Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Research article Special Issues

Solvability for the non-isothermal Kobayashi–Warren–Carter system

  • In this paper, a system of parabolic type initial-boundary value problems are considered. The system (S)ν is based on the non-isothermal model of grain boundary motion by [38], which was derived as an extending version of the "Kobayashi--Warren--Carter model" of grain boundary motion by [23]. Under suitable assumptions, the existence theorem of L2-based solutions is concluded, as a versatile mathematical theory to analyze various Kobayashi--Warren--Carter type models.

    Citation: Ken Shirakawa, Hiroshi Watanabe. Solvability for the non-isothermal Kobayashi–Warren–Carter system[J]. AIMS Mathematics, 2017, 2(1): 161-194. doi: 10.3934/Math.2017.1.161

    Related Papers:

    [1] Zhongying Liu, Yang Liu, Yiqi Jiang . Global solvability of 3D non-isothermal incompressible nematic liquid crystal flows. AIMS Mathematics, 2022, 7(7): 12536-12565. doi: 10.3934/math.2022695
    [2] Yasir Nadeem Anjam . The qualitative analysis of solution of the Stokes and Navier-Stokes system in non-smooth domains with weighted Sobolev spaces. AIMS Mathematics, 2021, 6(6): 5647-5674. doi: 10.3934/math.2021334
    [3] Joseph L. Shomberg . Well-posedness and global attractors for a non-isothermal viscous relaxationof nonlocal Cahn-Hilliard equations. AIMS Mathematics, 2016, 1(2): 102-136. doi: 10.3934/Math.2016.2.102
    [4] Bengisen Pekmen Geridonmez . RBF simulation of natural convection in a nanofluid-filled cavity. AIMS Mathematics, 2016, 1(3): 195-207. doi: 10.3934/Math.2016.3.195
    [5] Özge Arıbaş, İsmet Gölgeleyen, Mustafa Yıldız . On the solvability of an initial-boundary value problem for a non-linear fractional diffusion equation. AIMS Mathematics, 2023, 8(3): 5432-5444. doi: 10.3934/math.2023273
    [6] Santiago Cano-Casanova . Positive weak solutions for heterogeneous elliptic logistic BVPs with glued Dirichlet-Neumann mixed boundary conditions. AIMS Mathematics, 2023, 8(6): 12606-12621. doi: 10.3934/math.2023633
    [7] Doaa Filali, Mohammad Dilshad, Mohammad Akram . Generalized variational inclusion: graph convergence and dynamical system approach. AIMS Mathematics, 2024, 9(9): 24525-24545. doi: 10.3934/math.20241194
    [8] A.S. Hendy, R.H. De Staelen, A.A. Aldraiweesh, M.A. Zaky . High order approximation scheme for a fractional order coupled system describing the dynamics of rotating two-component Bose-Einstein condensates. AIMS Mathematics, 2023, 8(10): 22766-22788. doi: 10.3934/math.20231160
    [9] Yuanqiang Chen, Jihui Zheng, Jing An . A Legendre spectral method based on a hybrid format and its error estimation for fourth-order eigenvalue problems. AIMS Mathematics, 2024, 9(3): 7570-7588. doi: 10.3934/math.2024367
    [10] Yun-guang Lu, Xian-ting Wang, Richard De la cruz . Cauchy problem for isothermal system in a general nozzle with space-dependent friction. AIMS Mathematics, 2021, 6(6): 6482-6489. doi: 10.3934/math.2021381
  • In this paper, a system of parabolic type initial-boundary value problems are considered. The system (S)ν is based on the non-isothermal model of grain boundary motion by [38], which was derived as an extending version of the "Kobayashi--Warren--Carter model" of grain boundary motion by [23]. Under suitable assumptions, the existence theorem of L2-based solutions is concluded, as a versatile mathematical theory to analyze various Kobayashi--Warren--Carter type models.


    1. Introduction

    Let 0<T< be a constant of time, and let NN be a constant of spatial dimension such that 1N3. Let ΩRN be a bounded domain such that Γ:=Ω is smooth when N>1. Besides, let us denote by Q:=(0,T)×Ω the product space of the time-interval (0,T) and the spatial domain Ω, and similarly, let us set Σ:=(0,T)×Γ.

    In this paper, we fix a constant ν0, and consider the following system of initial-boundary value problems of parabolic types, denoted by (S)ν.

    (S)ν:

    {[uλ(w)]tΔu=f   in Q, DunΓ+n0(ufΓ)=0   on Σ, u(0,x)=u0(x),  xΩ; (1.1)
    {wtΔw+γ(w)+gw(w,η)+λ(w)u              +αw(w,η)|Dθ|+ν2βw(w,η)|Dθ|20 in Q,DwnΓ=0   on Σ,w(0,x)=w0(x),  xΩ; (1.2)
    {ηtΔη+gη(w,η)+αη(w,η)|Dθ|+ν2βη(w,η)|Dθ|2=0   in Q,DηnΓ=0   on Σ,η(0,x)=η0(x),  xΩ; (1.3)
    {α0(w,η)θtdiv(α(w,η)Dθ|Dθ|+2ν2β(w,η)Dθ)=0   in Q,(α(w,η)Dθ|Dθ|+2ν2β(w,η)Dθ)nΓ=0   on Σ,θ(0,x)=θ0(x),  xΩ. (1.4)

    Here, Du, Dw, Dη and Dθ denote, respectively, the (distributional) gradients of the unknowns u, w, η and θ on Ω. f=f(t,x) is the source term on Q, fΓ=fΓ(t,x) is the boundary source on Σ. u0=u0(x), w0=w0(x), η0=η0(x) and θ0=θ0(x) are given initial data on Ω. γ is the subdifferential of a proper lower semi-continuous (l.s.c.) and convex function γ=γ(w) on R. λ=λ(w), g=g(w,η), α0=α0(w,η), α=α(w,η) and β=β(w,η) are given real-valued functions, and the scripts "'", "w" and "η" denote differentials with respect to the corresponding variables. n0 is a given positive constant, and nΓ is the unit outer normal on Γ.

    The system (S)ν is based on the non-isothermal model of grain boundary motion by Warren et al. [36], which was derived as an extending version of the "Kobayashi-Warren-Carter model" of grain boundary motion by Kobayashi et al. [22,23]. Hence, the study of this paper is based on the previous works related to the Kobayashi-Warren-Carter model (e.g., [13,15,16,17,20,21,22,23,25,26,28,29,30,31,32,36,37,39]).

    According to the modeling method of [36], the system (S)ν is roughly configured as a coupled system of the heat equation in (1.1), and a gradient system {(1.2)-(1.4)} of the following governing energy, called free-energy:

    εν(u,w,η,θ):=12Ω|Dw|2dx+Ωγ(w)dx+Ωuλ(w)dx+12Ω|Dη|2dx+Ωg(w,η)dx+Ωα(w,η)d|Dθ|+Ωβ(w,η)|D(νθ)|2dx, (1.5)

    for [u,w,η,θ]L2(Ω)×H1(Ω)×H1(Ω)×BV(Ω) with νθH1(Ω).

    In this context, the unknown u=u(t,x) is the relative temperature with the critical degree 0, and the unknown w=w(t,x) is an order parameter to indicate the solidification order of the polycrystal. The term uλ(w) in (1.1) is the so-called enthalpy, and then the term λ(w) corresponds to the effect of the latent heat. The unknowns η=η(t,x) and θ=θ(t,x) are components of the vector field

    (t,x)Qη(t,x)[ cosθ(t,x),sinθ(t,x)]R2,

    which was adopted in [22,23] as a vectorial phase field to reproduce the crystalline orientation in Q. Here, the components η and θ are order parameters to indicate, respectively, the orientation order and angle of the grain. In particular, w and η are taken to satisfy the constraints 0w,η1 in Q, and the cases [w,η][1,1] and [w,η][0,0] are respectively assigned to "the solidified-oriented phase" and "the liquefied-disoriented phase" which correspond to two stable phases in physical.

    In view of these, we suppose that

    (g0) the function w ∈ [0, 1] 7→ λ(w) ∈ R is increasing, and if the temperature u is closed to the critical

    value, i.e. u ≈ 0, then the function

    [u,w,η]R2γ(w)+g(w,η)λ(w)u(,]

    has two minimums, around [1,1] and [0,0].

    Besides, referring to the previous works on phase transitions (e.g., [7,8,14,18,19,34,35]), we can exemplify the following settings as possible expressions of the functions λ, γ and g in the above (g0):

    (g1) (constrained setting by logarithmic function; cf. [14,34,35])

    {λ(w)=Lw, γ(w):=12(wlogw+(1w)log(1w))with  γ(0)=γ(1):=1,g(w,η):=L2(w12)2+c2(wη)2,    for  w,ηR,

    (g2) (setting with non-smooth constraint; cf. [7,8,18,19,35])

    {λ(w)=Lw,γ(w):=I[0,1](w),g(w,η):=L2(w12)2+c2(wη)2,   for  w,ηR.

    Here, L and c are positive constants, and R{0,} is the indicator function on the compact interval [0,1].

    Now, the objective of this study is to generalize the line of recent results [25,26,28,29,30,31,32,37,39], and to obtain an enhanced theory which enables the versatile analysis for Kobayashi-Warren-Carter type systems, under various situations. To this end, we set the goal of this paper to specify the assumptions, which can cover the settings as in (g1)-(g2), and can guarantee the validity of the following Main Theorem.

    Main Theorem: the existence theorem of the solution [u,w,η,θ] to the systems (S)ν, for any ν0, which behaves in the range of C([0,T];L2(Ω)4), with the L2-based sources fL2(0,T;L2(Ω)) and fΓL2(0,T;L2(Γ)).

    The main theorem is somehow to enhance the results [25,31,32] concerned with qualitative properties of isothermal/non-isothermal Kobayashi-Warren-Carter type systems.


    2. Preliminaries

    First we elaborate the notations which is used throughout this paper.

    Notation 1 (Real analysis). For arbitrary a0, b0 ∈ [-∞, ∞], we define

    Fix dN as a constant of dimension. Then, we denote by |x| and xy the Euclidean norm of xRd and the standard scalar product of x,yRd, respectively, as usual, i.e.:

    |x|:=x21++x2d and xy:=x1y1++xdyd for all x=[x1,,xd], y=[y1,,yd]Rd.

    The d-dimensional Lebesgue measure is denoted by Ld, and unless otherwise specified, the measure theoretical phrases, such as "a.e.", "dt", "dx", and so on, are with respect to the Lebesgue measure in each corresponding dimension. Also, in the observations on a smooth surface SRd, the phrase "a.e." is with respect to the Hausdorff measure in each corresponding Hausdorff dimension, and the area element on S is denoted by dS.

    For a (Lebesgue) measurable function f:B[,] on a Borel subset BRd, we denote by [f]+ and [f], respectively, the positive and negative parts of f, i.e.,

    [f]+(x):=f(x)0  and [f](x):=(f(x)0), a.e. xB.

    Notation 2 (Abstract functional analysis). For an abstract Banach space X, we denote by ||X the norm of X, and when X is a Hilbert space, we denote by (,)X its inner product. For a subset A of a Banach space X, we denote by int(A) and ¯A the interior and the closure of A, respectively.

    Fix 1<dN. Then, for a Banach space X, the topology of the product Banach space Xd is endowed with the norm:

    |z|Xd:=dk=1|zk|X, for z=[z1,,zd]Xd.

    However, if X is a Hilbert space, then the topology of the product Hilbert space Xd is endowed with the inner product:

    (z,˜z)Xd:=dk=1(zk,˜zk)X, for z=[z1,,zd]Xd and ˜z=[˜z1,,˜zd]Xd

    and hence, the norm in this case is provided by

    |z|Xd:=(z,z)Xd=(dk=1|zk|2X)1/2,for z=[z1,,zd]Xd.

    For a Banach space X, we denote the dual space by X. For a single-valued operator A:XX, we write

    Az=[Az1,,Azd][X]d for any z=[z1,,zd]Xd.

    For any proper lower semi-continuous (l.s.c. hereafter) and convex function Ψ defined on a Hilbert space X, we denote by D(Ψ) its effective domain, and denote by Ψ its subdifferential. The subdifferential Ψ is a set-valued map corresponding to a weak differential of Ψ, and it has a maximal monotone graph in the product Hilbert space X2. More precisely, for each z0X, the value Ψ(z0) is defined as the set of all elements z0X that satisfy the variational inequality

    (z0,zz0)XΨ(z)Ψ(z0) for any zD(Ψ)

    and the set D(Ψ):={zXΨ(z)} is called the domain of Ψ. We often use the notation "[z0,z0]Ψ in X2, " to mean "z0Ψ(z0) in X with z0D(Ψ)" by identifying the operator Ψ with its graph in X2.

    Notation 3 (Basic elliptic operators). Let V=H1(Ω) be a Hilbert space endowed with the inner product:

    (w,z)V:=Ωwzdx+n0ΓwzdΓ,  for [w,z]V2,

    and let CV>0 be the embedding constant of VL2(Ω).

    Let , be the duality pairing between V and the dual space V, and let F: VV be the duality mapping defined by

    Fw,z:=(w,z)V,   for [w,z]V2.

    Note that V forms a Hilbert space endowed with the inner product:

    (w,z)V:=w,F1z,  for [w,z](V)2.

    For any ϱL2(Ω) and any ϱΓL2(Γ), we can regard the vectorial function ϱ:=[ϱ,ϱΓ]L2(Ω)×L2(Γ) as an element of V, via the following variational form:

    ϱ,z:=(ϱ,z)L2(Ω)+n0(ϱΓ,z)L2(Γ)forzV. (2.1)

    Note that for any ϱ=[ϱ,ϱΓ]L2(Ω)×L2(Γ), the variational form (2.1) enables the following identification:

    Fω=ϱinV,iff.ωH2(Ω)and{Δω=ϱinL2(Ω)DωnΓ+n0(ωϱΓ)=0inL2(Γ).

    On this basis, the product space L2(Ω)×L2(Γ) can be regarded as a subspace of V, and the restriction F|H2(Ω):H2(Ω)L2(Ω)×L2(Γ) can be regarded as a bijective linear operator associated with the Laplacian, subject to Robin type boundary condition (cf. [24]).

    In the meantime, we denote by ΔN the Laplacian operator subject to the zero-Neumann boundary condition, i.e.,

    ΔN:zWN:={zH2(Ω)DznΓ=0 in L2(Γ)}L2(Ω)ΔzL2(Ω).

    Remark 1. We here show some representative examples of the subdifferentials, which is intimately related to our study.

    (Ex.1) The quadratic functional uL2(Ω)12|u|2L2(Ω) can be regarded as a proper l.s.c. and convex function on V, via the standard -extension, and then, the V-subdifferential of this function coincides with the duality map F:VV, i.e.:

    [u,u][12||2L2(Ω)] in [V]2, iff. uV and u=Fu in V.

    (Ex.2) Let dN, and let γ0:RdR be a convex function defined as

    y=[y1,,yd]Rdγ0(y):=γ1(y1)+γ2(y2)++γd(yd),

    by using proper l.s.c. and convex functions γk:R(,], for k=1,,d. Let Ψdγ0:L2(Ω)d(,] be a proper l.s.c. and convex function defined as:

    zL2(Ω)dΨdγ0(z):={12Ω|Dz|2RN×ddx+Ωγ0(z)dx,ifzH1(Ω)d, otherwise.

    Then, with regard to the subdifferential Ψdγ0[L2(Ω)d]2, it is known (see, e.g., [4,6]) that

    zL2(Ω)dΨdγ0(z)={zL2(Ω)d|z+ΔNzγ0(z)inRd,a.e.inΩ}ifzWdN,,otherwise.

    This fact is often summarized as Ψdγ0=ΔN+γ0 in [L2(Ω)d]2.

    Notation 4 (BV theory; cf. [2,3,11,12]). Let dN, and let URd be an open set. We denote by M(U) the space of all finite Radon measures on U. The space M(U) is known as the dual space of the Banach space C0(U), i.e., M(U)=C0(U), where C0(U) is the closure of the class of test functions Cc(U) in the topology of C(¯U).

    A function zL1(U) is called a function of bounded variation on U, iff. its distributional gradient Dz is a finite Radon measure on U, namely, DzM(U)d. Here, for any zBV(U), the Radon measure Dz is called the variation measure of z, and its total variation |Dz| is called the total variation measure of z. Additionally, for any zBV(U), it holds that

    |Dz|(U)=sup{UzdivφdxφC1c(U)d and |φ|1 on U}.

    The space BV(U) is a Banach space, endowed with the norm

    |z|BV(U):=|z|L1(U)+|Dz|(U) for any zBV(U)

    and we say that znz weakly- in BV(U), iff. zBV(U), {zn}n=1BV(U), znz in L1(U) and DznDz weakly- in M(U)d, as n.

    The space BV(U) has another topology, called "strict topology", which is provided by the following distance (cf.[2, Definition 3.14]):

    [φ,ψ]BV(U)2|φψ|L1(U)+||Dφ|(U)|Dψ|(U)|.

    In this regard, we say that znz strictly in BV(U) iff. zBV(U), {zn}n=1BV(U), znz in L1(U) and |Dzn|(U)|Dz|(U), as n.

    Specifically, when the boundary U is Lipschitz, the Banach space BV(U) is continuously embedded into Ld/(d1)(U) and compactly embedded into Lp(U) for any 1p<d/(d1) (see, e.g., [2, Corollary 3.49] or [3, Theorems 10.1.3-10.1.4]). Furthermore, if 1q<, then the space C(¯U) is dense in BV(U)Lq(U) for the intermediate convergence, i.e., for any zBV(U)Lq(U), there exists a sequence {zn}n=1C(¯U) such that znz in Lq(U) and strictly in BV(U), as n (see, e.g., [3, Definition 10.1.3 and Theorem 10.1.2]).

    Notation 5 (Weighted total variation; cf. [1,2]). For any nonnegative ϱH1(Ω)L(Ω) (i.e. any 0ϱH1(Ω)L(Ω)) and any zL2(Ω), we call the value Varϱ(z)[0,], defined as,

    Varϱ(v):=sup[Ωvdivϖdx|ϖL(Ω)Nwithacompactsupport,and|ϖ|ϱa.e.inΩ][0,],

    "the total variation of v weighted by ϱ", or the "weighted total variation" in short.

    Remark 2. Referring to the general theories (e.g., [1,2,5]), we can confirm the following facts associated with the weighted total variations.

    (Fact 1)(Cf.[5, Theorem 5]) For any 0ϱH1(Ω)L(Ω), the functional zL2(Ω)Varϱ(z)[0,] is a proper l.s.c. and convex function that coincides with the lower semi-continuous envelope of

    zW1,1(Ω)L2(Ω)Ωϱ|Dz|dx[0,).

    (Fact 2) (Cf. [1, Theorem 4.3] and [2, Proposition 5.48]) If 0ϱH1(Ω)L(Ω) and zBV(Ω)L2(Ω), then there exists a Radon measure |Dz|ϱM(Ω) such that

    |Dz|ϱ(Ω)=Ωd|Dz|ϱ=Varϱ(z),

    and

    {|Dz|ϱ(A)|ϱ|L(Ω)|Dz|(A)|Dz|ϱ(A)=inf{liminfnAϱ|D˜zn|dx|{˜zn}n=1W1.1(A)L2(A)suchthat˜znzinL2(A)asn} (2.2)

    for any open set AΩ.

    (Fact 3) If ϱH1(Ω)L(Ω), cϱ:=essinfxΩϱ>0, and zBV(Ω)L2(Ω), then for any open set AΩ, it follows that

    {|Dz|ϱ(A)cϱ|Dz|(A)foranyopensetAΩD(Varϱ)=BV(Ω)L2(Ω),andVarϱ(z)=sup{Ωzdiv(ϱϖ)dx|ϖL(Ω)Nwithacompactsupport,and|ϖ|1a.e.inΩ} (2.3)

    Moreover, the following properties can be inferred from (2.2)-(2.3):

    · |Dz|c=c|Dz| in M(Ω) for any constant c0 and zBV(Ω)L2(Ω);

    · |Dz|ϱ=ϱ|Dz|LN in M(Ω), if 0ϱH1(Ω)L(Ω) and zW1,1(Ω)L2(Ω).

    Notation 6 (Generalized weighted total variation; cf. [25, Section 2]). For any linebreak ϱH1(Ω)L(Ω) and any zBV(Ω)L2(Ω), we define a real-valued Radon measure [ϱ|Dz|]M(Ω), as follows:

    [ϱ|Dz|](B):=|Dz|[ϱ]+(B)|Dz|[ϱ](B) for any Borel set BΩ.

    Note that [ϱ|Dz|](Ω) can be configured as a generalized total variation of zBV(Ω)L2(Ω) by the possibly sign-changing weight ϱH1(Ω)L(Ω).

    Remark 3. With regard to the generalized weighted total variations, the following facts are verified in [25, Section 2].

    (Fact 4) (Strict approximation) Let ϱH1(Ω)L(Ω) and zBV(Ω)L2(Ω) be arbitrary fixed functions, and let {zn}n=1C(¯Ω) be a sequence such that

    znz in L2(Ω) and strictly in BV(Ω) as n.

    Then

    Ωϱ|Dzn|dxΩd[ϱ|Dz|] as n.

    (Fact 5) For any zBV(Ω)L2(Ω), the mapping

    ϱH1(Ω)L(Ω)Ωd[ϱ|Dz|]R

    is a linear functional, and moreover, if φH1(Ω)C(¯Ω) and ϱH1(Ω)L(Ω), then

    Ωd[φϱ|Dz|]=Ωφd[ϱ|Dz|].

    Finally, we mention the notion of functional convergences.

    Definition 1 (Mosco convergence; cf. [27]). Let X be an abstract Hilbert space. Let Ψ:X(,] be a proper l.s.c. and convex function, and let {Ψn}n=1 be a sequence of proper l.s.c. and convex functions Ψn:X(,], n=1,2,3,. We say that ΨnΨ on X, in the sense of Mosco, as n, iff. the following two conditions are fulfilled.

    The condition of lower bound: lim infnΨn(zn)Ψ(z), if zX, {zn}n=1X, and znz weakly in X as n.

    The condition of optimality: for any zD(Ψ), there exists a sequence {zn}n=1X such that znz in X and Ψn(zn)Ψ(z) as n.

    Definition 2 (Γ-convergence; cf. [9]). Let X be an abstract Hilbert space, Ψ:X(,] be a proper functional, and {Ψn}n=1 be a sequence of proper functionals Ψn:X(,], n=1,2,3,. We say that ΨnΨ on X, in the sense of Γ-convergence, as n, iff. the following two conditions are fulfilled.

    The condition of lower bound: lim infnΨn(zn)Ψ(z), if zX, {zn}n=1X, and znz (strongly) in X as n.

    The condition of optimality: for any zD(Ψ), there exists a sequence {zn}n=1X such that znz in X and Ψn(zn)Ψ(z) as n.

    Remark 4. Note that if the functionals are convex, then Mosco convergence implies Γ-convergence, i.e., the Γ-convergence of convex functions can be regarded as a weak version of Mosco convergence. Additionally, in the Γ-convergence of convex functions, we can see the following:

    (Fact 6) Let Ψ:X(,] and Ψn:X(,] be proper l.s.c. and convex functions on a Hilbert space X such that ΨnΨ on X, in the sense of Γ-convergence, as n. If it holds that:

    {X2,[zn,zn]ΨninX2,n=1,2,3,,znzinXandznzweaklyinX,asn

    then [z,z]Ψ in X2 and Ψn(zn)Ψ(z) as n.


    3. Main Theorem and the demonstration scenario

    Throughout the paper, we set the following assumptions.

    (A1) Let fL2(0,T;L2(Ω)) and fΓL2(0,T;L2(Γ)) be given functions, and let f:=[f,fΓ]L2(0,T;L2(Ω)×L2(Γ)) be a time-dependent vectorial function which is regarded as fL2(0,T;V), via (2.1) applied to ϱ=f(t) for a.e. t>0.

    (A2) Let λW2,loc(R) be a function, and let A>0 be a constant which is defined as:

    A:=14(1+C2V|λ|2W2,(0,1)),

    by using the embedding constant CV>0 of VL2(Ω).

    (A3) Let α0W1,loc(R2) and α,βC2(R2) be functions, such that:

    · α and β are convex on R2;

    · δ:=inf[α0(R2)α(R2)β(R2)]>0;

    · αη(w,0)0, βη(w,0)0, αη(w,1)0, and βη(w,1)0, for any w[0,1].

    (A4) Let γ : R(,] be a proper l.s.c. and convex function, such that D(γ)=[0,1].

    (A5) Let gC2(R2) be a function such that

    gη(w,0)0 and gη(w,1)0, for any w[0,1]

    (A6) There exists a constant c such that

    γ(w)+g(v)c, for any v=[w,η]R2

    (A7) Let [u0,v0,θ0]=[u0,w0,η0,θ0] is a quartet of initial data, such that:

    [u0,w0,η0,θ0]{D0:={[˜u,˜w,˜η,˜θ]|˜uL2(Ω),˜w,˜ηH1(Ω),˜θBV(Ω)L(Ω),and0˜w,˜η1a.e.inΩ},ifν=0D1:=D0[L2(Ω)×H1(Ω)×H1(Ω)×H1(Ω)],ifν>0.

    Now, for simplicity of description, we prepare the following notations:

    {G(u;v)=G(u;w,η):=g(w,η)+uλ(w),[g](v)=[g](w,η):=[gw(w,η),gη(w,η)],[G](u;v)=[G](u;w,η):=[gw(w,η)+uλ(w),gη(w,η)],

    and

    {[α](v)=[α](w,η):=[αw(w,η),αη(w,η)],[β](v)=[β](w,η):=[βw(w,η),βη(w,η)], for uR and v=[w,η]R2.

    For any ν0 and any v=[w,η][H1(Ω)L(Ω)]2, we define a proper l.s.c. and convex function Φν(v;) on L2(Ω) by letting:

    θL2(Ω)Φν(v;θ)=Φν(w,η;θ):={Ωd[α(v)|Dθ|]+Ωβ(v)|D(νθ)|2dx,ifθBV(Ω)andνθH1(Ω),, otherwise.

    Additionally, we set:

    B:=1+A2,byusingtheconstantAasin(A2), (3.1)

    and define a functional Fν on L2(Ω)4 by letting:

    [u,v,θ]=[u,w,η,θ]L2(Ω)4F(u,v,θ)=F(u,w,η,θ):=B|u|2L2(Ω)+Ψ2γ(v)+Ω(g(v)c)dx+Φν(v;θ), (3.2)

    where Ψ2γ is the convex function Ψdγ0 in Remark 1 in the case when d=2 and γ0=γ. The above functional Fν is a modified version of the free-energy as in (1.5), and the assumptions (A3)-(A6) guarantee the non-negativity of this functional, i.e. Fν0 on L2(Ω)4.

    Based on these, we define the solutions to the systems (S)ν, for ν0, as follows.

    Definition 3. For any ν0, a quartet [u,v,θ]=[u,w,η,θ]L2(0,T;L2(Ω)4) with v=[w,η] is called a solution to (S)ν, iff. [u,v,θ] fulfills the following (S1)-(S6).

    (S1) uW1,2(0,T;V)L(0,T;L2(Ω))L2(0,T;V)C([0,T];L2(Ω)).

    (S2) v=[w,η]W1,2(0,T;L2(Ω)2)L(0,T;H1(Ω)2), and 0w(t,x)1 and 0η(t,x)1, a.e. (t,x)Q.

    (S3) θW1,2(0,T;L2(Ω))L(Q), |Dθ()|(Ω)L(0,T), νθL(0,T;H1(Ω)), and |θ||θ0|L(Ω) a.e. in Q.

    (S4) u satisfies the following variational form:

    [uλ(w)]t(t),z+(Du(t),Dz)L2(Ω)N+n0(u(t),z)L2(Γ)=(f(t),z)L2(Ω)+n0(fΓ(t),z)L2(Γ), for any zV, and a.e. t(0,T)

    with the initial condition u(0)=u0 in L2(Ω).

    (S5) v=[w,η] satisfies the following two variational forms:

    (wt(t)+gw(v)(t)+u(t)λ(w(t)),w(t)φ)L2(Ω)+(Dw(t),D(w(t)φ))L2(Ω)N+Ωd[(w(t)φ)αw(v(t))|Dθ(t)|]+Ω(w(t)φ)βw(v(t))|D(νθ)(t)|2dx+Ωγ(w(t))dxΩγ(φ)dx, for any φH1(Ω)L(Ω) and a.e. t(0,T)

    and

    (ηt(t)+gη(v)(t),ψ)L2(Ω)+(Dη(t),Dψ)L2(Ω)N+Ωd[ψαη(v(t))|Dθ(t)|]+Ωψβη(v(t))|D(νθ)(t)|2dx=0,for any ψH1(Ω)L(Ω) and a.e. t(0,T)

    with the initial condition v(0)=[w(0),η(0)]=v0=[w0,η0] in L2(Ω)2.

    (S6) θ satisfies the following variational form:

    (α0(v(t))θt(t),θ(t)ω)L2(Ω)+Φν(v(t);θ(t))Φν(v(t);ω),for any ωD(Φν(v(t);)) and a.e. t(0,T)

    with the initial condition θ(0)=θ0 in L2(Ω).

    Remark 5. The variational identity in the above (S4) can be reformulated as:

    [uλ(w)]t(t)+Fu(t)=f(t)inV,fora.e.t(0,T). (3.3)

    Also, two variational forms in (S5) can be reduced to:

    (vt(t)+[G](u;v(t)),v(t)ϖ)L2(Ω)2+(Dv(t),D(v(t)ϖ))L2(Ω)N×2+Ωd[|Dθ(t)|(v(t)ϖ)[α](v(t))]+Ω|D(νθ)(t)|2(v(t)ϖ)[β](v(t))dx+Ωγ(v(t))dxΩγ(ϖ)dx,foranyϖ=[φ,ψ][H1(Ω)L(Ω)]2anda.e.t(0,T), (3.4)

    by using the identification

    γ(˜v):=γ(˜w),   for all ˜v=[˜w,˜η]R2,

    and by using the abbreviation:

    Ωd[|D˜θ|ϖ˜v]:=Ωd[φ˜w|D˜θ|]+Ωd[ψ˜η|D˜θ|],for˜v=[˜w,˜η],ϖ=[φ,ψ][H1(Ω)L(Ω)]2and˜θBV(Ω)L2(Ω) (3.5)

    Furthermore, the variational form in (S6) is equivalent to the following evolution equation:

    α0(v(t))θt(t)+Φν(v(t);θ(t))0inL2(Ω),a.e.t(0,T), (3.6)

    governed by the subdifferential Φν(v(t);)L2(Ω)2 of the time-dependent convex function Φν(v(t);), for t(0,T).

    Now, our Main Theorem is stated as follows.

    Main Theorem Let ν0 be a fixed constant. Then, under (A1)-(A7), the system (S)ν admits at least one solution [u,v,θ]=[u,w,η,θ]L2(0,T;L2(Ω)4) with v=[w,η].

    Remark 6. Note that the presence of mobilities α0=α0(w,η), α=α(w,η) and β=β(w,η) makes the uniqueness problems for the systems (S)ν, ν0, be quite tough. In fact, even if we overview the kindred works to this study, we can find only two cases [15, Theorem 2.2] and [40, Theorem 2.2] that obtained the uniqueness results under some restricted situations.

    Finally, we devote the remaining part of this Section to show the sketch of the demonstration scenario, since the proof of the Main Theorem is going to be extended.

    In this paper, the Main Theorem will be obtained as a consequence of some approximating approaches, and then, the approximating problems will be associated with the time-discretization versions of (3.3)-(3.6), under positive setting of the constant ν. Hence, when we consider the approximating problems, we suppose ν>0, and fix the constant of time-step h(0,1]. Also, we denote by [f]ex0L2(R;L2(Ω)), [fΓ]ex0L2(R;L2(Γ)) and [f]ex0L2(R;V) the zero-extensions of f, fΓ and f (=[f,fΓ]), respectively.

    On this basis, the approximating problem for our system (S)ν is denoted by (AP)νh, and stated as follows.

    (AP)νh: to find a sequence {[uνi,vνi,θνi]}i=1D1 with {vνi}i=1={[wνi,ηνi]}i=1, which fulfills that

    uνiuνi1hλ(wνi)wνiwνi1h+Fuνi=[fi]hinV, (3.7)
    1h(vνivνi1,vνiϖ)L2(Ω)2+(Dvνi,D(vνiϖ))L2(Ω)N×2+([G](uνi;vνi),vνiϖ)L2(Ω)2+Ωγ(vνi)dx+Ω(vνiϖ)(|Dθνi1|[α](vνi)+ν2|Dθνi1|2[β](vνi))dx (3.8)
    Ωγ(ϖ)dx,foranyϖ[H1(Ω)L(Ω)]2,α0(vνi)θνiθνi1h+Φν(vνi;θνi)0inL2(Ω), (3.9)

    for i=1,2,3,, starting from the initial data:

    [uν0,vν0,θν0]D1 with vν0=[wν0,ην0].

    In the context, for any iN, [fi]h=[fhi,fhΓ,i]L2(Ω)×L2(Γ) (V), consists of the components:

    fhi:=1hih(i1)h[f]ex0(τ)dτ in L2(Ω) and fhΓ,i:=1hih(i1)h[fΓ]ex0(τ)dτ in L2(Γ).

    Hence, before the proof of Main Theorem, it will be needed to verify the following theorem.

    Theorem 1 (Solvability of the approximating problem). There exists a small constant h1(0,1] such that if ν>0 and h(0,h1], then the approximating problem (AP)νh admits a unique solution {[uνi,vνi,θνi]}i=1D1, and moreover,

    A2h|uνiuνi1|2V+12h|vνivνi1|2L2(Ω)2+1h|α0(vνi)(θνiθνi1)|2L2(Ω)+h2|uνi|2V+F(uνi,vνi,θνi)F(uνi1,vνi1,θνi1)+h|[fi]h|2V,fori=1,2,3, (3.10)

    where A is the constant as in (A2).

    However, due to the presence of L1-terms ν2|Dθi1|2[β](vνi)L1(Ω)2, i=1,2,3,, in (3.8), the above Theorem 1 will not be a straightforward consequence of standard variational method, and in fact, this theorem will be obtained via further approximating approach by means of some relaxed systems for (AP)νh.

    In the observation of the relaxed system, we first fix a large constant M>(N+2)/2, and fix a small constant ε(0,1] as the relaxation index. Besides, we define

    DM:=D1[L2(Ω)×H1(Ω)×H1(Ω)×HM(Ω)],

    and for any ˜vL2(Ω)2, we define a relaxed functional Φνε(˜v;) for Φν(˜v;), by letting:

    θL2(Ω)Φνε(˜v;θ):={Φν(˜v;θ)+ε22|θ|2HM(Ω), ifθHM(Ω),,otherwise.

    Note that for any ˜vL2(Ω)2, the functional Φνε(˜v;) is proper l.s.c. and convex on L2(Ω), such that:

    D(Φνε(˜v;))=HM(Ω)W1,(Ω),

    and hence, the L2-subdifferential Φνε(˜v;) is a maximal monotone graph in L2(Ω)2.

    On this basis, we denote by (RX)ε the relaxed system for (AP)νh, and prescribe the system (RX)ε as follows.

    (RX)ε:to find a sequence {[uνε,i,vνε,i,θνε,i]}i=1DM with {vνε,i}i=1={[wνε,i,ηνε,i]}i=1, which fulfills that

    uνε,iuνε,i1hλ(wνε,i)wνε,iwνε,i1h+Fuνε,i=[fi]hinV, (3.11)
    vνε,ivνε,i1hΔNvνε,i+γ(vνε,i)+[G](uνε,i;vνε,i)+|Dθνε,i1|[α](vνε,i)+ν2|Dθνε,i1|2[β](vνε,i)0inL2(Ω)2, (3.12)
    α0(vνε,i)θνε,iθνε,i1h+Φνε(vνε,i;θνε,i)0inL2(Ω), (3.13)

    for i=1,2,3,, starting from the initial data:

    [uνε,0,vνε,0,θνε,0]DM with vνε,0=[wνε,0,ηνε,0].

    Then, we can see that

    |Dθνε,i1|L(Ω)  and ν2|Dθνε,i1|2[β](vνε,i)L(Ω)2, i=1,2,3,.

    It implies that the general theories of L2-subdifferentials will be available for the relaxed system (RX)ε.

    Thus, it will be needed to verify the following proposition, as the first task to proving the Main Theorem.

    Proposition 1. There exists a small constant h0(0,1], such that if h(0,h0], then the system (RX)ε admits a unique solution {[uνε,i,vνε,i,θνε,i]}i=1DM with {vνε,i}i=1= {[wνε,i,ηε,i]ν}i=1.

    In view of these, we set the demonstration scenario of the Main Theorem, by assigning the proofs of Proposition 1, Theorem 1 and Main Theorem to Sections 4, 5 and 6, respectively.


    4. Proof of Proposition 1

    Before we start the proof, we need to show some lemmas.

    Lemma 1. Let us put Δ:=[0,1]×[1,2]R2, and let us assume

    0<hh2:=12(1+|g|C2(Δ)+5|λ|2W2,(0,1)). (4.1)

    Let us fix f0V, [u0,η0,w0,θ0]L2(Ω)×H1(Ω)×H1(Ω)×W1,(Ω) and wH1(Ω), and let us assume that 0w0,w1 a.e. in Ω. Then, the following auxiliary system:

    uu0hλ(w)ww0h+Fu=f0inV, (4.2)
    ww0hΔNw+γ(w)+gw(w,η)+αw(w,η)|Dθ0|+ν2βw(w,η)|Dθ0|2λ(w)uinL2(Ω), (4.3)
    ηη0hΔNη+I[1,2](η)+gη(w,η)+αη(w,η)|Dθ0|+ν2βη(w,η)|Dθ0|2=0ΔinL2(Ω), (4.4)

    admits a unique solution [u,w,η]V×H1(Ω)2, where I[1.2] is the subdifferential of the indicator function I[1.2]:R{0,} on the compact interval [1,2], and this is an additional term to guarantee the boundedness of the range η(Ω) for the component η.

    Proof. First, we put:

    e:=uλ(w)w, e0:=u0λ(w)w0,and v0=[w0,η0],[˜w,˜η]Rγ(˜w,˜η):=γ(˜w)+I[1,2](˜η),

    and reformulate the system {(4.2)-(4.4)} as follows:

    ee0h+F(e+λ(w)w)=f0inV, (4.5)
    vv0h+Ψ2γ(v)+[g](w,η)+|Dθ0|[α](v)+ν2|Dθ0|2[β](v)[λ(w)(e+λ(w)w)0]inL2(Ω)2, (4.6)

    where Ψ2γ is the functional Ψdγ0 as in Remark 1 (Ex.2), in the case when d=2 and γ0=γ on R2, and Ψ2γ is the subdifferential of Ψ2γ in L2(Ω)2. Then, in the light of Remark 1, we can associate the auxiliary system {(4.2)-(4.4)} with a minimization problem for the following functional:

    [e,v]=[e,w,η]V×L2(Ω)2Ψ0(w;e,v)=Ψ0(w;e,w,η):={12h|ee0|2V+12h|vv0|2L2(Ω)2+12|e+λ(w)w|2L2(Ω)+Ψ2γ(v)+Ω(α(v)|Dθ0|+ν2β(v)|Dθ0|2)dx+Ωg(v)dx(f0,e)V,if[e,v]=[e,w,η]L2(Ω)×D(Ψ2γ),,otherwise, (4.7)

    via its stationary system {(4.5)-(4.6)}. Then, taking into account (A2)-(A6), (4.1) and (4.7), we find a positive constant C0, independent of the variables [e,v]=[e,w,η] and w, such that:

    Ψ0(w;e,v)C0(|e|2L2(Ω)+|v|2H1(Ω)21),forall[e,v]D(Ψ0(w;)). (4.8)

    Now, the above coercivity enables us to apply the standard minimization argument to Ψ0 (cf. [3,10]), and to obtain the solution [u,w,η]=[e+λ(w)w,w,η] to {(4.2)-(4.4)}, via the minimizer [e,v]=[e,w,η]V×H1(Ω)2 of Ψ0(w;), with v=[w,η]D(Ψ2γ).

    In the meantime, the uniqueness can be seen by using the standard method, i.e. by taking the difference of two solutions [ek,vk]=[ek,wk,ηk]V×L2(Ω)2 with vk=[wk,ηk]D(Ψ2γ), k=1,2, to the stationary system {(4.5)-(4.6)}. In fact, multiplying the both sides of the subtraction of (4.5) by e1e2 in V, multiplying the both sides of the subtraction of (4.6) by v1v2 in L2(Ω)2, and using (A2)-(A5), (4.8) and Schwartz's inequality, we have:

    1h|e1e2|2V+1h(1h|[g]|W1,(Δ)2)|v1v2|2L2(Ω)2+|D(v1v2)|2L2(Ω)N×2+|(e1e2)+λ(w)(w1w2)|2L2(Ω)0. (4.9)

    Since the assumption (4.1) implies (1h, we can deduce from (4.9) the uniqueness for the system {(4.2)-(4.4)}.

    Lemma 2. Let wH1(Ω) be the function as in Lemma 1, and let Ψ0(w;) be the functional on V×L2(Ω)2 given in (4.7). Also, let us take a sequence {wn}n=1H1(Ω) such that 0wn1 a.e. in Ω, for n=1,2,3,, and let us define a sequence {Ψ0(wn;)}n=1 of functionals on V×L2(Ω)2, by putting w=wn in (4.7), for n=1,2,3,. Besides, let us assume that:

    wnwinthepointwisesensea.e.inΩ,asn. (4.10)

    Then, Ψ0(wn;)Ψ0(w;) on V×L2(Ω)2, in the sense of Γ-convergence, as n.

    Proof. The condition of lower-bound will be seen, immediately, from the lower semi-continuity of the following functional (of 4-variables):

    [w,e,v]L2(Ω)×V×L2(Ω)2Ψ0(w;e,v)(,].

    The condition of optimality will be verified by taking the singleton {[e,v]} for any [e,v]D(Ψ0(w;)) =D(Ψ0(wn;)) for all n1 as the sequence corresponding to {zn}n=1 in Definition 2.

    Lemma 3. Under the assumptions as in the previous Lemmas 1-2, let us take the solution [e,v]=[e,w,η]V×H1(Ω)2 to the stationary system {(4.5)-(4.6)} with v=[w,η], and for any nN, let us denote by [en,vn]=[en,wn,ηn]V×H1(Ω)2 the solution to {(4.5)-(4.6)} with vn=[wn,ηn], when w=wn. Then, the assumption (4.10) implies that:

    [en,vn]=[en,wn,ηn][e,v]=[e,w,η]inV×L2(Ω)2,andweaklyinL2(Ω)×H1(Ω)2,asn (4.11)

    Proof. In the light of Lemma 1 (including the proof), we can see that:

    Ψ0(˜w;˜e,˜v)=Ψ0(˜w;˜e,˜w,˜η)Ψ0(˜w;0,0,0)C1:=12h(|e0|2V+|v0|2L2(Ω)2)+LN(Ω)(|γ(0)|+|g(0,0)|)+α(0,0)|θ0|W1,1(Ω)+ν2β(0,0)|θ0|2H1(Ω), (4.12)

    for any ˜wH1(Ω) with 0˜w1 a.e. in Ω, and

    solution [˜e,˜v]=[˜e,˜w,˜η] to {(4.5)-(4.6)} with ˜v=[˜w,˜η] when w=˜w.

    Since the constant C1 is independent of the choice of ˜w, the convergence (4.11) will be observed by taking into account (4.8), (4.12) and the uniqueness for {(4.5)-(4.6)}, and by applying Lemma (cf. [3,11]), and the general theories of the compact embeddings (cf. [3,11]) and the Γ convergence (cf. [9]).

    Lemma 4. Let h2 be the constant as in (4.1). Let f0V, u0L2(Ω), v0=[w0,η0]H1(Ω)2 and θ0W1,(Ω) be the functions as in Lemma 1. Then, for any h(0,h2], the following system:

    uu0hλ(w)ww0h+Fu=f0inV, (4.13)
    vv0hΔNv+[γ(w)0]+[g](v)+|Dθ0|[α](v)+ν2|Dθ0|2[β](v)[λ(w)u0]inL2(Ω)2, (4.14)

    admits at least one solution [u,v]=[u,w,η]V×H1(Ω)2 with v=[w,η].

    Proof. Let us set a compact set K1 in L2(Ω), by letting:

    K1:={˜wH1(Ω)|0˜w1 a.e. in Ω,and12h|˜ww0|2L2(Ω)+12|D˜w|2L2(Ω)NC1+|c|LN(Ω)+12h|e0|2V+h|f0|2V},

    and let us consider an operator P1:K1L2(Ω), which maps any wK1 to the component w of the solution [u,w,η]V×H1(Ω)×H1(Ω) to {(4.2)-(4.4)}. Then, on account of (A3), (A6), Lemma 3, (4.7) and (4.12), it will be seen that P1K1K1 and P1 is a continuous operator in the topology of L2(Ω). So, applying Schauder's fixed point theorem, we find a fixed point wK1 for P1, i.e. w=P1w in L2(Ω).

    Now, let us denote by [u,w,η]V×H1(Ω)×H1(Ω) the solution to {(4.2)-(4.4)}, involved in the fixed point w. Then, this triplet [u,w,η] must be a special solution to {(4.2)-(4.4)} such that w=w. Hence, our remaining task will be to show that

    0η1a.e.inΩ, (4.15)

    namely, the subdifferential I[1,2] in (4.4) will not affect for η.To this end, we check two inequalities:

    0η0h+gη(w,0)+|Dθ0|αη(w,0)+ν2|Dθ0|2βη(w,0)0inL2(Ω), (4.16)
    1η0h+gη(w,1)+|Dθ0|αη(w,1)+ν2|Dθ0|2βη(w,1)0inL2(Ω), (4.17)

    with use of the assumptions (A3), (A5) and 0η01 a.e. in Ω.

    On this basis, let us take the difference from (4.16) to (4.4) when η=η and w=w=w (resp. from (4.4) to (4.17) when η=η and w=w=w), and multiply the both sides by [η]+ (resp. by [η1]+). Then, with the assumptions (A3), (A5) and I[1,2](0)={0} (resp. I[1,2](1)={0}) in mind, it is inferred that:

    1h(1h|gηη|C(Δ))|[η]+|2L2(Ω)+|D[η]+|2L2(Ω)N0(resp.1h(1h|gηη|C(Δ))|[η1]+|2L2(Ω)+|D[η1]+|2L2(Ω)N0). (4.18)

    Since the assumption (4.1) implies 1h|gηη|C(Δ)12, we can deduce (4.15) from (4.18), and conclude that the triplet [u,v]=[u,η,w] with v:=[w,η] solves the system {(4.13)-(4.14)}.

    Lemma 5. Let f0V and θ0HM(Ω) be fixed functions, and let [u,v]=[u,w,η]V×H1(Ω)2 be a solution to the system {(4.13)-(4.14)} with v=[w,η]. Then, the inclusion

    α0(v)θθ0h+Φνε(v;θ)0inL2(Ω) (4.19)

    admits a unique solution θHM(Ω).

    Proof. We omit the proof, because this lemma is obtained, immediately, just as a direct consequence of [31, Lemma 3.4].

    Lemma 6. Under the assumption (4.1), let us take a quartet [u,v,θ]=[u,w,η,θ]DM with v=[w,η]H1(Ω)2, which solves the coupled system {(4.13)-(4.14), (4.19)}. Then, the following energy-inequality holds:

    A2h|uu0|2V+12h|vv0|2L2(Ω)2+1h|α0(v)(θθ0)|2L2(Ω)+h2|u|2V+Fνε(u,v,θ)Fνε(u0,v0,θ0)+h|f0|2V, (4.20)

    where A>0 is the constant as in (A2), and Fνε is the relaxed version of the functional Fν, defined as:

    [u,v,θ]=[u,w,η,θ]L2(Ω)4Fνε(u,v,θ)=Fνε(u,w,η,θ)=B|u|2L2(Ω)+Ψ2γ(v)+Ω(g(v)c)dx+Φνε(v;θ), (4.21)

    with the constant B=(1+A)/2 as in (3.1).

    Proof. First, let us rewrite the equation (4.13) as follows:

    (uu0,z)L2(Ω)+hFu,z=hf0,z+(λ(w)(ww0),z)L2(Ω),foranyzV, (4.22)

    and let us put z=u. Then, by using Schwarz's inequality, we have:

    12|u|2L2(Ω)+h2|u|2V12|u0|2L2(Ω)+h2|f0|2V+(λ(w)(ww0),u)L2(Ω). (4.23)

    Alternatively, if we rewrite the equation (4.13) to:

    1h(uu0,z)V+z,u=(f0,z)V+1h(λ(w)(ww0),z)V,

    for any zV,

    and put z=A(uu0)V, then we also see that:

    A2h|uu0|2V+A2|u|2L2(Ω)A2|u0|2L2(Ω)+Ah|f0|2V+14h|ww0|2L2(Ω). (4.24)

    Next, let us multiply the both sides of the inclusion (4.14) by vv0. Then, in the light of (A2)-(A5) and Taylor's theorem, we infer that:

    1h(1h2|g|C2([0,1]2))|vv0|2L2(Ω)2+12|Dv|2L2(Ω)N×2+Ωγ(w)dx+Ωg(v)dx+Ωα(v)|Dθ0|dx+ν2Ωβ(v)|Dθ0|2dx12|Dv0|2L2(Ω)N×2+Ωγ(w0)dx+Ωg(v0)dx.+Ωα(v0)|Dθ0|dx+ν2Ωβ(v0)|Dθ0|2dx(λ(w)(ww0),u)L2(Ω). (4.25)

    Furthermore, applying the both sides of (4.19) by θθ0, it follows that:

    1h|α0(v)(θθ0)|2L2(Ω)+Φνε(v;θ)Φνε(v;θ0). (4.26)

    Now, since (4.1) implies 1h2|g|C2([0,1]2)234, the energy-inequality (4.20) can be obtained by taking the sum of (4.23)-(4.26) with (A2) in mind.

    Lemma 7. By the restriction 1N3 of the spatial dimension, there exists a positive constant C2, such that under the notations and assumptions as in Lemma 6, the condition:

    C2h13(1+2(Fνε(u0,v0,θ0)+h|f0|2V)23)12,and0<hh2, (4.27)

    implies the uniqueness of the solution [u,v,θ]=[u,w,η,θ]DM to the system {(4.13)-(4.14), (4.19)} with v=[w,η].

    Proof. In the light of the uniqueness of θ as in Lemma 5, it is enough to focus only on the uniqueness for the component [u,v]=[u,w,η]V×H1(Ω)2 with v=[w,η]. To this end, we take two triplets [uk,vk]=[uk,wk,ηk]DM with vk=[wk,ηk], k=1,2, that fulfill (4.13)-(4.14).

    First, with the equivalence of (4.13) and (4.22) in mind, we take the difference between two variational forms (4.22) for uk, k=1,2, and put z=u1u2 in V. Then:

    |u1u2|2L2(Ω)+h|u1u2|2V=(λ(w1)w1λ(w2)w2,u1u2)L2(Ω)((λ(w1)λ(w2))w0,u1u2)L2(Ω),

    so that by using (A2) and Schwarz's inequality, we have:

    14|u1u2|2L2(Ω)+h|u1u2|2V3|λ|2W1,(0,1)|w1w2|2L2(Ω). (4.28)

    Secondly, let us take the difference between two inclusions (4.14) for vk=[wk,ηk], k=1,2, and multiply the both sides by v1v2 in L2(Ω)2. Then, by using (A2)-(A5) and Schwarz's inequality, it is computed that:

    1h(1h|[g]|C1([0,1]2)2)|v1v2|2L2(Ω)2+|D(v1v2)|2L2(Ω)N×2(λ(w1)u1λ(w2)u2,w1w2)L2(Ω)|λ|L(0,1)|u1u2|L2(Ω)|w1w2|L2(Ω)+(u1(λ(w1)λ(w2)),w1w2)L2(Ω)18|u1u2|2L2(Ω)+2|λ|2L(0,1)|w1w2|2L2(Ω)+|λ (4.29)

    Here, the dimensional restriction 1 \leq N \leq 3 and the assumption (4.27) enable to apply the analytic technique as in [19, Lemma 3.1], and to find a constant C_2^\circ > 0 , independent of h and triplets [u_0^\circ, v_0^\circ] and [u_k, v_k] , k=1, 2 , such that:

    |\lambda ''{|_{{L^\infty }(0,1)}}\int_\Omega | {u_1}||{w_1} - {w_2}{|^2}dx \le \frac{1}{2}|D({w_1} - {w_2})|_{{L^2}(\Omega )}^2 + C_2^ \circ (1 + |{u_1}|_V^{\frac{4}{3}})|{w_1} - {w_2}|_{{L^2}(\Omega )}^2. (4.30)

    Furthermore, under (4.27), the inequality (4.20) enables to derive that:

    \begin{array}{l} C_2^ \circ (1 + |{u_1}|_V^{\frac{4}{3}})|{w_1} - {w_2}|_{{L^2}(\Omega )}^2 = C_2^ \circ {h^{\frac{1}{3}}}({h^{\frac{2}{3}}} + {(h|{u_1}|_V^2)^{\frac{2}{3}}}) \cdot \frac{1}{h}|{w_1} - {w_2}|_{{L^2}(\Omega )}^2\\ \;\;\;\; \le C_2^ \circ {h^{\frac{1}{3}}}(1 + 2{({\mathscr{F}}_\varepsilon ^\nu (u_0^ \circ ,\mathit{\boldsymbol{v}}_0^ \circ ,\theta _0^ \circ ) + h|f_0^*|_{{V^*}}^2)^{\frac{2}{3}}}) \cdot \frac{1}{h}|{w_1} - {w_2}|_{{L^2}(\Omega )}^2\\ \;\;\;\; \le \frac{1}{{2h}}|{w_1} - {w_2}|_{{L^2}(\Omega )}^2. \end{array} (4.31)

    Now, taking sum of (4.28)-(4.29) with (4.30)-(4.31) in mind, we obtain that:

    \begin{array}{l} \frac{1}{8}|{u_1} - {u_2}|_{{L^2}(\Omega )}^2 + h|{u_1} - {u_2}|_V^2\\ + \frac{1}{h}\left( {\frac{1}{2} - h\left( {|g{|_{{C^2}({{[0,1]}^2})}} + 5|\lambda |_{{W^{2,\infty }}(0,1)}^2} \right)} \right)|{\mathit{\boldsymbol{v}}_1} - {\mathit{\boldsymbol{v}}_2}|_{{L^2}{{(\Omega )}^2}}^2\\ + \frac{1}{2}|D({\mathit{\boldsymbol{v}}_1} - {\mathit{\boldsymbol{v}}_2})|_{{L^2}{{(\Omega )}^{N \times 2}}}^2 \le 0. \end{array} (4.32)

    This implies the required uniqueness, because \frac{1}{2} -h (|g|_{C^2([0, 1]^2)^2} +5|\lambda|_{W^{2, \infty}(0, 1)}^2) > 0 follows from the assumption (4.1) and (4.27).

    Proof of Proposition 1. Let us take a positive constant h_{2}^{\circ} defined by (4.1). Let us set a positive constant h_0^\circ , so small to satisfy that:

    C_2^\circ (h_0^\circ)^{\frac{1}{3}} \bigl( 1 +2(F_\varepsilon^\nu(u_0^\circ, v_0^\circ, \theta_0^\circ) +h_0^\circ |[f^*]_0^{\rm ex}|_{L^2(0, T; V^*)}^2 )^{\frac{2}{3}} \bigr) \leq \frac{1}{2}, \mbox{ and } 0 < h_0^\circ \leq h_2^\circ.

    Then, from (4.20), it will be observed that:

    \begin{array}{l} C_2^ \circ {h^{\frac{1}{3}}}(1 + 2{({\mathscr{F}}_\varepsilon ^\nu (u_{\varepsilon ,i - 1}^\nu ,\mathit{\boldsymbol{v}}_{\varepsilon ,i - 1}^\nu ,\theta _{\varepsilon ,i - 1}^\nu ) + h|{[f_i^*]^h}|_{{V^*}}^2)^{\frac{2}{3}}})\\ \;\;\;\; \le C_2^ \circ {h^{\frac{1}{3}}}(1 + 2{({\mathscr{F}}_\varepsilon ^\nu (u_{\varepsilon ,i - 2}^\nu ,{\rm{ }}\mathit{\boldsymbol{v}}_{\varepsilon ,i - 2}^\nu ,\theta _{\varepsilon ,i - 2}^\nu ) + h(|{[f_i^*]^h}|_{{V^*}}^2 + |{[f_{i - 1}^*]^h}|_{{V^*}}^2))^{\frac{2}{3}}})\\ \;\;\;\; \le \cdots \le C_2^ \circ {h^{\frac{1}{3}}}(1 + 2{({\mathscr{F}}_\varepsilon ^\nu (u_{\varepsilon ,0}^\nu ,{\rm{ }}\mathit{\boldsymbol{v}}_{\varepsilon ,0}^\nu ,\theta _{\varepsilon ,0}^\nu ) + |[{f^*}]_0^{{\rm{ex}}}|_{{L^2}(0,T;{V^*})}^2)^{\frac{2}{3}}})\\ \;\;\;\; \le \frac{1}{2},{\rm{ }}\;{\rm{for}}\;{\rm{all}}\;0 < h \le h_2^ \circ \;( \le h_2^ \circ )\;{\rm{and}}\;i = 1,2,3, \ldots . \end{array} (4.33)

    In view of this, the Proposition 1 will be concluded by means of the following algorithm.

    (Step 0) Let h \in (0, h_0^\circ] , let i=1, and let [u_{\varepsilon, 0}^{\nu}, v_{\varepsilon, 0}^{\nu}, \theta_{\varepsilon, 0}^{\nu}] \in D_{M}.

    (Step 1) Obtain the quartet [u_{\varepsilon, i}^\nu, v_{\varepsilon, i}^\nu, \theta_{\varepsilon, i}^\nu] \in D_M as the unique solution to the system {(4.13)-(4.14), (4.19)}, by applying Lemmas 4-7 to the case when:

    \begin{align} & f_{0}^{*}={{[f_{i-1}^{*}]}^{h}}in\ {{V}^{*}}, u_{0}^{{}^\circ }=u_{\varepsilon, i-1}^{\nu }\ in\ {{L}^{2}}(\Omega ), \\ & v_{0}^{{}^\circ }=\text{ }v_{\varepsilon, i-1}^{\nu }\text{ }in\ {{H}^{1}}{{(\Omega )}^{2}}\ and\ \theta _{0}^{{}^\circ }=\theta _{\varepsilon, i-1}^{\nu }\ in\ {{H}^{M}}(\Omega ). \\ \end{align}

    (Step 2) Iterate the value of i and return to (Step 1).

    Note that (4.33) let the assumption h \in (0, h_0^\circ] be a uniform condition to make sense (Step 1), for all i=1, 2, 3, \dots


    5. Proof of Theorem 1

    Let us set h_1^\circ :=h_0^\circ i.e. the constant as in Proposition 1, and let us fix \nu > 0 , h \in (0, h_{1}^{\circ}] and the initial value [u_{0}^{\nu}, v_{0}^{\nu}, \theta_{0}^{\nu}]=[u_{0}^{\nu}, w_{0}^{\nu}, \eta_{0}^{\nu}, \theta_{0}^{\nu}] \in D_{1} with v_{0}^{\nu}=[w_{0}^{\nu}, \eta_{0}^{\nu}] . Besides, we recall the following lemmas obtained in [31].

    Lemma 8. (cf. [31, Lemma 4.1]) Assume v^\circ \in [H^1(\Omega) \cap L^\infty (\Omega)]^2 , \{ v_\varepsilon^\circ \}_{0 < \varepsilon \leq 1} \subset [H^1(\Omega) \cap L^\infty (\Omega)]^2 , and

    \left\{ \begin{align} & v_{\varepsilon }^{{}^\circ }\to {{v}^{{}^\circ }}in\ the\ pointwise\ sense\ a.e.\ in\ \Omega \ \varepsilon \downarrow 0, \\ & {{\{\text{ }v_{\varepsilon }^{{}^\circ }\}}_{0<\varepsilon \le 1}}\ is\ bounded\text{ }in\ {{L}^{\infty }}{{(\Omega )}^{2}}. \\ \end{align} \right.

    Then, for the sequence of convex functions \{ \Phi_{\varepsilon}^{\nu}(v_\varepsilon^\circ; \cdot\, ) \}_{0 < \varepsilon \leq 1} , it holds that \Phi_{\varepsilon}^{\nu}(v_\varepsilon^\circ; {}\cdot\, ) \to \Phi_\nu (v^\circ; \cdot\, ) on L^2(\Omega) , in the sense of Mosco, as \varepsilon \downarrow 0 .

    Lemma 9. (cf. [31, Lemma 4.2]) Assume that

    \left\{ \begin{align} & {{v}^{{}^\circ }}\in {{[{{H}^{1}}(\Omega )\cap {{L}^{\infty }}(\Omega )]}^{2}}, {{\{v_{\varepsilon }^{{}^\circ }\}}_{0<\varepsilon \le 1}}\subset {{[{{H}^{1}}(\Omega )\cap {{L}^{\infty }}(\Omega )]}^{2}}, \\ & {{\{v_{\varepsilon }^{{}^\circ }\}}_{0<\varepsilon \le 1}}\ is\text{ }bounded\text{ }in\ {{L}^{\infty }}{{(\Omega )}^{2}}, \\ & v_{\varepsilon }^{{}^\circ }\to {{v}^{{}^\circ }}\ in\text{ }the\text{ }po\operatorname{int}wise\text{ }sense, \text{ }a.e.\ in\text{ }\Omega \ as\ \varepsilon \downarrow 0, \\ \end{align} \right.

    and

    \left\{ \begin{align} & {{\theta }^{{}^\circ }}\in {{H}^{1}}(\Omega ), {{\{\theta _{\varepsilon }^{{}^\circ }\}}_{0<\varepsilon \le 1}}\subset {{H}^{1}}(\Omega ), \\ & \theta _{\varepsilon }^{{}^\circ }\to {{\theta }^{{}^\circ }}\ in\ {{L}^{2}}(\Omega )\ and\ \Phi _{\varepsilon }^{\nu }(v_{\varepsilon }^{{}^\circ };\theta _{\varepsilon }^{{}^\circ })\to {{\Phi }_{\nu }}({{v}^{{}^\circ }};{{\theta }^{{}^\circ }}), as\ \varepsilon \downarrow 0 \\ \end{align} \right.

    Then, \theta_\varepsilon^\circ \to \theta^\circ in H^1(\Omega) and {\varepsilon} \theta_\varepsilon^\circ \to 0 in H^M (\Omega) , as \varepsilon \downarrow 0 .

    Lemma 10. (cf. [31, Lemma 4.4]) Let v^{\circ} \in [H^{1}(\Omega) \cap L^{\infty}(\Omega)]^{2} and \check{\theta}_{0}^{\circ}, \hat{\theta}_{0}^{\circ} \in H^{1}(\Omega) be fixed functions, and let [\check{\theta}, \check{\theta}^{\ast}], [\hat{\theta}, \hat{\theta}^{\ast}] \in L^{2}(\Omega)^{2} be pairs of functions such that

    \left\{ \begin{array}{l} [\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\smile$}} \over \theta } ,{{\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\smile$}} \over \theta } }^ * }] \in \partial {\Phi _\nu }({\mathit{\boldsymbol{v}}^ \circ }; \cdot ){\rm{ in }}{L^2}{(\Omega )^2}{\rm{ and }}\frac{1}{h}{\alpha _0}({\mathit{\boldsymbol{v}}^ \circ })(\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\smile$}} \over \theta } - \mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\smile$}} \over \theta } _0^ \circ ) + {{\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\smile$}} \over \theta } }^ * } \le 0{\rm{ a}}{\rm{.e}}{\rm{. in }}\Omega ,\\ [\hat \theta ,{{\hat \theta }^ * }] \in \partial {\Phi _\nu }({\mathit{\boldsymbol{v}}^ \circ }; \cdot ){\rm{ in }}{L^2}{(\Omega )^2}{\rm{ and }}\frac{1}{h}{\alpha _0}({\mathit{\boldsymbol{v}}^ \circ })(\hat \theta - \hat \theta _0^ \circ ) + {{\hat \theta }^ * } \ge 0{\rm{ a}}{\rm{.e}}{\rm{. in }}\Omega , \end{array} \right. (5.1)

    respectively. Then, it follows that

    |\sqrt{\alpha_{0}(v^{\circ})}[\check{\theta}-\hat{\theta}]^{+}|_{L^{2}(\Omega)}^{2} \le |\sqrt{\alpha_{0}(v^{\circ})}[\check{\theta}_{0}^{\circ}-\hat{\theta}_{0}^{\circ}]^{+}|_{L^{2}(\Omega)}^{2},

    and therefore, if \check{\theta}_{0}^{\circ} \le \hat{\theta}_{0}^{\circ} in \Omega, then the inequality \check{\theta} \leq \hat{\theta} a.e. in \Omega also follows from (A3).

    Moreover, if the both inequalities in (5.1) hold as equalities, then:

    |\sqrt{\alpha_{0}(v^{\circ})}(\check{\theta}-\hat{\theta})|_{L^{2}(\Omega)}^{2} \le |\sqrt{\alpha_{0}(v^{\circ})}(\check{\theta}_{0}^{\circ}-\hat{\theta}_{0}^{\circ})|_{L^{2}(\Omega)}^{2},

    Based on these, we divide the proof of Theorem 1 in two parts: the part of existence; the part of uniqueness and energy inequality.

    The part of existence. Let \nu > 0 be a fixed constant. By Lemma 8, there exists a sequence \{\tilde{\theta}_{\varepsilon, 0}^{\nu} \}_{0 < \varepsilon \leq 1} \subset H^{M}(\Omega) such that

    \tilde{\theta}_{\varepsilon, 0}^{\nu} \to \theta_{0}^{\nu} \mbox{ in } H^{1}(\Omega) \mbox{ and } \Phi_{\varepsilon}^{\nu}(v_{0}^{\nu} ; \tilde{\theta}_{\varepsilon, 0}^{\nu}) \to \Phi_{\nu}(v_{0}^{\nu}; \theta_{0}^{\nu}) \mbox{ as } \varepsilon \downarrow 0.

    So, by virtue of Proposition 1 we can take a class \{ [\tilde{u}_{\varepsilon, i}^\nu, \tilde{\mathit{\pmb{v}}}_{\varepsilon, i}^\nu, \tilde{\theta}_{\varepsilon, i}^\nu] \, | \, i \in \mathbb{N}, ~\varepsilon \in (0, 1] \} consisting of solutions \{ [\tilde{u}_{\varepsilon, i}^\nu, \tilde{\mathit{\pmb{v}}}_{\varepsilon, i}^\nu, \tilde{\theta}_{\varepsilon, i}^\nu] \}_{i=1}^\infty=\{ [\tilde{u}_{\varepsilon, i}^\nu, \tilde{w}_{\varepsilon, i}^\nu, \tilde{\eta}_{\varepsilon, i}^\nu, \tilde{\theta}_{\varepsilon, i}^\nu] \}_{i=1}^\infty \subset {D_M} to (RX)_\varepsilon with \{ \tilde{\mathit{\pmb{v}}}_{\varepsilon, i}^\nu \}_{i=1}^\infty=\{ [\tilde{w}_{\varepsilon, i}^\nu, \tilde{\eta}_{\varepsilon, i}^\nu] \}_{i=1}^{\infty} , starting from the initial data [u_{\varepsilon, 0}^{\nu}, \mathit{\pmb{v}}_{\varepsilon, 0}^{\nu}, \theta_{\varepsilon, 0}^{\nu}]=[u_{0}^{\nu}, \mathit{\pmb{v}}_{0}^{\nu}, \tilde{\theta}_{\varepsilon, 0}^{\nu}] for 0 < \varepsilon \leq 1 . Then, with Lemma 6 and the algorithm (Step 0)-(Step 2) in mind, we remark the following energy-inequality:

    \begin{array}{l} \frac{{{A_*}}}{{2h}}|\tilde u_{\varepsilon ,i}^\nu - \tilde u_{\varepsilon ,i - 1}^\nu |_{{V^*}}^2 + \frac{1}{{2h}}|\mathit{\boldsymbol{\widetilde v}}_{\varepsilon ,i}^\nu - \mathit{\boldsymbol{\widetilde v}}_{\varepsilon ,i - 1}^\nu |_{{L^2}{{(\Omega )}^2}}^2 + \frac{1}{h}|\sqrt {{\alpha _0}(\mathit{\boldsymbol{\widetilde v}}_{\varepsilon ,i}^\nu )} (\tilde \theta _{\varepsilon ,i}^\nu - \tilde \theta _{\varepsilon ,i - 1}^\nu )|_{{L^2}(\Omega )}^2\\ + \frac{h}{2}|\tilde u_{\varepsilon ,i}^\nu |_V^2 + {\mathscr{F}}_\varepsilon ^\nu (\tilde u_{\varepsilon ,i}^\nu ,\mathit{\boldsymbol{\widetilde v}}_{\varepsilon ,i}^\nu ,\tilde \theta _{\varepsilon ,i}^\nu ) \le {\mathscr{F}}_\varepsilon ^\nu (\tilde u_{\varepsilon ,i - 1}^\nu ,\mathit{\boldsymbol{\widetilde v}}_{\varepsilon ,i - 1}^\nu ,\tilde \theta _{\varepsilon ,i - 1}^\nu ) + h|{[\mathit{\boldsymbol{f}}_i^*]^h}|_{{V^*}}^2,\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;{\rm{for all}}\;0 < \varepsilon \le 1\;{\rm{and}}\;i = 1,2,3, \ldots . \end{array} (5.2)

    In the light of (A3)-(A6), (4.21) and (5.2), the class \{ [\tilde{u}_{\varepsilon, i}^{\nu}, \tilde{\mathit{\pmb{v}}}_{\varepsilon, i}^{\nu}, \tilde{\theta}_{\varepsilon, i}^{\nu}]\ |\ i \in \mathbb{N}, ~\varepsilon \in (0, 1] \} is bounded in V \times H^{1}(\Omega)^{3}. Therefore, applying a diagonal argument and the general theories of compactness (cf. [3,11]), we find sequences \{\varepsilon_{n} \}_{n=1}^\infty \subset (0, 1] , \{[u_{i}^{\nu}, \mathit{\pmb{v}}_{i}^{\nu}, \theta_{i}^{\nu}] \}_{i=1}^\infty=\{ [u_{i}^{\nu}, w_{i}^{\nu}, \eta_{i}^{\nu}, \theta_{i}^{\nu}] \}_{i=1}^\infty \subset V \times H^{1}(\Omega)^{2} \times H^1(\Omega) , with \{ \mathit{\pmb{v}}_i^\nu \}_{i=1}^\infty=\{ [w_i^\nu, \eta_i^\nu] \}_{i=1}^\infty , such that

    \left\{ \begin{array}{l} 1 \ge {\varepsilon _1} > \cdots > {\varepsilon _n} \downarrow 0{\rm{ as }}n \to \infty ,\\ \tilde u_{i,n}^\nu : = \tilde u_{{\varepsilon _n},i}^\nu \to u_i^\nu {\rm{ in}}\;{L^2}(\Omega ),\;{\rm{ weakly}}\;{\rm{in}}\;\mathit{V}\;{\rm{as}}\;n \to \infty \\ \mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu : = \mathit{\boldsymbol{\widetilde v}}_{{\varepsilon _n},i}^\nu \to {\rm{ }}\mathit{\boldsymbol{v}}_i^\nu {\rm{ in}}\;{L^2}{(\Omega )^2},{\rm{ weakly}}\;{\rm{in }}{H^1}{(\Omega )^2},{\rm{ weakly - }} * {\rm{ in}}\;{L^\infty }{(\Omega )^2},\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;{\rm{and}}\;{\rm{in}}\;{\rm{the}}\;{\rm{pointwise}}\;{\rm{sense}}\;{\rm{a}}{\rm{.e}}{\rm{. in}}\;\Omega {\rm{, as}}\;n \to \infty ,\\ \tilde \theta _{i,n}^\nu \equiv \tilde \theta _{{\varepsilon _n},i}^\nu \to \theta _i^\nu {\rm{ in }}{L^2}(\Omega ),{\rm{ weakly}}\;{\rm{in }}{H^1}(\Omega )\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;{\rm{and}}\;{\rm{in}}\;{\rm{the}}\;{\rm{pointwise}}\;{\rm{sense}}\;{\rm{a}}{\rm{.e}}{\rm{. in }}\Omega {\rm{, as }}n \to \infty \\ \le w_i^\nu \le 1{\rm{ and }}0 \le \eta _i^\nu \le 1{\rm{ }}\;{\rm{a}}{\rm{.e}}{\rm{.in}}\;\Omega ;\;{\rm{for}}\;{\rm{all}}\;i = 0,1,2, \ldots . \end{array} \right. (5.3)

    Moreover, by (3.13), (5.3), Lemmas 8-9 and Remark 4 (Fact 6), we infer that

    \left\{ \begin{array}{l} [\theta _i^\nu , - \frac{1}{h}{\alpha _0}(\mathit{\boldsymbol{v}}_i^\nu )(\theta _i^\nu - \theta _{i - 1}^\nu )] \in \partial {\Phi _\nu }(\mathit{\boldsymbol{v}}_i^\nu ; \cdot ){\rm{ in}}\;{L^2}{(\Omega )^2},\\ \Phi _{{\varepsilon _n}}^\nu (\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu ;\tilde \theta _{i,n}^\nu ) \to {\Phi _\nu }(\mathit{\boldsymbol{v}}_i^\nu ;\theta _i^\nu ),\tilde \theta _{i,n}^\nu \to \theta _i^\nu {\rm{ in }}{H^1}(\Omega )\;\;\;\;\;for\;\mathit{i = }{\rm{0,1,2,}}\\ \;\;\;\;\;\;\;\;{\rm{and}}\;{\varepsilon _n}\tilde \theta _{i,n}^\nu \to 0\;{\rm{in}}\;{H^M}(\Omega ),{\rm{as}}\;n \to \infty \end{array} \right. \cdots (5.4)

    Also, since

    \begin{equation*} [c, 0] \in \partial\Phi_{\nu}(\mathit{\pmb{v}}_{i}^{\nu};\cdot) \ \ \mbox{ in } L^{2}(\Omega)^{2}, \mbox{ for all $c \in \mathbb{R}$ and $ i = 0, 1, 2, \dots $, } \end{equation*}

    it is observed that

    \begin{equation*} \theta_{i}^{\nu} \le |\theta_{i-1}^{\nu}|_{L^{\infty}(\Omega)} \ (\mbox{resp. } \theta_{i}^{\nu} \ge - |\theta_{i-1}^{\nu}|_{L^{\infty}(\Omega)}) \mbox{ a.e. in } \Omega, \mbox{ for any $i \in \mathbb{N}$, } \end{equation*}

    by applying Lemma 10 as the case when

    \begin{equation*} \left\{ \begin{array}{l} \mathit{\pmb{v}}^{\circ} = \mathit{\pmb{v}}_{i}^{\nu}, \\ \check{\theta}_{0}^{\circ} = \theta_{i-1}^{\nu}, ~ \hat{\theta}_{0}^{\circ} = |\theta_{i-1}^{\nu}|_{L^{\infty}(\Omega)} \ (\mbox{resp. } \check{\theta}_{0}^{\circ} = -|\theta_{i-1}^{\nu}|_{L^{\infty}(\Omega)}, ~\hat{\theta}_{0}^{\circ} = \theta_{i-1}^{\nu}), \\ [\check{\theta}, \check{\theta}^{\ast}] = [\theta_{i}^{\nu}, -\frac{1}{h}\alpha_{0}(\mathit{\pmb{v}}_{i}^{\nu})(\theta_{i}^{\nu}-\theta_{i-1}^{\nu})] \ (\mbox{resp. } [\check{\theta}, \check{\theta}^{\ast}] = [-|\theta_{i-1}^{\nu}|_{L^{\infty}(\Omega)}, 0]), \\ [\hat{\theta}, \hat{\theta}^{\ast}] = [|\theta_{i-1}^{\nu}|_{L^{\infty}(\Omega)}, 0] \ \left(\mbox{resp. } [\hat{\theta}, \hat{\theta}^{\ast}] = \bigl[\theta_{i}^{\nu}, -\frac{1}{h}\alpha_{0}(\mathit{\pmb{v}}_{i}^{\nu})(\theta_{i}^{\nu}-\theta_{i-1}^{\nu}) \bigr] \right). \end{array} \right. \end{equation*}

    Having in mind (A2)-(A5), (3.11)-(3.12) and (5.3)-(5.4), we can see that

    \begin{equation*} \begin{array}{ll} \frac{1}{h}(u_{i}^{\nu} - u_{i-1}^{\nu}, z)_{L^{2}(\Omega)} - \frac{1}{h}(\lambda'(w_{i}^{\nu})(w_{i}^{\nu}-w_{i-1}^{\nu}), z)_{L^{2}(\Omega)} +(u_i^\nu, z)_V \\ = \lim_{n \to \infty} \left[\frac{1}{h}(\tilde{u}_{i, n}^{\nu}-\tilde{u}_{i-1, n}^{\nu}, z)_{L^{2}(\Omega)}-\frac{1}{h}(\lambda'(\tilde{w}_{i, n}^{\nu})(\tilde{w}_{i, n}^{\nu}-\tilde{w}_{i-1, n}^{\nu}), z)_{L^{2}(\Omega)} +(\tilde{u}_{i, n}^{\nu}, z)_{V} \right] \\ = \langle [{\mathit{\pmb{f}}}_{i}^*]^{h}, z \rangle, \mbox{ for any $z \in V$ and $ i = 1, 2, 3, \dots $, } \end{array} \end{equation*}

    and

    \begin{array}{l} {(D\mathit{\boldsymbol{v}}_i^\nu ,D(\mathit{\boldsymbol{v}}_i^\nu - \varpi ))_{{L^2}{{(\Omega )}^{N \times 2}}}} + \int_\Omega \gamma (w_i^\nu ){\mkern 1mu} dx - \int_\Omega \gamma (\varphi ){\mkern 1mu} dx\\ \le \mathop {\lim \inf }\limits_{n \to \infty } {(D\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu ,D(\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu - \varpi ))_{{L^2}{{(\Omega )}^{N \times 2}}}} + \mathop {\lim \inf }\limits_{n \to \infty } \int_\Omega \gamma (\tilde w_{i,n}^\nu )dx - \int_\Omega \gamma (\varphi ){\mkern 1mu} dx\\ \le \mathop {\lim \sup }\limits_{n \to \infty } {(D\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu ,D(\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu - \varpi ))_{{L^2}{{(\Omega )}^{N \times 2}}}} + \mathop {\lim \inf }\limits_{n \to \infty } \int_\Omega \gamma (\tilde w_{i,n}^\nu )dx - \int_\Omega \gamma (\varphi ){\mkern 1mu} dx\\ \le - \mathop {\lim }\limits_{n \to \infty } \left( {\frac{1}{h}{{(\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu - \mathit{\boldsymbol{\widetilde v}}_{i - 1,n}^\nu ,\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu - \varpi )}_{{L^2}{{(\Omega )}^2}}} + \int_\Omega {[\nabla G](\tilde u_{i,n}^\nu ;\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu ) \cdot (\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu - \varpi )dx} } \right)\\ - \mathop {\lim }\limits_{n \to \infty } \int_\Omega {([\nabla \alpha ](} \mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu )|D\tilde \theta _{i - 1,n}^\nu | + {\nu ^2}[\nabla \beta ](\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu )|D\tilde \theta _{i - 1,n}^\nu {|^2}) \cdot (\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu - \varpi )dx\\ \le - \frac{1}{h}{(\mathit{\boldsymbol{v}}_i^\nu - \mathit{\boldsymbol{v}}_{i - 1}^\nu ,\mathit{\boldsymbol{v}}_i^\nu - \varpi )_{{L^2}{{(\Omega )}^2}}} - \int_\Omega {[\nabla G](u_i^\nu ;\mathit{\boldsymbol{v}}_i^\nu ) \cdot (\mathit{\boldsymbol{v}}_i^\nu - \varpi )dx} \\ - \int_\Omega {([\nabla \alpha ](\mathit{\boldsymbol{v}}_i^\nu )|D\theta _{i - 1}^\nu | + {\nu ^2}[\nabla \beta ](\mathit{\boldsymbol{v}}_i^\nu )|D\theta _{i - 1}^\nu {|^2})} \cdot (\mathit{\boldsymbol{v}}_i^\nu - \varpi )dx,\\ {\rm{for}}\;{\rm{any}}\;\varpi = [\varphi ,\psi ] \in {[{H^1}(\Omega ) \cap {L^\infty }(\Omega )]^2},{\rm{and}}\;i = 1,2,3, \ldots . \end{array} (5.5)

    The above calculations imply that the limiting sequence \{[u_{i}^{\nu}, \mathit{\pmb{v}}_{i}^{\nu}, \theta_{i}^{\nu}] \}_{i=1}^\infty is a solution to the approximating system (AP)_{h}^{\nu} .

    The part of uniqueness and energy inequality. By putting {\varpi}=\mathit{\pmb{v}}_i^\nu in (5.5), for i \in \mathbb{N} , one can see from (5.3) that:

    \begin{array}{l} |D\mathit{\boldsymbol{v}}_i^\nu |_{{L^2}{{(\Omega )}^{N \times 2}}}^2 \le \mathop {\lim \inf }\limits_{n \to \infty } |D\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu |_{{L^2}{{(\Omega )}^{N \times 2}}}^2 \le \mathop {\lim \sup }\limits_{n \to \infty } |D\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu |_{{L^2}{{(\Omega )}^{N \times 2}}}^2\\ \le \mathop {\lim }\limits_{n \to \infty } {(D\mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu ,D\mathit{\boldsymbol{v}}_i^\nu )_{{L^2}{{(\Omega )}^{N \times 2}}}} + \int_\Omega \gamma (w_i^\nu ){\mkern 1mu} dx - \mathop {\lim \inf }\limits_{n \to \infty } \int_\Omega \gamma (\tilde w_{i,n}^\nu ){\mkern 1mu} dx\\ \le |D{\rm{ }}\mathit{\boldsymbol{v}}_i^\nu |_{{L^2}{{(\Omega )}^{N \times 2}}}^2,\;{\rm{for}}\;i = 1,2,3, \ldots . \end{array} (5.6)

    By the convergences (5.3) and (5.6), the uniform convexity of L^2 -based topologies enable to say:

    \mathit{\boldsymbol{\widetilde v}}_{i,n}^\nu \to \mathit{\boldsymbol{v}}_i^\nu \;{\rm{in}}\;{H^1}{(\Omega )^2}\;{\rm{as}}\;n \to \infty ,{\rm{for}}\;i = 1,2,3, \ldots . (5.7)

    Hence, the energy-inequality (3.10) will be obtained, immediately, by putting \varepsilon=\varepsilon_n in (5.2), for n \in \mathbb{N} , and letting n \to \infty with (5.3) and (5.7) in mind.

    In the meantime, we note that the condition (4.33) is still available in the proof of Theorem 1. Also, the regularity \theta_0^\circ \in H^M (\Omega) will not necessary in the calculations (4.29)-(4.32), and the line of these calculations will work even if \theta_0^\circ \in H^1(\Omega) .

    In view of these, the verification part of the uniqueness for (AP)_h^\nu will be a slight modification of that as in (Step 1) in the previous section. Then, the principal modifications will be to replace the application parts of Lemma 5 and the energy-inequality (4.20), by those of Lemma 10 and (3.10), respectively.


    6. Proof of Main Theorem

    Let \nu \geq 0 be a fixed constant, and let h_{1}^\circ \in (0, 1] be the constant as in Theorem 1. Also, we refer to [31] to recall the following lemma.

    Lemma 11. (Γ-convergence; [31, Lemma 6.2]) Assume \mathit{\pmb{v}}^\bullet \in [H^1(\Omega) \cap L^\infty (\Omega)]^2 , \{ \mathit{\pmb{v}}_{\tilde{\nu}}^\bullet \}_{\tilde{\nu} > 0} \subset [H^1(\Omega) \cap L^\infty (\Omega)]^2 , and

    \left\{ \begin{align} & \mathit{\pmb{v}}_{{\tilde{\nu }}}^{\bullet }\to {{\mathit{\pmb{v}}}^{\bullet }}\ in\ the\ pointwise\ sense, \ a.e.\ in\ \Omega, as\ \tilde{\nu }\downarrow 0, \\ & {{\{\mathit{\pmb{v}}_{{\tilde{\nu }}}^{\bullet }\}}_{\tilde{\nu }>0}}\ is\ bounded\ in\ {{L}^{\infty }}{{(\Omega )}^{2}}. \\ \end{align} \right.

    Then, for the sequence of convex functions \{ \Phi_{\tilde{\nu}}(\mathit{\pmb{v}}_{\tilde{\nu}}^\bullet; {}\cdot\, ) \}_{\tilde{\nu} > 0} , it holds that \Phi_{\tilde{\nu}}(\mathit{\pmb{v}}_{\tilde{\nu}}^\bullet; \cdot\, ) \to \Phi_0(\mathit{\pmb{v}}^\bullet; {}\cdot\, ) on L^2(\Omega) , in the sense of {\mit \Gamma} -convergence, as \tilde{\nu} \downarrow 0 .

    Based on Lemma 11 and [31, Remark 6.1], we take a sequence \{ \vartheta_0^{\tilde{\nu}} \}_{\tilde{\nu} > 0} \subset H^1(\Omega) , such that:

    \begin{equation*} |\theta_0^{\tilde{\nu}}| \leq |\theta_0|_{L^\infty(\Omega)} \mbox{ a.e. in $ \Omega $, for any $ \tilde{\nu} > 0 $, } \end{equation*}

    and

    \begin{equation*} \left\{ \begin{array}{l} \vartheta_0^{\tilde{\nu}} \to \theta_0 \mbox{ in $ L^2(\Omega) $ and }\Phi_{\tilde{\nu}}(\mathit{\pmb{v}}_0; \vartheta_0^{\tilde{\nu}}) \to \Phi_0(\mathit{\pmb{v}}_0; \theta_0), \mbox{ as $ \tilde{\nu} \downarrow 0 $, if $ \nu = 0 $, } \\[1ex] \vartheta_0^{\tilde{\nu}} = \theta_0 \mbox{ in $ H^1(\Omega) $ for $ \tilde{\nu} > 0 $, if $ \nu > 0 $, } \end{array} \right. \end{equation*}

    and for any h \in (0, h_1^\circ] and any \tilde{\nu} \in (0, \nu +1] , let us take the solution \{[u_{i}^{\tilde{\nu}}, \mathit{\pmb{v}}_{i}^{\tilde{\nu}}, \theta_{i}^{\tilde{\nu}}]\}_{i=0}^{\infty} to (AP)_{h}^{\tilde{\nu}} with \{ \mathit{\pmb{v}}_i^{\tilde{\nu}} \}_{i=1}^\infty=\{ [w_i^{\tilde{\nu}}, \eta_i^{\tilde{\nu}}] \}_{i=1}^\infty , under the initial condition [u_{0}^{\tilde{\nu}}, \mathit{\pmb{v}}_{0}^{\tilde{\nu}}, \theta_{0}^{\tilde{\nu}}]=[u_{0}, \mathit{\pmb{v}}_{0}, \vartheta_{0}^{\tilde{\nu}}] \in D_{1} with \mathit{\pmb{v}}_0^{\tilde{\nu}}=[w_0^{\tilde{\nu}}, \eta_0^{\tilde{\nu}}]=[w_0, \eta_0] . As is easily seen,

    \begin{equation*} F_0^\nu := \sup_{0 < \tilde{\nu} \leq \nu +1} \mathscr{F}_{\tilde{\nu}}(u_0, \mathit{\pmb{v}}_0, \vartheta_0^{\tilde{\nu}}) < \infty. \end{equation*}

    For any h \in (0, h_1^\circ] and any \tilde{\nu} \in (0, \nu +1] , we define the following time-interpolations:

    \begin{equation*} \begin{array}{l} {\mathit{\pmb{f}}}_h^*(t) = [f_h(t), {\mathit{\pmb{f}}}_{\Gamma, h}(t)] := [{f}_i^*]^h = [f_i^h, f_{\Gamma, i}^h] \mbox{ in $ V^* $ (in $ L^2(\Omega) \times L^2(\Gamma) $), } \\ \hspace{5mm}\mbox{ for all } t \ge 0 \mbox{ and } 0 \le i \in \mathbb{Z} \mbox{ satisfying } t \in ((i-1)h, ih], \end{array} \end{equation*}

    and

    \left\{ \begin{array}{l} [\bar u_h^{{\kern 1pt} \tilde \nu }(t),\mathit{\boldsymbol{\overline v}} _h^{{\kern 1pt} \tilde \nu }(t),\bar \theta _h^{{\kern 1pt} \tilde \nu }(t)] = [\bar u_h^{{\kern 1pt} \tilde \nu }(t),\bar w_h^{{\kern 1pt} \tilde \nu }(t),\bar \eta _h^{{\kern 1pt} \tilde \nu }(t),\bar \theta _h^{{\kern 1pt} \tilde \nu }(t)]: = [u_i^{\tilde \nu },{\rm{ }}\mathit{\boldsymbol{v}}_i^{\tilde \nu },\theta _i^{\tilde \nu }]\;{\rm{in}}\;{L^2}{(\Omega )^4},\\ \;\;\;\;{\rm{for all}}\;t \ge 0\;{\rm{and}}\;0 \le i \in \mathbb{Z}\;{\rm{satisfying}}\;t \in ((i - 1)h,ih],\\ [\mathit{\underline u} _{{\kern 1pt} h}^{\tilde \nu }(t),\mathit{\boldsymbol{\underline v}} _{{\kern 1pt} h}^{\tilde \nu }(t),\underline \theta _{{\kern 1pt} h}^{\tilde \nu }(t)] = [\mathit{\underline u} _{{\kern 1pt} h}^{\tilde \nu }(t),\underline w _{{\kern 1pt} h}^{\tilde \nu }(t),\underline \eta _{{\kern 1pt} h}^{\tilde \nu }(t),\underline \theta _{{\kern 1pt} h}^{\tilde \nu }(t)]: = [u_{i - 1}^{\tilde \nu },{\rm{ }}\mathit{\boldsymbol{v}}_{i - 1}^{\tilde \nu },\theta _{i - 1}^{\tilde \nu }]\;{\rm{in}}\;{L^2}{(\Omega )^4},\\ \;\;\;\;{\rm{for all}}\;t \ge 0\;{\rm{and}}\;0 \le i \in \mathbb{Z}\;{\rm{satisfying}}\;t \in \left[ {(i - 1)h,ih} \right),\\ [\hat u_h^{{\kern 1pt} \tilde \nu }(t),\mathit{\boldsymbol{\widehat v}}_h^{{\kern 1pt} \tilde \nu }(t),\hat \theta _h^{{\kern 1pt} \tilde \nu }(t)] = [\hat u_h^{{\kern 1pt} \tilde \nu }(t),\hat w_h^{{\kern 1pt} \tilde \nu }(t),\hat \eta _h^{{\kern 1pt} \tilde \nu }(t),\hat \theta _h^{{\kern 1pt} \tilde \nu }(t)]\\ \;\;\;\;: = \frac{{ih - t}}{h}[u_{i - 1}^{\tilde \nu }(t),\mathit{\boldsymbol{v}}_{i - 1}^{\tilde \nu }(t),\theta _{i - 1}^{\tilde \nu }(t)] + \frac{{t - (i - 1)h}}{h}[u_i^{\tilde \nu },{\rm{ }}\mathit{\boldsymbol{v}}_i^{\tilde \nu },\theta _i^{\tilde \nu }]\;{\rm{in}}\;{L^2}{(\Omega )^4},\\ \;\;\;\;{\rm{ for all }}t \ge 0{\rm{ and }}0 \le i \in \mathbb{Z}{\rm{ satisfying }}t \in [(i - 1)h,ih).{\rm{ }} \end{array} \right. (6.1)

    Besides, we define:

    {{D}_{\nu }}({{\theta }_{0}}):=\left\{ \begin{align} & \left\{ [\tilde{u}, \widetilde{v}, \tilde{\theta }]\in {{D}_{0}}\left| |\tilde{\theta }{{|}_{{{L}^{\infty }}(\Omega )}}\le |{{\theta }_{0}}{{|}_{{{L}^{\infty }}(\Omega )}} \right. \right\}, \ \text{if}\ \ v=0, \\ & \left\{ [\tilde{u}, \widetilde{v}, \tilde{\theta }]\in {{D}_{1}}\left| |\tilde{\theta }{{|}_{{{L}^{\infty }}(\Omega )}}\le |{{\theta }_{0}}{{|}_{{{L}^{\infty }}(\Omega )}} \right. \right\}, \ \text{if}\ \ v>0, \\ \end{align} \right.\

    and we note that:

    \begin{equation*} \begin{array}{c} \bigl\{ [\overline{u}_{h}^{\, \tilde{\nu}}(t), \overline{\mathit{\pmb{v}}}_{h}^{\, \tilde{\nu}}(t), \overline{\theta}_{h}^{\, \tilde{\nu}}(t)], [\underline{u}_{\, h}^{\, \tilde{\nu}}(t), \underline{\mathit{\pmb{v}}}_{\, h}^{\, \tilde{\nu}}(t), \underline{\theta}_{\, h}^{\, \tilde{\nu}}(t)], [\widehat{u}_{h}^{\, \tilde{\nu}}(t), \widehat{\mathit{\pmb{v}}}_{h}^{\, \tilde{\nu}}(t), \widehat{\theta}_{h}^{\, \tilde{\nu}}(t)] \bigr\} \\[1ex] \subset D_\nu(\theta_0), \mbox{ for all $ t \geq 0 $, $ 0 < h \leq h_1^\circ $ and $ 0 < \tilde{\nu} \leq \nu +1 $.} \end{array} \end{equation*}

    Then, from the energy-inequality (3.10) in Theorem 1, it is checked that

    \begin{equation*} \begin{array}{c} \frac{A_*}{2} \int_s^t |(\widehat{u}_{h}^{\, \tilde{\nu}})_t|_{V^*} \, d\tau +\frac{1}{2}\int_{s}^{t}|(\widehat{\mathit{\pmb{v}}}_{h}^{\, \tilde{\nu}})_t(\tau)|_{L^{2}(\Omega)^{2}}^{2}d\tau + \int_{s}^{t} |{\textstyle \sqrt{\alpha_{0}(\overline{\mathit{\pmb{v}}}_{h}^{\, \tilde{\nu}})}}(\widehat{\theta}_{h}^{\, \tilde{\nu}})_t(\tau)|_{L^{2}(\Omega)}^{2}d\tau \\[2ex] + \frac{1}{2}\int_{s}^{t}|\overline{u}_{h}^{\, \tilde{\nu}}(\tau)|_{V}^{2}d\tau + \mathscr{F}_{\tilde{\nu}}(\overline{u}_{h}^{\, \tilde{\nu}}, \overline{\mathit{\pmb{v}}}_{h}^{\, \tilde{\nu}}, \overline{\theta}_{h}^{\, \tilde{\nu}})(t) \le \mathscr{F}_{\tilde{\nu}}(\underline{u}_{\, h}^{\tilde{\nu}}, \underline{\mathit{\pmb{v}}}_{\, h}^{\tilde{\nu}}, \underline{\theta}_{\, h}^{\tilde{\nu}})(s) + \int_{s}^{t} |{f}_h^*(\tau)|_{V^{\ast}}^{2} d\tau \\[2ex] \mbox{for all $0 \le s \le t \le T$, $ 0 < h \leq h_1^\circ $ and $ 0 < \tilde{\nu} \leq \nu +1 $, } \end{array} \end{equation*}

    and additionally, from (A1)-(A6) and (3.2), it follows that

    \begin{array}{l} {B_*}|\bar u_h^{{\kern 1pt} \tilde \nu }(t)|_{{L^2}(\Omega )}^2 + \frac{1}{2}|D\mathit{\boldsymbol{\overline v}} _h^{{\kern 1pt} \tilde \nu }(t)|_{{L^2}{{(\Omega )}^{N \times 2}}}^2 + {\delta _*}(|D\bar \theta _h^{{\kern 1pt} \tilde \nu }(t)|(\Omega ) + |D(\tilde \nu \bar \theta _h^{{\kern 1pt} \tilde \nu })(t)|_{{L^2}{{(\Omega )}^{N \times 2}}}^2)\\ \;\;\;\;\;\; \le \mathscr{F}{_{\tilde \nu }}(\bar u_h^{{\kern 1pt} \tilde \nu },\mathit{\boldsymbol{\overline v}} _h^{{\kern 1pt} \tilde \nu },\bar \theta _h^{{\kern 1pt} \tilde \nu })(t) \vee \mathscr{F}{_{\tilde \nu }}(\underline u _{{\kern 1pt} h}^{\tilde \nu },\mathit{\boldsymbol{\underline v}} _{{\kern 1pt} h}^{\tilde \nu },\underline \theta _{{\kern 1pt} h}^{\tilde \nu })(t)\\ \;\;\;\;\;\; \le F_ * ^\nu : = F_0^\nu + |{\mathit{\boldsymbol{f}}^*}|_{{L^2}(0,T;{V^*})}^2,\;{\rm{for}}\;{\rm{all}}\;0 \le t \le T\;{\rm{and}}\;0 < \tilde \nu \le \nu + 1. \end{array} (6.2)

    Based on these, one can see that:

    (\sharp 1) the class \{\widehat{u}_{h}^{\, \tilde{\nu}} \, | \, h \in (0, h_1^\circ], ~\tilde{\nu} \in (0, \nu +1] \} is bounded in the space W^{1, 2}(0, T; V^{\ast}) \cap C ([0, T]; L^{2}(\Omega)) \cap L^{2}(0, T; V).

    (\sharp 2) the class \{ \widehat{\mathit{\pmb{v}}}_{h}^{\, \tilde{\nu}} \, | \, h \in (0, h_1^\circ], ~\tilde{\nu} \in (0, \nu +1] \} is bounded in the space W^{1, 2}(0, T; L^{2}(\Omega)^{2}) \cap L^{\infty}(0, T; H^{1}(\Omega)^{2}) \cap L^\infty (Q)^2.

    (\sharp 3) the class \{ \widehat{\theta}_{h}^{\, \tilde{\nu}} \, | \, h \in (0, h_1^\circ], ~\tilde{\nu} \in (0, \nu +1] \} is bounded in the space W^{1, 2}(0, T; L^2(\Omega)) \cap L^\infty (Q) , and \{ \Phi_{\tilde{\nu}}(\mathit{\pmb{v}}_h^{\, \tilde{\nu}}; \theta_h^{\, \tilde{\nu}}) \, | \, h \in (0, h_1^\circ], ~\tilde{\nu} \in (0, \nu +1] \} is bounded in L^\infty (0, T) , i.e. \{ |D \overline{\theta}_h^{\, \tilde{\nu}}({}\cdot{})|(\Omega) \, | \, h \in (0, h_1^\circ], ~\tilde{\nu} \in (0, \nu +1] \} is bounded in L^\infty (0, T) , and \{ D (\tilde{\nu} \overline{\theta}_h^{\, \tilde{\nu}}) \, | \, h \in (0, h_1^\circ], ~\tilde{\nu} \in (0, \nu +1] \} is bounded in L^\infty (0, T; L^2(\Omega)^N) .

    Hence, by applying the general theories of compactness, as in [2,3,11,33], we find a quartet of functions [u, \mathit{\pmb{v}}, \theta]=[u, w, \eta, \theta] \in L^{2}(0, T; L^{2}(\Omega)^{4}) with \mathit{\pmb{v}}=[w, \eta] and sequences \{h_{n}\}_{n=1}^{\infty} \subset (0, h_1^\circ] and \{ \nu_n \}_{n=1}^\infty \subset (0, \nu +1] , with the subsequences:

    \begin{equation*} \left \{ \begin{array}{l} \{[\overline{u}_{n}, \overline{\mathit{\pmb{v}}}_{n}, \overline{\theta}_{n}]\}_{n=1}^{\infty} = \{[\overline{u}_{n}, \overline{w}_{n}, \overline{\eta}_{n}, \overline{\theta}_{n}]\}_{n=1}^{\infty} := \{[\overline{u}_{h_n}^{\nu_n}, \overline{\mathit{\pmb{v}}}_{h_n}^{\nu_n}, \overline{\theta}_{h_n}^{\nu_n}]\}_{n=1}^{\infty}, \\ \{[\underline{u}_{\, n}, \underline{\mathit{\pmb{v}}}_{\, n}, \underline{\theta}_{\, n}]\}_{n=1}^{\infty} = \{[\underline{u}_{\, n}, \underline{w}_{\, n}, \underline{\eta}_{\, n}, \underline{\theta}_{\, n}]\}_{n=1}^{\infty} := \{[\underline{u}_{\, h_n}^{\nu_n}, \underline{\mathit{\pmb{v}}}_{\, h_n}^{\nu_n}, \underline{\theta}_{\, h_n}^{\nu_n}]\}_{n=1}^{\infty}, \\ \{[\widehat{u}_{n}, \widehat{\mathit{\pmb{v}}}_{n}, \widehat{\theta}_{n}]\}_{n=1}^{\infty} = \{[\widehat{u}_{n}, \widehat{w}_{n}, \widehat{\eta}_{n}, \widehat{\theta}_{n}]\}_{n=1}^{\infty} := \{[\widehat{u}_{h_n}^{\nu_n}, \widehat{\mathit{\pmb{v}}}_{h_n}^{\nu_n}, \widehat{\theta}_{h_n}^{\nu_n}]\}_{n=1}^{\infty}, \\ \end{array}\right. \end{equation*}

    such that:

    h_1^ \circ \ge {h_1} > {h_2} > \cdots > {h_n} \downarrow 0\;{\rm{and}}\;{\nu _n} \to \nu ,{\rm{as}}\;n \to \infty , (6.3)
    \left\{ \begin{array}{l} u \in {W^{1,2}}(0,T;{V^ * }) \cap {L^\infty }(0,T;{L^2}(\Omega )) \cap {L^2}(0,T;V) \subset C([0,T];{L^2}(\Omega )),\\ \mathit{\boldsymbol{v}} \in {W^{1,2}}(0,T;{L^2}{(\Omega )^2}) \cap {L^\infty }(0,T;{H^1}{(\Omega )^2}) \cap {L^\infty }{(Q)^2},\\ \theta \in {W^{1,2}}(0,T;{L^2}(\Omega )) \cap {L^\infty }(Q),{\Phi _\nu }(\mathit{\boldsymbol{v}};\theta ) \in {L^\infty }(0,T),\\ [u(t),\mathit{\boldsymbol{v}}(t),\theta (t)] \in {D_\nu }({\theta _0}){\rm{ for all }}t \ge 0,\\ [u(0),\mathit{\boldsymbol{v}}(0),\theta (0)] = [{u_0},{\mathit{\boldsymbol{v}}_0},{\theta _0}]{\rm{ in }}{L^2}{(\Omega )^4}, \end{array} \right. (6.4)
    \left\{ \begin{array}{l} {{\hat u}_n} \to u\;{\rm{in}}\;{L^2}(I;{L^2}(\Omega )),{\rm{ weakly}}\;{\rm{in }}{W^{1,2}}(I;{V^ * }),\\ \;\;\;\;\;\;{\rm{weakly}} - *\;{\rm{in}}\;{L^\infty }(I;V),\\ {\mathit{\boldsymbol{\widehat v}}_n} \to \mathit{\boldsymbol{v}}\;{\rm{in}}\;C(\bar I;{L^2}{(\Omega )^2}),{\rm{weakly}}\;{\rm{in}}\;{W^{1,2}}(I;{L^2}{(\Omega )^2}),\\ \;\;\;\;\;\;{\rm{weakly}} - *\;{\rm{in}}\;{L^\infty }(I;{H^1}{(\Omega )^2})\;{\rm{and}}\;{\rm{weakly}} - *{\rm{in}}\;{L^\infty }{(Q)^2},\\ {{\hat \theta }_n} \to \theta \;{\rm{in}}\;C(\bar I;{L^2}(\Omega )),{\rm{weakly}}\;{\rm{in}}\;{W^{1,2}}(I;{L^2}(\Omega )),\\ \;\;\;\;\;\;{\rm{weakly}} - *\;{\rm{in}}\;{L^\infty }\left( Q \right),\\ {\nu _n}{{\hat \theta }_n} \to \nu \theta {\rm{weakly}}\;{\rm{in}}\;{L^2}(I;{H^1}(\Omega )), \end{array} \right. (6.5)
    \mathit{\boldsymbol{f}}_{{h_n}}^* \to {\rm{ }}{\mathit{\boldsymbol{f}}^*}\;{\rm{in}}\;{L^2}(I;{V^*})([{f_{{h_n}}},{f_{\Gamma ,{h_n}}}] \to [f,{f_\Gamma }]\;{\rm{in}}\;{L^2}(I;{L^2}(\Omega ) \times {L^2}(\Gamma ))), (6.6)

    as n \to \infty , for any open interval I \subset (0, T) , and

    \left\{ \begin{array}{l} {{\bar u}_n}(t) \to u(t)\;{\rm{and}}\;{\underline u _{{\kern 1pt} n}}(t) \to u(t)\;{\rm{in}}\;{L^2}(\Omega ),{\rm{weakly}}\;{\rm{in}}\;V,\\ {\mathit{\boldsymbol{\overline v}} _n}(t) \to \mathit{\boldsymbol{v}}(t)\;{\rm{and}}\;{\mathit{\boldsymbol{\underline v}} _{{\kern 1pt} n}}(t) \to \mathit{\boldsymbol{v}}(t)\;{\rm{in}}\;{L^2}{(\Omega )^2},{\rm{weakly}}\;{\rm{in}}\;{H^1}{(\Omega )^2}\\ \;\;\;\;\;\;\;\;{\rm{and}}\;{\rm{weakly}} - *\;{\rm{in}}\;{L^\infty }{(\Omega )^2},\\ {{\bar \theta }_n}(t) \to \theta (t)\;{\rm{in}}\;{L^2}(\Omega ),{\rm{weakly}}\;{\rm{in}}\; - *\;{\rm{in}}\;BV(\Omega ),\\ {\nu _n}{{\bar \theta }_n}(t) \to \nu \theta (t){\rm{ weakly}}\;{\rm{in}}\;{H^1}(\Omega ), \end{array} \right. (6.7)

    as n \to \infty for a.e. t \in (0, T).

    Now, we recall some lemmas which will act key-roles in the proof of Main Theorem.

    Lemma 12. Let I \subset (0, T) be an open interval, and let \nu \geq 0 and \{ \nu_n \}_{n=1}^\infty be the sequence as in (6.3). Let \zeta \in L^2(I; L^2(\Omega)) be a function such that

    \begin{equation*} |D \zeta({}\cdot{})|(\Omega) \in L^1(I) \mbox{ and } \nu \zeta \in L^{2}(I; H^{1}(\Omega)). \end{equation*}

    Then, there exists a sequence \{ \tilde{\zeta}_n \}_{n=1}^\infty \subset C^\infty (\overline{Q}) , such that:

    \begin{matrix} {{{\tilde{\zeta }}}_{n}}\to \zeta \text{ }in\ {{L}^{2}}(I;{{L}^{2}}(\Omega )), \int_{I}{\left| \int_{\Omega }{|}D{{{\tilde{\zeta }}}_{n}}\left. \left( t \right) \right|dx-\int_{\Omega }{d}|D\zeta (t)| \right|}dt\to 0, \\ and\ \ {{\nu }_{n}}{{{\tilde{\zeta }}}_{n}}\to \nu \zeta \text{ }in\ {{L}^{2}}(I;{{H}^{1}}(\Omega )), \ as\ n\to \infty . \\ \end{matrix}

    Proof. When \nu > 0 , the standard C^\infty -approximation in L^2(I; H^1(\Omega)) will correspond to the required sequence. Meanwhile, when \nu=0 , this lemma is verified by taking the C^\infty -approximation as in [25, Lemma 5]{MS14} and [29, Remark 2], and by applying the diagonal argument as in [25, Lemma 8].

    Lemma 13.Let I \subset (0, T) be any open interval. Assume that

    \left\{ \begin{align} & \varrho\in C(\bar{I};{{L}^{2}}(\Omega ))\cap {{L}^{\infty }}(I;{{H}^{1}}(\Omega )), \ \log \varrho\in {{L}^{\infty }}(I\times \Omega ) \\ & {{\varrho}_{n}}\in C(\bar{I};{{L}^{2}}(\Omega ))\cap {{L}^{\infty }}(I;{{H}^{1}}(\Omega )), \ \log {{\varrho}_{n}}\in {{L}^{\infty }}(I\times \Omega ), \ for\ n=1, 2, 3, \ldots \\ & {{\varrho}_{n}}(t)\to \varrho(t)\ in\ {{L}^{2}}(\Omega )\ and\ weakly\ in\ {{H}^{1}}(\Omega )\ as\ n\to \infty, for\ a.e.\ t\in I, \\ \end{align} \right.

    and

    \left\{ \begin{align} & \zeta \in {{L}^{2}}(I;{{L}^{2}}(\Omega ))\ with\ |D\zeta (\cdot )|(\Omega )\in {{L}^{1}}(I), \\ & \{{{\zeta }_{n}}\}_{n=1}^{\infty }\subset {{L}^{2}}(I;{{L}^{2}}(\Omega ))\ with\ \{|D{{\zeta }_{n}}(\cdot )|(\Omega )\}_{n=1}^{\infty }\subset {{L}^{1}}(I), \\ & {{\zeta }_{n}}(t)\to \zeta (t)\ in\ {{L}^{2}}(\Omega )\ as\ n\to \infty, \ a.e.\ t\in I. \\ \end{align} \right.

    Then the following items hold.

    (Ⅰ) The functions:

    \begin{equation*} t \in I \mapsto \int_\Omega d[\varrho(t)|D\zeta(t)|] \, dt \mbox{ and } t \in I \mapsto \int_\Omega d[\varrho_n(t)|D \zeta_n(t)|] \, dt, \mbox{ for $ n = 1, 2, 3, \dots $, } \end{equation*}

    are integrable, and

    \begin{equation*} \liminf_{n \to \infty}\int_I \int_\Omega d[\varrho_n(t) |D \zeta_n(t)|] \, dt \geq \int_I \int_\Omega d[\varrho(t) |D \zeta(t)|] \, dt. \end{equation*}

    (Ⅱ) If:

    \begin{equation*} \int_I \int_\Omega d [\varrho_n(t) |D \zeta_n(t)|] \, dt \to \int_I \int_\Omega d[\varrho(t) |D \zeta(t)|] \, dt \mbox{ as $ n \to \infty $} \end{equation*}

    and

    \left\{ \begin{align} & \omega \in {{L}^{\infty }}(I;{{H}^{1}}(\Omega ))\cap {{L}^{\infty }}(I\times \Omega ), \{{{\omega }_{n}}\}_{n=1}^{\infty }\subset {{L}^{\infty }}(I;{{H}^{1}}(\Omega ))\cap {{L}^{\infty }}(I\times \Omega ) \\ & \{{{\omega }_{n}}\}_{n=1}^{\infty }is\ a\ bounded\ sequence\ in\ {{L}^{\infty }}(I\times \Omega ), \\ & {{\omega }_{n}}(t)\to \omega (t)in\ {{L}^{2}}(\Omega )\ and\ weakly\ in\ {{H}^{1}}(\Omega )\ as\ n\to \infty, a.e.\ t\in I, \\ \end{align} \right.

    then

    \int_{I}{\int_{\Omega }{{{\omega }_{n}}}}(t)|D{{\zeta }_{n}}(t)\left| dx \right.\ dt\to \int_{I}{\int_{\Omega }{d}}[\omega (t)|D\zeta (t)|]\ \ as\ n\to \infty .

    Proof.This lemma is verified, immediately, as a consequence of [26, Lemmas 4.2-4.4] (see also [25, Section 2]).

    Proof of Main Theorem.  We show that the quartet [u, \mathit{\pmb{v}}, \theta]=[u, w, \eta, \theta] \in L^2(0, T; L^2(\Omega)^4) as in (6.4) fulfills the conditions (S1)-(S6) in Definition 3. Then, since (6.4) directly guarantees the conditions (S1)-(S3), we focus on the verifications of remaining (S4)-(S6).

    To this end, let us fix arbitrary open interval I \subset (0, T) , and let us review (3.7)-(3.9) and (6.1), to check that:

    \begin{array}{l} \int_I {\langle (} {{\hat u}_n}{)_t}(t),z\rangle {\mkern 1mu} dt + \int_I {({{\bar u}_n}(} t),z{)_V}{\mkern 1mu} dt = \int_I {{{(\lambda '({{\bar w}_n}(t)){{({{\hat w}_n})}_t}(t),z)}_{{L^2}(\Omega )}}} {\mkern 1mu} dt\\ \;\;\;\; + \int_I {\langle {\rm{ }}\mathit{\boldsymbol{f}}_{{h_n}}^*(} t),z\rangle {\mkern 1mu} dt,\;{\rm{for}}\;{\rm{any}}\;z \in V\;{\rm{and}}\;n = 1,2,3, \ldots , \end{array} (6.8)
    \begin{equation}\label{14-13} \begin{array}{l} \int_{I}((\widehat{\mathit{\pmb{v}}}_{n})_t(t), \overline{\mathit{\pmb{v}}}_{n}(t)-{\varpi})_{L^{2}(\Omega)^{2}} dt \\ \qquad + \int_{I} (D\overline{\mathit{\pmb{v}}}_{n}(t), D(\overline{\mathit{\pmb{v}}}_{n}(t)-{\varpi}))_{L^{2}(\Omega)^{N \times 2}} \, dt \\[2ex] \qquad + \int_{I} ([\nabla G](\overline{u}_{n};\overline{\mathit{\pmb{v}}}_{n})(t), \overline{\mathit{\pmb{v}}}_{n}(t)-{\varpi})_{L^{2}(\Omega)^{2}} dt \\[2ex] \qquad + \int_{I}\int_{\Omega} [\nabla\alpha](\overline{\mathit{\pmb{v}}}_{n}(t)) \cdot (\overline{\mathit{\pmb{v}}}_{n}(t)-{\varpi}) |D\underline{\theta}_{\,n}(t)| dxdt \\[2ex] \qquad + \nu_n^2 \int_{I}\int_{\Omega} [\nabla\beta](\overline{\mathit{\pmb{v}}}_{n}(t)) \cdot (\overline{\mathit{\pmb{v}}}_{n}(t)-{\varpi}) |D\underline{\theta}_{\,n}(t)|^{2} dxdt \\[2ex] \qquad + \int_{I}\int_{\Omega} \gamma(\overline{w}_{n}(t)) dxdt \le \int_{I}\int_{\Omega} \gamma(\varphi) dxdt, \\[3ex] \mbox{for any ${\varpi}=[\varphi,\psi] \in [H^{1}(\Omega) \cap L^{\infty}(\Omega)]^{2}$ and $ n = 1, 2, 3, \dots $,} \end{array} \end{equation} (6.9)

    and

    \begin{equation}\label{14-14} \begin{array}{l} \int_{I} (\alpha_{0}(\overline{\mathit{\pmb{v}}}_{n}(t))(\widehat{\theta}_{n})_{t}(t), \overline{\theta}_{n}(t)-\zeta(t))_{L^{2}(\Omega)} dt \\[2ex] \qquad +\int_I \int_\Omega \alpha(\overline{\mathit{\pmb{v}}}_n(t)) |D \overline{\theta}_n(t)| \, dx dt +\nu_n^2 \int_I \int_\Omega \beta(\overline{\mathit{\pmb{v}}}_n(t)) |D \overline{\theta}_n(t)|^2 \, dx dt \\[2ex] \leq \int_I \int_\Omega \alpha(\overline{\mathit{\pmb{v}}}_n(t)) |D \zeta(t)| \, dx dt +\nu_n^2 \int_I \int_\Omega \beta(\overline{\mathit{\pmb{v}}}_n(t)) |D \zeta(t)|^2 \, dx dt \\[2ex] \qquad \mbox{for any $\zeta \in L^{2}(I; H^{1}(\Omega))$ and $ n = 1, 2, 3, \dots $.} \end{array} \end{equation} (6.10)

    Now, let us first take the limit of (6.10) as n \to \infty . Then, from (A3), (\sharp 2)-(\sharp 3), (6.4)-(6.5), (6.7) and Lemma 13 (I), it is seen that

    \begin{equation*} \begin{array}{l} \int_{I} (\alpha_{0}(\mathit{\pmb{v}}(t))\theta_{t}(t), \theta(t)-\zeta(t))_{L^{2}(\Omega)} dt + \int_I \Phi_{\nu}(\mathit{\pmb{v}}(t); \theta(t)) \, dt \\[2ex] \hspace{5mm} \le \lim_{n \to \infty} \int_{I} (\alpha_{0}(\overline{\mathit{\pmb{v}}}_{n})(\widehat{\theta}_{n})_{t}(t), \overline{\theta}_{n}(t)-\zeta(t))_{L^{2}(\Omega)} \, dt \\[2ex] \hspace{10mm} + \liminf_{n \to \infty} \left[\int_I \int_\Omega \alpha(\overline{\mathit{\pmb{v}}}_n(t)) |D \overline{\theta}_n(t)| \, dx dt + \int_I \int_\Omega \beta(\overline{\mathit{\pmb{v}}}_n(t)) |D (\nu_n \overline{\theta}_n)(t)|^2 \, dx dt \right] \\ \hspace{5mm} \le \lim_{n \to \infty} \left[\int_I \int_\Omega \alpha(\overline{\mathit{\pmb{v}}}_n(t)) |D \zeta(t)| \, dx dt + \int_I \int_\Omega \beta(\overline{\mathit{\pmb{v}}}_n(t)) |D (\nu_n \zeta)(t)|^2 \, dx dt \right] \\[2ex] \hspace{5mm}= \int_I \Phi_{\nu}(\mathit{\pmb{v}}(t); \zeta(t)) \, dt, \mbox{ for any $\zeta \in L^{2}(I;H^{1}(\Omega))$.} \end{array} \end{equation*}

    Since the open interval I \subset (0, T) is arbitrary, the above inequality implies that

    \begin{equation*} \begin{array}{c} (\alpha_{0}(\mathit{\pmb{v}}(t))\theta_{t}(t), \theta(t) - \omega)_{L^{2}(\Omega)} + \Phi_{\nu}(\mathit{\pmb{v}}(t);\theta(t)) \le \Phi_{\nu}(\mathit{\pmb{v}}(t);\omega) \\[1ex] \mbox{for any $\omega \in H^{1}(\Omega)$ and a.e. $t \in (0, T)$.} \end{array} \end{equation*}

    Additionally, in the light of Remark 3 (Fact 4), we can say the above inequality holds for \omega \in BV (\Omega) \cap L^{2}(\Omega). Thus, (S6) is verified.

    Next, with (6.4) and Lemma 12 in mind, let us take a sequence \{ \tilde{\theta}_{n} \}_{n=1}^\infty \subset C^{\infty}(\overline{I \times \Omega}) such that

    \begin{equation*} \begin{array}{c} \tilde{\theta}_{n} \to \theta\ \ \mbox{ in } L^{2}(I; L^{2}(\Omega)), \ \ \int_I |D \tilde{\theta}_n| \, dx dt \to \int_I d|D \theta(t)| \, dt, \\[2ex] \nu_n \tilde{\theta}_n \to \nu \theta \mbox{ in $ L^2(I; H^1(\Omega)) $, as $ n \to \infty $.} \end{array} \end{equation*}

    Then, putting \zeta=\tilde{\theta}_{n} in (6.10) and letting n \to \infty , it is observed from (\sharp 2)-(\sharp 3), (6.4)-(6.5), (6.7) and Lemma 13 that:

    \begin{equation*} \begin{array}{l} \int_I \int_\Omega d[\alpha({\mathit{\pmb{v}}}(t)) |D {\theta}(t)|] \, dt + \int_I \int_\Omega \beta({\mathit{\pmb{v}}}(t)) |D (\nu {\theta})(t)|^2 \, dx dt \\[2ex] \qquad \leq \liminf_{n \to \infty} \int_I \int_\Omega \alpha(\overline{\mathit{\pmb{v}}}_n(t)) |D \overline{\theta}_n(t)| \, dx dt + \liminf_{n \to \infty} \int_I \int_\Omega \beta(\overline{\mathit{\pmb{v}}}_n(t)) |D (\nu_n \overline{\theta}_n)(t)|^2 \, dx dt \\[2ex] \qquad \leq \limsup_{n \to \infty} \left[\int_I \int_\Omega {\alpha(\overline{\mathit{\pmb{v}}}_n(t)) |D \overline{\theta}_n(t)|dx} \, dt + \int_I \int_\Omega \beta(\overline{\mathit{\pmb{v}}}_n(t)) |D (\nu_n \overline{\theta}_n)(t)|^2 \, dx dt \right] \\[2ex] \qquad \leq \lim_{n \to \infty} \left[\int_I \int_\Omega {\alpha(\overline{\mathit{\pmb{v}}}_n(t)) |D \tilde{\theta}_n(t)|dx} \, dt + \int_I \int_\Omega \beta(\overline{\mathit{\pmb{v}}}_n(t)) |D (\nu_n \tilde{\theta}_n)(t)|^2 \, dx dt \right] \\[2ex] \qquad \qquad - \lim_{n \to \infty} \int_{I} (\alpha_{0}(\overline{\mathit{\pmb{v}}}_{n})(\widehat{\theta}_{n})_{t}(t), \overline{\theta}_{n}(t)-\tilde{\theta}_{n}(t))_{L^{2}(\Omega)} \, dt \\[2ex] \qquad = \int_I \int_\Omega d[\alpha({\mathit{\pmb{v}}}(t)) |D {\theta}(t)|] \, dt + \int_I \int_\Omega \beta({\mathit{\pmb{v}}}(t)) |D (\nu {\theta})(t)|^2 \, dx dt. \end{array} \end{equation*}

    The above inequality implies that:

    \begin{equation}\label{14-18} \begin{matrix} \lim \\ n\to \infty \\ \end{matrix} \int_{I}\int_{\Omega} \alpha(\overline{\mathit{\pmb{v}}}_{n}(t))|D \overline{\theta}_{n}(t)| dxdt = \int_{I}\int_{\Omega} d[\alpha(\mathit{\pmb{v}}(t)) |D\theta(t)|]dt, \end{equation} (6.11)

    and

    \begin{equation}\label{14-18-1} \lim_{n \to \infty} \int_{I}\int_{\Omega} \beta(\overline{\mathit{\pmb{v}}}_{n}(t))|D (\nu_n \overline{\theta}_{n})(t)|^{2} dxdt = \int_{I}\int_{\Omega} \beta(\mathit{\pmb{v}}(t)) |D(\nu \theta)(t)|^2 \, dt. \end{equation} (6.12)

    By virtue of (\sharp 2)-(\sharp 3), (6.4)-(6.5), (6.7) and (6.11), we can apply Lemma 13 to see that:

    \begin{equation*} \int_{I}\int_{\Omega} |D \overline{\theta}_{n}(t)| dxdt \to \int_{I}\int_{\Omega} d|D \theta(t)| \, dt, \mbox{ as $n \to \infty$.} \end{equation*}

    Besides, (6.1)-(6.2) and (6.5) enable to check:

    \begin{equation*} \left| \int_{I}\int_{\Omega} |D \overline{\theta}_{n}| dxdt - \int_{I}\int_{\Omega} |D \underline{\theta}_{\, n}|dxdt \right| \le \frac{2F_{\ast}^{\nu}}{\delta_{\ast}}h_{n} \to 0, \mbox{ as $ n \to \infty $, } \end{equation*}

    and (6.5), (6.7) and the above convergence further enable to show that:

    \begin{equation}\label{14-19} \begin{array}{c} \begin{matrix} \lim \\ n\to \infty \\ \end{matrix} \int_{I}\int_{\Omega} (\overline{\mathit{\pmb{v}}}_n(t) -{\varpi}) \cdot [\nabla\alpha](\overline{\mathit{\pmb{v}}}_{n}(t))|D\underline{\theta}_{\,n}(t)| dxdt \\[2ex] = \int_{I}\int_{\Omega} d[ (\overline{\mathit{\pmb{v}}}_n(t) -{\varpi}) \cdot [\nabla\alpha](\mathit{\pmb{v}}(t))|D\theta(t)|]dt \mbox{ for any ${\varpi} \in [H^{1}(\Omega) \cap L^{\infty}(\Omega)]^{2}$,} \end{array} \end{equation} (6.13)

    by applying Lemma (13) (Ⅱ).

    Similarly, from (6.12) and the uniform convexity of L^2 -based topology, one can see that

    \begin{equation*} \left\{ \begin{array}{l} \sqrt{\beta(\overline{\mathit{\pmb{v}}}_n)} D ({\nu_{n}}\overline{\theta}_n) \to \sqrt{\beta(\mathit{\pmb{v}})} D ({\nu} \theta) \mbox{ in $ L^2(I; L^2(\Omega)^N) $, and hence} \\[1ex] D (\nu_n \overline{\theta}_n) \to D (\nu \theta) \mbox{ in $ L^2(I; L^2(\Omega)^N) $, as $ n \to \infty $.} \end{array} \right. \end{equation*}

    Besides, (6.1)-(6.2) and (6.5) enable to check:

    \left| \int_{I}\int_{\Omega} |D (\nu_n \overline{\theta}_{n})|^2 dxdt - \int_{I}\int_{\Omega} |D (\nu_n \underline{\theta}_{\, n})|^2 \, dxdt \right| \le \frac{2F_{\ast}^\nu} {\delta_{\ast}}h_{n} \to 0, \mbox{ as $ n \to \infty $, }

    and the above convergence further enables to show that:

    \left\{ \begin{array}{*{35}{l}} D({{\nu }_{n}}{{{\underset{\raise0.3em\hbox{$\smash{\scriptscriptstyle-}$}}{\theta }}}_{n}})\to D(\nu \theta )\text{ }\ in\ {{L}^{2}}(I;{{L}^{2}}{{(\Omega )}^{N}}),\text{and}\ \text{hence} \\ [1ex]({{\overline{\mathit{\pmb{v}}}}_{n}}-\varpi )\cdot [\nabla \beta ]({{\overline{\mathit{\pmb{v}}}}_{n}})D({{\nu }_{n}}{{{\underset{\raise0.3em\hbox{$\smash{\scriptscriptstyle-}$}}{\theta }}}_{n}}) \\ \ \ \ \ \ \to (\mathit{\pmb{v}}-\varpi )\cdot [\nabla \beta ](\mathit{\pmb{v}})D(\nu \theta )\text{ }in\ {{L}^{2}}(I;{{L}^{2}}{{(\Omega )}^{N}}), \\ \text{for}\ \text{any}\ \varpi \in {{[{{H}^{1}}(\Omega )\cap {{L}^{\infty }}(\Omega )]}^{2}},as\ n\to \infty . \\ \end{array} \right. (6.14)

    With (A2)-(A5), (\sharp 1)-(\sharp 3), (6.4)-(6.5), (6.7), (6.13)-(6.14) and lower semi-continuity of L^2-norm in mind, letting n \to \infty in (6.9) yields that:

    \begin{equation}\label{14-21} \begin{array}{l} \int_{I}(\mathit{\pmb{v}}_t(t), \mathit{\pmb{v}}(t)-{\varpi})_{L^{2}(\Omega)^{2}} dt + \int_{I} (D\mathit{\pmb{v}}(t), D(\mathit{\pmb{v}}(t)-{\varpi}))_{L^{2}(\Omega)^{N \times 2}} dt \\ \qquad + \int_{I}\int_{\Omega} \gamma(w(t)) dxdt + \int_{I} ([\nabla G](u(t);\mathit{\pmb{v}}(t)), \mathit{\pmb{v}}(t)-{\varpi})_{L^{2}(\Omega)^{2}} dt \\ \qquad + \int_{I}\int_{\Omega} d[(\mathit{\pmb{v}}(t)-{\varpi}) \cdot [\nabla\alpha](\mathit{\pmb{v}}(t))|D\theta(t)|] dt \\ \qquad + \int_{I}\int_{\Omega} [\nabla\beta](\mathit{\pmb{v}}(t)) \cdot (\mathit{\pmb{v}}(t)-{\varpi}) |\nabla(\nu \theta)|^{2} dxdt \\ \le \int_{I}\int_{\Omega} \gamma(\varphi) dxdt , \mbox{ for any $ {\varpi} = [\varphi,\psi] \in [H^{1}(\Omega) \cap L^{\infty}(\Omega)]^{2}$.} \end{array} \end{equation} (6.15)

    Finally, taking the limit of (6.8), and applying (6.5)-(6.7), one can see that:

    \begin{equation}\label{14-22} \begin{array}{c} \int_{I} \langle {u}_t(t), z \rangle \, dt + \int_{I} ({u}(t), z)_{V} \, dt = \int_I (\lambda'({w}(t)) {w}_t(t), z )_{L^2(\Omega)} \, dt \\[2ex] +\int_{I} \langle {\mathit{\pmb{f}}}^*(t), z \rangle \, dt, \mbox{ for any $z \in V$.} \end{array} \end{equation} (6.16)

    Since the open interval I \subset (0, T) is arbitrary, the conditions (S4)-(S5) will be verified by taking into account (6.4) and (6.15)-(6.16).


    Acknowledgments

    This research was supported by JSPS KAKENHI Grant-in-Aid for Scientific Research (C), 16K05224, No. 26400138 and Young Scientists (B), No. 25800086. The authors express their gratitude to an anonymous referees for reviewing the original manuscript and for many valuable comments that helped clarify and refine this paper.


    Conflict of Interest

    All authors declare no conflicts of interest in this paper.


    [1] M. Amar, G. Bellettini, A notion of total variation depending on a metric with discontinuous coeffcients. Ann. Inst. H. Poincaré Anal. Non Linéaire, 11 (1994), no. 1, 91-133.
    [2] L. Ambrosio, N. Fusco, D. Pallara, Functions of Bounded Variation and Free Discontinuity Problems. Oxford Mathematical Monographs. (2000).
    [3] H. Attouch, G. Buttazzo, G. Michaille, Variational Analysis in Sobolev and BV Spaces. Applications to PDEs and Optimization. MPS-SIAM Series on Optimization, 6. SIAM and MPS, (2006).
    [4] V. Barbu, Nonlinear semigroups and differential equations in Banach spaces. Editura Academiei Republicii Socialiste România, Noordhoff International Publishing, (1976).
    [5] G. Bellettini, G. Bouchitté, I. Fragalà, BV functions with respect to a measure and relaxation of metric integral functionals. J. Convex Anal., 6 (1999), no. 2, 349-366.
    [6] H. Brézis, Operateurs Maximaux Monotones et Semi-groupes de Contractions dans les Espaces de Hilbert. North-Holland Mathematics Studies, 5. Notas de Matemática (50). North-Holland Publishing and American Elsevier Publishing, (1973).
    [7] P. Colli, P. Laurençot, Weak solutions to the Penrose-Fife phase field model for a class of admissible heat flux laws. Phys. D, 111 (1998), 311-334.
    [8] P. Colli, J. Sprekels, Glob al solution to the Penrose-Fife phase-field model with zero interfacial energy and Fourier law. Adv. Math. Sci. Appl., 9 (1999), no. 1, 383-391.
    [9] G. Dal Maso, An Introduction to Γ-convergence. Progress in Nonlinear Differential Equations and their Applications, 8. Birkhäuser Boston, Inc., Boston, Ma, (1993).
    [10] I. Ekeland, R. Temam, Convex analysis and variational problems. Translated from the French. Corrected reprint of the 1976 English edition. Classics in Applied Mathematics, 28. SIAM, Philadelphia, (1999).
    [11] L. C. Evans, R. F. Gariepy, Measure Theory and Fine Properties of Functions. Studies in Advanced Mathematics. CRC Press, Boca Raton, (1992).
    [12] E. Giusti, Minimal Surfaces and Functions of Bounded Variation. Monographs in Mathematics, 80. Birkhäuser, (1984).
    [13] M.-H. Giga, Y. Giga, Very singular diffusion equations: second and fourth order problems. Jpn. J. Ind. Appl. Math., 27 (2010), no. 3, 323-345.
    [14] W. Horn, J. Sprekels, S. Zheng, Global existence of smooth solutions to the Penrose-Fife model for Ising ferromagnets. Adv. Math. Sci. Appl., 6 (1996), no. 1, 227-241.
    [15] A. Ito, N. Kenmochi, N. Yamazaki, A phase-field model of grain boundary motion. Appl. Math., 53 (2008), no. 5, 433-454.
    [16] A. Ito, N. Kenmochi, N. Yamazaki, Weak solutions of grain boundary motion model with singularity. Rend. Mat. Appl. (7), 29 (2009), no. 1, 51-63.
    [17] A. Ito, N.Kenmochi, N. Yamazaki, Global solvability of a model for grain boundary motion with constraint. Discrete Contin. Dyn. Syst. Ser. S, 5 (2012), no. 1, 127-146.
    [18] N. Kenmochi, : Systems of nonlinear PDEs arising from dynamical phase transitions. In: Phase transitions and hysteresis (Montecatini Terme, 1993), pp. 39-86, Lecture Notes in Math., 1584, Springer, Berlin, (1994).
    [19] N. Kenmochi, M. Kubo, Weak solutions of nonlinear systems for non-isothermal phase transitions. Adv. Math. Sci. Appl., 9 (1999), no. 1, 499-521.
    [20] N. Kenmochi, N. Yamazaki, Large-time behavior of solutions to a phase-field model of grain boundary motion with constraint. In: Current advances in nonlinear analysis and related topics, pp. 389-403, GAKUTO Internat. Ser. Math. Sci. Appl., 32, Gakkōtosho, Tokyo, (2010).
    [21] R. Kobayashi, Y. Giga, Equations with singular diffusivity. J. Statist. Phys., 95 (1999), 1187-1220.
    [22] R. Kobayashi, J. A. Warren, W. C. Carter, A continuum model of grain boundary. Phys. D, 140 (2000), no. 1-2, 141-150.
    [23] R. Kobayashi, J. A.Warren, W. C. Carter, Grain boundary model and singular diffusivity. In: Free Boundary Problems: Theory and Applications, pp. 283-294, GAKUTO Internat. Ser. Math. Sci. Appl., 14, Gakkōtosho, Tokyo, (2000).
    [24] J. L. Lions, E. Magenes, Non-Homogeneous Boundary Value Problems and Applications. Vol I. Springer-Verlag, New York-Heidelberg, (1972).
    [25] S. Moll, K. Shirakawa, Existence of solutions to the Kobayashi-Warren-Carter system. Calc. Var. Partial Differential Equations, 51 (2014), 621-656. DOI:10.1007/s00526-013-0689-2 doi: 10.1007/s00526-013-0689-2
    [26] S. Moll, K. Shirakawa, H.Watanabe, Energy dissipative solutions to the Kobayashi-Warren-Carter system. submitted.
    [27] U. Mosco, Convergence of convex sets and of solutions of variational inequalities. Advances in Math., 3 (1969), 510-585.
    [28] K. Shirakawa, H. Watanabe, Energy-dissipative solution to a one-dimensional phase field model of grain boundary motion. Discrete Contin. Dyn. Syst. Ser. S, 7 (2014), no. 1, 139-159. DOI:10.3934/dcdss.2014.7.139 doi: 10.3934/dcdss.2014.7.139
    [29] K. Shirakawa, H. Watanabe, Large-time behavior of a PDE model of isothermal grain boundary motion with a constraint. Discrete Contin. Dyn. Syst. 2015, Dynamical systems, differential equations and applications. 10th AIMS Conference. Suppl., 1009-1018.
    [30] K. Shirakawa, H. Watanabe, N. Yamazaki, Solvability of one-dimensional phase field systems associated with grain boundary motion. Math. Ann., 356 (2013), 301-330. DOI:10.1007/s00208-012-0849-2 doi: 10.1007/s00208-012-0849-2
    [31] K. Shirakawa, H. Watanabe, N. Yamazaki, Phase-field systems for grain boundary motions under isothermal solidifications. Adv. Math. Sci. Appl., 24 (2014), 353-400.
    [32] K. Shirakawa, H. Watanabe, N. Yamazaki, Mathematical analysis for a Warren-Kobayashi-Lobkovsky-Carter type system. RIMS Kôkyûroku, 1997 (2016), 64-85.
    [33] J. Simon, Compact sets in the space Lp(0; T; B). Ann. Mat. Pura Appl. (4), 146 (1987), 65-96.
    [34] J. Sprekels, S. Zheng, Global existence and asymptotic behaviour for a nonlocal phase-field model for non-isothermal phase transitions. J. Math. Anal. Appl., 279 (2003), 97-110.
    [35] A. Visintin, Models of phase transitions. Progress in Nonlinear Differential Equations and their Applications, 28, Birkhäuser, Boston, (1996).
    [36] J. A. Warren, R. Kobayashi, A. E. Lobkovsky, W. C. Carter, Extending phase field models of solidification to polycrystalline materials. Acta Materialia, 51 (2003), 6035-6058.
    [37] H.Watanabe, K. Shirakawa, Qualitative properties of a one-dimensional phase-field system associated with grain boundary. In: Current Advances in Applied Nonlinear Analysis and Mathematical Modelling Issues, pp. 301-328, GAKUTO Internat. Ser. Math. Sci. Appl., 36, Gakkōtosho, Tokyo, (2013).
    [38] James A.Warren, Ryo Kobayashi, Alexander E. Lobkovsky, W. Craig Carter, Extending phase field models of solidification to polycrystalline materials. Acta Materialia, 51 (2003), 60356058.
    [39] H. Watanabe, K. Shirakawa, Stability for approximation methods of the one-dimensional Kobayashi-Warren-Carter system. Mathematica Bohemica, 139 (2014), special issue dedicated to Equadiff 13, no. 2, 381-389.
    [40] N. Yamazaki, Global attractors for non-autonomous phase-field systems of grain boundary motion with constraint. Adv. Math. Sci. Appl. 23 (2013), no. 1, 267-296.
  • This article has been cited by:

    1. Hiroshi Watanabe, Ken Shirakawa, Energy‐dissipation in a coupled system of Allen–Cahn‐type equation and Kobayashi–Warren–Carter‐type model of grain boundary motion, 2020, 43, 0170-4214, 10138, 10.1002/mma.6684
  • Reader Comments
  • © 2017 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(5051) PDF downloads(859) Cited by(1)

Article outline

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog