Citation: Luca Altamore, Marzia Ingrassia, Stefania Chironi, Pietro Columba, Giuseppe Sortino, Ana Vukadin, Simona Bacarella. Pasta experience: Eating with the five senses—a pilot study[J]. AIMS Agriculture and Food, 2018, 3(4): 493-520. doi: 10.3934/agrfood.2018.4.493
[1] | Chao-Jen Li, Peiwen Li, Kai Wang, Edgar Emir Molina . Survey of Properties of Key Single and Mixture Halide Salts for Potential Application as High Temperature Heat Transfer Fluids for Concentrated Solar Thermal Power Systems. AIMS Energy, 2014, 2(2): 133-157. doi: 10.3934/energy.2014.2.133 |
[2] | Ayman B. Attya, T. Hartkopf . Wind Turbines Support Techniques during Frequency Drops — Energy Utilization Comparison. AIMS Energy, 2014, 2(3): 260-275. doi: 10.3934/energy.2014.3.260 |
[3] | Jialin Song, Haoyi Zhang, Yanming Zhang, Zhongjiao Ma, Mingfei He . Research progress on industrial waste heat recycling and seasonal energy storage. AIMS Energy, 2025, 13(1): 147-187. doi: 10.3934/energy.2025006 |
[4] | Joaquim Azevedo, Jorge Lopes . Energy harvesting from hydroelectric systems for remote sensors. AIMS Energy, 2016, 4(6): 876-893. doi: 10.3934/energy.2016.6.876 |
[5] | Tan Nguyen Tien, Quang Khong Vu, Vinh Nguyen Duy . Novel designs of thermoelectric generator for automotive waste heat recovery: A review. AIMS Energy, 2022, 10(4): 922-942. doi: 10.3934/energy.2022042 |
[6] | Azevedo Joaquim, Mendonça Fábio . Small scale wind energy harvesting with maximum power tracking. AIMS Energy, 2015, 2(3): 297-315. doi: 10.3934/energy.2015.3.297 |
[7] | Joanna McFarlane, Jason Richard Bell, David K. Felde, Robert A. Joseph III, A. Lou Qualls, Samuel Paul Weaver . Performance and Thermal Stability of a Polyaromatic Hydrocarbon in a Simulated Concentrating Solar Power Loop. AIMS Energy, 2014, 2(1): 41-70. doi: 10.3934/energy.2014.1.41 |
[8] | Nathnael Bekele, Wondwossen Bogale . Parametric study of a diffuser for horizontal axis wind turbine power augmentation. AIMS Energy, 2019, 7(6): 841-856. doi: 10.3934/energy.2019.6.841 |
[9] | Obafemi O. Olatunji, Stephen Akinlabi, Nkosinathi Madushele, Paul A. Adedeji, Ishola Felix . Multilayer perceptron artificial neural network for the prediction of heating value of municipal solid waste. AIMS Energy, 2019, 7(6): 944-956. doi: 10.3934/energy.2019.6.944 |
[10] | Daido Fujita, Takahiko Miyazaki . Techno-economic analysis on the balance of plant (BOP) equipment due to switching fuel from natural gas to hydrogen in gas turbine power plants. AIMS Energy, 2024, 12(2): 464-480. doi: 10.3934/energy.2024021 |
Let 0<T<∞ be a constant of time, and let N∈N be a constant of spatial dimension such that 1≤N≤3. Let Ω⊂RN 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 ν≥0, and consider the following system of initial-boundary value problems of parabolic types, denoted by (S)ν.
(S)ν:
{[u−λ(w)]t−Δu=f in Q, Du⋅nΓ+n0(u−fΓ)=0 on Σ, u(0,x)=u0(x), x∈Ω; | (1.1) |
{wt−Δw+∂γ(w)+gw(w,η)+λ′(w)u +αw(w,η)|Dθ|+ν2βw(w,η)|Dθ|2∋0 in Q,Dw⋅nΓ=0 on Σ,w(0,x)=w0(x), x∈Ω; | (1.2) |
{ηt−Δη+gη(w,η)+αη(w,η)|Dθ|+ν2βη(w,η)|Dθ|2=0 in Q,Dη⋅nΓ=0 on Σ,η(0,x)=η0(x), x∈Ω; | (1.3) |
{α0(w,η)θt−div(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), x∈Ω. | (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):=12∫Ω|Dw|2dx+∫Ωγ(w)dx+∫Ωuλ(w)dx+12∫Ω|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 νθ∈H1(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 u−λ(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)∈Q↦η(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 0≤w,η≤1 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]∈R2↦γ(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+(1−w)log(1−w))with γ(0)=γ(1):=1,g(w,η):=−L2(w−12)2+c2(w−η)2, for w,η∈R, |
(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(w−12)2+c2(w−η)2, for w,η∈R. |
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 ν≥0, which behaves in the range of C([0,T];L2(Ω)4), with the L2-based sources f∈L2(0,T;L2(Ω)) and fΓ∈L2(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.
First we elaborate the notations which is used throughout this paper.
Notation 1 (Real analysis). For arbitrary a0, b0 ∈ [-∞, ∞], we define
Fix d∈N as a constant of dimension. Then, we denote by |x| and x⋅y the Euclidean norm of x∈Rd and the standard scalar product of x,y∈Rd, respectively, as usual, i.e.:
|x|:=√x21+⋯+x2d and x⋅y:=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 S⊂Rd, 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 B⊂Rd, 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. x∈B. |
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<d∈N. 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=(d∑k=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:X→X∗, 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 z0∈X, the value ∂Ψ(z0) is defined as the set of all elements z∗0∈X that satisfy the variational inequality
(z∗0,z−z0)X≤Ψ(z)−Ψ(z0) for any z∈D(Ψ), |
and the set D(∂Ψ):={z∈X∣∂Ψ(z)≠∅} is called the domain of ∂Ψ. We often use the notation "[z0,z∗0]∈∂Ψ in X2, " to mean "z∗0∈∂Ψ(z0) in X with z0∈D(∂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:=∫Ω∇w⋅∇zdx+n0∫ΓwzdΓ, for [w,z]∈V2, |
and let CV>0 be the embedding constant of V⊂L2(Ω).
Let ⟨⋅,⋅⟩ be the duality pairing between V and the dual space V∗, and let F: V→V∗ 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∗,F−1z∗⟩, for [w∗,z∗]∈(V∗)2. |
For any ϱ∈L2(Ω) and any ϱΓ∈L2(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)forz∈V. | (2.1) |
Note that for any ϱ∗=[ϱ,ϱΓ]∈L2(Ω)×L2(Γ), the variational form (2.1) enables the following identification:
Fω=ϱ∗inV∗,iff.ω∈H2(Ω)and{−Δω=ϱinL2(Ω)Dω⋅nΓ+n0(ω−ϱΓ)=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:z∈WN:={z∈H2(Ω)Dz⋅nΓ=0 in L2(Γ)}⊂L2(Ω)↦Δz∈L2(Ω). |
Remark 1. We here show some representative examples of the subdifferentials, which is intimately related to our study.
(Ex.1) The quadratic functional u∈L2(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:V→V∗, i.e.:
[u,u∗]∈∂[12|⋅|2L2(Ω)] in [V∗]2, iff. u∈V and u∗=Fu in V∗. |
(Ex.2) Let d∈N, and let γ0:Rd→R be a convex function defined as
y=[y1,…,yd]∈Rd↦γ0(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:
z∈L2(Ω)d↦Ψdγ0(z):={12∫Ω|Dz|2RN×ddx+∫Ωγ0(z)dx,ifz∈H1(Ω)d∞, otherwise. |
Then, with regard to the subdifferential ∂Ψdγ0⊂[L2(Ω)d]2, it is known (see, e.g., [4,6]) that
z∈L2(Ω)d↦∂Ψdγ0(z)={z∗∈L2(Ω)d|z∗+ΔNz∈∂γ0(z)inRd,a.e.inΩ}ifz∈WdN,∅,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 d∈N, and let U⊂Rd 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 C∞c(U) in the topology of C(¯U).
A function z∈L1(U) is called a function of bounded variation on U, iff. its distributional gradient Dz is a finite Radon measure on U, namely, Dz∈M(U)d. Here, for any z∈BV(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 z∈BV(U), it holds that
|Dz|(U)=sup{∫Uzdivφdxφ∈C1c(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 z∈BV(U), |
and we say that zn→z weakly-∗ in BV(U), iff. z∈BV(U), {zn}∞n=1⊂BV(U), zn→z in L1(U) and Dzn→Dz 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↦|φ−ψ|L1(U)+||Dφ|(U)−|Dψ|(U)|. |
In this regard, we say that zn→z strictly in BV(U) iff. z∈BV(U), {zn}∞n=1⊂BV(U), zn→z 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/(d−1)(U) and compactly embedded into Lp(U) for any 1≤p<d/(d−1) (see, e.g., [2, Corollary 3.49] or [3, Theorems 10.1.3-10.1.4]). Furthermore, if 1≤q<∞, then the space C∞(¯U) is dense in BV(U)∩Lq(U) for the intermediate convergence, i.e., for any z∈BV(U)∩Lq(U), there exists a sequence {zn}∞n=1⊂C∞(¯U) such that zn→z 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 ϱ∈H1(Ω)∩L∞(Ω) (i.e. any 0≤ϱ∈H1(Ω)∩L∞(Ω)) and any z∈L2(Ω), we call the value Varϱ(z)∈[0,∞], defined as,
Varϱ(v):=sup[∫Ωvdivϖdx|ϖ∈L∞(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 0≤ϱ∈H1(Ω)∩L∞(Ω), the functional z∈L2(Ω)↦Varϱ(z)∈[0,∞] is a proper l.s.c. and convex function that coincides with the lower semi-continuous envelope of
z∈W1,1(Ω)∩L2(Ω)↦∫Ωϱ|Dz|dx∈[0,∞). |
(Fact 2) (Cf. [1, Theorem 4.3] and [2, Proposition 5.48]) If 0≤ϱ∈H1(Ω)∩L∞(Ω) and z∈BV(Ω)∩L2(Ω), then there exists a Radon measure |Dz|ϱ∈M(Ω) such that
|Dz|ϱ(Ω)=∫Ωd|Dz|ϱ=Varϱ(z), |
and
{|Dz|ϱ(A)≤|ϱ|L∞(Ω)|Dz|(A)|Dz|ϱ(A)=inf{liminfn→∞∫Aϱ|D˜zn|dx|{˜zn}∞n=1⊂W1.1(A)∩L2(A)suchthat˜zn→zinL2(A)asn→∞} | (2.2) |
for any open set A⊂Ω.
(Fact 3) If ϱ∈H1(Ω)∩L∞(Ω), cϱ:=essinfx∈Ωϱ>0, and z∈BV(Ω)∩L2(Ω), then for any open set A⊂Ω, it follows that
{|Dz|ϱ(A)≥cϱ|Dz|(A)foranyopensetA⊂ΩD(Varϱ)=BV(Ω)∩L2(Ω),andVarϱ(z)=sup{∫Ωzdiv(ϱϖ)dx|ϖ∈L∞(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 c≥0 and z∈BV(Ω)∩L2(Ω);
· |Dz|ϱ=ϱ|Dz|LN in M(Ω), if 0≤ϱ∈H1(Ω)∩L∞(Ω) and z∈W1,1(Ω)∩L2(Ω).
Notation 6 (Generalized weighted total variation; cf. [25, Section 2]). For any linebreak ϱ∈H1(Ω)∩L∞(Ω) and any z∈BV(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 B⊂Ω. |
Note that [ϱ|Dz|](Ω) can be configured as a generalized total variation of z∈BV(Ω)∩L2(Ω) by the possibly sign-changing weight ϱ∈H1(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 ϱ∈H1(Ω)∩L∞(Ω) and z∈BV(Ω)∩L2(Ω) be arbitrary fixed functions, and let {zn}∞n=1⊂C∞(¯Ω) be a sequence such that
zn→z in L2(Ω) and strictly in BV(Ω) as n→∞. |
Then
∫Ωϱ|Dzn|dx→∫Ωd[ϱ|Dz|] as n→∞. |
(Fact 5) For any z∈BV(Ω)∩L2(Ω), the mapping
ϱ∈H1(Ω)∩L∞(Ω)↦∫Ωd[ϱ|Dz|]∈R |
is a linear functional, and moreover, if φ∈H1(Ω)∩C(¯Ω) and ϱ∈H1(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 Ψn→Ψ on X, in the sense of Mosco, as n→∞, iff. the following two conditions are fulfilled.
The condition of lower bound: lim infn→∞Ψn(z∘n)≥Ψ(z∘), if z∘∈X, {z∘n}∞n=1⊂X, and z∘n→z∘ weakly in X as n→∞.
The condition of optimality: for any z∙∈D(Ψ), there exists a sequence {z∙n}∞n=1⊂X such that z∙n→z∙ in X and Ψn(z∙n)→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 Ψn→Ψ on X, in the sense of Γ-convergence, as n→∞, iff. the following two conditions are fulfilled.
The condition of lower bound: lim infn→∞Ψn(z∘n)≥Ψ(z∘), if z∘∈X, {z∘n}∞n=1⊂X, and z∘n→z∘ (strongly) in X as n→∞.
The condition of optimality: for any z∙∈D(Ψ), there exists a sequence {z∙n}∞n=1⊂X such that z∙n→z∙ in X and Ψn(z∙n)→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 Ψn→Ψ on X, in the sense of Γ-convergence, as n→∞. If it holds that:
{∈X2,[zn,z∗n]∈∂ΨninX2,n=1,2,3,…,zn→zinXandz∗n→z∗weaklyinX,asn→∞ |
then [z,z∗]∈∂Ψ in X2 and Ψn(zn)→Ψ(z) as n→∞.
Throughout the paper, we set the following assumptions.
(A1) Let f∈L2(0,T;L2(Ω)) and fΓ∈L2(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 f∗∈L2(0,T;V∗), via (2.1) applied to ϱ∗=f∗(t) for a.e. t>0.
(A2) Let λ∈W2,∞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 V⊂L2(Ω).
(A3) Let α0∈W1,∞loc(R2) and α,β∈C2(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 g∈C2(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]|˜u∈L2(Ω),˜w,˜η∈H1(Ω),˜θ∈BV(Ω)∩L∞(Ω),and0⩽˜w,˜η⩽1a.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 u∈R and v=[w,η]∈R2. |
For any ν≥0 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:
θ∈L2(Ω)↦Φν(v;θ)=Φν(w,η;θ):={∫Ωd[α(v)|Dθ|]+∫Ωβ(v)|D(νθ)|2dx,ifθ∈BV(Ω)andνθ∈H1(Ω),∞, otherwise. |
Additionally, we set:
B∗:=1+A∗2,byusingtheconstantA∗asin(A2), | (3.1) |
and define a functional Fν on L2(Ω)4 by letting:
\begin{array}{*{20}{l}} {[u,\mathit{\boldsymbol{v}},\theta ] = [u,w,\eta ,\theta ] \in {L^2}{{(\Omega )}^4} \mapsto \mathscr{F}(u,\mathit{\boldsymbol{v}},\theta ) = \mathscr{F}(u,w,\eta ,\theta )}\\ {: = {B_*}|u|_{{L^2}(\Omega )}^2 + \Psi _\gamma ^2(\mathit{\boldsymbol{v}}) + \int_\Omega {(g(} v) - {c_*})dx + {\Phi _\nu }(\mathit{\boldsymbol{v}};\theta ),} \end{array} | (3.2) |
where \Psi_{\gamma}^{2} is the convex function \Psi_{\gamma_{0}}^{d} in Remark 1 in the case when d=2 and \gamma_{0}=\gamma. The above functional \mathscr{F}_\nu 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. \mathscr{F}_\nu \geq 0 on L^2(\Omega)^4.
Based on these, we define the solutions to the systems (S)_\nu, for \nu \geq 0 , as follows.
Definition 3. For any \nu \geq 0 , a quartet [u, v, \theta]=[u, w, \eta, \theta] \in L^2(0, T; L^2(\Omega)^4) with v=[w, \eta] is called a solution to (S)_{\nu}, iff. [u, v, \theta] fulfills the following (S1)-(S6).
(S1) u \in W^{1, 2}(0, T; V^{\ast}) \cap L^{\infty}(0, T; L^{2}(\Omega)) \cap L^{2}(0, T; V) \subset C ([0, T]; L^2(\Omega)) .
(S2) v=[w, \eta] \in W^{1, 2}(0, T; L^2(\Omega)^{2}) \cap L^\infty (0, T; H^1(\Omega)^{2}) , and 0 \leq w (t, x) \leq 1 and 0 \leq \eta (t, x) \leq 1 , a.e. (t, x) \in Q .
(S3) \theta \in W^{1, 2}(0, T; L^2(\Omega)) \cap L^\infty (Q) , |D \theta (\, \cdot\, )|(\Omega) \in L^\infty (0, T) , \nu \theta \in L^\infty (0, T; H^1(\Omega)), and |\theta| \le |\theta_{0}|_{L^{\infty}(\Omega)} a.e. in Q.
(S4) u satisfies the following variational form:
\begin{array}{c} \langle [u-\lambda(w)]_{t}(t), z \rangle +(D u(t), D z)_{L^2(\Omega)^N} + n_{0} ( u(t), z)_{L^2(\Gamma)} \\ = (f(t), z)_{L^2(\Omega)} +n_0(f_\Gamma(t), z)_{L^2(\Gamma)}, \mbox{ for any $z \in V$, and a.e. $ t \in (0, T) $, } \end{array} |
with the initial condition u (0)=u_{0} in L^{2}(\Omega).
(S5) v=[w, \eta] satisfies the following two variational forms:
\begin{array}{l} \left( w_t(t) +g_w(v)(t) +u(t)\lambda'(w(t)), w(t) - \varphi \right)_{L^{2}(\Omega)} + (D w(t), D(w(t)-\varphi))_{L^{2}(\Omega)^{N}} \\[2ex] \qquad + \int_\Omega d[(w(t)-\varphi)\alpha_{w}(v(t)) |D \theta(t)|] +\int_\Omega (w(t) -\varphi) \beta_w(v(t)) |D(\nu \theta)(t)|^2 \, dx \\[2ex] \qquad +\int_\Omega \gamma(w(t)) \, dx \le \int_{\Omega}\gamma(\varphi) dx, \mbox{ for any $ \varphi \in H^1(\Omega) \cap L^{\infty}(\Omega) $ and a.e. $ t \in (0, T) $, } \end{array} |
and
\begin{array}{l} \left( \eta_t(t) +g_\eta(v)(t), \psi \right)_{L^{2}(\Omega)} + (D \eta(t), D \psi)_{L^{2}(\Omega)^{N}} \\[1ex] \qquad +\int_\Omega d \bigl[\psi \alpha_{\eta}(v(t)) |D \theta(t)| \bigr] +\int_\Omega \psi \beta_\eta(v(t)) |D (\nu \theta)(t)|^2 \, dx = 0, \\[2ex] \qquad \mbox{for any $ \psi \in H^1(\Omega) \cap L^{\infty}(\Omega) $ and a.e. $ t \in (0, T) $, } \end{array} |
with the initial condition v (0)=[w (0), \eta (0)]=v_0=[w_0, \eta_{0}] in L^{2}(\Omega)^2.
(S6) \theta satisfies the following variational form:
\begin{array}{c} \displaystyle (\alpha_0(v(t)) \theta_t(t), \theta(t) - \omega)_{L^2(\Omega)} + \Phi_{\nu}(v(t); \theta(t)) \le \Phi_{\nu}(v(t); \omega), \\[1ex] \displaystyle \mbox{for any $ \omega \in D(\Phi_{\nu}(v(t);{}\cdot\, )) $ and a.e. $ t \in (0, T) $, } \end{array} |
with the initial condition \theta (0)=\theta_{0} in L^{2}(\Omega).
Remark 5. The variational identity in the above (S4) can be reformulated as:
{[u - \lambda (w)]_t}(t) + Fu(t) = {\rm{ }}{\mathit{\boldsymbol{f}}^ * }(t)\;\;{\rm{ in }}{V^ * },\;{\rm{ }}for a.e.{\rm{ }}t \in (0,T). | (3.3) |
Also, two variational forms in (S5) can be reduced to:
\begin{array}{l} {({\mathit{\boldsymbol{v}}_t}(t) + [\nabla G](u;\mathit{\boldsymbol{v}}(t)),\mathit{\boldsymbol{v}}(t) - \varpi )_{{L^2}{{(\Omega )}^2}}}\\ + {(D\mathit{\boldsymbol{v}}(t),D(\mathit{\boldsymbol{v}}(t) - \varpi ))_{{L^2}{{(\Omega )}^{N \times 2}}}}\\ + \int_\Omega d [|D\theta (t)|(\mathit{\boldsymbol{v}}(t) - \varpi ) \cdot [\nabla \alpha ](\mathit{\boldsymbol{v}}(t))]\\ + \int_\Omega | D(\nu \theta )(t){|^2}(\mathit{\boldsymbol{v}}(t) - \varpi ) \cdot [\nabla \beta ](\mathit{\boldsymbol{v}}(t)){\mkern 1mu} dx\\ + \int_\Omega \gamma (\mathit{\boldsymbol{v}}(t)){\mkern 1mu} dx \le \int_\Omega \gamma (\varpi ){\mkern 1mu} dx,\\ for\;any\;\varpi = [\varphi ,\psi ] \in {[{H^1}(\Omega ) \cap {L^\infty }(\Omega )]^2}\;and\;a.e.t \in \left( {0,T} \right), \end{array} | (3.4) |
by using the identification
\gamma(\tilde{v}) := \gamma(\tilde{w}), \ \ \mbox{ for all } \tilde{v}=[\tilde{w}, \tilde{\eta}] \in \mathbb{R}^{2}, |
and by using the abbreviation:
\begin{array}{l} \int_\Omega d [|D\tilde \theta |\varpi \cdot \mathit{\boldsymbol{\widetilde v}}]: = \int_\Omega d [\varphi \tilde w|D\tilde \theta |] + \int_\Omega d [\psi \tilde \eta |D\tilde \theta |],\\ \mathit{for}\;\mathit{\boldsymbol{\widetilde v}} = [\tilde w,\tilde \eta ],{\rm{ }}\varpi = [\varphi ,\psi ] \in {[{H^1}(\Omega ) \cap {L^\infty }(\Omega )]^2}\;and\;\tilde \theta \in BV(\Omega ) \cap {L^2}(\Omega ) \end{array} | (3.5) |
Furthermore, the variational form in (S6) is equivalent to the following evolution equation:
{\alpha _0}(\mathit{\boldsymbol{v}}(t)){\theta _t}(t) + \partial {\Phi _\nu }(\mathit{\boldsymbol{v}}(t);\theta (t)) \ni 0{\rm{ }}\;in\;{L^2}(\Omega ),a.e.t \in (0,T), | (3.6) |
governed by the subdifferential \partial \Phi_{\nu}(v (t); {}\cdot\, ) \subset L^2(\Omega)^2 of the time-dependent convex function \Phi_{\nu}(v (t); {}\cdot\, ) , for t \in (0, T) .
Now, our Main Theorem is stated as follows.
Main Theorem Let \nu \ge 0 be a fixed constant. Then, under (A1)-(A7), the system (S)_{\nu} admits at least one solution [u, v, \theta]=[u, w, \eta, \theta] \in L^2(0, T; L^2(\Omega)^4) with v=[w, \eta] .
Remark 6. Note that the presence of mobilities \alpha_0=\alpha_0(w, \eta) , \alpha=\alpha (w, \eta) and \beta=\beta (w, \eta) makes the uniqueness problems for the systems (S)_\nu, \nu \geq 0 , 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 \nu . Hence, when we consider the approximating problems, we suppose \nu > 0 , and fix the constant of time-step h \in (0, 1] . Also, we denote by [f]_0^{\rm ex} \in L^2(\mathbb{R}; L^2(\Omega)) , [f_\Gamma]_0^{\rm ex} \in L^2(\mathbb{R}; L^2(\Gamma)) and [{f}^*]_0^{\rm ex} \in L^2(\mathbb{R}; V^*) the zero-extensions of f , f_\Gamma and {f}^* (=[f, f_\Gamma]) , respectively.
On this basis, the approximating problem for our system (S)_\nu is denoted by (AP)_h^\nu, and stated as follows.
(AP)_h^\nu: to find a sequence \{ [u_i^\nu, v_i^\nu, \theta_i^\nu] \}_{i=1}^\infty \subset D_1 with \{ v_i^\nu \}_{i=1}^\infty=\{ [w_i^\nu, \eta_i^\nu] \}_{i=1}^\infty , which fulfills that
\frac{{u_i^\nu - u_{i - 1}^\nu }}{h} - \lambda '(w_i^\nu )\frac{{w_i^\nu - w_{i - 1}^\nu }}{h} + Fu_i^\nu = {[\mathit{\boldsymbol{f}}_i^*]^h}{\rm{ }}\;in\;{V^*}, | (3.7) |
\begin{array}{l} \frac{1}{h}{(\mathit{\boldsymbol{v}}_i^\nu - \mathit{\boldsymbol{v}}_{i - 1}^\nu ,\mathit{\boldsymbol{v}}_i^\nu - \varpi )_{{L^2}{{(\Omega )}^2}}} + {(D\mathit{\boldsymbol{v}}_i^\nu ,D(\mathit{\boldsymbol{v}}_i^\nu - \varpi ))_{{L^2}{{(\Omega )}^{N \times 2}}}}\\ + {([\nabla G](u_i^\nu ;\mathit{\boldsymbol{v}}_i^\nu ),\mathit{\boldsymbol{v}}_i^\nu - \varpi )_{{L^2}{{(\Omega )}^2}}} + \int_\Omega \gamma (\mathit{\boldsymbol{v}}_i^\nu ){\mkern 1mu} dx\\ + \int_\Omega {(\mathit{\boldsymbol{v}}_i^\nu - \varpi )} \cdot (|D\theta _{i - 1}^\nu |[\nabla \alpha ](\mathit{\boldsymbol{v}}_i^\nu ) + {\nu ^2}|D\theta _{i - 1}^\nu {|^2}[\nabla \beta ](\mathit{\boldsymbol{v}}_i^\nu )){\mkern 1mu} dx \end{array} | (3.8) |
\begin{array}{l} \le \int_\Omega \gamma (\varpi ){\mkern 1mu} dx,{\rm{ for}}\;{\rm{any}}\;\varpi \in {[{H^1}(\Omega ) \cap {L^\infty }(\Omega )]^2},\\ {\alpha _0}(\mathit{\boldsymbol{v}}_i^\nu )\frac{{\theta _i^\nu - \theta _{i - 1}^\nu }}{h} + \partial {\Phi _\nu }(\mathit{\boldsymbol{v}}_i^\nu ;\theta _i^\nu ) \ni 0{\rm{ in}}\;{L^2}(\Omega ), \end{array} | (3.9) |
for i=1, 2, 3, \dots , starting from the initial data:
[u_0^\nu, v_0^\nu, \theta_0^\nu] \in D_1 \mbox{ with } v_0^\nu = [w_0^\nu, \eta_0^\nu]. |
In the context, for any i \in \mathbb{N} , [{f}_i^*]^h=[f_i^h, f_{\Gamma, i}^h] \in L^2(\Omega) \times L^2(\Gamma) (\subset V^*) , consists of the components:
f_i^h := \frac{1}{h} \int_{(i -1)h}^{ih} [f]_0^{\rm ex}(\tau) \, d\tau \mbox{ in $ L^2(\Omega) $ and } f_{\Gamma, i}^h := \frac{1}{h} \int_{(i -1)h}^{ih} [f_\Gamma]_0^{\rm ex}(\tau) \, d\tau \mbox{ in $ L^2(\Gamma) $.} |
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 h_1^\circ \in (0, 1] such that if \nu > 0 and h \in (0, h_1^\circ] , then the approximating problem (AP)^{\nu}_{h} admits a unique solution \{ [u_{i}^{\nu}, v_{i}^{\nu}, \theta_{i}^{\nu}] \}_{i=1}^\infty \subset D_1 , and moreover,
\begin{array}{l} \frac{{{A_*}}}{{2h}}|u_i^\nu - u_{i - 1}^\nu |_{{V^*}}^2 + \frac{1}{{2h}}|\mathit{\boldsymbol{v}}_i^\nu - \mathit{\boldsymbol{v}}_{i - 1}^\nu |_{{L^2}{{(\Omega )}^2}}^2 + \frac{1}{h}|\sqrt {{\alpha _0}(\mathit{\boldsymbol{v}}_i^\nu )} (\theta _i^\nu - \theta _{i - 1}^\nu )|_{{L^2}(\Omega )}^2 + \frac{h}{2}|u_i^\nu |_V^2\\ + \mathscr{F}(u_i^\nu ,\mathit{\boldsymbol{v}}_i^\nu ,\theta _i^\nu ) \le \mathscr{F}(u_{i - 1}^\nu ,\mathit{\boldsymbol{v}}_{i - 1}^\nu ,\theta _{i - 1}^\nu ) + h|{[\mathit{\boldsymbol{f}}_i^*]^h}|_{{V^*}}^2,for\;i = 1,2,3, \ldots \end{array} | (3.10) |
where A_* is the constant as in (A2).
However, due to the presence of L^1-terms \nu^2 |D \theta_{i -1}|^2 [\nabla \beta](v_i^\nu) \in L^1(\Omega)^2 , i=1, 2, 3, \dots , 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)_h^\nu.
In the observation of the relaxed system, we first fix a large constant M > (N +2)/2 , and fix a small constant \varepsilon \in (0, 1] as the relaxation index. Besides, we define
D_M := D_1 \cap [L^2(\Omega) \times H^1(\Omega) \times H^1(\Omega) \times H^M(\Omega)], |
and for any \tilde{v} \in L^2(\Omega)^2 , we define a relaxed functional \Phi_\varepsilon^\nu (\tilde{v}; {}\cdot\, ) for \Phi_\nu (\tilde{v}; {}\cdot\, ) , by letting:
\theta \in {L^2}(\Omega ) \mapsto \Phi _\varepsilon ^\nu (\widetilde {\text{v}};\theta ): = \left\{ {\begin{array}{*{20}{l}} {{\Phi _\nu }(\widetilde v;\theta ) + \frac{{{\varepsilon ^2}}}{2}|\theta |_{{H^M}(\Omega )}^2, {\text{ }}if\;\theta \in {H^M}(\Omega ), }&{} \\ {\infty, }&{{\text{otherwise}}{\text{.}}} \end{array}} \right. |
Note that for any \tilde{v} \in L^2(\Omega)^2 , the functional \Phi_\varepsilon^\nu (\tilde{v}; {}\cdot\, ) is proper l.s.c. and convex on L^2(\Omega) , such that:
D(\Phi_{\varepsilon}^\nu(\tilde{v};{}\cdot\, )) = H^{M}(\Omega) \subset W^{1, \infty}(\Omega), |
and hence, the L^2-subdifferential \partial \Phi_\varepsilon^\nu (\tilde{v}; {}\cdot\, ) is a maximal monotone graph in L^2(\Omega)^2 .
On this basis, we denote by (RX)_\varepsilon the relaxed system for (AP)_h^\nu, and prescribe the system (RX)_\varepsilon as follows.
(RX)_\varepsilon:to find a sequence \{ [u_{\varepsilon, i}^\nu, v_{\varepsilon, i}^\nu, \theta_{\varepsilon, i}^\nu] \}_{i=1}^\infty \subset D_M with \{ v_{\varepsilon, i}^\nu \}_{i=1}^\infty=\{ [w_{\varepsilon, i}^\nu, \eta_{\varepsilon, i}^\nu] \}_{i=1}^\infty , which fulfills that
\frac{{u_{\varepsilon ,i}^\nu - u_{\varepsilon ,i - 1}^\nu }}{h} - \lambda '(w_{\varepsilon ,i}^\nu )\frac{{w_{\varepsilon ,i}^\nu - w_{\varepsilon ,i - 1}^\nu }}{h} + Fu_{\varepsilon ,i}^\nu = {[\mathit{\boldsymbol{f}}_i^*]^h}{\rm{ }}\;{\rm{in}}\;{V^*}, | (3.11) |
\begin{array}{l} \frac{{\mathit{\boldsymbol{v}}_{\varepsilon ,i}^\nu - \mathit{\boldsymbol{v}}_{\varepsilon ,i - 1}^\nu }}{h} - {\Delta _N}\mathit{\boldsymbol{v}}_{\varepsilon ,i}^\nu + \partial \gamma (\mathit{\boldsymbol{v}}_{\varepsilon ,i}^\nu ) + [\nabla G](u_{\varepsilon ,i}^\nu ;\mathit{\boldsymbol{v}}_{\varepsilon ,i}^\nu )\\ + |D\theta _{\varepsilon ,i - 1}^\nu |[\nabla \alpha ](\mathit{\boldsymbol{v}}_{\varepsilon ,i}^\nu ) + {\nu ^2}|D\theta _{\varepsilon ,i - 1}^\nu {|^2}[\nabla \beta ](\mathit{\boldsymbol{v}}_{\varepsilon ,i}^\nu ) \ni 0\;{\rm{in}}\;{L^2}{(\Omega )^2}, \end{array} | (3.12) |
{\alpha _0}(\mathit{\boldsymbol{v}}_{\varepsilon ,i}^\nu )\frac{{\theta _{\varepsilon ,i}^\nu - \theta _{\varepsilon ,i - 1}^\nu }}{h} + \partial \Phi _\varepsilon ^\nu (\mathit{\boldsymbol{v}}_{\varepsilon ,i}^\nu ;\theta _{\varepsilon ,i}^\nu ) \ni 0{\rm{ }}\;{\rm{in}}\;{L^2}(\Omega ), | (3.13) |
for i=1, 2, 3, \dots , starting from the initial data:
[u_{\varepsilon, 0}^\nu, v_{\varepsilon, 0}^\nu, \theta_{\varepsilon, 0}^\nu] \in D_M \mbox{ with } v_{\varepsilon, 0}^\nu = [w_{\varepsilon, 0}^\nu, \eta_{\varepsilon, 0}^\nu]. |
Then, we can see that
|D \theta_{\varepsilon, i -1}^{\nu}| \in L^{\infty}(\Omega)\ \mbox{ and } \nu^2 |D \theta_{\varepsilon, i -1}^{\nu}|^{2}[\nabla\beta](v_{\varepsilon, i}^{\nu}) \in L^{\infty}(\Omega)^{2}, ~ i = 1, 2, 3, \dots. |
It implies that the general theories of L^{2}-subdifferentials will be available for the relaxed system (RX)_{\varepsilon}.
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 h_0^\circ \in (0, 1] , such that if h \in (0, h_0^\circ] , then the system (RX)_\varepsilon admits a unique solution \{ [u_{\varepsilon, i}^\nu, v_{\varepsilon, i}^\nu, \theta_{\varepsilon, i}^\nu] \}_{i=1}^\infty \subset D_M with \{ v_{\varepsilon, i}^\nu \}_{i=1}^\infty= \{ [w_{\varepsilon, i}^\nu, \eta_{\varepsilon, i}]^\nu \}_{i=1}^\infty .
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.
Before we start the proof, we need to show some lemmas.
Lemma 1. Let us put \Delta^\bullet :=[0, 1] \times [-1, 2] \subset \mathbb{R}^2 , and let us assume
0 < h \le h_2^ \circ : = \frac{1}{{2(1 + |g{|_{{C^2}({\Delta ^ \circ })}} + 5|\lambda |_{{W^{2,\infty }}(0,1)}^2)}}. | (4.1) |
Let us fix f_0^* \in V^*, [u_0^\circ, \eta_0^\circ, w_0^\circ, \theta_0^\circ] \in L^2(\Omega) \times H^1(\Omega)\times H^1(\Omega) \times W^{1, \infty}(\Omega) and w^\circ \in H^1(\Omega) , and let us assume that 0 \leq w_0^\circ, w^\circ \leq 1 a.e. in \Omega . Then, the following auxiliary system:
\frac{{u - u_0^ \circ }}{h} - \lambda '({w^ \circ })\frac{{w - w_0^ \circ }}{h} + Fu = f_0^*{\rm{ }}\;in\;{V^*}, | (4.2) |
\begin{array}{l} \frac{{w - w_0^ \circ }}{h} - {\Delta _N}w + \partial \gamma (w) + {g_w}(w,\eta )\\ + {\alpha _w}(w,\eta )|D\theta _0^ \circ | + {\nu ^2}{\beta _w}(w,\eta )|D\theta _0^ \circ {|^2} \ni - \lambda '({w^ \circ })u{\rm{ }}\;in\;{L^2}(\Omega ), \end{array} | (4.3) |
\begin{array}{l} \frac{{\eta - \eta _0^ \circ }}{h} - {\Delta _N}\eta + \partial {I_{[ - 1,2]}}(\eta ) + {g_\eta }(w,\eta )\\ + {\alpha _\eta }(w,\eta )|D\theta _0^ \circ | + {\nu ^2}{\beta _\eta }(w,\eta )|D\theta _0^ \circ {|^2} = 0{\rm{ }}\Delta \;in\;{L^2}(\Omega ), \end{array} | (4.4) |
admits a unique solution [u, w, \eta] \in V \times H^1(\Omega)^2 , where \partial I_{[-1.2]} is the subdifferential of the indicator function I_{[-1.2]} : \mathbb{R} \to \{ 0, \infty \} on the compact interval [-1, 2] , and this is an additional term to guarantee the boundedness of the range \eta (\Omega) for the component \eta.
Proof. First, we put:
\begin{align} & e:=u-{\lambda }'({{w}^{{}^\circ }})w, ~e_{0}^{{}^\circ }:=u_{0}^{{}^\circ }-{\lambda }'({{w}^{{}^\circ }})w_{0}^{{}^\circ }, and\ v_{0}^{{}^\circ }=[w_{0}^{{}^\circ }, \eta _{0}^{{}^\circ }], \\ & [\tilde{w}, \tilde{\eta }]\in \mathbb{R}\mapsto {{\gamma }^{\bullet }}(\tilde{w}, \tilde{\eta }):=\gamma (\tilde{w})+{{I}_{[-1,2]}}(\tilde{\eta }), \\ \end{align} |
and reformulate the system {(4.2)-(4.4)} as follows:
\frac{{e - e_0^ \circ }}{h} + F(e + \lambda '({w^ \circ })w) = f_0^*{\rm{ }}\;in\;{V^*}, | (4.5) |
\begin{array}{l} \frac{{\mathit{\boldsymbol{v}} - \mathit{\boldsymbol{v}}_0^ \circ }}{h} + \partial \Psi _{{\gamma ^ \bullet }}^2(\mathit{\boldsymbol{v}}) + [\nabla g](w,\eta )\\ + |D\theta _0^ \circ |[\nabla \alpha ](\mathit{\boldsymbol{v}}) + {\nu ^2}|D\theta _0^ \circ {|^2}[\nabla \beta ](\mathit{\boldsymbol{v}}) \ni \left[ {\begin{array}{*{20}{c}} { - \lambda '({w^ \circ })(e + \lambda '({w^ \circ })w)}\\ 0 \end{array}} \right]{\rm{ }}\;in\;{L^2}{(\Omega )^2}, \end{array} | (4.6) |
where \Psi_{\gamma^{\bullet}}^2 is the functional \Psi_{\gamma_0}^d as in Remark 1 (Ex.2), in the case when d=2 and \gamma_0=\gamma^\bullet on \mathbb{R}^2 , and \partial \Psi_{\gamma^{\bullet}}^2 is the subdifferential of \Psi_{\gamma^{\bullet}}^2 in L^2(\Omega)^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:
\begin{array}{l} [e,\mathit{\boldsymbol{v}}] = [e,w,\eta ] \in {V^*} \times {L^2}{(\Omega )^2} \mapsto \Psi _0^ \bullet ({w^ \circ };e,\mathit{\boldsymbol{v}}) = \Psi _0^ \bullet ({w^ \circ };e,w,\eta )\\ : = \left\{ \begin{array}{l} \frac{1}{{2h}}|e - e_0^ \circ |_{{V^*}}^2 + \frac{1}{{2h}}|\mathit{\boldsymbol{v}} - \mathit{\boldsymbol{v}}_0^ \circ |_{{L^2}{{(\Omega )}^2}}^2 + \frac{1}{2}|e + \lambda '({w^ \circ })w|_{{L^2}(\Omega )}^2\\ \;\;\;\; + \Psi _{{\gamma ^ \bullet }}^2(\mathit{\boldsymbol{v}}) + \int_\Omega ( \alpha (\mathit{\boldsymbol{v}})|D\theta _0^ \circ | + {\nu ^2}\beta (\mathit{\boldsymbol{v}})|D\theta _0^ \circ {|^2}){\mkern 1mu} dx\\ \;\;\;\; + \int_\Omega g (\mathit{\boldsymbol{v}}){\mkern 1mu} dx - {(f_0^*,e)_{{V^*}}},\\ \;\;\;\;{\rm{if}}\;[e,{\rm{ }}\mathit{\boldsymbol{v}}] = [e,w,\eta ] \in {L^2}(\Omega ) \times D(\Psi _{{\gamma ^ \bullet }}^2),\\ \infty ,{\rm{ otherwise,}} \end{array} \right. \end{array} | (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 C_0^\circ , independent of the variables [e, v]=[e, w, \eta] and w^\circ , such that:
\Psi _0^ \bullet ({w^ \circ };e,{\rm{ }}\mathit{\boldsymbol{v}}) \ge C_0^ \circ (|e|_{{L^2}(\Omega )}^2 + |\mathit{\boldsymbol{v}}|_{{H^1}{{(\Omega )}^2}}^2 - 1),{\rm{ }}\;{\rm{for}}\;{\rm{all}}[e,{\rm{ }}\mathit{\boldsymbol{v}}] \in D(\Psi _0^ \bullet ({w^ \circ }; \cdot {\mkern 1mu} )). | (4.8) |
Now, the above coercivity enables us to apply the standard minimization argument to \Psi_0^\bullet (cf. [3,10]), and to obtain the solution [u, w, \eta]=[e +\lambda'(w^\circ) w, w, \eta] to {(4.2)-(4.4)}, via the minimizer [e, v]=[e, w, \eta] \in V \times H^1(\Omega)^2 of \Psi_0^\bullet (w^\circ; {}\cdot\, ) , with v=[w, \eta] \in D (\Psi_{\gamma^\bullet}^2).
In the meantime, the uniqueness can be seen by using the standard method, i.e. by taking the difference of two solutions [e_k, v_k]=[e_k, w_k, \eta_k] \in V^* \times L^2(\Omega)^2 with v_k=[w_k, \eta_k] \in D (\Psi_\gamma^2) , k=1, 2 , to the stationary system {(4.5)-(4.6)}. In fact, multiplying the both sides of the subtraction of (4.5) by e_1 -e_2 in V^* , multiplying the both sides of the subtraction of (4.6) by v_1 -v_2 in L^2(\Omega)^2 , and using (A2)-(A5), (4.8) and Schwartz's inequality, we have:
\begin{array}{l} \frac{1}{h}|{e_1} - {e_2}|_{{V^*}}^2 + \frac{1}{h}\left( {1 - h|[\nabla g]{|_{{W^{1,\infty }}{{({\Delta ^ \bullet })}^2}}}} \right)|{\mathit{\boldsymbol{v}}_1} - {\mathit{\boldsymbol{v}}_2}|_{{L^2}{{(\Omega )}^2}}^2\\ + |D({\mathit{\boldsymbol{v}}_1} - {\mathit{\boldsymbol{v}}_2})|_{{L^2}{{(\Omega )}^{N \times 2}}}^2 + |({e_1} - {e_2}) + \lambda '({w^ \circ })({w_1} - {w_2})|_{{L^2}(\Omega )}^2 \le 0. \end{array} | (4.9) |
Since the assumption (4.1) implies (1-h%(|[\nabla g]{{|}_{{{W}^{1,\infty }}{{({{\Delta }^{\bullet }})}^{2}}}})\ge \frac{1}{2}, we can deduce from (4.9) the uniqueness for the system {(4.2)-(4.4)}.
Lemma 2. Let w^\circ \in H^1(\Omega) be the function as in Lemma 1, and let \Psi_0^\bullet (w^\circ; {}\cdot\, ) be the functional on V^* \times L^2(\Omega)^2 given in (4.7). Also, let us take a sequence \{ w_n^\circ \}_{n=1}^\infty \subset H^1(\Omega) such that 0 \leq w_n^\circ \leq 1 a.e. in \Omega , for n=1, 2, 3, \dots , and let us define a sequence \{ \Psi_0^\bullet (w_n^\circ; {}\cdot\, ) \}_{n=1}^\infty of functionals on V^* \times L^2(\Omega)^2 , by putting w^\circ=w_n^\circ in (4.7), for n=1, 2, 3, \dots . Besides, let us assume that:
w_n^ \circ \to {w^ \circ }{\rm{ }}in\;the\;pointwi\;sesense\;a.e.\;in\;\Omega ,as\;n \to \infty . | (4.10) |
Then, \Psi_0^\bullet (w_n^\circ; {}\cdot\, ) \to \Psi_0^\bullet (w^\circ; {}\cdot\, ) on V^* \times L^2(\Omega)^2 , in the sense of \Gamma -convergence, as n \to \infty .
Proof. The condition of lower-bound will be seen, immediately, from the lower semi-continuity of the following functional (of 4-variables):
[w^\circ, e, v] \in L^2(\Omega) \times V^* \times L^2(\Omega)^2 \mapsto \Psi_0^\bullet(w^\circ; e, v) \in (-\infty, \infty]. |
The condition of optimality will be verified by taking the singleton \{ [e, v] \} for any [e, v] \in D (\Psi_0^\bullet (w^\circ; {}\cdot\, )) =D (\Psi_0^\bullet (w_n^\circ; {}\cdot\, )) for all n \ge 1 as the sequence corresponding to \{ z_n^\bullet \}_{n=1}^\infty in Definition 2.
Lemma 3. Under the assumptions as in the previous Lemmas 1-2, let us take the solution [e, v]=[e, w, \eta] \in V \times H^1(\Omega)^{2} to the stationary system {(4.5)-(4.6)} with v=[w, \eta] , and for any n \in \mathbb{N} , let us denote by [e_n, v_n]=[e_n, w_n, \eta_n] \in V \times H^1(\Omega)^2 the solution to {(4.5)-(4.6)} with v_n=[w_n, \eta_n] , when w^\circ=w_n^\circ . Then, the assumption (4.10) implies that:
\begin{array}{l} [{e_n},{\mathit{\boldsymbol{v}}_n}] = [{e_n},{w_n},{\eta _n}] \to [e,\mathit{\boldsymbol{v}}] = [e,w,\eta ]\;in\;{V^*} \times {L^2}{(\Omega )^2},\\ \mathit{and weakly in}\;{L^2}(\Omega ) \times {H^1}{(\Omega )^2},as\;n \to \infty \end{array} | (4.11) |
Proof. In the light of Lemma 1 (including the proof), we can see that:
\begin{array}{l} \Psi _0^ \bullet ({{\tilde w}^ \circ };\tilde e,\mathit{\boldsymbol{\widetilde v}}) = \Psi _0^ \bullet ({{\tilde w}^ \circ };\tilde e,\tilde w,\tilde \eta ) \le \Psi _0^ \bullet ({{\tilde w}^ \circ };0,0,0)\\ \le C_1^ \circ : = \frac{1}{{2h}}(|e_0^ \circ |_{{V^*}}^2 + |\mathit{\boldsymbol{v}}_0^ \circ |_{{L^2}{{(\Omega )}^2}}^2) + {{\cal L}^N}(\Omega )(|\gamma (0)| + |g(0,0)|)\\ + \alpha (0,0)|\theta _0^ \circ {|_{{W^{1,1}}(\Omega )}} + {\nu ^2}\beta (0,0)|\theta _0^ \circ |_{{H^1}(\Omega )}^2, \end{array} | (4.12) |
for any \tilde{w}^\circ \in H^1(\Omega) with 0 \leq \tilde{w}^\circ \leq 1 a.e. in \Omega , and
solution [\tilde{e}, \tilde{v}]=[\tilde{e}, \tilde{w}, \tilde{\eta}] to {(4.5)-(4.6)} with \tilde{v}=[\tilde{w}, \tilde{\eta}] when w^\circ=\tilde{w}^\circ .
Since the constant C_1^\circ is independent of the choice of \tilde{w}^\circ , 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 \Gamma convergence (cf. [9]).
Lemma 4. Let h_2^\circ be the constant as in (4.1). Let f_0^* \in V^* , u_0^\circ \in L^2(\Omega) , v_0^\circ=[w_0^\circ, \eta_0^\circ] \in H^1(\Omega)^2 and \theta_0^\circ \in W^{1, \infty}(\Omega) be the functions as in Lemma 1. Then, for any h \in (0, h_2^\circ] , the following system:
\frac{{u - u_0^ \circ }}{h} - \lambda '(w)\frac{{w - w_0^ \circ }}{h} + Fu = f_0^*\;in\;{V^*}, | (4.13) |
\begin{array}{l} \frac{{\mathit{\boldsymbol{v}} - \mathit{\boldsymbol{v}}_0^ \circ }}{h} - {\Delta _N}\mathit{\boldsymbol{v}} + \left[ {\begin{array}{*{20}{c}} {\partial \gamma (w)}\\ 0 \end{array}} \right] + [\nabla g](\mathit{\boldsymbol{v}})\\ + |D\theta _0^ \circ |[\nabla \alpha ](\mathit{\boldsymbol{v}}) + {\nu ^2}|D\theta _0^ \circ {|^2}[\nabla \beta ](\mathit{\boldsymbol{v}}) \ni \left[ {\begin{array}{*{20}{c}} { - \lambda '(w)u}\\ 0 \end{array}} \right]{\rm{ }}\;in\;{L^2}{(\Omega )^2}, \end{array} | (4.14) |
admits at least one solution [u, v]=[u, w, \eta] \in V \times H^1(\Omega)^2 with v=[w, \eta] .
Proof. Let us set a compact set K_1^\bullet in L^2(\Omega) , by letting:
K_{1}^{\bullet }:=\left\{ \tilde{w}\in {{H}^{1}}(\Omega )\left| \begin{align} & 0\le \tilde{w}\le 1\ \text{a}\text{.e}\text{.}\ \text{in}\ \Omega, \text{and} \\ & \frac{1}{2h}|\tilde{w}-w_{0}^{{}^\circ }|_{{{L}^{2}}(\Omega )}^{2}+\frac{1}{2}|D\tilde{w}|_{{{L}^{2}}{{(\Omega )}^{N}}}^{2} \\ & \le C_{1}^{{}^\circ }+|{{c}_{*}}|{{\mathscr{L}}^{N}}(\Omega )+\frac{1}{2h}|e_{0}^{{}^\circ }|_{{{V}^{*}}}^{2}+h|f_{0}^{*}|_{{{V}^{*}}}^{2} \\ \end{align} \right. \right\}, |
and let us consider an operator P_1^\bullet : K_1^\bullet \to L^2(\Omega) , which maps any w^\circ \in K_1^\bullet to the component w of the solution [u, w, \eta] \in V \times H^1(\Omega) \times H^1(\Omega) to {(4.2)-(4.4)}. Then, on account of (A3), (A6), Lemma 3, (4.7) and (4.12), it will be seen that P_1^\bullet K_1^\bullet \subset K_1^\bullet and P_1^\bullet is a continuous operator in the topology of L^2(\Omega) . So, applying Schauder's fixed point theorem, we find a fixed point w^\bullet \in K_1^\bullet for P_1^\bullet , i.e. w^\bullet=P_1^\bullet w^\bullet in L^2(\Omega) .
Now, let us denote by [u^\bullet, w^\bullet, \eta^\bullet] \in V \times H^1(\Omega) \times H^1(\Omega) the solution to {(4.2)-(4.4)}, involved in the fixed point w^\bullet . Then, this triplet [u^\bullet, w^\bullet, \eta^\bullet] must be a special solution to {(4.2)-(4.4)} such that w^\bullet=w^\circ . Hence, our remaining task will be to show that
0 \le {\eta ^ \bullet } \le 1{\rm{ }}\;{\rm{a}}{\rm{.e}}{\rm{.in}}\;\Omega , | (4.15) |
namely, the subdifferential \partial I_{[-1, 2]} in (4.4) will not affect for \eta^\bullet .To this end, we check two inequalities:
\frac{{0 - \eta _0^ \circ }}{h} + {g_\eta }({w^ \bullet },0) + |D\theta _0^ \circ |{\alpha _\eta }({w^ \bullet },0) + {\nu ^2}|D\theta _0^ \circ {|^2}{\beta _\eta }({w^ \bullet },0) \le 0{\rm{ }}\;in\;{L^2}(\Omega ), | (4.16) |
\frac{{1 - \eta _0^ \circ }}{h} + {g_\eta }({w^ \bullet },1) + |D\theta _0^ \circ |{\alpha _\eta }({w^ \bullet },1) + {\nu ^2}|D\theta _0^ \circ {|^2}{\beta _\eta }({w^ \bullet },1) \le 0{\rm{ }}\;in\;{L^2}(\Omega ), | (4.17) |
with use of the assumptions (A3), (A5) and 0 \leq \eta_0^\circ \leq 1 a.e. in \Omega .
On this basis, let us take the difference from (4.16) to (4.4) when \eta=\eta^\bullet and w=w^\circ=w^\bullet (resp. from (4.4) to (4.17) when \eta=\eta^\bullet and w=w^\circ=w^\bullet ), and multiply the both sides by [-\eta^\bullet]^+ (resp. by [\eta^\bullet-1]^+ ). Then, with the assumptions (A3), (A5) and \partial I_{[-1, 2]}(0)=\{0\} (resp. \partial I_{[-1, 2]}(1)=\{0\} ) in mind, it is inferred that:
\begin{array}{l} \frac{1}{h}\left( {1 - h|{g_{\eta \eta }}{|_{C({\Delta ^ \bullet })}}} \right)|{[ - {\eta ^ \bullet }]^ + }|_{{L^2}(\Omega )}^2 + |D{[ - {\eta ^ \bullet }]^ + }|_{{L^2}{{(\Omega )}^N}}^2 \le 0\\ \left( {{\rm{resp}}{\rm{. }}\frac{1}{h}\left( {1 - h|{g_{\eta \eta }}{|_{C({\Delta ^ \bullet })}}} \right)|{{[{\eta ^ \bullet } - 1]}^ + }|_{{L^2}(\Omega )}^2 + |D{{[{\eta ^ \bullet } - 1]}^ + }|_{{L^2}{{(\Omega )}^N}}^2 \le 0} \right). \end{array} | (4.18) |
Since the assumption (4.1) implies 1 -h|g_{\eta \eta}|_{C (\Delta^\bullet)} \geq \frac{1}{2} , we can deduce (4.15) from (4.18), and conclude that the triplet [u^\bullet, v^\bullet]=[u^\bullet, \eta^\bullet, w^\bullet] with v^\bullet :=[w^\bullet, \eta^\bullet] solves the system {(4.13)-(4.14)}.
Lemma 5. Let f_0^* \in V^* and \theta_{0}^{\circ} \in H^{M}(\Omega) be fixed functions, and let [u, v]=[u, w, \eta] \in V \times H^{1}(\Omega)^2 be a solution to the system {(4.13)-(4.14)} with v=[w, \eta] . Then, the inclusion
{\alpha _0}(\mathit{\boldsymbol{v}})\frac{{\theta - \theta _0^ \circ }}{h} + \partial \Phi _\varepsilon ^\nu (\mathit{\boldsymbol{v}};\theta ) \ni 0{\rm{ in }}{L^2}(\Omega ) | (4.19) |
admits a unique solution \theta \in H^{M}(\Omega).
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, \theta]=[u, w, \eta, \theta] \in D_M with v=[w, \eta] \in H^1(\Omega)^2 , which solves the coupled system {(4.13)-(4.14), (4.19)}. Then, the following energy-inequality holds:
\begin{array}{l} \frac{{{A_*}}}{{2h}}|u - u_0^ \circ |_{{V^*}}^2 + \frac{1}{{2h}}|\mathit{\boldsymbol{v}} - \mathit{\boldsymbol{v}}_0^ \circ |_{{L^2}{{(\Omega )}^2}}^2 + \frac{1}{h}|\sqrt {{\alpha _0}(\mathit{\boldsymbol{v}})} (\theta - \theta _0^ \circ )|_{{L^2}(\Omega )}^2\\ + \frac{h}{2}|u|_V^2 + {\mathscr{F}}_\varepsilon ^\nu (u,{\rm{ }}\mathit{\boldsymbol{v}},\theta ) \le {\mathscr{F}}_\varepsilon ^\nu (u_0^ \circ ,{\rm{ }}\mathit{\boldsymbol{v}}_0^ \circ ,\theta _0^ \circ ) + h|f_0^*|_{{V^*}}^2, \end{array} | (4.20) |
where A_* > 0 is the constant as in (A2), and \mathscr{F}_\varepsilon^\nu is the relaxed version of the functional \mathscr{F}_\nu , defined as:
\begin{array}{l} [u,\mathit{\boldsymbol{v}},\theta ] = [u,w,\eta ,\theta ] \in {L^2}{(\Omega )^4} \mapsto {\mathscr{F}}_\varepsilon ^\nu (u,\mathit{\boldsymbol{v}},\theta ) = {\mathscr{F}}_\varepsilon ^\nu (u,w,\eta ,\theta )\\ = {B_*}|u|_{{L^2}(\Omega )}^2 + \Psi _\gamma ^2(\mathit{\boldsymbol{v}}) + \int_\Omega ( g(\mathit{\boldsymbol{v}}) - {c_*}){\mkern 1mu} dx + \Phi _\varepsilon ^\nu (\mathit{\boldsymbol{v}};\theta ), \end{array} | (4.21) |
with the constant B_*=(1 +A_*)/2 as in (3.1).
Proof. First, let us rewrite the equation (4.13) as follows:
\begin{array}{l} {(u - u_0^ \circ ,z)_{{L^2}(\Omega )}} + h\langle Fu,z\rangle = h\langle f_0^*,z\rangle \\ + {(\lambda '(w)(w - w_0^ \circ ),z)_{{L^2}(\Omega )}},{\rm{ for}}\;{\rm{any}}\;z \in V, \end{array} | (4.22) |
and let us put z=u . Then, by using Schwarz's inequality, we have:
\frac{1}{2}|u|_{{L^2}(\Omega )}^2 + \frac{h}{2}|u|_V^2 \le \frac{1}{2}|u_0^ \circ |_{{L^2}(\Omega )}^2 + \frac{h}{2}|f_0^*|_{{V^*}}^2 + {(\lambda '(w)(w - w_0^ \circ ),u)_{{L^2}(\Omega )}}. | (4.23) |
Alternatively, if we rewrite the equation (4.13) to:
\frac{1}{h} (u -u_0^\circ, z^*)_{V^*} + \langle z^*, u \rangle = (f_0^*, z^*)_{V^*} +\frac{1}{h} (\lambda'(w)(w -w_0^\circ), z^*)_{V^*}, |
for any z^* \in V^* ,
and put z^{\ast}=A_*(u -u_0^\circ) \in V, then we also see that:
\frac{{{A_*}}}{{2h}}|u - u_0^ \circ |_{{V^*}}^2 + \frac{{{A_*}}}{2}|u|_{{L^2}(\Omega )}^2 \le \frac{{{A_*}}}{2}|u_0^ \circ |_{{L^2}(\Omega )}^2 + {A_*}h|f_0^*|_{{V^*}}^2 + \frac{1}{{4h}}|w - w_0^ \circ |_{{L^2}(\Omega )}^2. | (4.24) |
Next, let us multiply the both sides of the inclusion (4.14) by v -v_0^\circ . Then, in the light of (A2)-(A5) and Taylor's theorem, we infer that:
\begin{array}{l} \frac{1}{h}\left( {1 - \frac{h}{2}|g{|_{{C^2}({{[0,1]}^2})}}} \right)|\mathit{\boldsymbol{v}} - \mathit{\boldsymbol{v}}_0^ \circ |_{{L^2}{{(\Omega )}^2}}^2 + \frac{1}{2}|D\mathit{\boldsymbol{v}}|_{{L^2}{{(\Omega )}^{N \times 2}}}^2 + \int_\Omega \gamma (w){\mkern 1mu} dx + \int_\Omega g (\mathit{\boldsymbol{v}}){\mkern 1mu} dx\\ + \int_\Omega \alpha (\mathit{\boldsymbol{v}})|D\theta _0^ \circ |{\mkern 1mu} dx + {\nu ^2}\int_\Omega \beta (\mathit{\boldsymbol{v}})|D\theta _0^ \circ {|^2}{\mkern 1mu} dx\\ \le \frac{1}{2}|D\mathit{\boldsymbol{v}}_0^ \circ |_{{L^2}{{(\Omega )}^{N \times 2}}}^2 + \int_\Omega \gamma (w_0^ \circ ){\mkern 1mu} dx + \int_\Omega g (\mathit{\boldsymbol{v}}_0^ \circ ){\mkern 1mu} dx.\\ + \int_\Omega \alpha (\mathit{\boldsymbol{v}}_0^ \circ )|D\theta _0^ \circ |{\mkern 1mu} dx + {\nu ^2}\int_\Omega \beta (\mathit{\boldsymbol{v}}_0^ \circ )|D\theta _0^ \circ {|^2}{\mkern 1mu} dx - {(\lambda '(w)(w - w_0^ \circ ),u)_{{L^2}(\Omega )}}. \end{array} | (4.25) |
Furthermore, applying the both sides of (4.19) by \theta -\theta_0^\circ , it follows that:
\frac{1}{h}|\sqrt {{\alpha _0}(\mathit{\boldsymbol{v}})} (\theta - \theta _0^ \circ )|_{{L^2}(\Omega )}^2 + \Phi _\varepsilon ^\nu (\mathit{\boldsymbol{v}};\theta ) \le \Phi _\varepsilon ^\nu (\mathit{\boldsymbol{v}};\theta _0^ \circ ). | (4.26) |
Now, since (4.1) implies 1 -\frac{h}{2}|g|_{C^2([0, 1]^2)^2} \geq \frac{3}{4} , 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 1 \leq N \leq 3 of the spatial dimension, there exists a positive constant C_2^\circ , such that under the notations and assumptions as in Lemma 6, the condition:
C_2^ \circ {h^{\frac{1}{3}}}(1 + 2{({\mathscr{F}}_\varepsilon ^\nu (u_0^ \circ ,{\rm{ }}\mathit{\boldsymbol{v}}_0^ \circ ,\theta _0^ \circ ) + h|f_0^*|_{{V^*}}^2)^{\frac{2}{3}}}) \le \frac{1}{2},{\rm{ and }}0 < h \le h_2^ \circ , | (4.27) |
implies the uniqueness of the solution [u, v, \theta]=[u, w, \eta, \theta] \in D_M to the system {(4.13)-(4.14), (4.19)} with v=[w, \eta] .
Proof. In the light of the uniqueness of \theta as in Lemma 5, it is enough to focus only on the uniqueness for the component [u, v]=[u, w, \eta] \in V \times H^1(\Omega)^2 with v=[w, \eta] . To this end, we take two triplets [u_k, v_k]=[u_k, w_k, \eta_k] \in D_M with v_k=[w_k, \eta_k] , 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 u_k , k=1, 2 , and put z=u_1 -u_2 in V . Then:
\begin{align} & |{{u}_{1}}-{{u}_{2}}|_{{{L}^{2}}(\Omega )}^{2}+h|{{u}_{1}}-{{u}_{2}}|_{V}^{2}={{({\lambda }'({{w}_{1}}){{w}_{1}}-{\lambda }'({{w}_{2}}){{w}_{2}}, {{u}_{1}}-{{u}_{2}})}_{{{L}^{2}}(\Omega )}} \\ & -{{(({\lambda }'({{w}_{1}})-{\lambda }'({{w}_{2}}))w_{0}^{{}^\circ }, {{u}_{1}}-{{u}_{2}})}_{{{L}^{2}}(\Omega )}}, \\ \end{align} |
so that by using (A2) and Schwarz's inequality, we have:
\frac{1}{4}|{u_1} - {u_2}|_{{L^2}(\Omega )}^2 + h|{u_1} - {u_2}|_V^2 \le 3|\lambda '|_{{W^{1,\infty }}(0,1)}^2|{w_1} - {w_2}|_{{L^2}(\Omega )}^2. | (4.28) |
Secondly, let us take the difference between two inclusions (4.14) for v_k=[w_k, \eta_k] , k=1, 2 , and multiply the both sides by v_1 -v_2 in L^2(\Omega)^2 . Then, by using (A2)-(A5) and Schwarz's inequality, it is computed that:
\begin{array}{l} \frac{1}{h}\left( {1 - h|[\nabla g]{|_{{C^1}{{({{[0,1]}^2})}^2}}}} \right)|{\mathit{\boldsymbol{v}}_1} - {\mathit{\boldsymbol{v}}_2}|_{{L^2}{{(\Omega )}^2}}^2 + |D({\mathit{\boldsymbol{v}}_1} - {\mathit{\boldsymbol{v}}_2})|_{{L^2}{{(\Omega )}^{N \times 2}}}^2\\ \le - {(\lambda '({w_1}){u_1} - \lambda '({w_2}){u_2},{w_1} - {w_2})_{{L^2}(\Omega )}}\\ \le |\lambda '{|_{{L^\infty }(0,1)}}|{u_1} - {u_2}{|_{{L^2}(\Omega )}}|{w_1} - {w_2}{|_{{L^2}(\Omega )}} + {({u_1}(\lambda '({w_1}) - \lambda '({w_2})),{w_1} - {w_2})_{{L^2}(\Omega )}}\\ \le \frac{1}{8}|{u_1} - {u_2}|_{{L^2}(\Omega )}^2 + 2|\lambda '|_{{L^\infty }(0,1)}^2|{w_1} - {w_2}|_{{L^2}(\Omega )}^2 + |\lambda ''{|_{{L^\infty }(0,1)}}\int_\Omega | {u_1}||{w_1} - {w_2}{|^2}{\mkern 1mu} dx. \end{array} | (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
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.
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).
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.
All authors declare no conflicts of interest in this paper.
[1] | Landry M, Lemieux S, Lapointe A et al. (2018) Is eating pleasure compatible with healthy eating? A qualitative study on Quebecers' perceptions. Appetite 125: 537–547. |
[2] |
Marty L, Chambaron S, Nicklaus S, et al. (2018) Learned pleasure from eating: An opportunity to promote healthy eating in children? Appetite 120: 265–274. doi: 10.1016/j.appet.2017.09.006
![]() |
[3] |
Berridge KC, Ho CY, Richard JM, et al. (2010) The tempted brain eats: Pleasure and desire circuits in obesity and eating disorders. Brain Res 1350: 43–46. doi: 10.1016/j.brainres.2010.04.003
![]() |
[4] | Asioli D, Wongprawns R, Pignatti E, et al. (2018) Can information affect sensory perceptions? Evidence from a survey on Italian organic food consumers. AIMS Agric Food 3: 327–344. |
[5] |
Bryant C, Barnett J (2018) Consumer acceptance of cultured meat: A systematic review. Meat Sci 143: 8–17. doi: 10.1016/j.meatsci.2018.04.008
![]() |
[6] |
Grunert KG, Sonntag WI, Glanz-Chanos V, et al. (2018) Consumer interest in environmental impact, safety, health and animal welfare aspects of modern pig production: Results of a cross-national choice experiment. Meat Sci 137: 123–129. doi: 10.1016/j.meatsci.2017.11.022
![]() |
[7] |
Kumpulainen T, Vainio A, Sandell M, et al. (2018) How young people in Finland respond to information about the origin of food products: The role of value orientations and product type. Food Qual Prefer 68: 173–182. doi: 10.1016/j.foodqual.2018.03.004
![]() |
[8] |
Santeramo FG, Carlucci D, De Devitiis B, et al. (2018) Emerging trends in European food, diets and food industry. Food Res Int 104: 39–47. doi: 10.1016/j.foodres.2017.10.039
![]() |
[9] |
Tey YS, Arsil P, Brindal M, et al. (2018) Personal values underlying ethnic food choice: Means-end evidence for Japanese food. J Ethnic Foods 5: 33–39. doi: 10.1016/j.jef.2017.12.003
![]() |
[10] |
Miceli A, Francesca N, Moschetti G, et al. (2015) The influence of addition of Borago officinalis with antibacterial activity on the sensory quality of fresh pasta. Int J Gastronomy Food Sci 2: 93–97. doi: 10.1016/j.ijgfs.2014.12.004
![]() |
[11] |
Boroski M, de Aguiar AC, Boeing JS, et al. (2011) Enhancement of pasta antioxidant activity with oregano and carrot leaf. Food Chem 125: 696–700. doi: 10.1016/j.foodchem.2010.09.068
![]() |
[12] |
Martini D, Ciccoritti R, Nicoletti I, et al. (2018) From seed to cooked pasta: Influence of traditional and non-conventional transformation processes on total antioxidant capacity and phenolic acid content. Int J Food Sci Nutr 69: 24–32. doi: 10.1080/09637486.2017.1336751
![]() |
[13] | AGRI-MARCHE Regione Marche Agricoltura, Italy. Aree Tematiche. Available from: http://www.agri.marche.it/Aree%20tematiche/ricerca%20e%20sperimentazione/pasta%20alimentare/Documentazione%20finale/Risultati%20finali.pdf. (Italian language). |
[14] | International Pasta Organization (IPO) (2016) Available from: http://www.internationalpasta.org/index.aspx?id=3. |
[15] | Associazione delle Industrie del Dolce e della Pasta Italiane (AIDEPI) (2017) L'industria italiana del dolce e della pasta: Andamento economico 2016. AIDEPI Roma. Available from: http://www.aidepi.it/associazione/pubblicazioni/115-andamento-economico-2016.html. (Italian language). |
[16] | Cox R, Skouteris H, Fuller-Tyszkiewicz M, et al. (2018) A qualitative exploration of coordinators' and carers' perceptions of the Healthy Eating, Active Living (HEAL) programme in residential care. Child Abuse Rev 27: 122–136. |
[17] | Belton B, Bush SR, Little DC (2017) Not just for the wealthy: Rethinking farmed fish consumption in the Global South. Global Food Secur 16: 85–92. |
[18] |
Desiere S, Hung Y, Verbeke W, et al. (2018) Assessing current and future meat and fish consumption in Sub-Sahara Africa: Learnings from FAO Food Balance Sheets and LSMS household survey data. Global Food Secur 16: 116–126. doi: 10.1016/j.gfs.2017.12.004
![]() |
[19] | Schreinemachers P, Simmons EB, Wopereis MC (2017) Tapping the economic and nutritional power of vegetables. Global Food Secur 16: 36–45. |
[20] |
Guillaume D (2011) The Mediterranean diet: A cultural journey. The Lancet 378: 766–767. doi: 10.1016/S0140-6736(11)61370-6
![]() |
[21] | Altamore L, Bacarella S, Columba P, et al. (2017) The Italian Consumers' Preferences for Pasta: Does Environment Matter? Chem Eng Trans 58: 859–864. |
[22] | UNESCO (2018) Intangible Cultural Heritage of Humanity, Mediterranean Diet. Available from: https://ich.unesco.org/en/RL/mediterranean-diet-00884. |
[23] | Bacarella S, Altamore L, Valdesi V, et al. (2015) Importance of food labeling as a means of information and traceability according to consumers. Adv Hortic Sci 29: 145–151. |
[24] |
Boukid F, Folloni S, Sforza S, et al. (2018) Current Trends in Ancient Grains Based Foodstuffs: Insights into Nutritional Aspects and Technological Applications. Compr Rev Food Sci Food Saf 17: 123–136. doi: 10.1111/1541-4337.12315
![]() |
[25] |
Dinu M, Whittaker A, Pagliai G, et al. (2018) Ancient wheat species and human health: Biochemical and clinical implications. J Nutr Biochem 52: 1–9. doi: 10.1016/j.jnutbio.2017.09.001
![]() |
[26] | Sortino G, Allegra A, Inglese P, et al. (2016) Influence of an evoked pleasant consumption context on consumers' hedonic evaluation for minimally processed cactus pear (Opuntia ficus-indica) fruit. Acta Hortic 1141: 327–334. |
[27] | Resmini P, Pagani MA (1983) Ultrastructure Studies of Pasta. A Review. Food Struct 2: 2. |
[28] |
Güler S, Köksel H, Ng PKW, et al. (2002) Effects of industrial pasta drying temperatures on starch properties and pasta quality. Food Res Int 35: 421–427. doi: 10.1016/S0963-9969(01)00136-3
![]() |
[29] | D'Egidio MG, Mariani BM, Nardi S, et al. (1990) Chemical and technological variables and their relationships: A predictive equation for pasta cooking quality. Cereal Chem 67: 275–281. |
[30] | Feillet P, Dexter JE, (1996) Quality requirements of durum wheat for semolina milling and pasta production, In: JE. Kruger RR. Matsuo JW. Dick (Eds.), Pasta and noodle technology, St. Paul, MN, USA: AACC Inc, 95–123. |
[31] |
Bustos MC, Perez GT, León AE (2011) Sensory and nutritional attributes of fibre-enriched pasta. LWT-Food Sci Technol 44:1429–1434. doi: 10.1016/j.lwt.2011.02.002
![]() |
[32] | Ares G, Varela P (2018) Methods in Consumer Research, Volume 1: New Approaches to Classic Methods. Woodhead Publishing, 1–582. |
[33] | Kotler P (1973) Atmospherics as a marketing tool. J Retailing 49: 48–64. |
[34] |
Helmefalk M, Hultén B (2017) Multi-sensory congruent cues in designing retail store atmosphere: Effects on shoppers' emotions and purchase behavior. J Retailing Consum Serv 38: 1–11. doi: 10.1016/j.jretconser.2017.04.007
![]() |
[35] |
Lee NY, Noble SM, Biswas D (2018) Hey big spender! A golden (color) atmospheric effect on tipping behavior. J Acad Mark Sci 46: 317–337. doi: 10.1007/s11747-016-0508-3
![]() |
[36] |
Vukadin A, Lemoine JF, Badot O (2016) Opportunities and risks of combining shopping experience and artistic elements in the same store: A contribution to the magical functions of the point of sale. J Mark Manage 32: 944–964. doi: 10.1080/0267257X.2016.1186106
![]() |
[37] |
Vukadin A, Wongkitrungrueng A, Assarut N (2018) When art meets mall: Impact on shopper responses. J Prod Brand Manage 27: 277–293. doi: 10.1108/JPBM-01-2017-1406
![]() |
[38] | Mehrabian A, Russell JA (1974) An approach to environmental psychology. MIT Press Cambridge MA. Behav Therap 7: 132–133. |
[39] |
Jang SS, Namkung Y (2009) Perceived quality, emotions, and behavioral intentions: Application of an extended Mehrabian-Russell model to restaurants. J Bus Res 62: 451–460. doi: 10.1016/j.jbusres.2008.01.038
![]() |
[40] |
Foster J, McLelland MA (2015) Retail atmospherics: The impact of a brand dictated theme. J Retailing Consum Serv 22: 195–205. doi: 10.1016/j.jretconser.2014.07.002
![]() |
[41] |
Hultén B (2011) Sensory marketing: The multi-sensory brand-experience concept. Eur Bus Rev 23: 256–273. doi: 10.1108/09555341111130245
![]() |
[42] |
Wiedmann KP, Labenz F, Haase J, et al. (2018) The power of experiential marketing: Exploring the causal relationships among multisensory marketing, brand experience, customer perceived value and brand strength. J Brand Manage 25: 101–118. doi: 10.1057/s41262-017-0061-5
![]() |
[43] | EUROSTAT (2017) Available from: https://ec.europa.eu/eurostat. |
[44] |
Morgan DL (2018) Living within blurry boundaries: The value of distinguishing between qualitative and quantitative research. J Mixed Meth Res 12: 268–279. doi: 10.1177/1558689816686433
![]() |
[45] |
Nyumba TO, Wilson K, Derrick CJ, et al. (2018) The use of focus group discussion methodology: Insights from two decades of application in conservation. Meth Ecol Evol 9: 20–32. doi: 10.1111/2041-210X.12860
![]() |
[46] | Chernev A (2018) Strategic marketing management (Ninth edition). Cerebellum Press Chicago IL. |
[47] | Dolnicar S, Grün B, Leisch F, (2018) Market Segmentation Analysis, In: Market Segmentation Analysis, Springer, Singapore, 11–22. |
[48] |
Chironi S, Bacarella S, Altamore L, et al. (2017) Quality Factors Influencing Consumer Demand for Small Fruit by Focus Group and Sensory Test. J Food Prod Mark 23: 857–872. doi: 10.1080/10454446.2017.1244791
![]() |
[49] | Stewart DW, Shamdasani PN (2014) Focus groups: Theory and practice (Vol. 20). Sage publications. |
[50] |
Kitzberger CSG, Scholz MBS, da Silva JBGD, et al. (2016) Free choice profiling sensory analysis to discriminate coffees. AIMS Agric Food 1: 455–469. doi: 10.3934/agrfood.2016.4.455
![]() |
[51] |
Streicher MC, Estes Z (2016) Multisensory interaction in product choice: Grasping a product affects choice of other seen products. J Consum Psychol 26: 558–565. doi: 10.1016/j.jcps.2016.01.001
![]() |
[52] |
Chironi S, Bacarella S, Altamore L, et al. (2017) Study of product repositioning for the Marsala Vergine DOC wine. Int J Entrepreneurship Small Bus 32: 118–138. doi: 10.1504/IJESB.2017.085982
![]() |
[53] | Ingrassia M, Chironi S, Allegra A, et al. (2017) Consumer preferences for fig fruit (Ficus carica L.) quality attributes and postharvest storage at low temperature by in-store survey and focus group. Acta Hortic 1173: 383–388. |
[54] | Ingrassia M, Bacarella S, Altamore L, et al. (2018) Consumer acceptance and primary drivers of liking for small fruits. Acta Hortic 1194: 1147–1154. |
[55] |
Thielecke F, Nugent A (2018) Contaminants in Grain-A Major Risk for Whole Grain Safety? Nutrients 10: 1213. doi: 10.3390/nu10091213
![]() |
[56] | Ennouari A, Sanchis V, Rahouti M, et al. (2018) Isolation and molecular identification of mycotoxin producing fungi in durum wheat from Morocco. J Mater Environ Sci 9: 1470–1479. |
[57] |
Bräuer J, Plösch R, Saft M, et al. (2018) Measuring object-oriented design principles: The results of focus group-based research. J Syst Software 140: 74–90. doi: 10.1016/j.jss.2018.03.002
![]() |
[58] | Chironi S, Sortino G, Allegra A, et al. (2017) Consumer assessment on sensory attributes of fresh table grapes cv "Italia" and "Red Globe" after long cold storage treatment. Chem Eng Trans 58: 421–426. |
[59] |
Ertl K, Goessler W (2018) Grains, whole flour, white flour, and some final goods: an elemental comparison. Eur Food Res Technol 244: 2065–2075. doi: 10.1007/s00217-018-3117-1
![]() |
[60] |
Piqueras-Fiszman B, Spence C (2012) The influence of the feel of product packaging on the perception of the oral-somatosensory texture of food. Food Qual Prefer 26: 67–73. doi: 10.1016/j.foodqual.2012.04.002
![]() |
[61] | Erenkol AD, Merve AK (2015) Sensory Marketing. J Administrative Sci Policy Stud 3: 1–26. |
[62] |
Madzharov AV, Block LG, Morrin M (2015) The cool scent of power: Effects of ambient scent on consumer preferences and choice behavior. J Mark 79: 83–96. doi: 10.1509/jm.13.0263
![]() |
[63] | Ingrassia M, Bacarella S, Columba P, et al. (2017) Traceability and Labelling of Food Products from the Consumer Perspective. Chem Eng Trans 58: 865–870. |
[64] | Caputo V, Sacchi G, Lagoudakis A, (2018) Traditional Food Products and Consumer Choices: A Review, In: Case Studies in the Traditional Food Sector, 47–87. |
[65] |
Bava L, Bacenetti J, Gislon G, et al. (2018) Impact assessment of traditional food manufacturing: The case of Grana Padano cheese. Sci Total Environ 626: 1200–1209. doi: 10.1016/j.scitotenv.2018.01.143
![]() |
[66] |
Cei L, Defrancesco E, Stefani G (2018) From Geographical Indications to Rural Development: A Review of the Economic Effects of European Union Policy. Sustainability 10: 1–21. doi: 10.3390/su10020001
![]() |
[67] | Eataly Net S.r.l (2018) Available from: https://www.eataly.com/. |
1. | Ibrahim A. Alnaser, Mohammed Yunus, Rami Alfattani, Turki Alamro, Multiple-Output Fracture Characteristics Optimization of Bi-material Interfaces for Composite Pipe Repair Using Swarm Intelligence Technique, 2021, 1547-7029, 10.1007/s11668-020-01086-3 |