞(R+),∣S∣)>0 is a constant. By letting ϵ→0+, then (8.15) reads
tα∫Ω∣un(x,t)−Un(x)∣dx≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)). | (8.16) |
By virtue of [24, Chapter V, Section 5.2.1, Theorem 1], then the semi-continuity of the total variation yields
tα∥u(⋅,t)−U(⋅)∥M+(Ω)≤lim infn→+∞tα∫Ω∣un(x,t)−Un(x)∣dx≤C(∥u0∥M+(Ω)+∥μ∥M+(Q)) | (8.17) |
for almost every t∈(0,T) and α>1. By considering to the limit as t→+∞ in the following inequality
∥u(⋅,t)−U(⋅)∥M+(Ω)≤Ctα(∥u0∥M+(Ω)+∥μ∥M+(Q)). |
Hence the statement (2.24) follows.
In this paper, we study the existence, uniqueness, decay estimates, and the asymptotic behavior of the Radon measure-valued solutions for a class of nonlinear parabolic equations with a source term and nonzero Neumann boundary conditions. To attain this, we use the natural approximation method, the definition of the weak solutions, and the properties of the Radon measure. Concerning the study of the existence and uniqueness of the solutions to the problem (P), we first show that the source term corresponding to the solution of the linear inhomogeneous heat equation with measure data is a unique Radon measure-valued. Moreover, we establish the decay estimates of these solutions by using the suitable test functions and the auxiliary functions. Finally, we analyze the asymptotic behavior of these solutions by establishing the decay estimate of the difference between the solution to the problem (P) and the solution to the steady state problem (E).
This work was partially supported by National Natural Sciences Foundation of China, grant No: 11571057
The authors declare no conflict of interest.
[1] | J. L. Vázquez, Measure-valued solutions and the phenomenon of blow-down in logarithmic diffusion, J. Math. Anal. Appl., 352 (2009), 515-547. |
[2] | S. Itǒ, Fundamental solutions of parabolic differential equations and boundary value problems, Jnp. J. Math., 27 (1957), 55-102. |
[3] | S. Itǒ, A boundary value problem of partial differential equations of parabolic type, Duke Math. J., 24 (1957), 299-312. |
[4] |
C. S. Kahane, On the asymptotic behavior of solutions of parabolic equations, Czechoslovak Math. J., 33 (1983), 262-285. doi: 10.21136/CMJ.1983.101876
![]() |
[5] | C. V. Pao, Nonlinear Parabolic and Elliptic Equations, New York: Plenum Press, 1992. |
[6] | M. Bertsch, F. Smarrazzo, A. Tesei, Pseudo-parabolic regularization of forward-backward parabolic equations: Power-type nonlinearities, J. Reine Angew. Math., 712 (2016), 51-80. |
[7] |
M. Bertsch, F. Smarrazzo, A. Tesei, Pseudo-parabolic regularization of forward-backward parabolic equations: A logarithmic nonlinearity, Anal. PDE, 6 (2013), 1719-1754. doi: 10.2140/apde.2013.6.1719
![]() |
[8] | L. C. Evance, Partial differential equations, Graduate Studies in Mathematics, American Mathematical Society, Providence, R.I., 2010. |
[9] |
M. M. Porzio, F. Smarrazzo, A. Tesei, Radon measure-valued solutions for a class of quasilinear parabolic equations, Arch. Ration. Mech. Anal., 210 (2013), 713-772. doi: 10.1007/s00205-013-0666-0
![]() |
[10] |
M. M. Porzio, F. Smarrazzo, A. Tesei, Radon measure-valued solutions of nonlinear strongly degenerate parabolic equations, Calc. Var. Partial Dif., 51 (2014), 401-437. doi: 10.1007/s00526-013-0680-y
![]() |
[11] |
F. Smarrazzo, A. Tesei, Degenerate regularization of forward-backward parabolic equations: The regularized problem, Arch. Ration. Mech. Anal., 204 (2012), 85-139. doi: 10.1007/s00205-011-0470-7
![]() |
[12] |
L. Orsina, M. M. Porzio, F. Smarrazzo, Measure-valued solutions of nonlinear parabolic equations with logarithmic diffusion, J. Evol. Equ., 15 (2015), 609-645. doi: 10.1007/s00028-015-0275-5
![]() |
[13] |
M. M. Porzio, F. Smarrazzo, Radon measure-valued solutions for some quasilinear degenerate elliptic equations, Ann. Mat. Pura Appl., 194 (2015), 495-532. doi: 10.1007/s10231-013-0386-y
![]() |
[14] |
F. Petitta, A. C. Ponce, A. Porretta, Approximation of diffuse measures for parabolic capacities, C. R. Math. Acad. Sci. Paris, 346 (2008), 161-166. doi: 10.1016/j.crma.2007.12.002
![]() |
[15] |
J. Droniou, A. Porretta, A. Prignet, Parabolic capacity and soft measures for nonlinear equations, Potential Anal., 19 (2003), 99-161. doi: 10.1023/A:1023248531928
![]() |
[16] |
F. Petitta, A. C. Ponce, A. Porretta, Diffuse measures and nonlinear parabolic equations, J. Evol. Equ., 11 (2011), 861-905. doi: 10.1007/s00028-011-0115-1
![]() |
[17] | J. L. Vázquez, The Porous Medium Equation: Mathematical Theory, Oxford: Oxford Mathematical Monographs, 2007. |
[18] |
J. R. Anderson, Local existence and uniqueness of solutions of degenerate parabolic equations, Commun. Part. Diff. Eq., 16 (1991), 105-143. doi: 10.1080/03605309108820753
![]() |
[19] |
E. DiBenedetto, Continuity of weak solutions to a general porous medium equation, Indiana Univ. Math. J., 32 (1983), 83-118. doi: 10.1512/iumj.1983.32.32008
![]() |
[20] |
J. Ding, B. Z. Guo, Global existence and blow-up solutions for quasilinear reaction-diffusion equations with a gradient term, Appl. Math. Lett., 24 (2011), 936-942. doi: 10.1016/j.aml.2010.12.052
![]() |
[21] | J. L. Lions, Quelques Méthodes de Résolutions des Problèmes aux Limites non Linéaires, Paris: Dunod., French, 1969. |
[22] | O. A. Ladyženskaja, V. A. Solonnikov, N. N. Ural'ceva, Linear and Quasilinear Equations of Parabolic Type, Translations of Mathematical Monographs, American Mathematical Society, Providence, R.I., 23 (1968). |
[23] | J. Simon, Compact sets in the space Lp(0,T;B), Ann. Mat. Pura Appl., 146 (1987), 65-96. |
[24] | L. C. Evans, R. F. Gariepy, Measure Theory and Fine Properties of Functions, Studies in Advanced Mathematics, CRC Press, Boca Raton, FL, 1992. |
[25] | M. Papi, M. M. Porzio, F. Smarrazzo, Existence of solutions to a class of weakly coercive diffusion equations with singular initial data, Adv. Differential Eq., 22 (2017), 893-963. |
[26] |
L. Boccardo, L. Moreno-Mérida, W1,1(Ω) Solutions of nonlinear problems with nonhomogeneous Neumann boundary conditions, Milan J. Math., 83 (2015), 279-293. doi: 10.1007/s00032-015-0235-0
![]() |
[27] | E. DiBenedetto, Degenerate Parabolic Equations, Universitext, New York: Springer-Verlag, 1993. |
[28] | H. Amann, Quasilinear parabolic system under nonlinear boundary conditions, Arch. Rational Mech. Anal., 92 (1986), 53-192. |
[29] | H. Amann, Parabolic evolution equations and nonlinear boundary conditions, J. Differ. Equations, 73 (1988), 201-269. |
[30] |
L. Dung, Remarks on Hölder continuity for parabolic equations and convergence to global attractors, Nonlinear Anal., 41 (2000), 921-941. doi: 10.1016/S0362-546X(98)00319-8
![]() |
[31] |
V. Bögelein, F. Duzaar, U. Gianazza, Porous medium type equations with measure data and potential estimates, SIAM J. Math. Anal., 45 (2013), 3283-3330. doi: 10.1137/130925323
![]() |
[32] |
V. Liskevich, I. I. Skrypnik, Pointwise estimates for solutions to the porous medium equation with measure as a forcing term, Israel J. Math., 194 (2013), 259-275. doi: 10.1007/s11856-012-0098-9
![]() |
[33] | U. Gianazza, Degenerate and singular porous medium type equations with measure data, Elliptic and Parabolic Equations, 119 (2015), 139-158. |
[34] | H. Brézis, W. A. Strauss, Semi-linear second-order elliptic equations in L1, J. Math. Soc. Japan, 25 (1973), 565-590. |
[35] | J. R. Cannon, The One-dimensional Heat Equations, Encyclopedia of Mathematics and its Applications, Addison-Wesley Publishing Company, 23 (1984). |
[36] |
G. Q. Chen, H. Frid, Extended divergence measure fields and the Euler equations for gas dynamic, Comm. Math, Phys., 236 (2003), 251-280. doi: 10.1007/s00220-003-0823-7
![]() |
[37] |
B. P. Andreianov, N. Igbida, Uniqueness for inhomogeneous Dirichlet problem for elliptic-parabolic equations, Proc. Roy. Soc. Edinburgh Sect. A, 137 (2007), 1119-1133 doi: 10.1017/S0308210505001290
![]() |
[38] |
L. Boccardo, J. M. Mazón, Existence of solutions and regularizing effect for some elliptic nonlinear problems with nonhomogeneous Neumann boundary conditions, Rev. Mat. Complut., 28 (2015), 263-280. doi: 10.1007/s13163-014-0162-6
![]() |
[39] | F. Audreu, N. Igbida, J. M. Mazón, J. Toledo, Degenerate elliptic equations with nonlinear boundary conditions and measure data, Ann. Sc. Norm. Super. Pisa. Cl. Sci., 8 (2009), 769-803. |
[40] | F. Audreu, J. M. Mazón, S. Segura De Léon, J. Toledo, Quasilinear elliptic and parabolic equations in L1 with nonlinear boundary conditions, Adv. Math. Sci. Appl., 7 (1997), 183-213. |
[41] |
M. M. Porzio, F. Smarrazzo, Radon measure-valued solutions for some quasilinear degenerate elliptic equations, Ann. Mat. Pura Appl., 194 (2015), 495-532. doi: 10.1007/s10231-013-0386-y
![]() |
[42] | S. Itô, The fundamental solution of the parabolic equation in a differentiable manifold. II, Osaka Math. J., 6 (1954), 167-185. |
[43] | A. Friedman, Partial Differential Equations of Parabolic Type, Prentice-Hall, Inc., Englewood Cliffs, N. J., 1964. |
[44] | M. Giaquinta, G. Modica, J. Souček, Cartesian Currents in the Calculus of Variations. I. Cartesian Currents, Berlin: Springer-Verlag, 1998. |
[45] | M. Bertsch, F. Smarrazzo, A. Terracina, A. Tesei, Radon measure-valued solutions of first order scalar conservation laws, Adv. Nonlinear Anal., 9 (2020), 65-107. |
[46] |
Q. S. Nkombo, F. Li, Radon Measure-valued solutions for nonlinear strongly degenerate parabolic equations with measure data, Eur. J. Pure Appl. Math., 14 (2021), 204-233. doi: 10.29020/nybg.ejpam.v14i1.3877
![]() |
[47] |
U. S. Fjordholm, S. Mishra, E. Tadmor, On the computation of measure-valued solutions, Acta Numer., 25 (2016), 567-679. doi: 10.1017/S0962492916000088
![]() |
[48] |
S. Lanthaler, S. Mishra, Computation of measure-valued solutions for the incompressible Euler equations, Math. Models Methods Appl. Sci., 25 (2015), 2043-2088. doi: 10.1142/S0218202515500529
![]() |
[49] | J. Droniou, T. Gallouët, R. Herbin, A finite volume scheme for a noncoercive elliptic equation with measure data, SIAM J. Numer. Anal., 41 (2003), 1997-2031. |
[50] |
R. Eymard, T. Gallouët, R. Herbin, A. Michel, Convergence of a finite volume scheme for nonlinear degenerate parabolic equations, Numer. Math., 92 (2002), 41-82. doi: 10.1007/s002110100342
![]() |
1. | Quincy Stévène Nkombo, Fengquan Li, Christian Tathy, Existence and asymptotic behavior of Radon measure-valued solutions for a class of nonlinear parabolic equations, 2021, 2021, 1687-1847, 10.1186/s13662-021-03668-3 |