This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\sauthor

Daiki Mizuno \saffiliationDivision of Mathematics and Informatics,
Graduate School of Science and Engineering, Chiba University,
1-33, Yayoi-cho, Inage-ku, 263-8522, Chiba, Japan \semail[email protected]

\tauthor

Ken Shirakawa \taffiliationDepartment of Mathematics, Faculty of Education, Chiba University
1–33 Yayoi-cho, Inage-ku, 263–8522, Chiba, Japan \temail[email protected]

\footcomment

 This work is supported by Grant-in-Aid for Scientific Research (C) No. 20K03672, JSPS. HA and KS are also partially supported by NSF grant DMS-2110263, the Air Force Office of Scientific Research (AFOSR) under Award NO: FA9550-22-1-0248, and the Office of Naval Research (ONR) under Award NO: N00014-24-1-2147.
AMS Subject Classification: 35A15, 35G50, 35G61, 35K70, 74N20.
Keywords: KWC-type system, pseudo-parabolic nature, existence, uniqueness, regularity, continuous-dependence

Well-posedness of a Pseudo-Parabolic KWC System
in Materials Science 11footnotemark: 1

Harbir Antil Department of Mathematical Sciences and
the Center for Mathematics and Artificial Intelligence, (CMAI),
George Mason University, Fairfax, VA 22030, USA
[email protected]

Abstract. The original KWC-system is widely used in materials science. It was proposed in [Kobayashi et al, Physica D, 140, 141–150 (2000)] and is based on the phase field model of planar grain boundary motion. This model suffers from two key challenges. Firstly, it is difficult to establish its relation to physics, in particular, a variational model. Secondly, it lacks uniqueness. The former has been recently studied within the realm of BV-theory. The latter only holds under various simplifications. This article introduces a pseudo-parabolic version of the KWC-system. A direct relationship with variational model (as gradient-flow) and uniqueness are established without making any unrealistic simplifications. Namely, this is the first KWC-system which is both physically and mathematically valid. The proposed model overcomes the well-known open issues.

1 Introduction

The Kobayashi–Warren–Carter (KWC) system consists of a set of non-smooth parabolic PDEs and is widely used in materials science [14, 15]. It is based on the phase field model of planar grain boundary motion. This model suffers from two fundamental challenges: (i) Physics: It is difficult to establish KWC-system as the gradient flow of a variational model; (ii) Mathematics: The solution to KWC-system are known to be unique only under special cases. Both of these challenges makes it difficult to rigorously use this model in practice or carry our new material design via optimization [5, 6].

This article aims to overcome both these challenges by introducing a pseudo-parabolic version of the KWC-system. Well-posedness (both existence and uniqueness) of the resulting system (S), which arise from gradient flow based on the KWC-energy

:[η,θ][L2(Ω)]2(η,θ):=12Ω|η|2𝑑x+ΩG(η)𝑑x\displaystyle\mathcal{F}:[\eta,\theta]\in[L^{2}(\Omega)]^{2}\mapsto\mathcal{F}(\eta,\theta):=\frac{1}{2}\int_{\Omega}|\nabla\eta|^{2}\,dx+\int_{\Omega}G(\eta)\,dx (1.1)
+Ωα(η)|Dθ|[0,],\displaystyle+\int_{\Omega}\alpha(\eta)|D\theta|\in[0,\infty],

is established. Namely, this article addresses open issues from previous works that deal with the KWC-system (cf. [18, 19, 22, 20, 21, 17]) and its regularized versions (cf. [13, 24, 7, 5, 6, 16]).

Next, we describe the system (S). Let 0<T<0<T<\infty be a fixed final time, and let N{1,2,3}N\in\{1,2,3\} denote the spatial dimension. Let ΩN\Omega\subset\mathbb{R}^{N} be an open bounded spatial domain with a boundary Γ:=Ω\Gamma:=\partial\Omega. When N>1N>1, Γ\Gamma is assumed to be sufficiently smooth, with the unit outer normal nΓn_{\Gamma}. Besides, we let Q:=(0,T)×ΩQ:=(0,T)\times\Omega and Σ:=(0,T)×Γ\Sigma:=(0,T)\times\Gamma. Then the pseudo-parabolic system denoted by (S), with two constants μ0\mu\geq 0 and ν>0\nu>0, is given by

(S):\displaystyle\mbox{(S)}:~{}~{}
{tηΔ(η+μ2tη)+g(η)+α(η(t))|θ|=u(t,x),for (t,x)Q,(η+μ2tη)nΓ=0,on Σ,η(0,x)=η0(x),for xΩ,\displaystyle\begin{cases}\partial_{t}\eta-\mathit{\Delta}\bigl{(}\eta+\mu^{2}\partial_{t}\eta\bigr{)}+g(\eta)+\alpha^{\prime}(\eta(t))|\nabla\theta|=u(t,x),~{}\mbox{for $(t,x)\in Q$,}\\[2.15277pt] \nabla\bigl{(}\eta+\mu^{2}\partial_{t}\eta\bigr{)}\cdot n_{\Gamma}=0,~{}\mbox{on $\Sigma$,}\ \\[2.15277pt] \eta(0,x)=\eta_{0}(x),~{}\mbox{for $x\in\Omega$,}\end{cases}
{α0(η)tθdiv(α(η)Dθ|Dθ|+ν2tθ)=v(t,x),for (t,x)Q,(α(η)Dθ|Dθ|+ν2tθ)nΓ=0on Σ,θ(0,x)=θ0(x),for xΩ.\displaystyle\begin{cases}\displaystyle\alpha_{0}(\eta)\partial_{t}\theta-\mathrm{div}\left(\alpha(\eta)\frac{D\theta}{|D\theta|}+\nu^{2}\nabla\partial_{t}\theta\right)=v(t,x),~{}\mbox{for $(t,x)\in Q$,}\\[4.30554pt] \bigl{(}\alpha(\eta)\frac{D\theta}{|D\theta|}+\nu^{2}\nabla\partial_{t}\theta\bigr{)}\cdot n_{\Gamma}=0~{}\mbox{on $\Sigma$,}\\[2.15277pt] \theta(0,x)=\theta_{0}(x),~{}\mbox{for $x\in\Omega$.}\end{cases}

Here, the unknowns η=η(t,x)\eta=\eta(t,x) and θ=θ(t,x)\theta=\theta(t,x) are order parameters that indicate the orientation order and orientation angle of the polycrystal body, respectively. Besides, η0=η0(x)\eta_{0}=\eta_{0}(x) and θ0=θ0(x)\theta_{0}=\theta_{0}(x) is the initial data. Moreover, u=u(t,x)u=u(t,x) and v=v(t,x)v=v(t,x) are the forcing terms. Additionally, α0=α0(η)\alpha_{0}=\alpha_{0}(\eta) and α=α(η)\alpha=\alpha(\eta) are fixed positive-valued functions to reproduce the mobilities of grain boundary motions. Finally, g=g(η)g=g(\eta) is a perturbation for the orientation order η\eta, having a nonnegative potential G=G(η)G=G(\eta), i.e. ddηG(η)=g(η)\frac{d}{d\eta}G(\eta)=g(\eta).

A generic form of the “KWC-system” is given by the evolution equation (cf. [14, 15]):

𝒜0(η(t))ddt[η(t)θ(t)]=δδ[η,θ](η(t),θ(t))+[u(t)v(t)]\displaystyle-\mathcal{A}_{0}\bigl{(}\eta(t)\bigr{)}\frac{d}{dt}\left[\begin{matrix}\eta(t)\\[4.30554pt] \theta(t)\end{matrix}\right]=\frac{\delta}{\delta[\eta,\theta]}\mathcal{F}(\eta(t),\theta(t))+\left[\begin{matrix}u(t)\\[4.30554pt] v(t)\end{matrix}\right] (1.2)
in [L2(Ω)]2[L^{2}(\Omega)]^{2}, for t(0,T)t\in(0,T),

which is motivated by the gradient flow of the free-energy, namely the KWC-energy (1.1), with a functional derivative δδ[η,θ]\frac{\delta}{\delta[\eta,\theta]}\mathcal{F}, and an unknown-dependent monotone operator 𝒜0(η)[L2(Ω)]2\mathcal{A}_{0}(\eta)\subset[L^{2}(\Omega)]^{2}. Here, the evolution equation (1.2) can be considered as the common root of the original KWC-system (cf. [14, 15]) and our system (S). Indeed, our system (S) is derived from the evolution equation (1.2), in the case when:

𝒜0(η):[η~θ~][H2(Ω)]2𝒜0(η)[η~θ~]:=[η~μ2Δη~α0(η)θ~ν2Δθ~][L2(Ω)]2,\displaystyle\mathcal{A}_{0}(\eta):\left[\begin{matrix}\widetilde{\eta}\\[4.30554pt] \widetilde{\theta}\end{matrix}\right]\in[H^{2}(\Omega)]^{2}\mapsto\mathcal{A}_{0}(\eta)\left[\begin{matrix}\widetilde{\eta}\\[4.30554pt] \widetilde{\theta}\end{matrix}\right]:=\left[\begin{matrix}\widetilde{\eta}-\mu^{2}\mathit{\Delta}\widetilde{\eta}\\[4.30554pt] \alpha_{0}(\eta)\widetilde{\theta}-\nu^{2}\mathit{\Delta}\widetilde{\theta}\end{matrix}\right]\in[L^{2}(\Omega)]^{2}, (1.3)
for each ηL2(Ω)\eta\in L^{2}(\Omega), subject to the zero-Neumann boundary condition,

while the original KWC-system corresponds to the case when μ=ν=0\mu=\nu=0.

In recent years, the principal issue has been to clarify the variational structure (representation) of the functional derivative δδ[η,θ]\frac{\delta}{\delta[\eta,\theta]}\mathcal{F} of the nonsmooth and nonconvex energy \mathcal{F} in (1.1). The positive answer to the issue was obtained in [18, 19, 22], by means of BV-theory (cf. [4, 11, 12, 8, 2, 3]), and this work has provided a basis of the study of KWC-system, e.g. the existence and large-time behavior [18, 19], the observations under other boundary conditions [20, 21], the time-periodic solution [17], and so on.

However, the uniqueness of solutions has been a significant challenge, due to the velocity term α0(η)tθ\alpha_{0}(\eta)\partial_{t}\theta and the singular diffusion flux α(η)Dθ|Dθ|\alpha(\eta)\frac{D\theta}{|D\theta|}, both of which depend on the unknown-dependent mobilities. Therefore, previous researchers have implemented the following modifications to the modelling framework (1.1) and (1.3):

\bullet

resetting α0\alpha_{0} to be a function which is independent of η\eta (effectively a constant);

\bullet

modifying the free-energy functional to a more relaxed form:

ε:[η,θ][L2(Ω)]2ε(η,θ)=(η,θ)+ε22Ω|θ|2𝑑x,\displaystyle\mathcal{F}_{\varepsilon}:[\eta,\theta]\in[L^{2}(\Omega)]^{2}\mapsto\mathcal{F}_{\varepsilon}(\eta,\theta)=\mathcal{F}(\eta,\theta)+\frac{\varepsilon^{2}}{2}\int_{\Omega}|\nabla\theta|^{2}\,dx,
with a small constant ε>0\varepsilon>0.

These modifications have been pivotal in addressing the uniqueness challenges (cf. [13, 24]), and several advanced issues, such as the optimal control problems (see [7, 5, 6, 16]).

In light of this, we can expect that the pseudo-parabolic nature of our system will effectively address the uniqueness challenge. This is due to the positive constants μ\mu and ν\nu in (1.3), which are expected to bring a smoothing effect for the regularity of solution. In addition, it is also crucial from a mathematical perspective to clarify the similarities and differences between our pseudo-parabolic system and the original parabolic KWC-system.

Consequently, we set the goal to prove the following two Main Theorems, concerned with the well-posedness of our pseudo-parabolic system (S), i.e. the evolution equation (1.2) under (1.1) and (1.3).

Main Theorem 1.

Existence and regularity of solution to (S).

Main Theorem 2.

Uniqueness of solution to (S), and continuous dependence with respect to the initial data and forcings.

These Main Theorems will provide the positive answer to our earlier expectation regarding the effectiveness of the pseudo-parabolic nature of our system in resolving the uniqueness issue. Also, the Main Theorems will focus on two conflicting properties: the singularity in the diffusion flux α(η)Dθ|Dθ|\alpha(\eta)\frac{D\theta}{|D\theta|}; and the smoothing effect encouraged by the Laplacian in (1.3). This conflicting situation will be clarified by the differences in regularity between components: ηW1,2(0,T;H2(Ω))\eta\in W^{1,2}(0,T;H^{2}(\Omega)); and θW1,2(0,T;H1(Ω))L(0,T;H2(Ω))\theta\in W^{1,2}(0,T;H^{1}(\Omega))\cap L^{\infty}(0,T;H^{2}(\Omega)); in the Main Theorem 1. Moreover, the results of this paper will form a fundamental part of the optimization problem in grain boundary motion, which will be exploring in a forthcoming paper.

Outline: Preliminaries are given in Section 1, and on this basis, the Main Theorems are stated in Section 2. For the proofs of Main Theorems, we prepare Section 3 to set up an approximation method for (S). Based on these, the Main Theorems are proved in Section 4, by means of the auxiliary results obtained in Section 3.

2 Preliminaries

We begin by prescribing the notations used throughout this paper.

Notations in real analysis. We define:

rs:=max{r,s} and rs:=min{r,s}, for all r,s[,],\displaystyle r\vee s:=\max\{r,s\}~{}\mbox{ and }~{}r\wedge s:=\min\{r,s\},\mbox{ for all $r,s\in[-\infty,\infty]$,}

and especially, we write:

[r]+:=r0 and [r]:=(r0), for all r[,].\displaystyle[r]^{+}:=r\vee 0~{}\mbox{ and }~{}[r]^{-}:=-(r\wedge 0),\mbox{ for all $r\in[-\infty,\infty]$.}

Additionally, for any M>0M>0, let 𝒯M:[M,M]\mathcal{T}_{M}:\mathbb{R}\longrightarrow[-M,M] be the truncation operator, defined as:

𝒯M:r(r(M))M[M,M].\mathcal{T}_{M}:r\in\mathbb{R}\mapsto(r\vee(-M))\wedge M\in[-M,M].

Let dd\in\mathbb{N} be a fixed dimension. We denote by |y||y| and yzy\cdot z the Euclidean norm of ydy\in\mathbb{R}^{d} and the scalar product of y,zdy,z\in\mathbb{R}^{d}, respectively, i.e.,

|y|:=y12++yd2 and yz:=y1z1++ydzd, for all y=[y1,,yd],z=[z1,,zd]d.\begin{array}[]{c}|y|:=\sqrt{y_{1}^{2}+\cdots+y_{d}^{2}}\mbox{ \ and \ }y\cdot z:=y_{1}z_{1}+\cdots+y_{d}z_{d},\\[4.30554pt] \mbox{ for all $y=[y_{1},\ldots,y_{d}],~{}z=[z_{1},\ldots,z_{d}]\in\mathbb{R}^{d}$.}\end{array}

Besides, we let:

𝔹d:={yd|y|<1} and 𝕊d1:={yd|y|=1}.\displaystyle\mathbb{B}^{d}:=\left\{\begin{array}[]{l|l}y\in\mathbb{R}^{d}&|y|<1\end{array}\right\}~{}\mbox{ and }~{}\mathbb{S}^{d-1}:=\left\{\begin{array}[]{l|l}y\in\mathbb{R}^{d}&|y|=1\end{array}\right\}.

We denote by d\mathcal{L}^{d} the dd-dimensional Lebesgue measure, and we denote by d\mathcal{H}^{d} the dd-dimensional Hausdorff measure. In particular, the measure theoretical phrases, such as “a.e.”, “dtdt”, and “dxdx”, and so on, are all with respect to the Lebesgue measure in each corresponding dimension. Also on a Lipschitz-surface SS, the phrase “a.e.” is with respect to the Hausdorff measure in each corresponding Hausdorff dimension. In particular, if SS is C1C^{1}-surface, then we simply denote by dSdS the area-element of the integration on SS.

For a Borel set EdE\subset\mathbb{R}^{d}, we denote by χE:d{0,1}\chi_{E}:\mathbb{R}^{d}\longrightarrow\{0,1\} the characteristic function of EE. Additionally, for a distribution ζ\zeta on an open set in d\mathbb{R}^{d} and any i{1,,d}i\in\{1,\dots,d\}, let iζ\partial_{i}\zeta be the distributional differential with respect to ii-th variable of ζ\zeta. As well as we consider, the differential operators, such as ,div,2\nabla,\ \operatorname{div},\ \nabla^{2}, and so on, in distributional sense.

Abstract notations. (cf. [9, Chapter II]) For an abstract Banach space XX, we denote by ||X|\cdot|_{X} the norm of XX, and denote by ,X\langle\cdot,\cdot\rangle_{X} the duality pairing between XX and its dual XX^{*}. In particular, when XX is a Hilbert space, we denote by (,)X(\cdot,\cdot)_{X} the inner product of XX.

For two Banach spaces XX and YY, let (X;Y)\mathscr{L}(X;Y) be the Banach space of bounded linear operators from XX into YY.

For Banach spaces X1,,XdX_{1},\dots,X_{d} with 1<d1<d\in\mathbb{N}, let X1××XdX_{1}\times\dots\times X_{d} be the product Banach space endowed with the norm ||X1××Xd:=||X1++||Xd|\cdot|_{X_{1}\times\cdots\times X_{d}}:=|\cdot|_{X_{1}}+\cdots+|\cdot|_{X_{d}}. However, when all X1,,XdX_{1},\dots,X_{d} are Hilbert spaces, X1××XdX_{1}\times\dots\times X_{d} denotes the product Hilbert space endowed with the inner product (,)X1××Xd:=(,)X1++(,)Xd(\cdot,\cdot)_{X_{1}\times\cdots\times X_{d}}:=(\cdot,\cdot)_{X_{1}}+\cdots+(\cdot,\cdot)_{X_{d}} and the norm ||X1××Xd:=(||X12++||Xd2)12|\cdot|_{X_{1}\times\cdots\times X_{d}}:=\bigl{(}|\cdot|_{X_{1}}^{2}+\cdots+|\cdot|_{X_{d}}^{2}\bigr{)}^{\frac{1}{2}}. In particular, when all X1,,XdX_{1},\dots,X_{d} coincide with a Banach space YY, the product space X1××XdX_{1}\times\dots\times X_{d} is simply denoted by [Y]d[Y]^{d}.

Basic notations. Let 0<T<0<T<\infty be a fixed constant of time, and let N{1,2,3}N\in\{1,2,3\} is a fixed dimension. Let ΩN\Omega\subset\mathbb{R}^{N} be a bounded domain with a boundary Γ:=Ω\Gamma:=\partial\Omega, and when N>1N>1, Γ\Gamma has CC^{\infty}-regularity with the unit outer normal nΓn_{\Gamma}. Additionally, as notations of base spaces, we let:

H:=L2(Ω),V:=H1(Ω),:=L2(0,T;H), and 𝒱:=L2(0,T;V).H:=L^{2}(\Omega),\ V:=H^{1}(\Omega),~{}\mathscr{H}:=L^{2}(0,T;H),\mbox{ and }\mathscr{V}:=L^{2}(0,T;V).

Let W0H2(Ω)W_{0}\subset H^{2}(\Omega) be the closed linear subspace of HH, given by:

W0:={zH2(Ω)|znΓ=0 on Γ}.W_{0}:=\{z\in H^{2}(\Omega)\,|\,\nabla z\cdot n_{\Gamma}=0\mbox{ on }\Gamma\}.

Let ANA_{N} be a differential operator, defined as:

AN:zW0HANz:=ΔzH.A_{N}:\,z\in W_{0}\subset H\mapsto A_{N}z:=-\mathit{\Delta}z\in H.

It is well-known that ANH×HA_{N}\subset H\times H is linear, positive, and self-adjoint, and the domain W0W_{0} is a Hilbert space, endowed with the inner product:

(z1,z2)W0:=(z1,z2)H+(ANz1,z2)H(=(z1,z2)V), for zkW0k=1,2.\displaystyle(z_{1},z_{2})_{W_{0}}:=(z_{1},z_{2})_{H}+(A_{N}z_{1},z_{2})_{H}~{}\bigl{(}=(z_{1},z_{2})_{V}\bigr{)},\mbox{ for $z_{k}\in W_{0}$, $k=1,2$.}

Moreover, there exists a positive constant C0C_{0} such that:

|z|H2(Ω)2C0(|ANz|H2+|z|H2), for all zW0.\displaystyle|z|_{H^{2}(\Omega)}^{2}\leq C_{0}\bigl{(}|A_{N}z|_{H}^{2}+|z|_{H}^{2}\bigr{)},~{}\mbox{ for all $z\in W_{0}$.} (2.1)

Notations for the time-discretization. Let τ>0\tau>0 be a constant of the time step-size, and let {ti}i=0[0,)\{t_{i}\}_{i=0}^{\infty}\subset[0,\infty) be the time sequence defined as:

ti:=iτ,i=0,1,2,.t_{i}:=i\tau,\ i=0,1,2,\ldots.

Let XX be a Banach space. Then, for any sequence {[ti,zi]}i=0[0,)×X\{[t_{i},z_{i}]\}_{i=0}^{\infty}\subset[0,\infty)\times X, we define the forward time-interpolation [z¯]τLloc([0,);X)[\overline{z}]_{\tau}\in L^{\infty}_{\mathrm{loc}}([0,\infty);X), the backward time-interpolation [z¯]τLloc([0,);X)[\underline{z}]_{\tau}\in L^{\infty}_{\mathrm{loc}}([0,\infty);X) and the linear time-interpolation [z]τWloc1,2([0,);X)[z]_{\tau}\in W^{1,2}_{\mathrm{loc}}([0,\infty);X), by letting:

{[z¯]τ(t):=χ(,0]z0+i=1χ(ti1,ti](t)zi,[z¯]τ(t):=i=0χ(ti,ti+1](t)zi,[z]τ(t):=i=1χ[ti1,ti)(t)(tti1τzi+titτzi1),inX,fort0,\left\{\begin{aligned} &[\overline{z}]_{\tau}(t):=\chi_{(-\infty,0]}z_{0}+\sum_{i=1}^{\infty}\chi_{(t_{i-1},t_{i}]}(t)z_{i},\\ &[\underline{z}]_{\tau}(t):=\sum_{i=0}^{\infty}\chi_{(t_{i},t_{i+1}]}(t)z_{i},\\ &[z]_{\tau}(t):=\sum_{i=1}^{\infty}\chi_{[t_{i-1},t_{i})}(t)\left(\frac{t-t_{i-1}}{\tau}z_{i}+\frac{t_{i}-t}{\tau}z_{i-1}\right),\end{aligned}\right.~{}{\rm in}~{}X,\ {\rm for}~{}t\geq 0,

respectively.

In the meantime, for any q[1,)q\in[1,\infty) and any ζLlocq([0,);X)\zeta\in L_{\mathrm{loc}}^{q}([0,\infty);X), we denote by {ζi}i=0X\{\zeta_{i}\}_{i=0}^{\infty}\subset X the sequence of time-discretization data of ζ\zeta, defined as:

ζ0:=0 in X, and ζi:=1τti1tiζ(ς)𝑑ς in X,   for i=1,2,3,.\displaystyle\zeta_{0}:=0\mbox{ in $X$, and }\zeta_{i}:=\frac{1}{\tau}\int_{t_{i-1}}^{t_{i}}\zeta(\varsigma)\,d\varsigma~{}\mbox{ in $X$, ~{} for $i=1,2,3,\dots$.} (2.2a)
As is easily checked, the time-interpolations [ζ¯]τ,[ζ¯]τLlocq([0,);X)[\overline{\zeta}]_{\tau},[\underline{\zeta}]_{\tau}\in L^{q}_{\mathrm{loc}}([0,\infty);X) for the above {ζi}i=0\{\zeta_{i}\}_{i=0}^{\infty} fulfill that:
[ζ¯]τζ and [ζ¯]τζ in Llocq([0,);X), as τ0.\displaystyle[\overline{\zeta}]_{\tau}\to\zeta\mbox{ and }[\underline{\zeta}]_{\tau}\to\zeta\mbox{ in $L^{q}_{\mathrm{loc}}([0,\infty);X)$, as $\tau\downarrow 0$.} (2.2b)

3 Main results

In this paper, the main assertions are discussed under the following assumptions.

  • (A1)

    μ>0\mu>0 and ν>0\nu>0 are fixed constants.

  • (A2)

    g:g:\mathbb{R}\longrightarrow\mathbb{R} is a locally Lipschitz continuous function with a nonnegative primitive GC1()G\in C^{1}(\mathbb{R}). Moreover, gg satisfies the following condition:

    lim infξg(ξ)=,lim supξg(ξ)=.\liminf_{\xi\downarrow-\infty}g(\xi)=-\infty,\ \limsup_{\xi\uparrow\infty}g(\xi)=\infty.
  • (A3)

    α0:(0,)\alpha_{0}:\mathbb{R}\longrightarrow(0,\infty) is a locally Lipschitz continuous function, and α:[0,)\alpha:\mathbb{R}\longrightarrow[0,\infty) is a C1C^{1}-class convex function, such that:

    α(0)=0 and δα:=infα0()>0.\alpha^{\prime}(0)=0\mbox{ and }\delta_{\alpha}:=\inf\alpha_{0}(\mathbb{R})>0.
  • (A4)

    u,vLloc2([0,);H)u,v\in L_{\rm loc}^{2}([0,\infty);H), and uL(Q)u\in L^{\infty}(Q).

  • (A5)

    The initial data [η0,θ0][\eta_{0},\theta_{0}] belong to the class [W0]2[H2(Ω)]2[W_{0}]^{2}\subset[H^{2}(\Omega)]^{2}.

Now, the main results are stated as follows.

Main Theorem 1 (Existence and regularity).

Under the assumptions (A1)–(A5), the system (S) admits a solution [η,θ][]2[\eta,\theta]\in[\mathscr{H}]^{2} in the following sense.

  • (S0)

    [η,θ][W1,2(0,T;W0)L(Q)]×[W1,2(0,T;V)×L(0,T;W0)][\eta,\theta]\in\bigl{[}W^{1,2}(0,T;W_{0})\cap L^{\infty}(Q)\bigr{]}\times\bigl{[}W^{1,2}(0,T;V)\times L^{\infty}(0,T;W_{0})\bigr{]}.

  • (S1)

    η\eta solves the following variational identity:

    (tη(t),φ)H+((η+μ2tη)(t),φ)[H]N+(g(η(t)),φ)H\displaystyle(\partial_{t}\eta(t),\varphi)_{H}+(\nabla(\eta+\mu^{2}\partial_{t}\eta)(t),\nabla\varphi)_{[H]^{N}}+(g(\eta(t)),\varphi)_{H}
    +(α(η(t))|θ(t)|,φ)H=(u(t),φ)H,\displaystyle+(\alpha^{\prime}(\eta(t))|\nabla\theta(t)|,\varphi)_{H}=(u(t),\varphi)_{H},
     for any φV, and a.e. t(0,T).\displaystyle\mbox{ for any }\varphi\in V,\mbox{ and a.e. }t\in(0,T).
  • (S2)

    θ\theta solves the following variational inequality:

    ((α0(η)tθ)(t),θ(t)ψ)H+ν2(tθ(t),(θ(t)ψ))[H]N\displaystyle\bigl{(}(\alpha_{0}(\eta)\partial_{t}\theta)(t),\theta(t)-\psi\bigr{)}_{H}+\nu^{2}(\nabla\partial_{t}\theta(t),\nabla(\theta(t)-\psi))_{[H]^{N}}
    +Ωα(η(t))|θ(t)|𝑑xΩα(η(t))|ψ|𝑑x+(v(t),θ(t)ψ)H,\displaystyle+\int_{\Omega}\alpha(\eta(t))|\nabla\theta(t)|\,dx\leq\int_{\Omega}\alpha(\eta(t))|\nabla\psi|\,dx+(v(t),\theta(t)-\psi)_{H},
     for any ψV, and a.e. t(0,T).\displaystyle\mbox{ for any }\psi\in V,\mbox{ and a.e. }t\in(0,T).
  • (S3)

    [η(0),θ(0)]=[η0,θ0][\eta(0),\theta(0)]=[\eta_{0},\theta_{0}] in [H]2[H]^{2}.

Main Theorem 2 (Uniqueness and continuous dependence).

Under the assumptions (A1)–(A5), let [ηk,θk],k=1,2[\eta^{k},\theta^{k}],\,k=1,2 be two solutions to (S) with two initial values [η0k,θ0k][\eta_{0}^{k},\theta_{0}^{k}] and two forcings [uk,vk][u^{k},v^{k}], k=1,2k=1,2. Then, there exists a constant C1=C1(ν)>0C_{1}=C_{1}(\nu)>0, depending on ν\nu, such that:

J(t)C1(ν)(J(0)+|u1u2|2+|v1v2|2),\displaystyle J(t)\leq C_{1}(\nu)\bigl{(}J(0)+|u^{1}-u^{2}|_{\mathscr{H}}^{2}+|v^{1}-v^{2}|_{\mathscr{H}}^{2}\bigr{)},
 for any t[0,T] and any T>0,\displaystyle\mbox{ for any }t\in[0,T]\mbox{ and any }T>0,

where

J(t):=\displaystyle J(t):= |(η1η2)(t)|H2+μ2|(η1η2)(t)|[H]N2+|α0(η1)(θ1θ2)(t)|H2\displaystyle|(\eta^{1}-\eta^{2})(t)|_{H}^{2}+\mu^{2}|\nabla(\eta^{1}-\eta^{2})(t)|_{[H]^{N}}^{2}+|\sqrt{\alpha_{0}(\eta^{1})}(\theta^{1}-\theta^{2})(t)|_{H}^{2}
+ν2|(θ1θ2)(t)|[H]N2, for any t0.\displaystyle\quad+\nu^{2}|\nabla(\theta^{1}-\theta^{2})(t)|_{[H]^{N}}^{2},\ \mbox{ for any }t\geq 0.

4 Approximating method

In the Main Theorems, the solution to (S) will be obtained by means of the time-discretization method. In this light, let τ(0,1)\tau\in(0,1) be a constant of the time-step size, and let ε(0,1)\varepsilon\in(0,1) be a relaxation constant. Based on this, we adopt the following time-discretization scheme (AP)ετ{}_{\tau}^{\varepsilon}, as our approximating problem of (S).

(AP)ετ{}_{\tau}^{\varepsilon}: To find {[ηi,θi]}i=0[W0]2\{[\eta_{i},\theta_{i}]\}_{i=0}^{\infty}\subset[W_{0}]^{2} satisfying:

ηiεηi1ετ+AN(ηiε+μ2τ(ηiεηi1ε))+g(𝒯Mηiε)\displaystyle\frac{\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}}{\tau}+A_{N}\left(\eta_{i}^{\varepsilon}+\frac{\mu^{2}}{\tau}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})\right)+g(\mathcal{T}_{M}\eta_{i}^{\varepsilon}) (4.1)
+α(𝒯Mηiε)γε(θiε)=ui in H,\displaystyle+\alpha^{\prime}(\mathcal{T}_{M}\eta_{i}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon})=u_{i}\mbox{ in }H,
α0(𝒯Mηi1ε)τ(θiεθi1ε)div(α~M(η)γε(θiε))\displaystyle\frac{\alpha_{0}(\mathcal{T}_{M}\eta_{i-1}^{\varepsilon})}{\tau}(\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon})-\operatorname{div}\bigl{(}\widetilde{\alpha}_{M}(\eta)\nabla\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon})\bigr{)} (4.2)
+ν2τAN(θiεθi1ε)=vi in H,\displaystyle+\frac{\nu^{2}}{\tau}A_{N}(\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon})=v_{i}\mbox{ in }H,
for i=1,2,3,, subject to [η0ε,θ0ε]=[η0,θ0] in [H]2.\displaystyle\mbox{for }i=1,2,3,\dots,\mbox{ subject to }[\eta_{0}^{\varepsilon},\theta_{0}^{\varepsilon}]=[\eta_{0},\theta_{0}]\mbox{ in }[H]^{2}.

In this context, γεC0,1(N)C(N)\gamma_{\varepsilon}\in C^{0,1}(\mathbb{R}^{N})\cap C^{\infty}(\mathbb{R}^{N}) is a smooth approximation of the Euclidean norm ||C0,1(N)|\cdot|\in C^{0,1}(\mathbb{R}^{N}), defined as:

γε:yNγε(y):=ε2+|y|2[0,).\gamma_{\varepsilon}:y\in\mathbb{R}^{N}\mapsto\gamma_{\varepsilon}(y):=\sqrt{\varepsilon^{2}+|y|^{2}}\in[0,\infty).

Also, we define an approximating free-energy ε\mathcal{F}_{\varepsilon} on [H]2[H]^{2}, by setting:

ε:[η,θ][H]2ε(η,θ):=12Ω|η|2𝑑x+ΩG~M(η)\displaystyle\mathcal{F}_{\varepsilon}:[\eta,\theta]\in[H]^{2}\mapsto\mathcal{F}_{\varepsilon}(\eta,\theta):=\frac{1}{2}\int_{\Omega}|\nabla\eta|^{2}\,dx+\int_{\Omega}\widetilde{G}_{M}(\eta) (4.3)
+Ωα~M(η)γε(Dθ), for any 0<ε<1.\displaystyle+\int_{\Omega}\widetilde{\alpha}_{M}(\eta)\gamma_{\varepsilon}(D\theta),\ \mbox{ for any }0<\varepsilon<1.

where α~MC1,1()\widetilde{\alpha}_{M}\in C^{1,1}(\mathbb{R}) and G~MC1,1()\widetilde{G}_{M}\in C^{1,1}(\mathbb{R}) are nonnegative primitives of α𝒯MW1,()\alpha\circ\mathcal{T}_{M}\in W^{1,\infty}(\mathbb{R}) and G𝒯MW1,()G\circ\mathcal{T}_{M}\in W^{1,\infty}(\mathbb{R}), respectively. Finally, ui,viu_{i},v_{i} are given as in (2.2).

The solution to (AP)ετ{}_{\tau}^{\varepsilon} is given as follows.

Definition 1.

The sequence of functions {[ηiε,θiε]}i=0\{[\eta_{i}^{\varepsilon},\theta_{i}^{\varepsilon}]\}_{i=0}^{\infty} is called a solution to (AP)ετ{}_{\tau}^{\varepsilon} iff. {[ηiε,θiε]}i=0[W0]2\{[\eta_{i}^{\varepsilon},\theta_{i}^{\varepsilon}]\}_{i=0}^{\infty}\subset[W_{0}]^{2}, and [ηiε,θiε][\eta_{i}^{\varepsilon},\theta_{i}^{\varepsilon}] fulfills (4.1) and (4.2) for any i=1,2,3,i=1,2,3,\dots.

In this paper, the following theorem will plays an important role for the proof of Main Theorems.

Theorem 1 (Solvability of the approximating problem).

There exists a sufficiently small constant τ0(0,1)\tau_{0}\in(0,1) such that for any τ(0,τ0)\tau\in(0,\tau_{0}) and ε(0,1)\varepsilon\in(0,1), (AP)ετ{}_{\tau}^{\varepsilon} admits a unique solution {[ηiε,θiε]}i=0\{[\eta_{i}^{\varepsilon},\theta_{i}^{\varepsilon}]\}_{i=0}^{\infty}. Additionally, the following energy inequality holds:

14τ|ηiεηi1ε|H2+μ2τ|(ηiεηi1ε)|[H]N2+δα2τ|θiεθi1ε|H2\displaystyle\frac{1}{4\tau}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H}^{2}+\frac{\mu^{2}}{\tau}|\nabla(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{[H]^{N}}^{2}+\frac{\delta_{\alpha}}{2\tau}|\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}|_{H}^{2} (4.4)
+ν2τ|(θiεθi1ε)|[H]N2+ε(ηiε,θiε)\displaystyle+\frac{\nu^{2}}{\tau}|\nabla(\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon})|_{[H]^{N}}^{2}+\mathcal{F}_{\varepsilon}(\eta_{i}^{\varepsilon},\theta_{i}^{\varepsilon})
ε(ηi1ε,θi1ε)+τ2|ui|H2+τ2δα|vi|H2, for any i=1,2,3,.\displaystyle\leq\mathcal{F}_{\varepsilon}(\eta_{i-1}^{\varepsilon},\theta_{i-1}^{\varepsilon})+\frac{\tau}{2}|u_{i}|_{H}^{2}+\frac{\tau}{2\delta_{\alpha}}|v_{i}|_{H}^{2},\ \mbox{ for any }i=1,2,3,\dots.

Theorem 1 is proved through several lemmas.

Lemma 4.1.

For arbitrary θ~V\widetilde{\theta}\in V, η~0W0\widetilde{\eta}_{0}\in W_{0} and u~H\widetilde{u}\in H, we consider the following elliptic problem:

1τ(ηη~0)+AN(η+μ2τ(ηη~0))+g(𝒯Mη)+α(𝒯Mη)γε(θ~)=u~, in H.\frac{1}{\tau}(\eta-\widetilde{\eta}_{0})+A_{N}\left(\eta+\frac{\mu^{2}}{\tau}(\eta-\widetilde{\eta}_{0})\right)+g(\mathcal{T}_{M}\eta)+\alpha^{\prime}(\mathcal{T}_{M}\eta)\gamma_{\varepsilon}(\nabla\widetilde{\theta})=\widetilde{u},\mbox{ in }H. (4.5)

Then, there exists a small constant τ1(0,1)\tau_{1}\in(0,1), depending only on |g|L(M,M)|g^{\prime}|_{L^{\infty}(-M,M)}, and for any 0<τ<τ10<\tau<\tau_{1}, the elliptic problem (4.5) admits a unique solution ηW0\eta\in W_{0}.

Proof.

First, for any ηV\eta^{\dagger}\in V, we define a functional Υ:H(,]\Upsilon:H\longrightarrow(-\infty,\infty] as follows:

Υ:zHΥ(z):={12τΩ|z|2𝑑x+12Ω|z|2𝑑x+μ22τΩ|(zη~0)|2𝑑x+Ωg~(𝒯Mη)z𝑑x+Ωα~M(z)γε(θ~)𝑑xΩu~z𝑑x, if zV,, otherwise.\Upsilon:z\in H\mapsto\Upsilon(z):=\left\{\begin{aligned} &\frac{1}{2\tau}\int_{\Omega}|z|^{2}\,dx+\frac{1}{2}\int_{\Omega}|\nabla z|^{2}\,dx+\frac{\mu^{2}}{2\tau}\int_{\Omega}|\nabla(z-\widetilde{\eta}_{0})|^{2}\,dx\\ &\quad+\int_{\Omega}\widetilde{g}(\mathcal{T}_{M}\eta^{\dagger})z\,dx+\int_{\Omega}\widetilde{\alpha}_{M}(z)\gamma_{\varepsilon}(\nabla\widetilde{\theta})\,dx\\ &\quad-\int_{\Omega}\widetilde{u}z\,dx,\ \mbox{ if }z\in V,\\ &\infty,\mbox{ otherwise}.\end{aligned}\right.

As is easily checked, Υ\Upsilon is proper, l.s.c., strictly convex, and coercive, and its unique minimizer solves the following elliptic equation:

1τ(ηη~0)+AN(η+μ2τ(ηη~0))+g(𝒯Mη)+α(𝒯Mη)γε(θ~)=u~, in H.\frac{1}{\tau}(\eta-\widetilde{\eta}_{0})+A_{N}\left(\eta+\frac{\mu^{2}}{\tau}(\eta-\widetilde{\eta}_{0})\right)+g(\mathcal{T}_{M}\eta^{\dagger})+\alpha^{\prime}(\mathcal{T}_{M}\eta)\gamma_{\varepsilon}(\nabla\widetilde{\theta})=\widetilde{u},\mbox{ in }H. (4.6)

Now, we define an operator Sτ:VH2(Ω)S_{\tau}:V\longrightarrow H^{2}(\Omega) which maps any ηV\eta^{\dagger}\in V to the unique solution to (4.6), and consider the smallness condition of τ\tau for SS to be contractive. Here, let ηk:=SτηkH2(Ω)\eta_{k}:=S_{\tau}\eta_{k}^{\dagger}\in H^{2}(\Omega), k=1,2k=1,2. By taking differences of (4.6), multiplying both sides by η1η2\eta_{1}-\eta_{2} and applying Young’s inequality, we see from (A1) and (A2) that:

1μ22τ|η1η2|V2τ|g|L(M,M)22|η1η2|H2.\displaystyle\frac{1\wedge\mu^{2}}{2\tau}|\eta_{1}-\eta_{2}|_{V}^{2}\leq\frac{\tau|g^{\prime}|_{L^{\infty}(-M,M)}^{2}}{2}|\eta_{1}^{\dagger}-\eta_{2}^{\dagger}|_{H}^{2}.

Therefore, if we assume that

0<τ<τ1:=(1μ2|g|L(M,M)2)12,0<\tau<\tau_{1}:=\left(\frac{1\wedge\mu^{2}}{|g|_{L^{\infty}(-M,M)}^{2}}\right)^{\frac{1}{2}}, (4.7)

then the mapping SτS_{\tau} becomes a contraction mapping from VV into itself. Therefore, applying Banach’s fixed point theorem, we find a unique fixed point η~V\widetilde{\eta}\in V of SτS_{\tau} under the condition (4.7). The identity Sτη~=η~S_{\tau}\widetilde{\eta}=\widetilde{\eta} implies that η~\widetilde{\eta} is the unique solution to (4.5). ∎

Lemma 4.2.

For arbitrary η~H2(Ω)\widetilde{\eta}\in H^{2}(\Omega), θ~0W0\widetilde{\theta}_{0}\in W_{0} and v~H\widetilde{v}\in H, we consider the following elliptic equation:

{α0(𝒯Mη~)θθ~0τdiv(α~M(η~)γε(θ)+ν2τ(θθ~0))=v~ a.e. in Ω,θnΓ=0 a.e. on Γ.\left\{\begin{aligned} &\alpha_{0}(\mathcal{T}_{M}\widetilde{\eta})\frac{\theta-\widetilde{\theta}_{0}}{\tau}-\operatorname{div}\left(\widetilde{\alpha}_{M}(\widetilde{\eta})\nabla\gamma_{\varepsilon}(\nabla\theta)+\frac{\nu^{2}}{\tau}\nabla(\theta-\widetilde{\theta}_{0})\right)=\widetilde{v}\mbox{ a.e. in }\Omega,\\ &\nabla\theta\cdot n_{\Gamma}=0\mbox{ a.e. on }\Gamma.\end{aligned}\right. (4.8)

Then, for any 0<τ<10<\tau<1, (4.8) admits a unique solution θW0\theta\in W_{0}.

Proof.

Let us consider a proper, l.s.c., strictly convex, and coercive function Υ\Upsilon_{*} defined as follows:

Υ:zHΥ(z):={12τΩα0(𝒯Mη~)|zθ~0|2𝑑x+Ωα~M(η~)γε(z)𝑑x+ν22τΩ|(zθ~0)|2𝑑xΩv~z𝑑x, if zV,, otherwise. \Upsilon_{*}:z\in H\mapsto\Upsilon_{*}(z):=\left\{\begin{aligned} &\frac{1}{2\tau}\int_{\Omega}\alpha_{0}(\mathcal{T}_{M}\widetilde{\eta})|z-\widetilde{\theta}_{0}|^{2}\,dx+\int_{\Omega}\widetilde{\alpha}_{M}(\widetilde{\eta})\gamma_{\varepsilon}(\nabla z)\,dx\\ &\quad+\frac{\nu^{2}}{2\tau}\int_{\Omega}|\nabla(z-\widetilde{\theta}_{0})|^{2}\,dx-\int_{\Omega}\widetilde{v}z\,dx,\ \mbox{ if }z\in V,\\ &\infty,\mbox{ otherwise. }\end{aligned}\right.

As is discussed in [1, Theorem 1], the unique minimizer θ~\widetilde{\theta} of Υ\Upsilon_{*} solves (4.8), and θ~\widetilde{\theta} belongs to W0W_{0}. ∎

Proof of theorem 1.

Let us fix any τ(0,τ1)\tau\in(0,\tau_{1}) and any ε(0,1)\varepsilon\in(0,1). Then, for any ii\in\mathbb{N}, we can obtain θiεW0\theta_{i}^{\varepsilon}\in W_{0} by applying lemma 4.2 in the case that:

η~=ηi1ε,θ~0=θi1ε and v~=vi in H.\widetilde{\eta}=\eta_{i-1}^{\varepsilon},\ \widetilde{\theta}_{0}=\theta_{i-1}^{\varepsilon}\mbox{ and }\widetilde{v}=v_{i}\mbox{ in }H.

Moreover, for any ii\in\mathbb{N}, the component ηiεW0\eta_{i}^{\varepsilon}\in W_{0} can be obtained by applying lemma 4.1 in the case that:

θ~=θiε,η~0:=ηi1ε and u~=ui in H.\widetilde{\theta}=\theta_{i}^{\varepsilon},\ \widetilde{\eta}_{0}:=\eta_{i-1}^{\varepsilon}\mbox{ and }\widetilde{u}=u_{i}\mbox{ in }H.

Thus we can find the unique solution {[ηiε,θiε]}i=0[W0]2\{[\eta_{i}^{\varepsilon},\theta_{i}^{\varepsilon}]\}_{i=0}^{\infty}\subset[W_{0}]^{2} to (AP)ετ{}_{\tau}^{\varepsilon}.

Next, we verify the inequality (4.4). Multiplying both sides of (4.1) with ηiεηi1ε\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}, we see that:

12τ|ηiεηi1ε|H2+μ2τ|(ηiεηi1ε)|[H]N2+12|ηiε|[H]N212|ηi1ε|[H]N2\displaystyle\frac{1}{2\tau}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H}^{2}+\frac{\mu^{2}}{\tau}|\nabla(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{[H]^{N}}^{2}+\frac{1}{2}|\nabla\eta_{i}^{\varepsilon}|_{[H]^{N}}^{2}-\frac{1}{2}|\nabla\eta_{i-1}^{\varepsilon}|_{[H]^{N}}^{2} (4.9)
+(g(𝒯Mηiε),ηiεηi1ε)H+(α(𝒯Mηiε)γε(θiε),ηiεηi1ε)Hτ2|ui|H2,\displaystyle+\bigl{(}g(\mathcal{T}_{M}\eta_{i}^{\varepsilon}),\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}\bigr{)}_{H}+\bigl{(}\alpha^{\prime}(\mathcal{T}_{M}\eta_{i}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon}),\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}\bigr{)}_{H}\leq\frac{\tau}{2}|u_{i}|_{H}^{2},
for i=1,2,3,.\displaystyle\mbox{for }i=1,2,3,\dots.

via the following computations:

(ηiε,(ηiεηi1ε))[H]N12(|ηiε|[H]N2|ηi1ε|[H]N2),\displaystyle(\nabla\eta_{i}^{\varepsilon},\nabla(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}))_{[H]^{N}}\geq\frac{1}{2}\bigl{(}|\nabla\eta_{i}^{\varepsilon}|_{[H]^{N}}^{2}-|\nabla\eta_{i-1}^{\varepsilon}|_{[H]^{N}}^{2}\bigr{)},
μ2τ((ηiεηi1ε),(ηiεηi1ε))[H]N=μ2τ|(ηiεηi1ε)|[H]N2,\displaystyle\frac{\mu^{2}}{\tau}(\nabla(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}),\nabla(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}))_{[H]^{N}}=\frac{\mu^{2}}{\tau}|\nabla(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{[H]^{N}}^{2},

and

(ui,ηiεηi1ε)H12τ|ηiεηi1ε|H2+τ2|ui|H2.\displaystyle(u_{i},\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})_{H}\leq\frac{1}{2\tau}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H}^{2}+\frac{\tau}{2}|u_{i}|_{H}^{2}.

In addition, by using (A1), it is obtained that:

(g(𝒯Mηiε),ηiεηi1ε)HΩG~M(ηiε)𝑑xΩG~M(ηi1ε)𝑑x\displaystyle\bigl{(}g(\mathcal{T}_{M}\eta_{i}^{\varepsilon}),\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}\bigr{)}_{H}\geq\int_{\Omega}\widetilde{G}_{M}(\eta_{i}^{\varepsilon})\,dx-\int_{\Omega}\widetilde{G}_{M}(\eta_{i-1}^{\varepsilon})\,dx (4.10)
+(g(𝒯Mηiε)g(𝒯Mηi1ε),ηiεηi1ε)H12|g|L(M,M)|ηiεηi1ε|H2\displaystyle+\bigl{(}g(\mathcal{T}_{M}\eta_{i}^{\varepsilon})-g(\mathcal{T}_{M}\eta_{i-1}^{\varepsilon}),\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}\bigr{)}_{H}-\frac{1}{2}|g^{\prime}|_{L^{\infty}(-M,M)}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H}^{2}
ΩG~M(ηiε)𝑑xΩG~M(ηi1ε)𝑑x32|g|L(M,M)|ηiεηi1ε|H2,\displaystyle\geq\int_{\Omega}\widetilde{G}_{M}(\eta_{i}^{\varepsilon})\,dx-\int_{\Omega}\widetilde{G}_{M}(\eta_{i-1}^{\varepsilon})\,dx-\frac{3}{2}|g^{\prime}|_{L^{\infty}(-M,M)}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H}^{2},
 for i=1,2,3,,\displaystyle\mbox{ for }i=1,2,3,\dots,

and by the convexity of α~M\widetilde{\alpha}_{M},

(α(𝒯Mηiε)γε(θiε),ηiεηi1ε)HΩα~M(ηiε)γε(θiε)𝑑x\displaystyle\bigl{(}\alpha^{\prime}(\mathcal{T}_{M}\eta_{i}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon}),\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}\bigr{)}_{H}\geq\int_{\Omega}\widetilde{\alpha}_{M}(\eta_{i}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon})\,dx (4.11)
Ωα~M(ηi1ε)γε(θiε)𝑑x, for i=1,2,3,.\displaystyle-\int_{\Omega}\widetilde{\alpha}_{M}(\eta_{i-1}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon})\,dx,\mbox{ for }i=1,2,3,\dots.

On account of (4.9)–(4.11), it is inferred that:

(123τ2|g|L(M,M))1τ|ηiεηi1ε|H2+μ2τ|(ηiεηi1ε)|[H]N2\displaystyle\left(\frac{1}{2}-\frac{3\tau}{2}|g^{\prime}|_{L^{\infty}(-M,M)}\right)\frac{1}{\tau}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H}^{2}+\frac{\mu^{2}}{\tau}|\nabla(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{[H]^{N}}^{2} (4.12)
+12|ηiε|[H]N2+ΩG~M(ηiε)𝑑x+Ωα~M(ηiε)γε(θiε)𝑑x\displaystyle+\frac{1}{2}|\nabla\eta_{i}^{\varepsilon}|_{[H]^{N}}^{2}+\int_{\Omega}\widetilde{G}_{M}(\eta_{i}^{\varepsilon})\,dx+\int_{\Omega}\widetilde{\alpha}_{M}(\eta_{i}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon})\,dx
12|ηi1ε|[H]N2+ΩG~M(ηi1ε)𝑑x+Ωα~M(ηi1ε)γε(θiε)𝑑x+τ2|ui|H2,\displaystyle\leq\frac{1}{2}|\nabla\eta_{i-1}^{\varepsilon}|_{[H]^{N}}^{2}+\int_{\Omega}\widetilde{G}_{M}(\eta_{i-1}^{\varepsilon})\,dx+\int_{\Omega}\widetilde{\alpha}_{M}(\eta_{i-1}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon})\,dx+\frac{\tau}{2}|u_{i}|_{H}^{2},
 for i=1,2,3,.\displaystyle\mbox{ for }i=1,2,3,\dots.

On the other hand, by multiplying both sides of (4.2) by θiεθi1ε\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}, and using (A3) and the convexity of γε\gamma_{\varepsilon}, we have

δα2τ|θiεθi1ε|H2+ν2τ|(θiεθi1ε)|[H]N2+Ωα~M(ηi1ε)γε(θiε)𝑑x\displaystyle\frac{\delta_{\alpha}}{2\tau}|\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}|_{H}^{2}+\frac{\nu^{2}}{\tau}|\nabla(\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon})|_{[H]^{N}}^{2}+\int_{\Omega}\widetilde{\alpha}_{M}(\eta_{i-1}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon})\,dx (4.13)
Ωα~M(ηi1ε)γε(θi1ε)𝑑x+τ2δα|vi|H2, for i=1,2,3,,\displaystyle\leq\int_{\Omega}\widetilde{\alpha}_{M}(\eta_{i-1}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i-1}^{\varepsilon})\,dx+\frac{\tau}{2\delta_{\alpha}}|v_{i}|_{H}^{2},\ \mbox{ for }i=1,2,3,\dots,

via the following computation:

1τ(α0(𝒯Mηi1)(θiεθi1ε),θiεθi1ε)Hδατ|θiεθi1ε|H2,\displaystyle\frac{1}{\tau}(\alpha_{0}(\mathcal{T}_{M}\eta_{i-1})(\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}),\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon})_{H}\geq\frac{\delta_{\alpha}}{\tau}|\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}|_{H}^{2},
(α~M(ηi1ε)γε(θiε),(θiεθi1ε))[H]N\displaystyle(\widetilde{\alpha}_{M}(\eta_{i-1}^{\varepsilon})\nabla\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon}),\nabla(\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}))_{[H]^{N}}\qquad\qquad
Ωα~M(ηi1ε)γε(θiε)𝑑xΩα~M(ηi1ε)γε(θi1ε)𝑑x,\displaystyle\qquad\qquad\qquad\geq\int_{\Omega}\widetilde{\alpha}_{M}(\eta_{i-1}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon})\,dx-\int_{\Omega}\widetilde{\alpha}_{M}(\eta_{i-1}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i-1}^{\varepsilon})\,dx,

and

(ui,θiεθi1ε)Hδα2τ|θiεθi1ε|H2+τ2δα|ui|H2.\displaystyle(u_{i},\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon})_{H}\leq\frac{\delta_{\alpha}}{2\tau}|\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}|_{H}^{2}+\frac{\tau}{2\delta_{\alpha}}|u_{i}|_{H}^{2}.

Now, let us set τ0\tau_{0} as

τ0:=min{τ1,16|g|L(M,M)}, with the constant τ1 as in lemma 4.1.\displaystyle\tau_{0}:=\min\left\{\tau_{1},\frac{1}{6|g^{\prime}|_{L^{\infty}(-M,M)}}\right\},\mbox{ with the constant $\tau_{1}$ as in \lx@cref{creftype~refnum}{Lem_AP1}.}

Then, from (4.12) and (4.13), we obtain that:

14τ|ηiεηi1ε|H2+μ2τ|(ηiεηi1ε)|[H]N2+δα2τ|θiεθi1ε|H2\displaystyle\frac{1}{4\tau}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H}^{2}+\frac{\mu^{2}}{\tau}|\nabla(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{[H]^{N}}^{2}+\frac{\delta_{\alpha}}{2\tau}|\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}|_{H}^{2}
+ν2τ|(θiεθi1ε)|[H]N2+ε(ηiε,θiε)\displaystyle\quad+\frac{\nu^{2}}{\tau}|\nabla(\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon})|_{[H]^{N}}^{2}+\mathcal{F}_{\varepsilon}(\eta_{i}^{\varepsilon},\theta_{i}^{\varepsilon})
ε(ηi1ε,θi1ε)+τ2|ui|H2+τ2δα|vi|H2, for i=1,2,3,.\displaystyle\quad\leq\mathcal{F}_{\varepsilon}(\eta_{i-1}^{\varepsilon},\theta_{i-1}^{\varepsilon})+\frac{\tau}{2}|u_{i}|_{H}^{2}+\frac{\tau}{2\delta_{\alpha}}|v_{i}|_{H}^{2},\ \mbox{ for }i=1,2,3,\dots.

Thus, we conclude theorem 1. ∎

5 Proofs of Main Theorems

In this section, we will provide proofs of Main Theorems. We set

nτ:=min{n|nτT}.n_{\tau}:=\min\{n\in\mathbb{N}\,|\,n\tau\geq T\}. (5.1)

Additionally, under the notations as in theorem 1, we invoke (2.2a) and (2.2b), and take a small constant τ(0,τ0)\tau_{*}\in(0,\tau_{0}), such that:

τi=1nτ(|ui|H2+|vi|H2)|u|2+|v|2+1, whenever τ(0,τ).\tau\sum_{i=1}^{n_{\tau}}\bigl{(}|u_{i}|_{H}^{2}+|v_{i}|_{H}^{2}\bigr{)}\leq|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1,\mbox{ whenever }\tau\in(0,\tau_{*}).

5.1 Proof of Main Theorem 1

Before we deal with the proof, we will prepare some lemmas. Hereafter, based on (A1), (A3), (A4) and (A5), we set the constant M>0M>0 of truncation, so large to satisfy that:

M|η0|L(Ω),g(M)|u|L(Q), and g(M)|u|L(Q).M\geq|\eta_{0}|_{L^{\infty}(\Omega)},\ g(M)\geq|u|_{L^{\infty}(Q)},\mbox{ and }g(-M)\leq-|u|_{L^{\infty}(Q)}. (5.2)

Then, it immediately follows that:

G~M(η0)=G(η0), and α~M(η0)=α(η0).\widetilde{G}_{M}(\eta_{0})=G(\eta_{0}),\mbox{ and }\widetilde{\alpha}_{M}(\eta_{0})=\alpha(\eta_{0}). (5.3)
Lemma 5.1.

Let τ(0,τ)\tau\in(0,\tau_{*}). Then, there exists a constant C2>0C_{2}>0, independent of ε\varepsilon and τ\tau, such that:

1τi=1nτ|ηiεηi1ε|H2(Ω)2C2(|η0|H2(Ω)2+|θ0|V2+|u|2+|v|2+1).\displaystyle\frac{1}{\tau}\sum_{i=1}^{n_{\tau}}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H^{2}(\Omega)}^{2}\leq C_{2}\bigl{(}|\eta_{0}|_{H^{2}(\Omega)}^{2}+|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}. (5.4)
Proof.

First, from the definition of ε\mathcal{F}_{\varepsilon} (4.3), (5.3), and the embedding W0L(Ω)W_{0}\subset L^{\infty}(\Omega) under N3N\leq 3, it is seen that:

ε(η0,θ0)\displaystyle\mathcal{F}_{\varepsilon}(\eta_{0},\theta_{0}) =12Ω|η0|2𝑑x+ΩG~M(η0)𝑑x+Ωα~M(η0)γε(θ0)𝑑x\displaystyle=\frac{1}{2}\int_{\Omega}|\nabla\eta_{0}|^{2}\,dx+\int_{\Omega}\widetilde{G}_{M}(\eta_{0})\,dx+\int_{\Omega}\widetilde{\alpha}_{M}(\eta_{0})\gamma_{\varepsilon}(\nabla\theta_{0})\,dx
=12Ω|η0|2𝑑x+ΩG(η0)𝑑x+Ωα(η0)γε(θ0)𝑑x\displaystyle=\frac{1}{2}\int_{\Omega}|\nabla\eta_{0}|^{2}\,dx+\int_{\Omega}G(\eta_{0})\,dx+\int_{\Omega}\alpha(\eta_{0})\gamma_{\varepsilon}(\nabla\theta_{0})\,dx
12|η0|V2+|G(η0)|L1(Ω)+|α(η0)|L2(Ω)2+N(Ω)+|θ0|V2.\displaystyle\leq\frac{1}{2}|\eta_{0}|_{V}^{2}+|G(\eta_{0})|_{L^{1}(\Omega)}+|\alpha(\eta_{0})|_{L^{2}(\Omega)}^{2}+\mathcal{L}^{N}(\Omega)+|\theta_{0}|_{V}^{2}.

Hence,

CF:=supε(0,1)ε(η0,θ0)<.C_{F}:=\sup_{\varepsilon\in(0,1)}\mathcal{F}_{\varepsilon}(\eta_{0},\theta_{0})<\infty.

Also, from (4.4),(5.1), theorem 1, and Hölder’s inequality, it is observed that:

|θiε|V22|θ0|V2+2(i=1nτ|θiεθi1ε|V)2\displaystyle|\theta_{i}^{\varepsilon}|_{V}^{2}\leq 2|\theta_{0}|_{V}^{2}+2\left(\sum_{i=1}^{n_{\tau}}|\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}|_{V}\right)^{2} (5.5)
\displaystyle\leq~{} 2|θ0|V2+2(T+1)i=1nτ1τ|θiεθi1ε|V2\displaystyle 2|\theta_{0}|_{V}^{2}+2(T+1)\sum_{i=1}^{n_{\tau}}\frac{1}{\tau}|\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}|_{V}^{2}
\displaystyle\leq~{} 2|θ0|V2+4(T+1)δαν2(ε(η0,θ0)+12(1δα)(|u|2+|v|2+1))\displaystyle 2|\theta_{0}|_{V}^{2}+\frac{4(T+1)}{\delta_{\alpha}\wedge\nu^{2}}\left(\mathcal{F}_{\varepsilon}(\eta_{0},\theta_{0})+\frac{1}{2(1\wedge\delta_{\alpha})}\bigl{(}|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}\right)
\displaystyle\leq~{} 4(CF+1)(T+1)1δα2ν4(|θ0|V2+|u|2+|v|2+1), for any i=1,2,3,,nτ.\displaystyle\frac{4(C_{F}+1)(T+1)}{1\wedge\delta_{\alpha}^{2}\wedge\nu^{4}}\bigl{(}|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)},\ \mbox{ for any }i=1,2,3,\dots,n_{\tau}.

Next, we verify the estimate (5.4). Multiplying the both side of (4.1) by Δ(ηiεηi1ε)-\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}) and applying Young’s inequality, it can be seen that:

3μ24τ|Δ(ηiεηi1ε)|H212(|Δηi1ε|H2|Δηiε|H2)\displaystyle\frac{3\mu^{2}}{4\tau}|\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{H}^{2}\leq\frac{1}{2}\bigl{(}|\mathit{\Delta}\eta_{i-1}^{\varepsilon}|_{H}^{2}-|\mathit{\Delta}\eta_{i}^{\varepsilon}|_{H}^{2}\bigr{)} (5.6)
+(g(𝒯Mηiε),Δ(ηiεηi1ε))H+(α(𝒯Mηiε)γε(θiε),Δ(ηiεηi1ε))H\displaystyle\quad+(g(\mathcal{T}_{M}\eta_{i}^{\varepsilon}),\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}))_{H}+(\alpha^{\prime}(\mathcal{T}_{M}\eta_{i}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon}),\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}))_{H}
+τμ2|ui|H2, for i=1,2,3,,nτ,\displaystyle\quad+\frac{\tau}{\mu^{2}}|u_{i}|_{H}^{2},\ \mbox{ for }i=1,2,3,\dots,n_{\tau},

via the following calculations:

(Δηiε,Δ(ηiεηi1ε))H12(|Δηiε|H2|Δηi1ε|H2),\displaystyle(-\mathit{\Delta}\eta_{i}^{\varepsilon},-\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}))_{H}\geq\frac{1}{2}\bigl{(}|\mathit{\Delta}\eta_{i}^{\varepsilon}|_{H}^{2}-|\mathit{\Delta}\eta_{i-1}^{\varepsilon}|_{H}^{2}\bigr{)},

and

(ui,Δ(ηiεηi1ε))Hμ24τ|Δ(ηiεηi1ε)|H2+τμ2|ui|H2.\displaystyle(u_{i},-\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}))_{H}\leq\frac{\mu^{2}}{4\tau}|\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{H}^{2}+\frac{\tau}{\mu^{2}}|u_{i}|_{H}^{2}.
(g(𝒯Mηiε),Δ(ηiεηi1ε))H\displaystyle(g(\mathcal{T}_{M}\eta_{i}^{\varepsilon}),\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}))_{H} (5.7)
τμ2|g|L(M,M)2N(Ω)+μ24τ|Δ(ηiεηi1ε)|H2,\displaystyle\quad\leq\frac{\tau}{\mu^{2}}|g|_{L^{\infty}(-M,M)}^{2}\mathcal{L}^{N}(\Omega)+\frac{\mu^{2}}{4\tau}|\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{H}^{2},

and

(α(𝒯Mηiε)γε(θ),Δ(ηiεηi1ε))H\displaystyle(\alpha^{\prime}(\mathcal{T}_{M}\eta_{i}^{\varepsilon})\gamma_{\varepsilon}(\nabla\theta),\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}))_{H} (5.8)
τμ2|α|L(M,M)2Ω(ε2+|θiε|2)𝑑x+μ24τ|Δ(ηiεηi1ε)|H2\displaystyle\quad\leq\frac{\tau}{\mu^{2}}|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}\int_{\Omega}\bigl{(}\varepsilon^{2}+|\nabla\theta_{i}^{\varepsilon}|^{2}\bigr{)}\,dx+\frac{\mu^{2}}{4\tau}|\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{H}^{2}
τμ2|α|L(M,M)2(N(Ω)+|θiε|V2)+μ24τ|Δ(ηiεηi1ε)|H2,\displaystyle\quad\leq\frac{\tau}{\mu^{2}}|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}\bigl{(}\mathcal{L}^{N}(\Omega)+|\theta_{i}^{\varepsilon}|_{V}^{2}\bigr{)}+\frac{\mu^{2}}{4\tau}|\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{H}^{2},
 for i=1,2,3,,nτ.\displaystyle\qquad\qquad\qquad\qquad\qquad\mbox{ for }i=1,2,3,\dots,n_{\tau}.

On account of (5.5)–(5.8), we infer that:

μ24τ|Δ(ηiεηi1ε)|H2\displaystyle\frac{\mu^{2}}{4\tau}|\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{H}^{2} (5.9)
12(|Δηi1ε|H2|Δηiε|H2)+τN(Ω)μ2(|g|L(M,M)2+|α|L(M,M))\displaystyle\quad\leq\frac{1}{2}\bigl{(}|\mathit{\Delta}\eta_{i-1}^{\varepsilon}|_{H}^{2}-|\mathit{\Delta}\eta_{i}^{\varepsilon}|_{H}^{2}\bigr{)}+\frac{\tau\mathcal{L}^{N}(\Omega)}{\mu^{2}}\bigl{(}|g|_{L^{\infty}(-M,M)}^{2}+|\alpha^{\prime}|_{L^{\infty}(-M,M)}\bigr{)}
+4τ(CF+1)(T+1)|α|L(M,M)2μ2(1δα2ν4)(|θ0|V2+|u|2+|v|2+1)\displaystyle\qquad+\frac{4\tau(C_{F}+1)(T+1)|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}}{\mu^{2}(1\wedge\delta_{\alpha}^{2}\wedge\nu^{4})}\bigl{(}|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}
12(|Δηi1ε|H2|Δηiε|H2)+τC~3(|θ0|V2+|u|2+|v|2+1),\displaystyle\quad\leq\frac{1}{2}\bigl{(}|\mathit{\Delta}\eta_{i-1}^{\varepsilon}|_{H}^{2}-|\mathit{\Delta}\eta_{i}^{\varepsilon}|_{H}^{2}\bigr{)}+\tau\widetilde{C}_{3}\bigl{(}|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)},
 for i=1,2,3,,nτ,\displaystyle\qquad\qquad\qquad\qquad\qquad\mbox{ for }i=1,2,3,\dots,n_{\tau},

with

C~3:=4(CF+N(Ω)+1)(T+1)(|g|L(M,M)2+|α|L(M,M)2+1)μ2(1δα2ν4).\widetilde{C}_{3}:=\frac{4(C_{F}+\mathcal{L}^{N}(\Omega)+1)(T+1)\bigl{(}|g|_{L^{\infty}(-M,M)}^{2}+|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}+1\bigr{)}}{\mu^{2}(1\wedge\delta_{\alpha}^{2}\wedge\nu^{4})}.

Hence, taking the sum of (5.9) with respect to i=1,2,3,,nτi=1,2,3,\dots,n_{\tau}, one can deduce from (2.1), (4.4), and (5.9) that:

1τi=1nτ|ηiεηi1ε|H2(Ω)2C0τi=1nτ(|AN(ηiεηi1ε)|H2+|ηiεηi1ε|H2)\displaystyle\frac{1}{\tau}\sum_{i=1}^{n_{\tau}}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H^{2}(\Omega)}^{2}\leq\frac{C_{0}}{\tau}\sum_{i=1}^{n_{\tau}}\bigl{(}|A_{N}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{H}^{2}+|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H}^{2}\bigr{)}
=C0τi=1nτ(|Δ(ηiεηi1ε)|H2+|ηiεηi1ε|H2)\displaystyle\quad=\frac{C_{0}}{\tau}\sum_{i=1}^{n_{\tau}}\bigl{(}|\mathit{\Delta}(\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon})|_{H}^{2}+|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H}^{2}\bigr{)}
2C0μ2|Δη0|H2+4C0C~3(T+1)μ2(|θ0|V2+|u|2+|v|2+1)\displaystyle\quad\leq\frac{2C_{0}}{\mu^{2}}|\mathit{\Delta}\eta_{0}|_{H}^{2}+\frac{4C_{0}\widetilde{C}_{3}(T+1)}{\mu^{2}}\bigl{(}|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}
+4C0(ε(η0,θ0)+12(1δα)(|u|2+|v|2+1))\displaystyle\quad\qquad+4C_{0}\left(\mathcal{F}_{\varepsilon}(\eta_{0},\theta_{0})+\frac{1}{2(1\wedge\delta_{\alpha})}\bigl{(}|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}\right)
2NC0μ2|η0|H2(Ω)2+4C0C~3(T+1)μ2(|θ0|V2+|u|2+|v|2+1)\displaystyle\quad\leq\frac{2NC_{0}}{\mu^{2}}|\eta_{0}|_{H^{2}(\Omega)}^{2}+\frac{4C_{0}\widetilde{C}_{3}(T+1)}{\mu^{2}}\bigl{(}|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}
+4C0(CF+1)1δα(|u|2+|v|2+1)\displaystyle\quad\qquad+\frac{4C_{0}(C_{F}+1)}{1\wedge\delta_{\alpha}}\bigl{(}|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}
C2(|η0|H2(Ω)2+|θ0|V2+|u|2+|v|2+1),\displaystyle\quad\leq C_{2}\bigl{(}|\eta_{0}|_{H^{2}(\Omega)}^{2}+|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)},

where

C2:=4NC0(T+1)(C~3+CF+1)1μ2δα.\displaystyle C_{2}:=\frac{4NC_{0}(T+1)(\widetilde{C}_{3}+C_{F}+1)}{1\wedge\mu^{2}\wedge\delta_{\alpha}}.

Thus we conclude lemma 5.1. ∎

Lemma 5.2.

There exist a small time-step size τ(0,τ)\tau_{**}\in(0,\tau_{*}) and a constant C4>0C_{4}>0 such that for any τ(0,τ)\tau\in(0,\tau_{**}), the following estimate holds:

|θiε|H2(Ω)2C4(|η0|H2(Ω)2+|θ0|H2(Ω)2+|u|2+|v|2+1)2,\displaystyle|\theta_{i}^{\varepsilon}|_{H^{2}(\Omega)}^{2}\leq C_{4}\bigl{(}|\eta_{0}|_{H^{2}(\Omega)}^{2}+|\theta_{0}|_{H^{2}(\Omega)}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}^{2}, (5.10)
for i=1,2,3,,nτ.\displaystyle\mbox{for }i=1,2,3,\dots,n_{\tau}.

We can prove this lemma 5.2 by using the following technical lemma, obtained in [1].

Lemma 5.3.

(cf. [1, Lemma 3.2]) Let us fix ε>0\varepsilon>0, wW0w\in W_{0}, and αL(Ω)V\alpha^{\circ}\in L^{\infty}(\Omega)\cap V. Then, for any L|α|L(Ω)L\geq|\alpha^{\circ}|_{L^{\infty}(\Omega)}, there exists a constant C5(L)>0C_{5}(L)>0, depending only on LL, and being independent of ε\varepsilon and ww, such that:

(div(αγε(w)),Δw)H|2w|[H]N×N2C5(L)(|α|V2+1)(|w|V2+1).\bigl{(}\operatorname{div}(\alpha^{\circ}\nabla\gamma_{\varepsilon}(\nabla w)),\mathit{\Delta}w\bigr{)}_{H}\geq-|\nabla^{2}w|_{[H]^{N\times N}}^{2}-C_{5}(L)(|\alpha^{\circ}|_{V}^{2}+1)(|w|_{V}^{2}+1).
Proof of lemma 5.2.

First, we note that lemma 5.1 leads to the boundedness of {ηiε}i=0nτ\{\eta_{i}^{\varepsilon}\}_{i=0}^{n_{\tau}} in H2(Ω)H^{2}(\Omega), with the following estimate:

|ηiε|H2(Ω)2\displaystyle|\eta_{i}^{\varepsilon}|_{H^{2}(\Omega)}^{2} 2|η0|H2(Ω)2+2(i=1nτ|ηiεηi1ε|H2(Ω))2\displaystyle\leq 2|\eta_{0}|_{H^{2}(\Omega)}^{2}+2\left(\sum_{i=1}^{n_{\tau}}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H^{2}(\Omega)}\right)^{2} (5.11)
2|η0|H2(Ω)2+2(T+1)i=1nτ1τ|ηiεηi1ε|H2(Ω)2\displaystyle\leq 2|\eta_{0}|_{H^{2}(\Omega)}^{2}+2(T+1)\sum_{i=1}^{n_{\tau}}\frac{1}{\tau}|\eta_{i}^{\varepsilon}-\eta_{i-1}^{\varepsilon}|_{H^{2}(\Omega)}^{2}
2(T+1)(C2+1)(|η0|H2(Ω)2+|θ0|V2+|u|2+|v|2+1),\displaystyle\leq 2(T+1)(C_{2}+1)\bigl{(}|\eta_{0}|_{H^{2}(\Omega)}^{2}+|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)},
 for i=1,2,3,,nτ.\displaystyle\qquad\qquad\qquad\ \mbox{ for }i=1,2,3,\dots,n_{\tau}.

Moreover, by (5.11) and continuous embedding from H2(Ω)H^{2}(\Omega) to L(Ω)L^{\infty}(\Omega) under N3N\leq 3, we see that α~M(ηiε)L(Ω)V\widetilde{\alpha}_{M}(\eta_{i}^{\varepsilon})\in L^{\infty}(\Omega)\cap V for any i=1,2,3,,nτi=1,2,3,\dots,n_{\tau}, with the following estimates hold:

|α~M(ηiε)|L(Ω)2\displaystyle|\widetilde{\alpha}_{M}(\eta_{i}^{\varepsilon})|_{L^{\infty}(\Omega)}^{2} 2α(0)2+2|α|L(M,M)2|ηiε|L(Ω)2\displaystyle\leq 2\alpha(0)^{2}+2|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}|\eta_{i}^{\varepsilon}|_{L^{\infty}(\Omega)}^{2} (5.12)
2α(0)2+2(CH2L)2|α|L(M,M)2|ηiε|H2(Ω)2,\displaystyle\leq 2\alpha(0)^{2}+2(C_{H^{2}}^{L^{\infty}})^{2}|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}|\eta_{i}^{\varepsilon}|_{H^{2}(\Omega)}^{2},

and

|α~M(ηiε)|V2N(Ω)|α~M(ηiε)|L(Ω)2+|α|L(M,M)2|ηiε|[H]N2\displaystyle|\widetilde{\alpha}_{M}(\eta_{i}^{\varepsilon})|_{V}^{2}\leq\mathcal{L}^{N}(\Omega)|\widetilde{\alpha}_{M}(\eta_{i}^{\varepsilon})|_{L^{\infty}(\Omega)}^{2}+|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}|\nabla\eta_{i}^{\varepsilon}|_{[H]^{N}}^{2} (5.13)
2N(Ω)(α(0)2+(CH2L)2|α|L(M,M)2|ηiε|H2(Ω)2)+|α|L(M,M)2|ηiε|H2(Ω)2\displaystyle\leq 2\mathcal{L}^{N}(\Omega)\bigl{(}\alpha(0)^{2}+(C_{H^{2}}^{L^{\infty}})^{2}|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}|\eta_{i}^{\varepsilon}|_{H^{2}(\Omega)}^{2}\bigr{)}+|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}|\eta_{i}^{\varepsilon}|_{H^{2}(\Omega)}^{2}
C~6(|ηiε|H2(Ω)2+1), for i=1,2,3,,nτ,\displaystyle\leq\widetilde{C}_{6}\bigl{(}|\eta_{i}^{\varepsilon}|_{H^{2}(\Omega)}^{2}+1\bigr{)},\ \mbox{ for }i=1,2,3,\dots,n_{\tau},

where CH2LC_{H^{2}}^{L^{\infty}} is a constant of the embedding from H2(Ω)H^{2}(\Omega) to L(Ω)L^{\infty}(\Omega), and

C~6:=2(N(Ω)α(0)2+N(Ω)(CH2L)2+1)(|α|L(M,M)2+1),\widetilde{C}_{6}:=2\bigl{(}\mathcal{L}^{N}(\Omega)\alpha(0)^{2}+\mathcal{L}^{N}(\Omega)(C_{H^{2}}^{L^{\infty}})^{2}+1\bigr{)}\bigl{(}|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}+1\bigr{)},

Next, we verify the inequality (5.10). Let us consider to multiply the both sides of (4.2) by Δθiε-\mathit{\Delta}\theta_{i}^{\varepsilon}.

By applying Young’s inequality, we have:

ν2τ(Δ(θiεθi1ε),Δθiε)Hν22τ(|Δθiε|H2|Δθi1ε|H2),\displaystyle\frac{\nu^{2}}{\tau}\bigl{(}-\mathit{\Delta}(\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}),-\mathit{\Delta}\theta_{i}^{\varepsilon}\bigr{)}_{H}\geq\frac{\nu^{2}}{2\tau}\bigl{(}|\mathit{\Delta}\theta_{i}^{\varepsilon}|_{H}^{2}-|\mathit{\Delta}\theta_{i-1}^{\varepsilon}|_{H}^{2}\bigr{)}, (5.14)

and

(vi,Δθiε)H12|Δθiε|H2+12|vi|H2, for i=1,2,3,,nτ.\displaystyle(v_{i},-\mathit{\Delta}\theta_{i}^{\varepsilon})_{H}\leq\frac{1}{2}|\mathit{\Delta}\theta_{i}^{\varepsilon}|_{H}^{2}+\frac{1}{2}|v_{i}|_{H}^{2},\ \mbox{ for }i=1,2,3,\dots,n_{\tau}. (5.15)

Moreover, from (4.4) and (A3), we see that:

1τ(α0(𝒯Mηi1ε)(θiεθi1ε),Δθiε)H\displaystyle\frac{1}{\tau}\bigl{(}\alpha_{0}(\mathcal{T}_{M}\eta_{i-1}^{\varepsilon})(\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}),-\mathit{\Delta}\theta_{i}^{\varepsilon}\bigr{)}_{H} (5.16)
12τ2|α0|L(M,M)2|θiεθi1ε|H212|Δθiε|H2,\displaystyle\qquad\geq-\frac{1}{2\tau^{2}}|\alpha_{0}|_{L^{\infty}(-M,M)}^{2}|\theta_{i}^{\varepsilon}-\theta_{i-1}^{\varepsilon}|_{H}^{2}-\frac{1}{2}|\mathit{\Delta}\theta_{i}^{\varepsilon}|_{H}^{2},
|α0|L(M,M)22δα1τ(ε(ηiε,θiε)ε(ηi1ε,θi1ε))12|Δθiε|H2,\displaystyle\qquad\geq\frac{|\alpha_{0}|_{L^{\infty}(-M,M)}^{2}}{2\delta_{\alpha}}\cdot\frac{1}{\tau}(\mathcal{F}_{\varepsilon}(\eta_{i}^{\varepsilon},\theta_{i}^{\varepsilon})-\mathcal{F}_{\varepsilon}(\eta_{i-1}^{\varepsilon},\theta_{i-1}^{\varepsilon}))-\frac{1}{2}|\mathit{\Delta}\theta_{i}^{\varepsilon}|_{H}^{2},
 for i=1,2,3,,nτ.\displaystyle\qquad\qquad\qquad\qquad\mbox{ for }i=1,2,3,\dots,n_{\tau}.

Using (2.1), (5.11)–(5.13), and applying lemma 5.3 to the case that:

α=α~M(ηi1ε) and w=θiε, for each i{1,2,,nτ},\displaystyle\alpha^{\circ}=\widetilde{\alpha}_{M}(\eta_{i-1}^{\varepsilon})\mbox{ and }w=\theta_{i}^{\varepsilon},\mbox{ for each }i\in\{1,2,\ldots,n_{\tau}\},

and

L\displaystyle L =2α(0)2+2(CH2L)2|α|L(M,M)2\displaystyle=2\alpha(0)^{2}+2(C_{H^{2}}^{L^{\infty}})^{2}|\alpha^{\prime}|_{L^{\infty}(-M,M)}^{2}\cdot
2(T+1)(C2+1)(|η0|H2(Ω)2+|θ0|V2+|u|2+|v|2+1),\displaystyle\cdot 2(T+1)(C_{2}+1)\bigl{(}|\eta_{0}|_{H^{2}(\Omega)}^{2}+|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)},

it is observed that:

(div(α~M(ηi1ε)γε(θiε)),Δθiε)H\displaystyle\quad\bigl{(}\operatorname{div}(\widetilde{\alpha}_{M}(\eta_{i-1}^{\varepsilon})\nabla\gamma_{\varepsilon}(\nabla\theta_{i}^{\varepsilon})),\mathit{\Delta}\theta_{i}^{\varepsilon}\bigr{)}_{H} (5.17)
|2θiε|[H]N×N2C5(L)(|α~M(ηi1ε)|V2+1)(|θiε|V2+1)\displaystyle\geq-|\nabla^{2}\theta_{i}^{\varepsilon}|_{[H]^{N\times N}}^{2}-C_{5}(L)(|\widetilde{\alpha}_{M}(\eta_{i-1}^{\varepsilon})|_{V}^{2}+1)(|\theta_{i}^{\varepsilon}|_{V}^{2}+1)
C0(|ANθiε|H2+|θiε|H2)\displaystyle\geq-C_{0}\bigl{(}|A_{N}\theta_{i}^{\varepsilon}|_{H}^{2}+|\theta_{i}^{\varepsilon}|_{H}^{2}\bigr{)}
C5(L)(C~6(|ηi1ε|H2(Ω)2+1)+1)(|θiε|V2+1)\displaystyle\qquad-C_{5}(L)\bigl{(}\widetilde{C}_{6}\bigl{(}|\eta_{i-1}^{\varepsilon}|_{H^{2}(\Omega)}^{2}+1\bigr{)}+1\bigr{)}\bigl{(}|\theta_{i}^{\varepsilon}|_{V}^{2}+1\bigr{)}
C0|Δθiε|H2(C5(L)+C0)(C~6+1)(|θiε|V2+1)(|ηi1ε|H2(Ω)2+1),\displaystyle\geq-C_{0}|\mathit{\Delta}\theta_{i}^{\varepsilon}|_{H}^{2}-(C_{5}(L)+C_{0})(\widetilde{C}_{6}+1)\bigl{(}|\theta_{i}^{\varepsilon}|_{V}^{2}+1\bigr{)}\bigl{(}|\eta_{i-1}^{\varepsilon}|_{H^{2}(\Omega)}^{2}+1\bigr{)},
 for i=1,2,3,,nτ.\displaystyle\qquad\qquad\qquad\qquad\mbox{ for }i=1,2,3,\dots,n_{\tau}.

Now, by using (5.14)–(5.17), we will obtain that:

1τ(XiXi1)C~7(Xi+Fi), for i=1,2,3,,nτ,\frac{1}{\tau}(X_{i}-X_{i-1})\leq\widetilde{C}_{7}(X_{i}+F_{i}),\ \mbox{ for }i=1,2,3,\dots,n_{\tau}, (5.18)

with

{Xi:=ν2|Δθiε|H2+|α0|L(M,M)2δαε(ηiε,θiε),Fi:=(|θiε|V2+|vi|H2+1)(|ηi1ε|H2(Ω)2+1), for i=1,2,3,,nτ,\displaystyle\begin{cases}\displaystyle X_{i}:=\nu^{2}|\mathit{\Delta}\theta_{i}^{\varepsilon}|_{H}^{2}+\frac{|\alpha_{0}|_{L^{\infty}(-M,M)}^{2}}{\delta_{\alpha}}\mathcal{F}_{\varepsilon}(\eta_{i}^{\varepsilon},\theta_{i}^{\varepsilon}),\\[8.61108pt] \displaystyle F_{i}:=\bigl{(}|\theta_{i}^{\varepsilon}|_{V}^{2}+|v_{i}|_{H}^{2}+1\bigr{)}\bigl{(}|\eta_{i-1}^{\varepsilon}|_{H^{2}(\Omega)}^{2}+1\bigr{)},\end{cases}\mbox{ for }i=1,2,3,\dots,n_{\tau},

and

C~7:=4(C5(L)+C0+1)(C~6+1)1ν24.\widetilde{C}_{7}:=\frac{4(C_{5}(L)+C_{0}+1)(\widetilde{C}_{6}+1)}{1\wedge\nu^{2}}\geq 4.

Here, let us take τ(0,τ)\tau_{**}\in(0,\tau_{*}) satisfying:

τ<min{τ,12C~7}, and in particular, 1τC~7>12.\tau_{**}<\min\left\{\tau_{*},\frac{1}{2\widetilde{C}_{7}}\right\},\mbox{ and in particular, }1-\tau_{**}\widetilde{C}_{7}>\frac{1}{2}.

Then, applying the discrete version of Gronwall’s lemma (cf. [10, Section 3.1]) to (5.18), one can see from (5.5), (5.11), and (5.18) that:

XiC~7e2C~7(T+1)(X0+τi=1nτFi)\displaystyle\quad X_{i}\leq\widetilde{C}_{7}e^{2\widetilde{C}_{7}(T+1)}\left(X_{0}+\tau\sum_{i=1}^{n_{\tau}}F_{i}\right) (5.19)
C~7e2C~7(T+1)(ν2|Δθ0|H2+|α0|L(M,M)2CFδα)\displaystyle\leq\widetilde{C}_{7}e^{2\widetilde{C}_{7}(T+1)}\left(\nu^{2}|\mathit{\Delta}\theta_{0}|_{H}^{2}+\frac{|\alpha_{0}|_{L^{\infty}(-M,M)}^{2}C_{F}}{\delta_{\alpha}}\right)
+4C~7e2C~7(T+1)(CF+3)(T+1)21δα2ν4(|θ0|V2+|u|2+|v|2+1)\displaystyle\,\quad+\frac{4\widetilde{C}_{7}e^{2\widetilde{C}_{7}(T+1)}(C_{F}+3)(T+1)^{2}}{1\wedge\delta_{\alpha}^{2}\wedge\nu^{4}}\bigl{(}|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}\cdot
2(T+1)(C2+2)(|η0|H2(Ω)2+|θ0|V2+|u|2+|v|2+1)\displaystyle\qquad\quad\cdot 2(T+1)(C_{2}+2)\bigl{(}|\eta_{0}|_{H^{2}(\Omega)}^{2}+|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}
C~8(|η0|H2(Ω)2+|θ0|H2(Ω)2+|u|2+|v|2+1)2, for i=1,2,3,,nτ,\displaystyle\leq\widetilde{C}_{8}\bigl{(}|\eta_{0}|_{H^{2}(\Omega)}^{2}+|\theta_{0}|_{H^{2}(\Omega)}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}^{2},\ \mbox{ for }i=1,2,3,\dots,n_{\tau},

where

C~8:=8C~7e2C~7(T+1)(CF+3)(T+1)3(C2+Nν2+|α0|L(M,M)+2)1δα2ν4.\widetilde{C}_{8}:=\frac{8\widetilde{C}_{7}e^{2\widetilde{C}_{7}(T+1)}(C_{F}+3)(T+1)^{3}(C_{2}+N\nu^{2}+|\alpha_{0}|_{L^{\infty}(-M,M)}+2)}{1\wedge\delta_{\alpha}^{2}\wedge\nu^{4}}.

In the light of (2.1) (5.5), and (5.19), we arrive at:

|θiε|H2(Ω)2C0(|ANθiε|H2+|θiε|H2)=C0(|Δθiε|H2+|θiε|H2)\displaystyle\quad|\theta_{i}^{\varepsilon}|_{H^{2}(\Omega)}^{2}\leq C_{0}\bigl{(}|A_{N}\theta_{i}^{\varepsilon}|_{H}^{2}+|\theta_{i}^{\varepsilon}|_{H}^{2}\bigr{)}=C_{0}\bigl{(}|\mathit{\Delta}\theta_{i}^{\varepsilon}|_{H}^{2}+|\theta_{i}^{\varepsilon}|_{H}^{2}\bigr{)}
C0ν2Xi+4C0(CF+1)(T+1)1δα2ν4(|θ0|V2+|u|2+|v|2+1)\displaystyle\leq\frac{C_{0}}{\nu^{2}}X_{i}+\frac{4C_{0}(C_{F}+1)(T+1)}{1\wedge\delta_{\alpha}^{2}\wedge\nu^{4}}\bigl{(}|\theta_{0}|_{V}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}
4C0(CF+C~8+1)(T+1)1δα2ν4(|η0|H2(Ω)2+|θ0|H2(Ω)2+|u|2+|v|2+1)2,\displaystyle\leq\frac{4C_{0}(C_{F}+\widetilde{C}_{8}+1)(T+1)}{1\wedge\delta_{\alpha}^{2}\wedge\nu^{4}}\bigl{(}|\eta_{0}|_{H^{2}(\Omega)}^{2}+|\theta_{0}|_{H^{2}(\Omega)}^{2}+|u|_{\mathscr{H}}^{2}+|v|_{\mathscr{H}}^{2}+1\bigr{)}^{2},
 for i=1,2,3,,nτ.\displaystyle\qquad\qquad\qquad\qquad\qquad\mbox{ for }i=1,2,3,\dots,n_{\tau}.

Thus we conclude lemma 5.2 with the constant:

C4=4C0(CF+C~8+1)(T+1)1δα2ν4.C_{4}=\frac{4C_{0}(C_{F}+\widetilde{C}_{8}+1)(T+1)}{1\wedge\delta_{\alpha}^{2}\wedge\nu^{4}}.

Next, we confirm the comparison principle for single pseudo-parabolic equation, which will play a key-role in the LL^{\infty}-estimate of the component η\eta.

Lemma 5.4.

We assume that η1,η2W1,2(0,T;W0)\eta^{1},\eta^{2}\in W^{1,2}(0,T;W_{0}), η01,η02W0\eta_{0}^{1},\eta_{0}^{2}\in W_{0}, θ~𝒱\widetilde{\theta}\in\mathscr{V}, u~\widetilde{u}\in\mathscr{H}, and

{(1)i1(tηiΔ(ηi+μ2tηi)+g(𝒯Mηi)+α(𝒯Mηi)|θ~|)u~, a.e. in Q,ηi(0)=η0i, a.e. in Ω.\left\{\begin{aligned} &(-1)^{i-1}\left(\partial_{t}\eta^{i}-\mathit{\Delta}(\eta^{i}+\mu^{2}\partial_{t}\eta^{i})+g(\mathcal{T}_{M}\eta^{i})+\alpha^{\prime}(\mathcal{T}_{M}\eta^{i})|\nabla\widetilde{\theta}|\right)\\ &\qquad\leq\widetilde{u},\mbox{ a.e. in }Q,\\ &\eta^{i}(0)=\eta_{0}^{i},\mbox{ a.e. in }\Omega.\end{aligned}\right. (5.20)

Then, there exists a constant C9>0C_{9}>0 such that:

|[η1η2]+(t)|V2C9|[η01η02]+|V2, for any t[0,T].\bigl{|}[\eta^{1}-\eta^{2}]^{+}(t)\bigr{|}_{V}^{2}\leq C_{9}\bigl{|}[\eta_{0}^{1}-\eta_{0}^{2}]^{+}\bigr{|}_{V}^{2},\mbox{ for any }t\in[0,T].
Proof.

Taking the difference of two inequality (5.20) for ηi\eta^{i}, i=1,2i=1,2, and multiplying the both sides by [η1η2]+(t)[\eta^{1}-\eta^{2}]^{+}(t), we see that:

12ddt|[η1η2]+(t)|H2+μ22ddt|[η1η2]+(t)|[H]N2\displaystyle\frac{1}{2}\frac{d}{dt}\bigl{|}[\eta^{1}-\eta^{2}]^{+}(t)\bigr{|}_{H}^{2}+\frac{\mu^{2}}{2}\frac{d}{dt}\bigl{|}\nabla[\eta^{1}-\eta^{2}]^{+}(t)\bigr{|}_{[H]^{N}}^{2} (5.21)
+|[η1η2]+(t)|[H]N2+(g(𝒯Mη1(t))g(𝒯Mη2(t)),[η1η2]+(t))H\displaystyle\qquad+\bigl{|}\nabla[\eta^{1}-\eta^{2}]^{+}(t)\bigr{|}_{[H]^{N}}^{2}+(g(\mathcal{T}_{M}\eta^{1}(t))-g(\mathcal{T}_{M}\eta^{2}(t)),[\eta^{1}-\eta^{2}]^{+}(t))_{H}
+Ω((α(𝒯Mη1(t))α(𝒯Mη2(t)))|θ(t)|,[η1η2]+(t)dx\displaystyle\qquad+\int_{\Omega}\bigl{(}(\alpha^{\prime}(\mathcal{T}_{M}\eta^{1}(t))-\alpha^{\prime}(\mathcal{T}_{M}\eta^{2}(t))\bigr{)}|\nabla\theta(t)|,[\eta^{1}-\eta^{2}]^{+}(t)\,dx
0, for a.e. t(0,T).\displaystyle\leq 0,\ \mbox{ for a.e. }t\in(0,T).

Here, from the assumption (A1), it is deduced that:

(g(𝒯Mη1(t))g(𝒯Mη2(t)),[η1η2]+(t))H\displaystyle(g(\mathcal{T}_{M}\eta^{1}(t))-g(\mathcal{T}_{M}\eta^{2}(t)),[\eta^{1}-\eta^{2}]^{+}(t))_{H}
|g|L(M,M)|[η1η2]+(t)|H2, for a.e. t(0,T).\displaystyle\qquad\geq-|g^{\prime}|_{L^{\infty}(-M,M)}\bigl{|}[\eta^{1}-\eta^{2}]^{+}(t)\bigr{|}_{H}^{2},\ \mbox{ for a.e. }t\in(0,T). (5.22)

Also, by the monotonicity of α𝒯M\alpha^{\prime}\circ\mathcal{T}_{M}, we can say that:

(α(𝒯Mη1)α(𝒯Mη2))[η1η2]+0, a.e. in Q.\bigl{(}\alpha^{\prime}(\mathcal{T}_{M}\eta^{1})-\alpha^{\prime}(\mathcal{T}_{M}\eta^{2})\bigr{)}[\eta^{1}-\eta^{2}]^{+}\geq 0,\ \mbox{ a.e. in }Q.

Hence one can see that:

Ω((α(𝒯Mη1(t))α(𝒯Mη2(t)))|θ(t)|,[η1η2]+(t)dx\displaystyle\int_{\Omega}\bigl{(}(\alpha^{\prime}(\mathcal{T}_{M}\eta^{1}(t))-\alpha^{\prime}(\mathcal{T}_{M}\eta^{2}(t))\bigr{)}|\nabla\theta(t)|,[\eta^{1}-\eta^{2}]^{+}(t)\,dx
0, for a.e. t(0,T).\displaystyle\geq 0,\ \mbox{ for a.e. }t\in(0,T). (5.23)

Now, in the light of (5.21)–(5.23), it is deduced that:

ddt(|[η1η2]+(t)|H2+μ2|[η1η2]+(t)|[H]N2)\displaystyle\frac{d}{dt}\left(\bigl{|}[\eta^{1}-\eta^{2}]^{+}(t)\bigr{|}_{H}^{2}+\mu^{2}\bigl{|}\nabla[\eta^{1}-\eta^{2}]^{+}(t)\bigr{|}_{[H]^{N}}^{2}\right)
2|g|L(M,M)|[η1η2]+(t)|H2, for a.e. t(0,T).\displaystyle\qquad\leq 2|g^{\prime}|_{L^{\infty}(-M,M)}\bigl{|}[\eta^{1}-\eta^{2}]^{+}(t)\bigr{|}_{H}^{2},\ \mbox{ for a.e. }t\in(0,T).

Applying Gronwall’s inequality, we arrive at:

|[η1η2]+(t)|H2+μ2|[η1η2]+(t)|[H]N2\displaystyle\bigl{|}[\eta^{1}-\eta^{2}]^{+}(t)\bigr{|}_{H}^{2}+\mu^{2}\bigl{|}\nabla[\eta^{1}-\eta^{2}]^{+}(t)\bigr{|}_{[H]^{N}}^{2} (5.24)
e2T|g|L(M,M)(|[η01η02]+|H2+μ2|[η01η02]+|[H]N2),\displaystyle\ \leq e^{2T|g^{\prime}|_{L^{\infty}(-M,M)}}\left(\bigl{|}[\eta_{0}^{1}-\eta_{0}^{2}]^{+}\bigr{|}_{H}^{2}+\mu^{2}\bigl{|}\nabla[\eta_{0}^{1}-\eta_{0}^{2}]^{+}\bigr{|}_{[H]^{N}}^{2}\right),
 for any t[0,T].\displaystyle\qquad\qquad\qquad\qquad\mbox{ for any }t\in[0,T].

(5.24) finishes the proof of lemma 5.4 with the constant:

C9:=1+μ21μ2e2T|g|L(M,M).C_{9}:=\frac{1+\mu^{2}}{1\wedge\mu^{2}}e^{2T|g^{\prime}|_{L^{\infty}(-M,M)}}.

The proof of 1.

Let us take ε(0,1)\varepsilon\in(0,1), τ(0,τ)\tau\in(0,\tau_{**}) where τ\tau_{**} is given in lemma 5.2. As a consequence of lemmas 5.1, 5.2 and 1, the following boundedness are derived:

  \bullet

{[ηε]τ|ε(0,1),τ(0,τ)}\{[\eta^{\varepsilon}]_{\tau}\,|\,\varepsilon\in(0,1),\tau\in(0,\tau_{**})\} is bounded in W1,2(0,T;W0)W^{1,2}(0,T;W_{0}),

  \bullet

{[η¯ε]τ|ε(0,1),τ(0,τ)}\{[\overline{\eta}^{\varepsilon}]_{\tau}\,|\,\varepsilon\in(0,1),\tau\in(0,\tau_{**})\}, {[η¯ε]τ|ε(0,1),τ(0,τ)}\{[\underline{\eta}^{\varepsilon}]_{\tau}\,|\,\varepsilon\in(0,1),\tau\in(0,\tau_{**})\} is bounded in L(0,T;W0)L^{\infty}(0,T;W_{0}),

  \bullet

{[θε]τ|ε(0,1),τ(0,τ)}\{[\theta^{\varepsilon}]_{\tau}\,|\,\varepsilon\in(0,1),\tau\in(0,\tau_{**})\} is bounded in L(0,T;W0)L^{\infty}(0,T;W_{0}) and in W1,2(0,T;V)W^{1,2}(0,T;V),

  \bullet

{[θ¯ε]τ|ε(0,1),τ(0,τ)}\{[\overline{\theta}^{\varepsilon}]_{\tau}\,|\,\varepsilon\in(0,1),\tau\in(0,\tau_{**})\}, {[θ¯ε]τ|ε(0,1),τ(0,τ)}\{[\underline{\theta}^{\varepsilon}]_{\tau}\,|\,\varepsilon\in(0,1),\tau\in(0,\tau_{**})\} is bounded in L(0,T;W0)L^{\infty}(0,T;W_{0}).

Therefore, by applying Aubin’s type compactness theory (cf. [23, Corollary 4]), we can find sequences {εn}n=1(0,1)\{\varepsilon_{n}\}_{n=1}^{\infty}\subset(0,1), {τn}n=1(0,τ)\{\tau_{n}\}_{n=1}^{\infty}\subset(0,\tau_{**}) and a pair of functions [η,θ][]2[\eta,\theta]\in[\mathscr{H}]^{2} such that εn0\varepsilon_{n}\searrow 0 and τn0\tau_{n}\searrow 0, as nn\to\infty, we obtain the following convergences as nn\to\infty:

{ηn:=[ηεn]τnη in C([0,T];V), and weakly in W1,2(0,T;W0),θn:=[θεn]τnθ in C([0,T];V), weakly in W1,2(0,T;V),and weakly- in L(0,T;W0).\left\{\begin{aligned} &\eta_{n}:=[\eta^{\varepsilon_{n}}]_{\tau_{n}}\to\eta\mbox{ in }C([0,T];V),\mbox{ and weakly in }W^{1,2}(0,T;W_{0}),\\ &\theta_{n}:=[\theta^{\varepsilon_{n}}]_{\tau_{n}}\to\theta\mbox{ in }C([0,T];V),\mbox{ weakly in }W^{1,2}(0,T;V),\\ &\qquad\mbox{and weakly-$*$ in }L^{\infty}(0,T;W_{0}).\end{aligned}\right. (5.25)

Besides, having in mind:

{max{|[η¯εn]τnηn|V,|[η¯εn]τnηn|V}Δi,τ|tηn|V𝑑t<,max{|[θ¯εn]τnθn|V,|[θ¯εn]τnθn|V}Δi,τ|tθn|V𝑑t<,\left\{\begin{aligned} &\max\left\{|[\overline{\eta}^{\varepsilon_{n}}]_{\tau_{n}}-\eta_{n}|_{V},\,|[\underline{\eta}^{\varepsilon_{n}}]_{\tau_{n}}-\eta_{n}|_{V}\right\}\leq\int_{\Delta_{i,\tau}}|\partial_{t}\eta_{n}|_{V}\,dt<\infty,\\ &\max\left\{|[\overline{\theta}^{\varepsilon_{n}}]_{\tau_{n}}-\theta_{n}|_{V},\,|[\underline{\theta}^{\varepsilon_{n}}]_{\tau_{n}}-\theta_{n}|_{V}\right\}\leq\int_{\Delta_{i,\tau}}|\partial_{t}\theta_{n}|_{V}\,dt<\infty,\end{aligned}\right.

we can derive that:

{η¯n:=[η¯εn]τnη and η¯n:=[η¯εn]τnη in L(0,T;V), and weakly- in L(0,T;W0),θ¯n:=[θ¯εn]τnθ and θ¯n:=[θ¯εn]τnθ in L(0,T;V), and weakly- in L(0,T;W0).\left\{\begin{aligned} &\overline{\eta}_{n}:=[\overline{\eta}^{\varepsilon_{n}}]_{\tau_{n}}\to\eta\mbox{ and }\underline{\eta}_{n}:=[\underline{\eta}^{\varepsilon_{n}}]_{\tau_{n}}\to\eta\mbox{ in }L^{\infty}(0,T;V),\mbox{ and }\\ &\mbox{weakly-$*$ in }L^{\infty}(0,T;W_{0}),\\ &\overline{\theta}_{n}:=[\overline{\theta}^{\varepsilon_{n}}]_{\tau_{n}}\to\theta\mbox{ and }\underline{\theta}_{n}:=[\underline{\theta}^{\varepsilon_{n}}]_{\tau_{n}}\to\theta\mbox{ in }L^{\infty}(0,T;V),\mbox{ and }\\ &\mbox{weakly-$*$ in }L^{\infty}(0,T;W_{0}).\end{aligned}\right. (5.26)

Now, we verify that the limiting pair [η,θ][\eta,\theta] satisfies (S0)–(S3). Let us take an arbitrary open interval I(0,T)I\subset(0,T). Then, in the light of (4.1), (4.2), and the convexity of γε\gamma_{\varepsilon}, the sequences as in (5.25) and (5.26) should fulfill the following two variational forms:

I(tηn(t),φ)𝑑t+I((η¯n+μ2tηn)(t),φ)[H]N𝑑t\displaystyle\int_{I}(\partial_{t}\eta_{n}(t),\varphi)\,dt+\int_{I}(\nabla(\overline{\eta}_{n}+\mu^{2}\partial_{t}\eta_{n})(t),\nabla\varphi)_{[H]^{N}}\,dt (5.27)
+I(g(𝒯Mη¯n(t)),+α(𝒯Mη¯n(t))γε(θ¯n(t)),φ)H𝑑t=I([u¯]τn(t),φ)H𝑑t,\displaystyle\quad+\int_{I}(g(\mathcal{T}_{M}\,\overline{\eta}_{n}(t)),+\alpha^{\prime}(\mathcal{T}_{M}\,\overline{\eta}_{n}(t))\gamma_{\varepsilon}(\nabla\overline{\theta}_{n}(t)),\varphi)_{H}\,dt=\int_{I}([\overline{u}]_{\tau_{n}}(t),\varphi)_{H}\,dt,
for all φV\varphi\in V, and n=1,2,3,,nτn=1,2,3,\dots,n_{\tau},

and

I((α0(𝒯Mη¯n)tθn)(t),θ¯n(t)ψ)H𝑑t+IΩα(η¯n(t))γε(θ¯n(t))𝑑x𝑑t\displaystyle\int_{I}\bigl{(}(\alpha_{0}(\mathcal{T}_{M}\underline{\eta}_{n})\partial_{t}\theta_{n})(t),\overline{\theta}_{n}(t)-\psi\bigr{)}_{H}\,dt+\int_{I}\int_{\Omega}\alpha(\underline{\eta}_{n}(t))\gamma_{\varepsilon}(\nabla\overline{\theta}_{n}(t))\,dxdt
+ν2I(tθn(t),(θ¯n(t)ψ))[H]N𝑑t\displaystyle\quad+\nu^{2}\int_{I}(\nabla\partial_{t}\theta_{n}(t),\nabla(\overline{\theta}_{n}(t)-\psi))_{[H]^{N}}\,dt (5.28)
IΩα(η¯n(t))γε(ψ)𝑑x𝑑t+I([v¯]τn(t),θ¯n(t)ψ)H𝑑t,\displaystyle\quad\leq\int_{I}\int_{\Omega}\alpha(\underline{\eta}_{n}(t))\gamma_{\varepsilon}(\nabla\psi)\,dxdt+\int_{I}([\overline{v}]_{\tau_{n}}(t),\overline{\theta}_{n}(t)-\psi)_{H}\,dt,
for all ψV\psi\in V, and n=1,2,3,,nτn=1,2,3,\dots,n_{\tau}.

On this basis, having in mind (5.25), (A1), (A2), and the fact:

γε|| uniformly on N, as ε0,\gamma_{\varepsilon}\to|\cdot|\mbox{ uniformly on }\mathbb{R}^{N},\mbox{ as }\varepsilon\to 0,

letting nn\to\infty in (5.27) and (5.28) yields that:

I(tη(t),φ)𝑑t+I((η+μ2tη)(t),φ)[H]N𝑑t\displaystyle\int_{I}(\partial_{t}\eta(t),\varphi)\,dt+\int_{I}(\nabla(\eta+\mu^{2}\partial_{t}\eta)(t),\nabla\varphi)_{[H]^{N}}\,dt (5.29)
+I(g(𝒯Mη(t))+α(𝒯Mη(t))|θ(t)|,φ)H𝑑t=I(u(t),φ)H𝑑t,\displaystyle\quad+\int_{I}(g(\mathcal{T}_{M}\,\eta(t))+\alpha^{\prime}(\mathcal{T}_{M}\,\eta(t))|\nabla\theta(t)|,\varphi)_{H}\,dt=\int_{I}(u(t),\varphi)_{H}\,dt,
 for any φV,\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\mbox{ for any }\varphi\in V,

and

I((α0(𝒯Mη)tθ)(t),θ(t)ψ)H𝑑t+IΩα(η(t))|θ(t)|𝑑x𝑑t\displaystyle\int_{I}\bigl{(}(\alpha_{0}(\mathcal{T}_{M}\eta)\partial_{t}\theta)(t),\theta(t)-\psi\bigr{)}_{H}\,dt+\int_{I}\int_{\Omega}\alpha(\eta(t))|\nabla\theta(t)|\,dxdt
+ν2I(tθ(t),(θ(t)ψ))[H]N𝑑t\displaystyle\quad+\nu^{2}\int_{I}(\nabla\partial_{t}\theta(t),\nabla(\theta(t)-\psi))_{[H]^{N}}\,dt
IΩα(η(t))|ψ|𝑑x𝑑t+I(v(t),θ(t)ψ)H𝑑t, for any ψV,\displaystyle\quad\leq\int_{I}\int_{\Omega}\alpha(\eta(t))|\nabla\psi|\,dxdt+\int_{I}(v(t),\theta(t)-\psi)_{H}\,dt,\ \mbox{ for any }\psi\in V,

respectively. Since I(0,T)I\subset(0,T) is arbitrary, [η,θ][\eta,\theta] should satisfy (S1) and (S2).

Next, let us verify ηL(Q)\eta\in L^{\infty}(Q). By (5.2), the following inequalities can be obtained:

{tMΔ(M+μ2tM)+g(M)+α(M)|θ(t)|u,t(M)Δ((M)+μ2t(M))+g(M)+α(M)|θ(t)|u,\displaystyle\left\{\begin{aligned} &\partial_{t}M-\mathit{\Delta}(M+\mu^{2}\partial_{t}M)+g(M)+\alpha^{\prime}(M)|\nabla\theta(t)|\geq u,\\ &\partial_{t}(-M)-\mathit{\Delta}((-M)+\mu^{2}\partial_{t}(-M))+g(-M)+\alpha^{\prime}(-M)|\nabla\theta(t)|\leq u,\end{aligned}\right.
a.e. in Q.\displaystyle\mbox{a.e. in }Q.

Hence, applying lemma 5.4 to the case when

[η1,η2,θ~,u~]=[η,M,θ,u],[η01,η02]=[η0,M],[\eta^{1},\eta^{2},\widetilde{\theta},\widetilde{u}]=[\eta,M,\theta,u],~{}~{}[\eta_{0}^{1},\eta_{0}^{2}]=[\eta_{0},M],

and

[η1,η2,θ~,u~]=[M,η,θ,u],[η01,η02]=[M,η0],[\eta^{1},\eta^{2},\widetilde{\theta},\widetilde{u}]=[-M,\eta,\theta,u],~{}~{}[\eta_{0}^{1},\eta_{0}^{2}]=[-M,\eta_{0}],

we arrive at:

{|[ηM]+(t)|V2C9|[η0M]+|V2=0,|[Mη]+(t)|V2C9|[Mη0]+|V2=0, for any t[0,T],\displaystyle\left\{\begin{aligned} &\bigl{|}[\eta-M]^{+}(t)\bigr{|}_{V}^{2}\leq C_{9}\bigl{|}[\eta_{0}-M]^{+}\bigr{|}_{V}^{2}=0,\\ &\bigl{|}[-M-\eta]^{+}(t)\bigr{|}_{V}^{2}\leq C_{9}\bigl{|}[-M-\eta_{0}]^{+}\bigr{|}_{V}^{2}=0,\end{aligned}\right.\mbox{ for any }t\in[0,T], (5.30)

respectively. This implies that:

|η(t)|L(Ω)M, for any t[0,T].|\eta(t)|_{L^{\infty}(\Omega)}\leq M,\ \mbox{ for any }t\in[0,T].

Finally, due to (5.25), the condition (S3) as in 1 is immediately confirmed as follows:

η(0)=limnηn(0)=η0 and θ(0)=limnθn(0)=θ0 in H.\eta(0)=\lim_{n\to\infty}\eta_{n}(0)=\eta_{0}\mbox{ and }\theta(0)=\lim_{n\to\infty}\theta_{n}(0)=\theta_{0}\mbox{ in }H.

Thus, we conclude that [η,θ][\eta,\theta] is a solution to (S).

5.2 Proof of 2

Let [ηk,θk],k=1,2[\eta^{k},\theta^{k}],\,k=1,2, be the solutions to (S) corresponding to initial values η0k,θ0k\eta_{0}^{k},\,\theta_{0}^{k} and forcings uk,vku^{k},\,v^{k}, k=1,2k=1,2. Let us set

M0:=|η1|L(Q)|η2|L(Q), and δ(ν):=1δαν2,M_{0}:=|\eta^{1}|_{L^{\infty}(Q)}\vee|\eta^{2}|_{L^{\infty}(Q)},\mbox{ and }\delta_{*}(\nu):=1\wedge\delta_{\alpha}\wedge\nu^{2}, (5.31)

and take the difference between the variational formulas for ηk,k=1,2\eta^{k},\,k=1,2, and put φ:=(η1η2)(t)\varphi:=(\eta^{1}-\eta^{2})(t). Then, by using (A1), the monotonicity of α\alpha^{\prime}, and Young’s inequality, we see that:

12ddt(|(η1η2)(t)|H2+μ2|(η1η2)(t)|[H]N2)\displaystyle\frac{1}{2}\frac{d}{dt}\bigl{(}|(\eta^{1}-\eta^{2})(t)|_{H}^{2}+\mu^{2}|\nabla(\eta^{1}-\eta^{2})(t)|_{[H]^{N}}^{2}\bigr{)} (5.32)
|g|L(M0,M0)|(η1η2)(t)|H2\displaystyle\quad\qquad-|g|_{L^{\infty}(-M_{0},M_{0})}|(\eta^{1}-\eta^{2})(t)|_{H}^{2}
Ωα(η1(t))(η2η1)(t)|θ1(t)|𝑑x\displaystyle\quad\leq\int_{\Omega}\alpha^{\prime}(\eta^{1}(t))(\eta^{2}-\eta^{1})(t)|\nabla\theta^{1}(t)|\,dx
+Ωα(η2(t))(η1η2)(t)|θ2(t)|𝑑x\displaystyle\quad\qquad+\int_{\Omega}\alpha^{\prime}(\eta^{2}(t))(\eta^{1}-\eta^{2})(t)|\nabla\theta^{2}(t)|\,dx
+((η1η2)(t),(u1u2)(t))H\displaystyle\quad\qquad+\bigl{(}(\eta^{1}-\eta^{2})(t),(u^{1}-u^{2})(t)\bigr{)}_{H}
Ω|α(η1(t))α(η2(t))||(θ1θ2)(t)|𝑑x\displaystyle\quad\leq\int_{\Omega}|\alpha(\eta^{1}(t))-\alpha(\eta^{2}(t))||\nabla(\theta^{1}-\theta^{2})(t)|\,dx
+|(η1η2)(t)|H|(u1u2)(t)|H\displaystyle\quad\qquad+|(\eta^{1}-\eta^{2})(t)|_{H}|(u^{1}-u^{2})(t)|_{H}
|α|L(M0,M0)2(|(η1η2)(t)|H2+|(θ1θ2)(t)|[H]N2)\displaystyle\quad\leq\frac{|\alpha^{\prime}|_{L^{\infty}(-M_{0},M_{0})}}{2}\left(|(\eta^{1}-\eta^{2})(t)|_{H}^{2}+|\nabla(\theta^{1}-\theta^{2})(t)|_{[H]^{N}}^{2}\right)
+12|(η1η2)(t)|H2+12|(u1u2)(t)|H2, for a.e. t>0.\displaystyle\quad\qquad+\frac{1}{2}|(\eta^{1}-\eta^{2})(t)|_{H}^{2}+\frac{1}{2}|(u^{1}-u^{2})(t)|_{H}^{2},\ \mbox{ for a.e. }t>0.

On the other hand, by putting ψ=θ2\psi=\theta^{2} in the variational inequality for θ1\theta^{1}, and ψ=θ1\psi=\theta^{1} in the one for θ2\theta^{2}, and by taking the sum of two inequalities, we have:

(α0(η1)tθ1(t)α0(η2)tθ2(t),(θ1θ2)(t))H\displaystyle\bigl{(}\alpha_{0}(\eta^{1})\partial_{t}\theta^{1}(t)-\alpha_{0}(\eta^{2})\partial_{t}\theta^{2}(t),(\theta^{1}-\theta^{2})(t)\bigr{)}_{H} (5.33)
+12ddt(ν2|(θ1θ2)(t)|[H]N2)\displaystyle\quad\qquad+\frac{1}{2}\frac{d}{dt}\bigl{(}\nu^{2}|\nabla(\theta^{1}-\theta^{2})(t)|_{[H]^{N}}^{2}\bigr{)}
Ωα(η1(t))(|θ2(t)||θ1(t)|)𝑑x\displaystyle\quad\leq\int_{\Omega}\alpha(\eta^{1}(t))\bigl{(}|\nabla\theta^{2}(t)|-|\nabla\theta^{1}(t)|\bigr{)}\,dx
+Ωα(η2(t))(|θ1(t)||θ2(t)|)𝑑x\displaystyle\quad\qquad+\int_{\Omega}\alpha(\eta^{2}(t))\bigl{(}|\nabla\theta^{1}(t)|-|\nabla\theta^{2}(t)|\bigr{)}\,dx
+((θ1θ2)(t),(v1v2)(t))H\displaystyle\quad\qquad+\bigl{(}(\theta^{1}-\theta^{2})(t),(v^{1}-v^{2})(t)\bigr{)}_{H}
Ω|α(η1(t))α(η2(t))|(θ1θ2)(t)|dx\displaystyle\quad\leq\int_{\Omega}|\alpha(\eta^{1}(t))-\alpha(\eta^{2}(t))|\nabla(\theta^{1}-\theta^{2})(t)|\,dx
+|(θ1θ2)(t)|H|(v1v2)(t)|H\displaystyle\quad\qquad+|(\theta^{1}-\theta^{2})(t)|_{H}|(v^{1}-v^{2})(t)|_{H}
|α|L(M0,M0)2(|(η1η2)(t)|H2+|(θ1θ2)(t)|[H]N2)\displaystyle\quad\leq\frac{|\alpha^{\prime}|_{L^{\infty}(-M_{0},M_{0})}}{2}\left(|(\eta^{1}-\eta^{2})(t)|_{H}^{2}+|\nabla(\theta^{1}-\theta^{2})(t)|_{[H]^{N}}^{2}\right)
+12|(θ1θ2)(t)|H2+12|(v1v2)(t)|H2, a.e. t>0.\displaystyle\quad\qquad+\frac{1}{2}|(\theta^{1}-\theta^{2})(t)|_{H}^{2}+\frac{1}{2}|(v^{1}-v^{2})(t)|_{H}^{2},\ \mbox{ a.e. }t>0.

Here, we can compute the first term in (5.33) as follows:

(α0(η1(t))tθ1(t)α0(η2(t))tθ2(t),(θ1θ2)(t))H\displaystyle\quad\bigl{(}\alpha_{0}(\eta^{1}(t))\partial_{t}\theta^{1}(t)-\alpha_{0}(\eta^{2}(t))\partial_{t}\theta^{2}(t),(\theta^{1}-\theta^{2})(t)\bigr{)}_{H} (5.34)
=12ddt|α0(η1(t))(θ1θ2)(t)|H212Ωα0(η1(t))tη1(t)|(θ1θ2)(t)|2dx\displaystyle=\frac{1}{2}\frac{d}{dt}|\sqrt{\alpha_{0}(\eta^{1}(t))}(\theta^{1}-\theta^{2})(t)|_{H}^{2}-\frac{1}{2}\int_{\Omega}\alpha_{0}^{\prime}(\eta^{1}(t))\partial_{t}\eta^{1}(t)|(\theta^{1}-\theta^{2})(t)|^{2}\,dx
+Ω(α0(η1(t))α0(η2(t)))tθ2(t)(θ1θ2)(t)dx.\displaystyle\quad+\int_{\Omega}(\alpha_{0}(\eta^{1}(t))-\alpha_{0}(\eta^{2}(t)))\partial_{t}\theta^{2}(t)(\theta^{1}-\theta^{2})(t)\,dx.

Also, by using (5.31), the continuous embedding from H1(Ω)H^{1}(\Omega) to L4(Ω)L^{4}(\Omega) under N3N\leq 3, and Young’s inequality, one can see that:

Ωα0(η1(t))tη1(t)|(θ1θ2)(t)|2dx\displaystyle-\int_{\Omega}\alpha_{0}^{\prime}(\eta^{1}(t))\partial_{t}\eta^{1}(t)|(\theta^{1}-\theta^{2})(t)|^{2}\,dx (5.35)
\displaystyle\geq |α0|L(M0,M0)|tη1(t)|H|(θ1θ2)(t)|L4(Ω)2\displaystyle-|\alpha_{0}^{\prime}|_{L^{\infty}(-M_{0},M_{0})}|\partial_{t}\eta^{1}(t)|_{H}|(\theta_{1}-\theta^{2})(t)|_{L^{4}(\Omega)}^{2}
\displaystyle\geq (CH1L4)2|α0|L(M0,M0)|tη1(t)|H|(θ1θ2)(t)|V2\displaystyle-(C_{H^{1}}^{L^{4}})^{2}|\alpha_{0}^{\prime}|_{L^{\infty}(-M_{0},M_{0})}|\partial_{t}\eta^{1}(t)|_{H}|(\theta_{1}-\theta^{2})(t)|_{V}^{2}
\displaystyle\geq (CH1L4)2|α0|L(M0,M0)δ(ν)|tη1(t)|H\displaystyle-\frac{(C_{H^{1}}^{L^{4}})^{2}|\alpha_{0}^{\prime}|_{L^{\infty}(-M_{0},M_{0})}}{\delta_{*}(\nu)}|\partial_{t}\eta^{1}(t)|_{H}\cdot
(|α0(η1(t))(θ1θ2)(t)|H2+ν2|(θ1θ2)(t)|[H]N2),\displaystyle\qquad\cdot\bigl{(}|\sqrt{\alpha_{0}(\eta^{1}(t))}(\theta^{1}-\theta^{2})(t)|_{H}^{2}+\nu^{2}|\nabla(\theta^{1}-\theta^{2})(t)|_{[H]^{N}}^{2}\bigr{)},

and

Ω(α0(η1(t))α0(η2(t)))tθ2(t)(θ1θ2)(t)dx\displaystyle\int_{\Omega}(\alpha_{0}(\eta^{1}(t))-\alpha_{0}(\eta^{2}(t)))\partial_{t}\theta^{2}(t)(\theta^{1}-\theta^{2})(t)\,dx (5.36)
\displaystyle\geq |α0|L(M0,M0)|tθ2(t)|L4(Ω)|(η1η2)(t)|H|(θ1θ2)(t)|L4(Ω)\displaystyle-|\alpha_{0}^{\prime}|_{L^{\infty}(-M_{0},M_{0})}|\partial_{t}\theta^{2}(t)|_{L^{4}(\Omega)}|(\eta^{1}-\eta^{2})(t)|_{H}|(\theta^{1}-\theta^{2})(t)|_{L^{4}(\Omega)}
\displaystyle\geq (CH1L4)2|α0|L(M0,M0)|tθ2(t)|V|(η1η2)(t)|H|(θ1θ2)(t)|V\displaystyle-(C_{H^{1}}^{L^{4}})^{2}|\alpha_{0}^{\prime}|_{L^{\infty}(-M_{0},M_{0})}|\partial_{t}\theta^{2}(t)|_{V}|(\eta^{1}-\eta^{2})(t)|_{H}|(\theta^{1}-\theta^{2})(t)|_{V}
\displaystyle\geq (CH1L4)2|α0|L(M0,M0)2|tθ2(t)|V(|(η1η2)(t)|H2+|(θ1θ2)(t)|V2)\displaystyle-\frac{(C_{H^{1}}^{L^{4}})^{2}|\alpha_{0}^{\prime}|_{L^{\infty}(-M_{0},M_{0})}}{2}|\partial_{t}\theta^{2}(t)|_{V}\bigl{(}|(\eta^{1}-\eta^{2})(t)|_{H}^{2}+|(\theta^{1}-\theta^{2})(t)|_{V}^{2}\bigr{)}
\displaystyle\geq (CH1L4)2|α0|L(M0,M0)2δ(ν)|tθ2(t)|V\displaystyle-\frac{(C_{H^{1}}^{L^{4}})^{2}|\alpha_{0}^{\prime}|_{L^{\infty}(-M_{0},M_{0})}}{2\delta_{*}(\nu)}|\partial_{t}\theta^{2}(t)|_{V}\cdot
(|(η1η2)(t)|H2+|α0(η1(t))(θ1θ2)(t)|H2+ν2|(θ1θ2)(t)|[H]N2),\displaystyle\quad\cdot\bigl{(}|(\eta^{1}-\eta^{2})(t)|_{H}^{2}+|\sqrt{\alpha_{0}(\eta^{1}(t))}(\theta^{1}-\theta^{2})(t)|_{H}^{2}+\nu^{2}|\nabla(\theta^{1}-\theta^{2})(t)|_{[H]^{N}}^{2}\bigr{)},
 for a.e. t>0,\displaystyle\qquad\qquad\qquad\qquad\qquad\mbox{ for a.e. }t>0,

where CH1L4C_{H^{1}}^{L^{4}} is a constant of the continuous embedding from H1(Ω)H^{1}(\Omega) to L4(Ω)L^{4}(\Omega).

Therefore, putting

J1(t)\displaystyle J_{1}(t) :=|(u1u2)(t)|H2+|(v1v2)(t)|H2, for t0,\displaystyle:=|(u^{1}-u^{2})(t)|_{H}^{2}+|(v^{1}-v^{2})(t)|_{H}^{2},\ \mbox{ for }t\geq 0,

and

C~10:=2(|α|L(M0,M0)+|g|L(M0,M0)+(CH1L4)2|α0|L(M0,M0)+1),\widetilde{C}_{10}:=2\bigl{(}|\alpha^{\prime}|_{L^{\infty}(-M_{0},M_{0})}+|g^{\prime}|_{L^{\infty}(-M_{0},M_{0})}+(C_{H^{1}}^{L^{4}})^{2}|\alpha_{0}^{\prime}|_{L^{\infty}(-M_{0},M_{0})}+1\bigr{)},

it is deduced from (5.32)–(5.36) that:

ddtJ(t)C~10δ(ν)(|tη1(t)|H+|tθ2(t)|V+1)J(t)+J1(t), a.e. t>0,\frac{d}{dt}J(t)\leq\frac{\widetilde{C}_{10}}{\delta_{*}(\nu)}\bigl{(}|\partial_{t}\eta^{1}(t)|_{H}+|\partial_{t}\theta^{2}(t)|_{V}+1\bigr{)}J(t)+J_{1}(t),\ \mbox{ a.e. }t>0, (5.37)

Applying Gronwall’s lemma in (5.37), it can be obtained that for any T>0T>0,:

J(t)\displaystyle J(t) C1(ν)(J(0)+0TJ1(s)𝑑s), for any t[0,T].\displaystyle\leq C_{1}(\nu)\left(J(0)+\int_{0}^{T}J_{1}(s)\,ds\right),\ \mbox{ for any }t\in[0,T].

with

C1(ν):=exp(C~10Tδ(ν)(|tη1|+|tθ2|𝒱+T))C_{1}(\nu):=\exp\left(\frac{\widetilde{C}_{10}\sqrt{T}}{\delta_{*}(\nu)}\bigl{(}|\partial_{t}\eta^{1}|_{\mathscr{H}}+|\partial_{t}\theta^{2}|_{\mathscr{V}}+\sqrt{T}\bigr{)}\right)

Thus, we finish the proof of 2. ∎

References

  • [1] T. Aiki, D. Mizuno, and K. Shirakawa, A class of initial-boundary value problems governed by pseudo-parabolic weighted total variation flows, Advances in Mathematical Sciences and Applications, 32 (2023), pp. 311–341, https://mcm-www.jwu.ac.jp/~aikit/AMSA/pdf/abstract/2023/Top_2023_015.pdf.
  • [2] M. Amar and G. Bellettini, A notion of total variation depending on a metric with discontinuous coefficients, Ann. Inst. H. Poincaré Anal. Non Linéaire, 11 (1994), pp. 91–133, https://doi.org/10.1016/S0294-1449(16)30197-4.
  • [3] M. Amar, V. De Cicco, and N. Fusco, A relaxation result in BV for integral functionals with discontinuous integrands, ESAIM Control Optim. Calc. Var., 13 (2007), pp. 396–412, https://doi.org/10.1051/cocv:2007015.
  • [4] L. Ambrosio, N. Fusco, and D. Pallara, Functions of Bounded Variation and Free Discontinuity Problems, Oxford Mathematical Monographs, The Clarendon Press, Oxford University Press, New York, 2000.
  • [5] H. Antil, S. Kubota, K. Shirakawa, and N. Yamazaki, Optimal control problems governed by 1-D Kobayashi-Warren-Carter type systems, Math. Control Relat. Fields, 11 (2021), pp. 253–289, https://doi.org/10.3934/mcrf.2020036.
  • [6] H. Antil, S. Kubota, K. Shirakawa, and N. Yamazaki, Constrained optimization problems governed by PDE models of grain boundary motions, Adv. Nonlinear Anal., 11 (2022), pp. 1249–1286, https://doi.org/10.1515/anona-2022-0242.
  • [7] H. Antil, K. Shirakawa, and N. Yamazaki, A class of parabolic systems associated with optimal controls of grain boundary motions, Adv. Math. Sci. Appl., 27 (2018), pp. 299–336.
  • [8] H. Attouch, G. Buttazzo, and G. Michaille, Variational analysis in Sobolev and BV spaces, vol. 17 of MOS-SIAM Series on Optimization, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA; Mathematical Optimization Society, Philadelphia, PA, second ed., 2014, https://doi.org/10.1137/1.9781611973488, Applications to PDEs and optimization.
  • [9] H. Brézis, Opérateurs Maximaux Monotones et Semi-groupes de Contractions dans les Espaces de Hilbert, North-Holland Publishing Co., Amsterdam-London; American Elsevier Publishing Co., Inc., New York, 1973. North-Holland Mathematics Studies, No. 5. Notas de Matemática (50).
  • [10] E. Emmrich, Discrete versions of gronwall’s lemma and their application to the numerical analysis of parabolic problems, Tech. Report 637, Institute of Mathematics, Technische Universität Berlin, “http://www3.math.tu-berlin.de/preprints/files/Preprint-637-1999.pdf”, 1999.
  • [11] L. C. Evans and R. F. Gariepy, Measure Theory and Fine Properties of Functions, Textbooks in Mathematics, CRC Press, Boca Raton, FL, revised ed., 2015.
  • [12] E. Giusti, Minimal Surfaces and Functions of Bounded Variation, vol. 80 of Monographs in Mathematics, Birkhäuser Verlag, Basel, 1984, https://doi.org/10.1007/978-1-4684-9486-0.
  • [13] A. Ito, N. Kenmochi, and N. Yamazaki, A phase-field model of grain boundary motion, Appl. Math., 53 (2008), pp. 433–454, https://doi.org/10.1007/s10492-008-0035-8.
  • [14] R. Kobayashi, J. A. Warren, and W. C. Carter, A continuum model of grain boundaries, Phys. D, 140 (2000), pp. 141–150, https://doi.org/10.1016/S0167-2789(00)00023-3.
  • [15] R. Kobayashi, J. A. Warren, and W. C. Carter, Grain boundary model and singular diffusivity, in Free boundary problems: theory and applications, II (Chiba, 1999), vol. 14 of GAKUTO Internat. Ser. Math. Sci. Appl., Gakkōtosho, Tokyo, 2000, pp. 283–294.
  • [16] S. Kubota, R. Nakayashiki, and K. Shirakawa, Optimal control problems for 1D parabolic state-systems of KWC types with dynamic boundary conditions, Adv. Math. Sci. Appl., 29 (2020), pp. 583–637.
  • [17] S. Kubota and K. Shirakawa, Periodic solutions to kobayashi–warren–carter systems, Advances in Mathematical Sciences and Applications, 32 (2023), pp. 511–553, https://mcm-www.jwu.ac.jp/~aikit/AMSA/pdf/abstract/2023/Top_2023_022.pdf.
  • [18] S. Moll and K. Shirakawa, Existence of solutions to the Kobayashi–Warren–Carter system, Calc. Var. Partial Differential Equations, 51 (2014), pp. 621–656, https://doi.org/10.1007/s00526-013-0689-2.
  • [19] S. Moll, K. Shirakawa, and H. Watanabe, Energy dissipative solutions to the Kobayashi–Warren–Carter system, Nonlinearity, 30 (2017), pp. 2752–2784, https://doi.org/10.1088/1361-6544/aa6eb4.
  • [20] S. Moll, K. Shirakawa, and H. Watanabe, Kobayashi-Warren-Carter type systems with nonhomogeneous Dirichlet boundary data for crystalline orientation, Nonlinear Anal., 217 (2022), pp. Paper No. 112722, 44, https://doi.org/10.1016/j.na.2021.112722.
  • [21] R. Nakayashiki and K. Shirakawa, Kobayashi-warren-carter system of singular type under dynamic boundary condition, Discrete and Continuous Dynamical Systems – S, 16 (2023), pp. 3746–3783, https://doi.org/10.3934/dcdss.2023162.
  • [22] K. Shirakawa, H. Watanabe, and N. Yamazaki, Solvability of one-dimensional phase field systems associated with grain boundary motion, Math. Ann., 356 (2013), pp. 301–330, https://doi.org/10.1007/s00208-012-0849-2.
  • [23] J. Simon, Compact sets in the space Lp(0,T;B)L^{p}(0,T;B), Ann. Mat. Pura Appl. (4), 146 (1987), pp. 65–96, https://doi.org/10.1007/BF01762360.
  • [24] N. Yamazaki, Global attractors for non-autonomous phase-field systems of grain boundary motion with constraint, Adv. Math. Sci. Appl., 23 (2013), pp. 267–296.