
Cocoa (Theobroma cacao L.), indigenous to the tropical forests of the Americas, is renowned not only as the primary raw material for chocolate and its derivatives (cocoa liquor and butter) but also as a rich source of phytonutrients with beneficial health effects. Current research has elucidated that within the post-harvest process, fermentation stands as the critical stage for the formation of the principal biochemical quality markers in cocoa, known as polyphenols. These compounds contribute to the bitterness and astringency that constitute the complex flavor profile of chocolate; however, their excessive presence can be organoleptically undesirable. A high phenolic content (>10%) is associated with insufficient fermentation and certain varieties of ordinary cocoa, thereby serving as a discriminatory parameter between fine-flavor cocoa (Nacional) and bulk cocoa (CCN-51). Beyond their technological significance, these components have garnered substantial scientific interest, as polyphenol consumption is associated with potential protective effects against the development of non-communicable chronic diseases (including diabetes, cancer, and atherosclerosis), attributable to their potent antioxidant properties. In this context, the objective of this study was to evaluate the impact of fermentation time on the antioxidant capacity (AC) and total polyphenol content (TPC) in the principal Ecuadorian cocoa varieties (i.e., CCN-51 clone and Nacional). Pilot-scale fermentation experiments demonstrated significant variations in antioxidant capacity (CCN-51 clone: 785.61 to 1852.78 and Nacional: 564.32 to 1428.60 µmol TE/g) and total polyphenol content (CCN-51 clone: 52.92 to 162.82; Nacional: 40.55 to 157.50 mg gallic acid/g). Both parameters decreased markedly throughout the process, with the CCN-51 clone exhibiting greater retention.
Citation: Orbe Chamorro Mayra, Luis- Armando Manosalvas-Quiroz, Nicolás Pinto Mosquera, Iván Samaniego. Effect of fermentation parameters on the antioxidant activity of Ecuadorian cocoa (Theobroma cacao L.)[J]. AIMS Agriculture and Food, 2024, 9(3): 872-886. doi: 10.3934/agrfood.2024047
[1] | Aziz Belmiloudi . Cardiac memory phenomenon, time-fractional order nonlinear system and bidomain-torso type model in electrocardiology. AIMS Mathematics, 2021, 6(1): 821-867. doi: 10.3934/math.2021050 |
[2] | Binaya Tuladhar, Hana M. Dobrovolny . Testing the limits of cardiac electrophysiology models through systematic variation of current. AIMS Mathematics, 2020, 5(1): 140-157. doi: 10.3934/math.2020009 |
[3] | Folashade Agusto, Daniel Bond, Adira Cohen, Wandi Ding, Rachel Leander, Allis Royer . Optimal impulse control of West Nile virus. AIMS Mathematics, 2022, 7(10): 19597-19628. doi: 10.3934/math.20221075 |
[4] | Pierluigi Colli, Gianni Gilardi, Jürgen Sprekels . Distributed optimal control of a nonstandard nonlocal phase field system. AIMS Mathematics, 2016, 1(3): 225-260. doi: 10.3934/Math.2016.3.225 |
[5] | Biwen Li, Qiaoping Huang . Synchronization issue of uncertain time-delay systems based on flexible impulsive control. AIMS Mathematics, 2024, 9(10): 26538-26556. doi: 10.3934/math.20241291 |
[6] | Li Wan, Qinghua Zhou, Hongbo Fu, Qunjiao Zhang . Exponential stability of Hopfield neural networks of neutral type with multiple time-varying delays. AIMS Mathematics, 2021, 6(8): 8030-8043. doi: 10.3934/math.2021466 |
[7] | Kunting Yu, Yongming Li . Adaptive fuzzy control for nonlinear systems with sampled data and time-varying input delay. AIMS Mathematics, 2020, 5(3): 2307-2325. doi: 10.3934/math.2020153 |
[8] | N. Jayanthi, R. Santhakumari, Grienggrai Rajchakit, Nattakan Boonsatit, Anuwat Jirawattanapanit . Cluster synchronization of coupled complex-valued neural networks with leakage and time-varying delays in finite-time. AIMS Mathematics, 2023, 8(1): 2018-2043. doi: 10.3934/math.2023104 |
[9] | Hongguang Fan, Jihong Zhu, Hui Wen . Comparison principle and synchronization analysis of fractional-order complex networks with parameter uncertainties and multiple time delays. AIMS Mathematics, 2022, 7(7): 12981-12999. doi: 10.3934/math.2022719 |
[10] | Ravi P. Agarwal, Snezhana Hristova . Stability of delay Hopfield neural networks with generalized proportional Riemann-Liouville fractional derivative. AIMS Mathematics, 2023, 8(11): 26801-26820. doi: 10.3934/math.20231372 |
Cocoa (Theobroma cacao L.), indigenous to the tropical forests of the Americas, is renowned not only as the primary raw material for chocolate and its derivatives (cocoa liquor and butter) but also as a rich source of phytonutrients with beneficial health effects. Current research has elucidated that within the post-harvest process, fermentation stands as the critical stage for the formation of the principal biochemical quality markers in cocoa, known as polyphenols. These compounds contribute to the bitterness and astringency that constitute the complex flavor profile of chocolate; however, their excessive presence can be organoleptically undesirable. A high phenolic content (>10%) is associated with insufficient fermentation and certain varieties of ordinary cocoa, thereby serving as a discriminatory parameter between fine-flavor cocoa (Nacional) and bulk cocoa (CCN-51). Beyond their technological significance, these components have garnered substantial scientific interest, as polyphenol consumption is associated with potential protective effects against the development of non-communicable chronic diseases (including diabetes, cancer, and atherosclerosis), attributable to their potent antioxidant properties. In this context, the objective of this study was to evaluate the impact of fermentation time on the antioxidant capacity (AC) and total polyphenol content (TPC) in the principal Ecuadorian cocoa varieties (i.e., CCN-51 clone and Nacional). Pilot-scale fermentation experiments demonstrated significant variations in antioxidant capacity (CCN-51 clone: 785.61 to 1852.78 and Nacional: 564.32 to 1428.60 µmol TE/g) and total polyphenol content (CCN-51 clone: 52.92 to 162.82; Nacional: 40.55 to 157.50 mg gallic acid/g). Both parameters decreased markedly throughout the process, with the CCN-51 clone exhibiting greater retention.
The heart is an electrically controlled mechanical pump which drives blood flow through the circulatory system vessels (through deformation of its walls), where electrical impulses trigger mechanical contraction (of various chambers of heart) and whose dysfunction is incompatible with life. The electrical system of a normal heart is highly organized in a steady rhythmic pattern. This normal heartbeat is called sinus rhythm. Irregular or abnormal heartbeats, called arrhythmias, are caused by a change in the propagation and/or formation of electrical impulses, that regulate a steady heartbeat, causing a heartbeat that is too fast or too slow, that can remain stable or become chaotic (irregular and disorganized). Many times, arrhythmias are harmless and can occur in healthy people without heart disease; however, some of these rhythms can be serious and require special and efficiency treatments. Fibrillation is one type of arrhythmia and is considered the most serious cardiac rhythm disturbance. It occurs when the heart beats with rapid, erratic electrical impulses (highly disorganized almost chaotic activation). This causes the heart's chambers to quiver (or fibrillate) uselessly instead of contracting normally. Then the heart loses its ability to pump enough blood through the circulatory system. The treatment therapy of these diseases, when it becomes troublesome or when it can present a danger, often uses electrical impulses to stabilize cardiac function and restore the sinus rhythm, by implanting the patients with active cardiac devices (electrotherapy). For example, in case of cardiac rhythms that are too slow, the devices transmit electronic impulses and ensure that periodic contractions of heart are maintained at a hemodynamically sufficient rate; and in the case of a fast heart rate or irregular, the devices monitor the heart rate and, if needed, treat episodes of tachyarrhythmia (including tachycardia and/or fibrillation) by transmitting automatically impulses to either give defibrillation shocks or cause overstimulation (via an ICDs*) or synchronize the contraction of left and right ventricles. Although ICD electrotherapy has been shown to be an effective treatment against lethal cardiac arrhythmias, it remains a highly non-optimal therapy since the administrated strong shocks required for defibrillation can cause significant extra-cardiac stimulation, resulting in (physical and psychological) pains and long-term tissue damage. It is then necessary to optimize the defibrillation shock impulse in order to achieve the lowest energy necessary to successfully cardiovert a patient and, consequently, a maximum result with minimal detrimental side effects.
*The so-called implantable cardioverter defibrillators
Then, efficient tools for the assistance of patient specific treatment of cardiac disorders is of great scientific and socio-economical interest. The evaluation of the bioelectrical activity in the heart is a very complex process which uses different phenomenological mechanism and subject to various perturbations, and physiological and pathophysiological variations. Consequently, this has greatly emphasized the need for methodologies capable of predicting, understanding and optimizing different complex phenomena occurring in these fields, despite different sources of uncertainty like the absence of complete or reliable data (e.g., stimulus currents, measurement data), neglected dynamics, or intrinsic physical variability. The challenge here is e.g., to reduce the uncertainty and increase the reliability of model predictions in treatment of cardiac disease.
The goal of the present paper is to investigate minimax control problems for a bidomain type system, commonly used for modeling the propagation of electrophysiological waves in the myocardium, with disturbances (perturbation or noise) and controls in which multiple time-varying delays appear in the state system. The objective of a minimax control is to compensate the undesirable effects of system disturbances through control actions such that a cost function achieves its minimum for the worst disturbances: i.e. to find the best control which takes into account the worst-case disturbance. From the standpoint of our specific application, the main goal is to regulate and stabilize the optimal external applied current via transmembrane potential sensor.
Tissue-level cardiac electrophysiology, which can provide a bridge between electrophysiological cell models at smaller scales, and tissue mechanics, metabolism and blood flow at larger scales, is usually modeled using the coupled bidomain equations, originally derived in [67], which represent a homogenization of the intracellular and extracellular medium, where electrical currents are governed by Ohm's law (see also e.g. [44] for a review and an introduction to this field). The model was modified and extended to include heart tissue surrounded by a conductive bath or a conductive body (see e.g. [56] and [65]). From mathematical viewpoint, the classical bidomain system (Figure 1) is commonly formulated in terms of intracellular and extracellular electrical potentials of anisotropic cardiac tissue (macroscale), ϕi and ϕe, (or, equivalently, extracellular potential ϕe and the transmembrane voltage ϕ=φi−φe) coupled with cellular state variables u describing cellular membrane dynamics. This is a system of non-linear partial differential equations (PDEs) coupled with ordinary differential equations (ODEs), in the physical region Ω (occupied by excitable cardiac tissue, which is an open, bounded, and connected subset of d-dimensional Euclidean space Rd, d≤3). The PDEs describe the propagation of the electrical potentials and ODEs describe the electrochemical processes.
Time delays in signal transmission are inevitable and a small delay can affect considerably the resulting electrical activity in heart and thus the cardiac disorders therapeutic treatment. It is then necessary to introduce the impact of delays on dynamical behaviors of such a system. Delay terms can lead to change the stability of dynamics and give rise to highly complex behavior including oscillations and chaos. Motivated by above discussions, to take into account the effect of time-delays in propagation of electrophysiological waves in heart, together with other critical cardiac material parameters, we have developed a new bidomain model by incorporating multiple time delays in [6].
In this new model, in order to take into account the influence of time-delays in signal transmission and inward movement of u into the cell which prolongs the depolarization phase of action potential, classical bidomain model has been modified by using multiple time-delays functions in operators representing the ionic activity in myocardium. More precisely, the derived system, is a nonlinear coupled reaction-diffusion model in shape of a set of delay differential equations (DDE) coupled with a set of delay partial differential equations, in the heart's spatial domain Ω which is a bounded open subset with a sufficiently regular boundary Γ=∂Ω, and during the final fixed time horizon T>0, as follows (for more detail to derive this model see [6])
cm∂ϕ∂t+I(.;ϕ,u)−div(Ki∇ϕ)=div(Ki∇φe)+H(.;ϕτ,uτ)+Ii,inQ=Ω×(0,T)−div((Ke+Ki)∇φe)=div(Ki∇ϕ)+I,inQ∂u∂t+G(.;ϕ,u)=E(.;ϕτ,uτ),inQsubject to initial and past conditionsϕ(.,t=0)=ϕ0,u(.,t=0)=u0,inΩϕ=ϕpast,u=upast,inQ0=Ω×[−δ(0),0[and boundary conditions(Ki∇(ϕ+φe))⋅n=0,onΣ=∂Ω×(0,T)(Ke∇φe)⋅n=0,onΣ | (1.1) |
where ϕ=φi−φe, φe and φi are the transmembrane, extracellular and intracellular potentials, respectively; Ki and Ke are the conductivity tensors describing anisotropic intracellular and extracellular conductive media and cm(x)=κCm(x)>0, where Cm is the membrane capacitance per unit area and κ is the surface area-to-volume ratio. The tissue is assumed to be passive, so the capacitance Cm can be assumed to be not a function of the state variables. The function cm is assumed to be space variable and satisfies 0<c_m≤cm=b2m≤¯cm (where c_m and ¯cm are positive constants). The electrophysiological ionic state u describes a cumulative way of the effects of the ion transport through the cell membranes (which describe e.g., the dynamics of ion-channel and ion concentrations in different cellular compartments). The operator I=κIion, where the nonlinear operator Iion describes the sum of transmembrane ionic currents across cell membrane with u. The nonlinear operator G is representing the ionic activity in myocardium. Functional forms for I and G are determined by an electrophysiological cell model (which can found in the CellMl Repository†). The source terms are Ii=κfi, Ie=κfe and I=−Ii−Ie, where fi and fe describe intracellular and extracellular stimulation currents, respectively. The operators H and E are time-delay operators and the functions ϕτ and uτ are delayed states corresponding to ϕ and u respectively, and n is the outward normal to Γ=∂Ω. Here, the unknowns are the potentials ϕ, φe and a single ionic variable u (e.g. gating variable, concentration, etc.).
In absence of a grounded electrode, the bidomain equations are a naturally singular problem since φe only appears in the equations and boundary conditions through its gradient. Moreover, the state φe is only defined up to a constant. Such problems have compatibility conditions determining whether there are any solution to the PDEs. This is easily found by integrating the second equation of (1.1) over the domain and using the divergence theorem with boundary conditions. Then the following conservation of the total current is derived (a.e.in(0,T))
∫ΩIdx=0. | (1.2) |
Consequently, we must choose I such that the compatibility condition (1.2) is satisfied. Moreover, the function φe is defined within a class of equivalence, regardless of a time-dependent function. This function can be fixed, for example by setting the Gauge condition (a.e. in (0, T))
∫Ωφedx=0a.e.in(0,T). | (1.3) |
Remark 1.1. 1. Condition (1.3) is a common condition for pressure in fluid mechanics (in Navier-Stokes systems).
2. The functions Ki, Ke, H, G and cm depend on the fiber extension ratio.
3. If we assume that I is only dependent on time and is of the form
I(x,t)=θ(t)(χΩ1(x)−χΩ2(x)), | (1.4) |
where χΩi is the characteristic function of set Ωi, i=1,2, then condition (1.2) is satisfied if mes(Ω1)=mes(Ω2). The support regions Ω1 and Ω2 can be considered to represent an anode (positive electrode) and a cathode (negative electrode) respectively.
In recent years, various problems concerning biological rhythmic phenomena and delayed processes have been studied (see e.g., [8,13,14,16,20,21,22,23,38,42,43,55,60,61,64,72] and the references therein). For problems associated with bidomain models with time-delay, the literature is limited, to our knowledge, to [6,30]. Concerning problems associated with bidomain models without time-delays various methods and technique, as evolution variational inequalities approach, semi-group theory, Faedo-Galerkin method and others, the studies of the well-posedness of solutions have been derived in the literature (see e.g., [9,15,18,27,69] and the references therein); for development of multiscale mathematical and computational modeling of bioelectrical activity in myocardial tissue and their numerical simulations, which are based on methods as finite difference method, finite element method or lattice Boltzmann method, have been receiving a significant amount of attention (see e.g., [4,17,24,25,26,28,29,31,32,33,34,36,40,44,46,57,63,65,70,71] and the references therein), with a particular attention to the formation of cardiac disorders (as arrhythmias) and their therapeutic treatment (see e.g., [3,41,58,68] and the references therein). For control problems associated with the electrocardiology, we can mention [2,9,19,45,53].
The new feature introduced in this work concerns the study of nonlinear minimax control problem for a bidomain model with time-delays of cardiac tissue electrophysiology system, in order to take into account the influence of noises in data. The minimax control problem and the necessary optimality conditions are new for these types of equations studied here. This study is motivated by the applications, for example, in determining the best optimal current to be applied (taking into account the influence of disturbances in data), so that the peaks in the transmembrane potential are damped. In this context, it is possible to consider the specific application of implantable Cardioverter defibrillators, which are used to treat patients with life-threatening ventricular arrhythmias, in order to maximize both cardiac performance and additionally the lifetime of the device. Our approach is based on the results of existence and characterization of saddle points in infinite-dimensional (for more details see [10] and for minimax control see [11,12]).
The paper is organized as follows. In next section, first we give some preliminaries and well-posedness of the state equations results. Then some regularity results of the solution as well as the input-to-state stability estimate are derived, under extra assumptions. In Section 3, first we formulate the minimax control problem and we study rigorously the Fréchet differentiability of the solution operator of the problem. Second, we study the minimax control problem corresponding to obtain the saddle point of cost function J. The functional J is depending on disturbance and control in the domain Ω over the time interval under consideration [0,T]. We prove the existence of an optimal solution and give necessary optimality conditions. The optimality system is corresponding to identify the gradient of the cost function that is necessary to develop a numerical computation in order to solve the minimax control problem.
We use the standard notation for Sobolev spaces (see [1]), denoting the norm of Wm,p(Ω) (m∈IN, p∈[1,∞]) by ‖.‖Wm,p. In the special case p=2, we use Hm(Ω) instead of Wm,2(Ω). The duality pairing of a Banach space X with its dual space X′ is given by ⟨.,.⟩X′,X. For a Hilbert space Y the inner product is denoted by (.,.)Y and the inner product in L2(Ω) is denoted by (.,.). For any pair of real numbers r,s≥0, we introduce the Sobolev space Hr,s(Q) defined by Hr,s(Q)=L2(0,T;Hr(Ω))∩Hs(0,T;L2(Ω)), which is a Hilbert space normed by
(‖v‖2L2(0,T;Hr(Ω))+‖v‖2Hs(0,T;L2(Ω)))1/2, |
where Hs(0,T;L2(Ω)) denotes the Sobolev space of order s of functions defined on (0,T) and taking values in L2(Ω), and defined by, for θ∈(0,1),s=(1−θ)m with m an integer, (see e.g., [48]) Hs(0,T;L2(Ω))=[Hm(0,T,L2(Ω)),L2(Q)]θ, Hm(0,T;L2(Ω))={v∈L2(Q)|∂jv∂tj∈L2(Q),for 1≤j≤m}. For a given Banach space X, with norm ‖.‖X, of functions integrable on Ω, we define its subspace X|IR={u∈X,∫Ωu=0} that is a Banach space with norm ‖.‖X, and we denote by [u] the projection of u∈X on X|IR such that [u]=u−1mes(Ω)∫Ωudx (with mes(Ω) standing for Lebesgue measure of the domain Ω). Finally, we introduce the spaces:
● H=L2(Ω) and V=H1(Ω) endowed with their usual norms,
● U=V|IR.
We will denote by V′ (resp. U′) the dual of V (resp. of U). We have the following continuous embeddings (see e.g. [1,47]), where p≥2 if d=2 and 2≤p≤6 if d=3, p′ is such that 1p′+1p=1
V⊂H⊂V′, U⊂H|IR⊂U′,V⊂Lp(Ω)⊂H≡(H)′⊂Lp′(Ω)⊂V′ | (2.1) |
and the injections V⊂H and U⊂H|IR are compact. We can now introduce the following spaces: H(Q)=L∞(0,T;L2(Ω)), V(Q)=L2(0,T;V), ˜V(Q)=L2(0,T;U) and, for q>1, the space Wq(Q)={w∈V(Q)|∂w∂t∈Lq(0,T,V′)}.
Remark 2.1. If u∈Wq(Q)∩H(Q), then u is a weakly continuous function on [0,T] with values in L2(Ω) i.e. u∈Cw([0,T];L2(Ω)) (see e.g. [47]).
Remark 2.2. Let Ω⊂IRm, m≥1, be an open and bounded set with a smooth boundary and q be a nonnegative integer. We have the following results (see e.g. [1])
(i) Hq(Ω)⊂Lp(Ω), ∀p∈[1,2mm−2q], with continuous embedding (with the exception that if 2q=m, then p∈[1,+∞[ and if 2q>m, then p∈[1,+∞]).
(ii) (Gagliardo-Nirenberg inequalities) There exists C>0 such that
‖v‖Lp≤C‖v‖θHq‖v‖1−θL2,∀v∈Hq(Ω), |
where 0≤θ<1 and p=2mm−2θq (with the exception that if q−m/2 is a nonnegative integer, then θ is restricted to 0).
Remark 2.3. The spaces W(i)=H1(0,T;Hi−2(Ω))∩L2(0,T;Hi(Ω)) satisfy the following embedding:
(i) W(i), for i=1,3, is compactly embedded into L2(0,T,Hi−1(Ω)) (see e.g. [66]).
(ii) W(i)⊂C0([0,T];Hi−1(Ω)), for i=1,3 (see e.g. [48]).
Definition 2.1. A real valued function H defined on D×IRq, q≥1, is a Carathéodory function iff H(.;v) is measurable for all v∈IRq and H(y;.) is continuous for almost all y∈D.
Lemma 2.1. (Poincaré–Wirtinger inequality) Assume that 1≤p≤∞ and that Ω is a bounded connected open subset of IRd with a sufficiently regular boundary ∂Ω (e.g., a Lipschitz boundary). Then there exists a Poincaré constant C, depending only on Ω and p, such that for every function u in Sobolev space W1,p(Ω)
‖[u]‖Lp(Ω)≤C‖∇u‖Lp(Ω). |
Remark 2.4. From the Poincaré–Wirtinger inequality, we can deduce that the H1 semi-norm and the H1 norm are equivalent in the space U.
Our study involves the following fundamental inequalities, which are repeated here for review:
(ⅰ) Hölder's inequality: ∫DΠi=1,kfidx≤Πi=1,k‖fi‖Lqi(D),
where ‖fi‖Lqi(D)=(∫D∣fi∣qidx)1/qiand∑1≤i≤k1qi=1.
(ⅱ) Young's inequality (∀a,b>0 and ϵ>0): ab≤ϵpap+ϵ−q/pqbq,forp,q∈]1,+∞[and1p+1q=1.
(ⅲ) Minkowski's integral inequality:
[∫Ω(∫t0∣f(x,s)∣ds)pdx]1/p≤∫t0(∫Ω∣f(x,s)∣pdx)1/pds,forp∈]1,+∞[andt>0. |
(ⅳ) Gronwall's Lemma:
Ifdψdt≤g(t)ψ(t)+h(t),∀t≥0thenψ(t)≤ψ(0)exp(∫t0g(s)ds)+∫t0h(s)exp(∫tsg(τ)dτ)ds,∀t≥0. |
Finally, we denote by L(A;B) the set of linear and continuous operators from a vectorial space A into a vectorial space B, and by R∗ the adjoint operator to a linear operator R between Banach spaces.
From now on, we assume that the following assumptions hold for the nonlinear operators and tensor functions appearing in our model.
(H1) We assume that the conductivity tensor functions Kθ∈W1,∞(¯Ω), θ∈{i,e} are symmetric, positive definite matrix functions and that they are uniformly elliptic, i.e., there exist constants 0<K1<K2 such that
K1‖ψ‖2≤ψTKθψ≤K2‖ψ‖2in¯Ω,∀ψ∈IRd. | (2.2) |
Remark 2.5. We can emphasize a specificity of the tensors Ke and Ki (see e.g., [29]).
1. The tensors Ke(x) and Ki(x) have the same basis of eigenvectors Q(x)=(qk(x))1≤k≤d in IRd, which reflect the organization of muscle in fibers, and consequently Ki(x)=Q(x)Λi(x)Q(x)T and Ke(x)=Q(x)Λe(x)Q(x)T, where Λi(x)=diag((λi,k)1≤k≤d) and Λe(x)=diag((λe,k)1≤k≤d).
2. The muscle fibers are tangent to Γ so that (for θ∈{i,e}) : Kθn=λθ,dn,a.e.,inΓ, with λθ,d(x)≥λ>0, λ a constant.
The operators I and G which describe electrophysiological behavior of the system can be taken as follows (affine functions with respect to u)
I(x,t;ϕ,u)=I0(x,t;ϕ)+I1(x,t;ϕ)u,G(x,t;ϕ,u)=I2(x,t;ϕ)+ℏ(x,t)u, | (2.3) |
where ℏ is a sufficiently regular function. Moreover, the operators I0, I1 and I2 appearing in I and G, are supposed to satisfy the following assumptions.
(H2) The operators I0, I1 and I2 are Carathéodory functions from (Ω×IR)×IR into IR and continuous on ϕ (as in Belmiloudi [9]). Furthermore, for some p≥2 if d=2 and p∈[2,6] if d=3 (for more details see [18]), the following requirements hold
(ⅰ) there exist constants βi≥0(i=1,…,6) such that for any v∈IR
|I0(.;v)|≤β1+β2|v|p−1, | (2.4) |
|I1(.;v)|≤β3+β4|v|p/2−1, | (2.5) |
|I2(.;v)|≤β5+β6|v|p/2. | (2.6) |
(ⅱ) there exist constants μ1>0,μ2>0,μ3≥0,μ4≥0 such that for any (v,w)∈IR2:
μ1vI(.;v,w)+wG(.;v,w)≥μ2|v|p−μ3(μ1|v|2+|w|2)−μ4. | (2.7) |
In order to assure the uniqueness of solution we assume that
(H3) The Nemytskii operators I and G satisfy Carathéodory conditions and there exists some μ>0 such the operator Fμ:IR2→IR2 defined by
Fμ(.;v)=(μ(I(.;v))G(.;v)),∀v=(v,w)∈IR2, | (2.8) |
satisfies a one-sided Lipschitz condition (see e.g. Seidman et al. [62], Belmiloudi [14]): there exists a constant CL>0 such that (∀vi=(vi,wi)∈IR2,i=1,2)
(Fμ(.;v1)−Fμ(.;v2))⋅(v1−v2)≥−CL‖v1−v2‖2. | (2.9) |
Finally, we assume that the operators H and E which describe multiple time-delays related to ϕ and u are defined as in Belmiloudi [14] i.e.,
H(x,t;ϕτ,uτ)=n1∑k=1ak(x,t)ϕ(x,t−ξk(t))+n2∑l=1bl(x,t)u(x,t−ηl(t)),E(x,t;ϕτ,uτ)=n1∑k=1ck(x,t)ϕ(x,t−ξk(t))+n2∑l=1dl(x,t)u(x,t−ηl(t)), | (2.10) |
where ak,ck, bl and dl (for 1≤k≤n1 and 1≤l≤n2) are C∞ functions. For the functions ξk and ηl (for 1≤k≤n1 and 1≤l≤n2), we suppose that (as in [14]):
(RC) t∈[0,T)→(rk(t)=t−ξk(t),1≤k≤n1) and t∈[0,T)→(pl(t)=t−ηl(t),1≤l≤n2) are strictly increasing functions and (ξk(t),1≤k≤n1) and (ηl(t),1≤l≤n2) are C1 non-negative functions on [0,T). So we have the existence of inverse functions (ek)k of (rk)k and (ql)l of (pl)l, respectively. We also define the following subdivision: s−1=−δ(0)=max, s_0 = 0 and \forall j\in \mbox{IN}^{{*}} , s_j = \min\limits_{1\leq k\leq n_{1}}\min\limits_{1\leq l\leq n_{2}}\left(e_{k}(s_{j-1}), q_{l}(s_{j-1})\right) , and we denote T_{j} as T_{j} = s_{j}-s_{j-1} , \forall j\in \mbox{IN}^{{*}} . We introduce the following notations: I_{j} = (s_{-1}, s_{j}) and {\cal Q}_{j} = \Omega\times I_{j} for j\in \mbox{IN}^{} .
Remark 2.6. According to hypotheses (RC), we prove easily that:
(i) the sequences (s_{j})_{j\in \mathit{IN}^{{}}} is strictly increasing and s_{j}\leq T, \forall j\geq 0,
(ii) for j\geq 2 , if t\in(s_{j-1}, s_{j}) then \forall i = 1, n , r_{i}(t)\leq s_{j-1} , p_{i}(t)\leq s_{j-1},
(iii) if t\in(s_{0}, s_{1}) then \forall i = 1, n , r_{i}(t)\in (s_{-1}, s_{0}) , p_{i}(t)\in (s_{-1}, s_{0}).
Remark 2.7. The functions a_k, c_{k}, b_{k} and d_k are diffusion coefficients which represent the strength of each associated time-delay. A zero coefficient means that the associated previous state doesn't impact the system. Time-delays come from biological inhomogeneous properties of heart region. Electrical waves go through muscles, bones or fat which induce time-delays in their interaction in regards to ionical channels behavior.
Lemma 2.2. ([6]) Assume that F_{\mu} is differentiable with respect to (\phi, u) and denote by \lambda_1(\phi, u) \leq \lambda_2(\phi, u) the eigenvalues of the symmetrical part of Jacobian matrix \nabla F_\mu(\phi, u) :
Q_{\mu}(\phi, u) = \frac{1}{2}\left(\nabla F_\mu(\phi, u)^T + \nabla F_\mu(\phi, u) \right). |
If there exist a constant C_F independent of \phi and u such as:
\begin{equation} C_F \leq \lambda_1(\phi, u) \leq \lambda_2(\phi, u), \end{equation} | (2.11) |
then F_{\mu} satisfies the hypothesis (H3).
Lemma 2.3. ([6]) Let assumptions (2.3), (H1) and (H2) be fulfilled. For (\phi, u)\in L^p(\Omega) \times \mathbb{H} and a.e., t , there exist constants C_{i} > 0 ( i = 1, 6 ) such that
\begin{eqnarray} \Vert \mathcal{I}(., t;\phi, u)\Vert_{L^{p^\prime}(\Omega)} &\leq& C_1+C_2\Vert\phi\Vert^{p/p^\prime}_{L^{p}(\Omega)} +C_3\Vert u\Vert^{2/p^\prime}_{\mathbb{H}}, \end{eqnarray} | (2.12) |
\begin{eqnarray} \Vert \mathcal{G}(., t;\phi, u)\Vert_{L^{2}(\Omega)} &\leq& C_4+C_5\Vert\phi\Vert^{p/2}_{L^{p}(\Omega)} +C_6\Vert u\Vert_{\mathbb{H}}, \end{eqnarray} | (2.13) |
where p' is such that \frac{1}{p}+\frac{1}{p'} = 1 .
In the sequel we will always denote C some positive constant which may be different at each occurrence.
We now define the following forms
\begin{equation} \begin{array}{rcl} A_i(\psi, v) = \int\limits_\Omega \mathcal{K}_i\nabla\psi\cdot\nabla v d{\bf{x}}, \; \; \; A_e(\psi, v) = \int\limits_\Omega \mathcal{K}_e\nabla\psi\cdot\nabla v d{\bf{x}}. \end{array} \end{equation} | (2.14) |
Proposition 2.1. (i) A_i and A_e are symmetric bilinear continuous forms on \mathbb{V} and \mathbb{U} , respectively.
(ii) A_i and A_e are coercive on \mathbb{V} and \mathbb{U} , respectively (we denote by \alpha_i and \alpha_e their coercivity coefficients).
Proof. (ⅰ) and (ⅱ) are easily obtained providing that properties of tensors \mathcal{K}_i and \mathcal{K}_e and (2.2) are satisfied.
We can now write the weak formulation of problem (1.1) (for all v\in\mathbb{V}, \; v_e\in\mathbb{U} and \rho\in\mathbb{H} )
\begin{equation} { \begin{array}{lcr} \big\langle \mathfrak{c}_{m}\frac{\partial \phi}{\partial t}, v\big\rangle_{\mathbb{V}', \mathbb{V}}+ \int_\Omega \mathcal{I}(.;\phi, u)v d{\bf{x}} + A_i(\phi+\varphi_e, v) = \langle I_{i}, v\rangle_{\mathbb{V}', \mathbb{V}}+ \int_\Omega \mathcal{H}(., \phi_\tau, u_\tau) v d{\bf{x}}, \\ A_i(\phi + \varphi_e, v_e)+ A_e(\varphi_e, v_e) = \langle I, v_{e}\rangle_{\mathbb{V}', \mathbb{V}}, \\ \big( \frac{\partial u}{\partial t}, \rho\big)_{\mathbb{H}}+\int_\Omega \mathcal{G}(.;\phi, u)\rho d{\bf{x}} = \int_\Omega \mathcal{E}(.;\phi_\tau, u_\tau)\rho d{\bf{x}}, \\ \phi(., t = 0) = \phi_{0}, \; \; u(.t = 0) = u_{0}, \\ \phi(., t^\prime) = \phi_{past}(., t^\prime), \; \; u(., t^\prime) = u_{past}(., t^\prime) , \; \; t^\prime \in [-\delta(0), 0[. \end{array} } \end{equation} | (2.15) |
Theorem 2.1. ([18]) Let g \in \mathbb{V}' and \varphi \in \mathbb{U} be given. The variational equations
\begin{equation} (A_i + A_e)(\underline{\varphi}_e, v_e) +A_i(\varphi, v_e) = 0, \; \; \forall v_e \in \mathbb{U} \end{equation} | (2.16) |
and
\begin{equation} (A_i + A_e)(\overline{\varphi}_e, v_e) = \langle g, v_{e}\rangle_{\mathbb{V}', \mathbb{V}}, \; \; \forall v_e \in \mathbb{U} \end{equation} | (2.17) |
have unique solutions \underline{\varphi}_e, \overline{\varphi}_e \in \mathbb{U} . Moreover we have that the operator \underline{A}_i: (\varphi, v)\in(\mathbb{U})^{2} \to \underline{A}_i(\varphi, v) = {A}_i(\varphi, v)+{A}_i(\underline{\varphi}_e, v) is symmetric bilinear continuous forms on \mathbb{U} .
Introduce the following spaces for S_{0} < S_{f} be fixed real values (where {\cal Q}_{S} = \Omega\times (S_{0}, S_{f}) , p\geq 2 and \frac{1}{p}+\frac{1}{p'} = 1 )
\begin{equation*} \begin{array}{lcr} \mathbb{D}_{p}(S_{0}, S_{f}) = L^{p'}({\cal Q}_{S})+L^{2}(S_{0}, S_{f}; \mathbb{V}')\subset L^{p'}(S_{0}, S_{f}; \mathbb{V}'), \\ \mathbb{W}_{p}(S_{0}, S_{f}) = \big\{u\in L^{p}({\cal Q}_{S})\cap L^{2}(S_{0}, S_{f}; \mathbb{V}) \; \text{such that}\; \frac{\partial u}{\partial t} \in \mathbb{D}_{p}(S_{0}, S_{f})\big\}. \end{array} \end{equation*} |
Lemma 2.4. ([6,18]) Let \pi_{m} be a sequence converging toward \pi in \mathbb{W}_{p}(S_{0}, S_{f}) weakly and in L^{2}({\cal Q}_{S}) strongly and V_{m} be a sequence converging toward V in L^{2}({\cal Q}_{S})\cap H^{1}(S_{0}, S_{f}; \mathbb{H}) weakly. Then we have the following convergence results:
(i) {\cal I}_{0}(.; \pi_{m}) \rightharpoonup {\cal I}_{0}(.; \pi) weakly in L^{p'}({\cal Q}_{S})
(ii) \mathcal{I}_{2}(.; \pi_{m}) \rightharpoonup \mathcal{I}_{2}(.; \pi) weakly in L^{2}({\cal Q}_{S})
(iii) {\cal I}_{1}(.; \pi_{m})V_{m} \rightharpoonup {\cal I}_{1}(.; \pi)V weakly in L^{2}({\cal Q}_{S}) .
The considered functions \mathcal{I}_{i} , in this paper, include the three classical type models in which these assumptions are satisfied (for the proof, we use similar arguments as in [18]) namely the Rogers-McCulloch [51] (RM), Fitz-Hugh-Nagumo [37] (FHN) and Aliev-Panfilov [54](LAP) models as follows. The function \mathcal{I}_{0} is defined by a cubic reaction term of the form \mathcal{I}_{0}(.; v) = b_{1}(.)v(v-r)(v-1), and the functions \mathcal{I}_{1} and \mathcal{I}_{2} are given by
\begin{equation*} \begin{array}{lcr} \text{(a) for RM type model}\; : \mathcal{I}_{1}(.;v) = b_{2}(.)v, \; \; \mathcal{I}_{2}(;, v) = -b_{3}(.)v, \\ \text{(b) for FHN type model}\; : \mathcal{I}_{1}(.;v) = b_{2}(.), \; \; \mathcal{I}_{2}(;, v) = -b_{3}(.)v, \\ \text{(c) for LAP type model}\; : \mathcal{I}_{1}(.;v) = b_{2}(.)v, \; \; \mathcal{I}_{2}(;, v) = b_{3}(.)v(r+1-v), \end{array} \end{equation*} |
where b_{i}\in W^{1, \infty}({\cal Q}) , i = 1, 3 , are sufficiently regular functions from {\cal Q} into \mbox{IR}^{{+, *}} and r\in[0, 1] . We obtain easily the following Lemma.
Lemma 2.5. The following properties hold:
1. For all v_{1} , v_{2} in \mathit{\mbox{IR}}^{} we have
\begin{equation*} \begin{array}{lcr} \mathcal{I}_{0}(.;v_{1})-\mathcal{I}_{0}(.;v_{2}) \quad = b_{1}(v_{1}-v_{2})\left(v_{1}^{2}+v_{2}^{2}+v_{1}v_{2}-(r+1)(v_{1}+v_{2})+r\right) \mathit{\; \text{and}}\;\\ \mathit{\text{(a) for RM type model}}\;:\; \; \mathcal{I}_{1}(.;v_{1})-\mathcal{I}_{1}(.;v_{2}) = b_{2}(v_{1}-v_{2}), \; \; \mathcal{I}_{2}(.;v_{1})-\mathcal{I}_{2}(.;v_{2}) = -b_{3}(v_{1}-v_{2}), \\ \mathit{\text{(b) for FHN type model}}\;:\; \; \mathcal{I}_{1}(.;v_{1})-\mathcal{I}_{1}(.;v_{2}) = 0, \; \; \mathcal{I}_{2}(.;v_{1})-\mathcal{I}_{2}(.;v_{2}) = -b_{3}(v_{1}-v_{2}), \\ \mathit{\text{(c) for LAP type model}}\;:\; \; \mathcal{I}_{1}(.;v_{1})-\mathcal{I}_{1}(.;v_{2}) = b_{2}(v_{1}-v_{2}), \\ \quad \mathcal{I}_{2}(.;v_{1})-\mathcal{I}_{2}(.;v_{2}) = b_{3}(v_{1}-v_{2})((r+1)-v_{1}-v_{2}). \end{array} \end{equation*} |
2. The partial derivative of the function \mathcal{I}_{0} is given by \frac{\partial \mathcal{I}_{0}}{\partial v}(.; v) = b_{1}\left(3v^{2}-2(r+1)v+r\right) and these of the functions \mathcal{I}_{1} and \mathcal{I}_{2} are given by
\begin{equation*} \begin{array}{lcr} \mathit{\text{(a) for RM type model}}\;: \frac{\partial \mathcal{I}_{1}}{\partial v}(.;v) = b_{2}, \; \; \frac{\partial \mathcal{I}_{2}}{\partial v}(.;v) = -b_{3}, \\ \mathit{\text{(b) for FHN type model}}\;: \frac{\partial \mathcal{I}_{1}}{\partial v}(.;v) = 0, \; \; \frac{\partial \mathcal{I}_{2}}{\partial v}(.;v) = -b_{3}, \\ \mathit{\text{(c) for LAP type model}}\;: \frac{\partial \mathcal{I}_{1}}{\partial v}(.;v) = b_{2}, \; \; \frac{\partial \mathcal{I}_{2}}{\partial v}(.;v) = b_{3}(r+1-2v). \end{array} \end{equation*} |
Remark 2.8. According to Lemma 2.5, the partial derivatives of {\cal I} and {\cal G} are given by
(a) for RM type model:
\begin{equation} \begin{array}{lcr} \frac{\partial {\cal I}}{\partial \phi} = b_{1}(3\phi^{2}-2(1+r)\phi+r)+b_{2}u, \; \frac{\partial {\cal I}}{\partial u} = b_{2}\phi, \; \; \frac{\partial {\cal G}}{\partial \phi} = -b_{3}, \; \frac{\partial {\cal G}}{\partial u} = \hbar, \end{array} \end{equation} | (2.18) |
(b) for FHN type model:
\begin{equation} \begin{array}{lcr} \frac{\partial {\cal I}}{\partial \phi} = b_{1}(3\phi^{2}-2(1+r)\phi+r), \; \frac{\partial {\cal I}}{\partial u} = b_{2}, \; \; \frac{\partial {\cal G}}{\partial \phi} = -b_{3}, \; \frac{\partial {\cal G}}{\partial u} = \hbar, \end{array} \end{equation} | (2.19) |
(c) for LAP type model:
\begin{equation} \begin{array}{lcr} \frac{\partial {\cal I}}{\partial \phi} = b_{1}(3\phi^{2}-2(1+r)\phi+r)+b_{2}u, \; \frac{\partial {\cal I}}{\partial u} = b_{2}\phi, \; \; \frac{\partial {\cal G}}{\partial \phi} = b_{3}(r+1-2\phi), \; \frac{\partial {\cal G}}{\partial u} = \hbar. \end{array} \end{equation} | (2.20) |
Consequently, {\cal I}_{i} (for i = 0, 2 ) and the partial derivatives of {\cal I} and {\cal G} for this three models are of the form
\begin{equation} \begin{array}{lcr} {\cal I}_{0} = \hbar_{11}\phi^{3}-\hbar_{12}\phi^{2}+\hbar_{13}\phi, \; \; {\cal I}_{1} = \epsilon_{1}\hbar_{21}\phi+\hbar_{22}, {\cal I}_{2} = -\epsilon_{2} \hbar_{31}\phi^{2}+(2\epsilon_{2} -1)\hbar_{32}\phi, \\ \frac{\partial {\cal I}}{\partial \phi} = \frac{\partial {\cal I}_{0}}{\partial \phi}+u \frac{\partial {\cal I}_{1}}{\partial \phi} = 3\hbar_{11}\phi^{2}-2\hbar_{12}\phi+\hbar_{13}+\epsilon_{1}\hbar_{21}u, \; \; \; \frac{\partial {\cal I}}{\partial u} = {\cal I}_{1} = \epsilon_{1}\hbar_{21}\phi+\hbar_{22}, \\ \frac{\partial {\cal G}}{\partial \phi} = \frac{\partial {\cal I}_{2}}{\partial \phi} = -2\epsilon_{2} \hbar_{31}\phi+(2\epsilon_{2} -1)\hbar_{32}, \; \; \; \; \frac{\partial {\cal G}}{\partial u} = \hbar, \end{array} \end{equation} | (2.21) |
where \hbar_{ij} and \hbar are sufficiently regular and bounded functions from {\cal Q} into [h_{0}, +\infty[, with h_{0}\in \mbox{IR}^{{+, *}} and (\epsilon_{1}, \epsilon_{2})\in \big\{ (1, 0), (0, 0), (1, 1) \big\} .
For delay operators we have the following estimates.
Lemma 2.6. Let (v, \rho) be in (L^{q}(0, T; L^{\sigma}(\Omega)))^{2} , with \sigma, q\in [1, \infty[, such that on the domain {\cal Q}_{0} , (v, \rho) = (v_{past}, \rho_{past})\in (L^{q}(-\delta(0), 0;L^{\sigma}(\Omega)))^{2} . Then the following estimates hold.
(i) There exists a constant C_{\infty, 0} > 0 (depending on \Vert a_{k} \Vert_{\infty}, \Vert c_{k} \Vert_{\infty} , \Vert b_{l} \Vert_{\infty}, \Vert d_{l} \Vert_{\infty} , 1\leq k\leq n_{1} , 1\leq l\leq n_{2} ) such that
\begin{equation} \begin{array}{lcr} \left\| { \mathcal{H}(v_{\tau}, \rho_{\tau})} \right\| _{L^{\sigma}(\Omega)} \leq C_{\infty, 0} \left(\sum\limits_{k = 1}^{n_{1}}\Vert v(., r_{k}(t)) \Vert_{L^{\sigma}(\Omega)}+\sum\limits_{l = 1}^{n_{2}}\Vert \rho(., p_{l}(t))\Vert_{L^{\sigma}(\Omega)}\right), \\ \left\| { \mathcal{E}(v_{\tau}, \rho_{\tau})} \right\| _{L^{\sigma}(\Omega)} \leq C_{\infty, 0} \left(\sum\limits_{k = 1}^{n_{1}}\Vert v(., r_{k}(t)) \Vert_{L^{\sigma}(\Omega)}+\sum\limits_{l = 1}^{n_{2}}\Vert \rho(., p_{l}(t))\Vert_{L^{\sigma}(\Omega)}\right). \end{array} \end{equation} | (2.22) |
(ii) There exists a constant C_{\infty, 1} > 0 (depending on \Vert a_{k} \Vert_{\infty}, \Vert c_{k} \Vert_{\infty} , \Vert a'_{k} \Vert_{\infty}, \Vert c'_{k} \Vert_{\infty} , \Vert b_{l} \Vert_{\infty}, \Vert d_{l} \Vert_{\infty} , \Vert b'_{l} \Vert_{\infty}, \Vert d'_{l} \Vert_{\infty} , 1\leq k\leq n_{1} , 1\leq l\leq n_{2} ) such that
\begin{equation} \begin{array}{lcr} \left\| { \mathcal{H}(v_{\tau}, \rho_{\tau})} \right\| _{L^{q}(0, t;L^{\sigma}(\Omega))} \leq C_{\infty, 1}\Big(\Vert v \Vert_{L^{q}(0, t;L^{\sigma}(\Omega))}+\Vert \rho \Vert_{L^{q}(0, t;L^{\sigma}(\Omega))}\\ \quad +\Vert v_{past}\Vert_{L^{q}(-\delta(0), 0;L^{\sigma}(\Omega))}+\Vert \rho_{past}\Vert_{L^{q}(-\delta(0), 0;L^{\sigma}(\Omega))}\Big), \\ \left\| { \mathcal{E}(v_{\tau}, \rho_{\tau})} \right\| _{L^{q}(0, t;L^{\sigma}(\Omega))} \leq C_{\infty, 1}\Big(\Vert v \Vert_{L^{q}(0, t;L^{\sigma}(\Omega))}+\Vert \rho \Vert_{L^{q}(0, t;L^{\sigma}(\Omega))}\\ \quad +\Vert v_{past}\Vert_{L^{q}(-\delta(0), 0;L^{\sigma}(\Omega))}+\Vert \rho_{past}\Vert_{L^{q}(-\delta(0), 0;L^{\sigma}(\Omega))}\Big), \end{array} \end{equation} | (2.23) |
Proof. (ⅰ) According to regularity of (a_{k})_{1\leq k\leq n_{1}} , (c_{k})_{1\leq k\leq n_{1}} , (b_{l})_{1\leq l\leq n_{2}} , (c_{l})_{1\leq l\leq n_{2}} and to Remark 2.6, we obtain (for 1\leq k\leq n_{1} , 1\leq l\leq n_{2} and T\geq t\geq 0 )
\begin{equation} \begin{array}{lcr} \int_\Omega \mid \tilde{a}_{k}(x, t)v({\bf x}, r_{k}(t))\mid^{\sigma} d{\bf x} \leq \Vert \tilde{a}_{k} \Vert_{\infty}^{\sigma}\Vert v(., r_{k}(t)) \Vert_{L^{\sigma}(\Omega)}^{\sigma}, \; \; \text{(for $\tilde{a}_{k} = a_{k}$ or $c_{k}$)}\\ \int_\Omega \mid \tilde{b}_{l}(x, t)\rho({\bf x}, p_{l}(t))\mid^{\sigma}d{\bf x}\leq \Vert \tilde{b}_{l} \Vert_{\infty}^{\sigma}\Vert \rho(., p_{l}(t)) \Vert_{L^{\sigma}(\Omega)}^{\sigma}, \; \; \text{(for $\tilde{b}_{l} = b_{l}$ or $d_{l}$)}\\ \end{array} \end{equation} | (2.24) |
Then, from the expression of \mathcal{H} and \mathcal{E} , we can deduce that
\begin{equation} \begin{array}{lcr} \left\| { \mathcal{H}(v_{\tau}(., t), \rho_{\tau}(., t))} \right\| _{L^{\sigma}(\Omega)}\leq D_{1, 0}\left(\sum\limits_{k = 1}^{n_{1}}\left\| { v(., r_{k}(t))} \right\| _{L^{\sigma}(\Omega)}+ \sum\limits_{l = 1}^{n_{2}}\left\| { \rho(., p_{l}(t))} \right\| _{L^{\sigma}(\Omega)}\right), \\ \left\| { \mathcal{E}(v_{\tau}(., t), \rho_{\tau}(., t))} \right\| _{L^{\sigma}(\Omega)}\leq D_{2, 0}\left(\sum\limits_{k = 1}^{n_{1}}\left\| { v(., r_{k}(t))} \right\| _{L^{\sigma}(\Omega)}+ \sum\limits_{l = 1}^{n_{2}}\left\| { \rho(., p_{l}(t))} \right\| _{L^{\sigma}(\Omega)}\right), \end{array} \end{equation} | (2.25) |
where {D_{1,0}} = \max (\mathop {\max }\limits_{1 \le k \le {n_1}} {\left\| {{a_k}} \right\|_\infty },\mathop {\max }\limits_{1 \le l \le {n_2}} {\left\| {{b_l}} \right\|_\infty }),{D_{2,0}} = \max (\mathop {\max }\limits_{1 \le k \le {n_1}} {\left\| {{c_k}} \right\|_\infty },\mathop {\max }\limits_{1 \le l \le {n_2}} {\left\| {{d_l}} \right\|_\infty }) .
(ⅱ) Setting \theta = r_{k}(s) (resp. \theta = p_{l}(s) ), we have s = e_{k}(\theta) (resp. s = q_{l}(\theta) ) and then ds = e'_{k}(\theta)d\theta (resp. ds = q'_{l}(\theta)d\theta ). So
\begin{equation} \begin{array}{lcr} \Vert v(., r_{k}(.)) \Vert^{q}_{L^{q}(0, t;L^{\sigma}(\Omega))} = \int_{0}^{t} \Vert v(., r_{k}(s)) \Vert^{q}_{L^{\sigma}(\Omega)}ds\leq \Vert e'_{k}\Vert_{\infty}\left(\int_{-\xi_{k}(0)}^{t-\xi_{k}(t)} \Vert v(., \theta) \Vert^{q}_{L^{\sigma}(\Omega)}d\theta \right), \\ \Vert \rho(., r_{k}(.)) \Vert^{q}_{L^{q}(0, t;L^{\sigma}(\Omega))} = \int_{0}^{t} \Vert \rho(., p_{l}(s)) \Vert^{q}_{L^{\sigma}(\Omega)}ds\leq \Vert q'_{l}\Vert_{\infty}\left(\int_{-\eta_{l}(0)}^{t-\eta_{l}(t)} \Vert \rho(., \theta) \Vert^{q}_{L^{\sigma}(\Omega)}d\theta \right). \end{array} \end{equation} | (2.26) |
Since -\delta(0)\leq -\xi_{k}(0) , -\delta(0)\leq -\eta_{k}(0) , t-\xi_{k}(t)\leq t and t-\eta_{k}(t)\leq t we can deduce that (since v = v_{past}, \; \; \rho = \rho_{past}, \; \; \text{on}\; {\cal Q}_{0} )
\begin{equation} \begin{array}{lcr} \Vert v(., r_{k}(.)) \Vert^{q}_{L^{q}(0, t;L^{\sigma}(\Omega))}\leq \Vert e'_{k}\Vert_{\infty}\left(\int_{0}^{t} \Vert v(., \theta) \Vert^{q}_{L^{\sigma}(\Omega)}d\theta+\int_{-\delta(0)}^{0} \Vert v_{past}(., s) \Vert^{q}_{L^{\sigma}(\Omega)} ds\right)\\ \quad \leq \max\limits_{1\leq k\leq n_{1}}\left(\Vert e'_{k}\Vert_{\infty}\right)\left(\Vert v \Vert^{q}_{L^{q}(0, t;L^{\sigma}(\Omega))}+ \Vert v_{past}\Vert^{q}_{L^{q}(-\delta(0), 0;L^{\sigma}(\Omega))}\right), \\ \Vert \rho(., r_{k}(.)) \Vert^{q}_{L^{q}(0, t;L^{\sigma}(\Omega))}\leq \Vert q'_{l}\Vert_{\infty}\left(\int_{0}^{t} \Vert \rho(., \theta) \Vert^{q}_{L^{\sigma}(\Omega)}d\theta+ \int_{-\delta(0)}^{0} \Vert \rho_{past}(., s) \Vert^{q}_{L^{\sigma}(\Omega)} ds\right)\\ \quad \leq \max\limits_{1\leq l\leq n_{2}}\left(\Vert q'_{l}\Vert_{\infty}\right)\left(\Vert \rho \Vert^{q}_{L^{q}(0, t;L^{\sigma}(\Omega))}+ \Vert \rho_{past}\Vert^{q}_{L^{q}(-\delta(0), 0;L^{\sigma}(\Omega))}\right), \end{array} \end{equation} | (2.27) |
and then, from (2.25) and Jensen inequality, we can deduce the result (ⅱ) of Lemma. This completes the proof.
For the sake of simplicity, we shall write \mathcal{I}_{i}(\psi) , {\cal I}(\psi, v) and {\cal G}(\psi, v) in place of \mathcal{I}_{i}(x, t; \psi) , {\cal I}(x, t; \psi, v) and {\cal G}(x, t; \psi, v) , respectively (for i = 0, 2 ).
The results of this section concern the existence, uniqueness and regularity of solution of (1.1).
Theorem 2.2. ([6]) Let assumptions (H1)-(H3) and (RC) be fulfilled. Let be given (\phi_{0}, u_{0})\in (L^{2}(\Omega))^{2} , (\phi_{past}, u_{past})\in (L^{2}({\cal Q}_{0}))^{2} and (I_{i}, I)\in \big(L^2(0, T; \mathbb{V}')\big)^{2} . Then there exists a solution (\phi, \varphi_{e}, u) of (2.15) verifying : \phi\in L^2(0, T; \mathbb{V})\cap L^p(\mathcal{Q})\cap L^\infty(0, T; \mathbb{H}), \; \; \varphi_e \in L^2(0, T; \mathbb{U}) \mathit{\; \text{and}}\; u\in C^{0}([0, T]; \mathbb{H}) with the following a priori estimate
\begin{equation} \begin{array}{lcr} \Vert \phi \Vert_{L^{2}(0, T;\mathbb{V})\cap L^\infty(0, T;\mathbb{H})}^2 + \Vert u\Vert_{L^\infty(0, T;\mathbb{H})}^2 + \Vert \varphi_{e}\Vert_{L^{2}(0, T;\mathbb{U})}^2 \leq C\Big(1+\Vert I_{i}\Vert_{L^2(0, T;\mathbb{V}')}^2 + \Vert I\Vert_{L^2(0, T;\mathbb{V}')}^2 \\ \quad + \Vert \phi_{past}\Vert_{L^{2}({\cal Q}_{0})}^2+\Vert u_{past}\Vert_{L^{2}({\cal Q}_{0})}^2+\Vert \phi_{0}\Vert_\mathbb{H}^2 + \Vert u_{0}\Vert_\mathbb{H}^2\Big). \end{array} \end{equation} | (2.28) |
Moreover the Lipschitz continuity relation is satisfied, i.e., for any element (\phi_{0j}, u_{0j})\in (L^{2}(\Omega))^{2} , (I^{(j)}_{i}, I^{(j)})\in (L^2(0, T; \mathbb{V}'))^{2} and (\phi_{j, past}, u_{j, past})\in (L^{2}({\cal Q}_{0}))^{2} , for j = 1, 2 , we have
\begin{equation} \begin{array}{lcr} \Vert \phi_{1}-\phi_{2} \Vert_{L^{2}(0, T;\mathbb{V})\cap L^\infty(0, T;\mathbb{H})}^2 + \Vert u_{1}-u_{2}\Vert_{L^\infty(0, T;\mathbb{H})}^2 + \Vert \varphi_{e, 1}-\varphi_{e, 2} \Vert_{L^{2}(0, T;\mathbb{U})}^2 \\ \quad \leq C\big(\Vert I^{(1)}_{i}-I^{(2)}_{i}\Vert_{L^2(0, T;\mathbb{V}')}^2 + \Vert I^{(1)}-I^{(2)}\Vert_{L^2(0, T;\mathbb{V}')}^2 + \Vert \phi_{1, past}-\phi_{2, past}\Vert_{L^{2}({\cal Q}_{0})}^2\\ \quad +\Vert u_{1, past}-u_{2, past}\Vert_{L^{2}({\cal Q}_{0})}^2+\Vert \phi_{01}-\phi_{02} \Vert_\mathbb{H}^2 + \Vert u_{01}-u_{02} \Vert_\mathbb{H}^2\big), \end{array} \end{equation} | (2.29) |
where (\phi_{j}, u_{j}, \varphi_{e, j}) is solution of (2.15), which corresponds to data (\phi_{0j}, u_{0j}) , (\phi_{j, past}, u_{j, past}) and (I^{(j)}_{i}, I^{(j)}) .
Theorem 2.3. Consider the case of p = 4 . Assume that (u_{0}, u_{past}, \phi_{past}, \phi_{0}) is given such that (\phi_{past}, u_{past})\in \big(L^{2}(-\delta(0), 0;L^{3}(\Omega))\big)^{2} , u_{0}\in L^{3}(\Omega) and \phi(t\! = \!0)\! = \! \varphi_{i}(t\! = \! 0)-\varphi_{e}(t\! = \! 0)\! = \! \varphi_{i}^{(0)}-\varphi_{e}^{(0)}\! = \! \phi_{0} with (\phi_{0}, \varphi_{e}^{(0)})\in (L^{2}(\Omega))^{2} .
(i) If I_{i}\in L^2({\cal Q}) and I\in L^2({\cal Q}) , we have u belongs even to C^{0}([0, T], L^{3}(\Omega)) and it holds that
\begin{equation} \begin{array}{l} \left\| { u(., t)} \right\| _{L^{3}(\Omega)}\leq C\Big(1+\left\| { \phi_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}+\left\| { u_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}\\ \quad + \Vert I_{i}\Vert_{L^2({\cal Q})}+ \Vert I\Vert_{L^2({\cal Q})}+\left\| { u_{0}} \right\| _{L^{3}(\Omega)}+\Vert \phi_{0}\Vert_{L^{2}(\Omega)}\Big). \end{array} \end{equation} | (2.30) |
(ii) Moreover if \mathcal{I}_{2} satisfies the following assumption
(H4) there exist constants \beta_{i}\geq 0 ( i = 7, ..., 9 ) such that, for any (v, w)\in \mathit{\mbox{IR}}^{{2}} ,
\mid \mathcal{I}_{2}(.;v)- \mathcal{I}_{2}(.;w)\mid \leq \mid v-w\mid \big(\beta_{7}+\beta_{8}\mid v\mid +\beta_{9}\mid w\mid \big), |
we have for any element (I^{(j)}_{i}, I^{(j)})\in (L^2({\cal Q}))^{2} , for j = 1, 2 ,
\begin{equation} \begin{array}{lcr} \left\| { u(., t)} \right\| _{L^{3}(\Omega)}\leq C\left(\Vert I_{i}\Vert_{L^2({\cal Q})}+ \Vert I\Vert_{L^2({\cal Q})}\right), \end{array} \end{equation} | (2.31) |
where (\phi_{j}, u_{j}, \varphi_{e, j}) is solution of (2.15), which corresponds to data (\phi_{0}, u_{0}) , (\phi_{past}, u_{past}) and (I^{(j)}_{i}, I^{(j)}) , and \phi = \phi_{1}-\phi_{2} , u = u_{1}-u_{2} , \varphi_{e} = \varphi_{e, 1}-\varphi_{e, 2} , I = I^{(1)}-I^{(2)} , I_{i} = I_{i}^{(1)}-I_{i}^{(2)} .
(iii) Assume now that (\phi_{0}, \varphi_{e}^{(0)}) \in (H^{1}(\Omega))^{2} and the primitive \tilde{\mathcal{I}}_{0} of \mathcal{I}_{0} satisfies the assumptions
(H5) there exist constants \beta_{i}\geq 0 ( i = 10, ..., 15 ) such that, for any v\in \mathit{\mbox{IR}}^{} ,
\begin{equation*} \begin{array}{lcr} \tilde{\mathcal{I}}_{0}(v)\geq \beta_{10}\mid v\mid^{4}-\beta_{11}\mid v\mid^{2}, \; \; \; \; \frac{\partial \tilde{\mathcal{I}}_{0}}{\partial t}(v)\geq \beta_{12}\mid v\mid^{4}-\beta_{13}\mid v\mid^{2}, \\ \mid \tilde{\mathcal{I}}_{0}(v)\mid \leq \beta_{14}+\beta_{15}\mid v\mid^{4}. \end{array} \end{equation*} |
Then
(a) if I_{i}\in L^{2}({\cal Q}) and I is in the space
U_{c} = \big\{v\in L^{2}({\cal Q}) \mathit{\; \; \text{such that}}\; \; \frac{\partial v}{\partial t}\in L^{2}({\cal Q})\big\}\subset C^{0}([0, T]; L^{2}(\Omega)) \mathit{\; \text{(see Remark 2.3)}}, |
then (\phi, \varphi_{e}) \in \big(L^{\infty}(0, T; H^{1}(\Omega))\big)^{2} , \frac{\partial \phi}{\partial t}\in L^{2}({\cal Q}) and u\in C^{0}([0, T]; L^{3}(\Omega)) .
(b) Moreover if \mathcal{I}_{0} and \mathcal{I}_{1} satisfy the following assumption
(H6) there exist constants \beta_{i}\geq 0 ( i = 16, ..., 19 ) such that, for any (v, w)\in \mathit{\mbox{IR}}^{{2}} ,
\begin{equation*} \begin{array}{lcr} \mid \mathcal{I}_{0}(.;v)- \mathcal{I}_{0}(.;w)\mid \leq \mid v-w\mid \left(\beta_{16}+\beta_{17}\mid v\mid^{2} +\beta_{18}\mid w\mid^{2}\right), \\ \mid \mathcal{I}_{1}(.;v)- \mathcal{I}_{1}(.;w)\mid \leq \beta_{19}\mid v-w\mid, \end{array} \end{equation*} |
we have, for any element (I^{(j)}_{i}, I^{(j)})\in L^2({\cal Q})\times U_{c} (for j = 1, 2 ),
\begin{equation} \begin{array}{lcr} \left\| { \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}({\cal Q})}+ \left\| { \frac{\partial u}{\partial t}} \right\| ^{2}_{L^{2}(0, T;L^{3}(\Omega))}+\left\| { \phi} \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { \varphi_e} \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}\\ \quad \leq C\left(\left\| { I_{i} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { I } \right\| ^{2}_{U_{c}}\right). \end{array} \end{equation} | (2.32) |
where (\phi_{j}, u_{j}, \varphi_{e, j}) is solution of (2.15), which corresponds to data (\phi_{0}, u_{0}) , (\phi_{past}, u_{past}) and (I^{(j)}_{i}, I^{(j)}) , and \phi = \phi_{1}-\phi_{2} , u = u_{1}-u_{2} , \varphi_{e} = \varphi_{e, 1}-\varphi_{e, 2} , I = I^{(1)}-I^{(2)} , I_{i} = I_{i}^{(1)}-I_{i}^{(2)} .
Proof. (ⅰ) Since u satisfies the equation
\begin{equation} \frac{\partial u }{\partial t} = -\mathcal{I}_{2}(.;\phi)-\hbar u+\mathcal{E}(\phi_{\tau}, u_{\tau}), \; \text{in}\; {\cal Q} \end{equation} | (2.33) |
where \mathcal{E}(\phi_{\tau}, u_{\tau}) = \sum\limits_{k = 1}^{n_{1}}\! c_k({\bf{x}}, t) \phi({\bf{x}}, t-\xi_{k}(t))+\sum\limits_{l = 1}^{n_{2}}\! d_l({\bf{x}}, t) u({\bf{x}}, t-\eta_{l}(t)) , then we have (for all t )
\begin{equation} \begin{array}{lcr} u({\bf{x}}, t) \! = \! u_{0}-\int_{0}^{t}\mathcal{I}_{2}({\bf{x}}, s;\phi)ds-\int_{0}^{t}\hbar u({\bf{x}}, s)ds+\sum\limits_{k = 1}^{n_{1}}\! \int_{0}^{t} c_k({\bf{x}}, s)\phi({\bf{x}}, s-\xi_{k}(s))ds\\ \quad +\sum\limits_{l = 1}^{n_{2}}\! \int_{0}^{t}d_l({\bf{x}}, s) u({\bf{x}}, s-\eta_{l}(s))ds. \end{array} \end{equation} | (2.34) |
Consequently (as in (2.26))
\begin{equation} \begin{array}{lcr} u({\bf{x}}, t) \! = \! u_{0}-\int_{0}^{t}\mathcal{I}_{2}({\bf{x}}, s;\phi)ds-\int_{0}^{t}\hbar u({\bf{x}}, s)ds+\sum\limits_{k = 1}^{n_{1}}\! \int_{-\xi_{k}(0)}^{t-\xi_{k}(t)}e'_{k}(\theta)c_k({\bf{x}}, e_k(\theta)) \phi({\bf{x}}, \theta)d\theta\\ \quad +\sum\limits_{l = 1}^{n_{2}}\! \int_{-\eta_{l}(0)}^{t-\eta_{l}(t)}q'_{l}(\theta)d_l({\bf{x}}, q_{l}(\theta)) u({\bf{x}}, \theta)d\theta. \end{array} \end{equation} | (2.35) |
Since -\delta(0) \leq -\xi_{k}(0) , -\delta(0) \leq -\eta_{k}(0) , t -\xi_{k}(t) \leq t and t -\eta_{k}(t) \leq t we can deduce that (according to the regularity of c_{k} , d_{l} and \hbar )
\begin{equation} \begin{array}{lcr} \mid u({\bf{x}}, t)\mid \leq C\Big(\mid u_{0}\mid+\int_{0}^{t}\mid\mathcal{I}_{2}({\bf{x}}, s;\phi)\mid ds+\int_{0}^{t}\mid u({\bf{x}}, s)\mid ds+ \int_{-\delta(0)}^{t}\mid \phi({\bf{x}}, \theta)\mid d\theta\\ \quad + \int_{-\delta(0)}^{t}\mid u({\bf{x}}, \theta)\mid d\theta\Big) \end{array} \end{equation} | (2.36) |
and then (since from the assumption (2.6) we have \mid \mathcal{I}_{2}(.; \phi) \mid \leq \beta_5 + \beta_6\mid \phi \mid^{2} )
\begin{equation} \begin{array}{lcr} \mid u({\bf{x}}, t)\mid \leq C\Big(1+\int_{0}^{t}\mid \phi({\bf{x}}, s)\mid^{2} ds+\int_{0}^{t}\mid u({\bf{x}}, s)\mid ds+ \int_{0}^{t}\mid \phi({\bf{x}}, \theta)\mid d\theta\\ \quad +\mid u_{0}\mid+ \int_{-\delta(0)}^{0}\mid \phi_{past}({\bf{x}}, \theta)\mid d\theta+ \int_{-\delta(0)}^{0}\mid u_{past}({\bf{x}}, \theta)\mid d\theta\Big). \end{array} \end{equation} | (2.37) |
This implies
\begin{equation} \begin{array}{lcr} \left(\int_{\Omega}\mid u({\bf{x}}, t)\mid^{3} d{\bf{x}}\right)^{1/3} \leq C\left(1+\left\| { u_{0}} \right\| _{L^{3}(\Omega)}+\Big[\int_{\Omega}\big(\int_{0}^{t}\mid u({\bf{x}}, s)\mid ds\big)^{3}d{\bf{x}}\Big]^{1/3} \right.\\ \quad \left.+\Big[\int_{\Omega}\big(\int_{0}^{t}\mid \phi({\bf{x}}, s)\mid^{2} ds\big)^{3}d{\bf{x}}\Big]^{1/3}+ \Big[\int_{\Omega}\big(\int_{0}^{t}\mid \phi({\bf{x}}, \theta)\mid d\theta \big)^{3}d{\bf{x}}\Big]^{1/3} \right. \\ \quad \left.+ \Big[\int_{\Omega}\big(\int_{-\delta(0)}^{0}\mid \phi_{past}({\bf{x}}, \theta)\mid d\theta\big)^{3}d{\bf{x}}\Big]^{1/3}+ \Big[\int_{\Omega}\big(\int_{-\delta(0)}^{0}\mid u_{past}({\bf{x}}, \theta)\mid d\theta\big)^{3}d{\bf{x}}\Big]^{1/3}\right) \end{array} \end{equation} | (2.38) |
and then (using Minkowski inequality)
\begin{equation} \begin{array}{lcr} \left(\int_{\Omega}\mid u({\bf{x}}, t)\mid^{3} d{\bf{x}}\right)^{1/3} \leq C\left(1+\left\| { u_{0}} \right\| _{L^{3}(\Omega)}+\int_{0}^{t}\big(\int_{\Omega}\mid u({\bf{x}}, s)\mid^{3} d{\bf{x}}\big)^{1/3}ds \right.\\ \quad \left.+\int_{0}^{t}\big(\int_{\Omega}\mid \phi({\bf{x}}, s)\mid^{6} d{\bf{x}}\big)^{1/3}ds + \int_{0}^{t}\big(\int_{\Omega}\mid \phi({\bf{x}}, \theta)\mid^{3} d{\bf{x}}\big)^{1/3}d\theta \right.\\ \quad \left.+ \int_{-\delta(0)}^{0}\big(\int_{\Omega}\mid \phi_{past}({\bf{x}}, \theta)\mid^{3} d{\bf{x}}\big)^{1/3}d\theta + \int_{-\delta(0)}^{0}\big(\int_{\Omega}\mid u_{past}({\bf{x}}, \theta)\mid^{3} d{\bf{x}}\big)^{1/3}d\theta\right). \end{array} \end{equation} | (2.39) |
Since \phi \in L^{2}(0, T, H^{1}(\Omega))\subset L^{2}(0, T, L^{r}(\Omega)) ( r\in [1, 6] ), then
\begin{equation} \begin{array}{lcr} \left\| { u(., t)} \right\| _{L^{3}(\Omega)}\leq C_{1}\left(1+\left\| { u_{0}} \right\| _{L^{3}(\Omega)}+\left\| { \phi_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}+\left\| { u_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}\right)\\ \quad + C_{2}\left\| { \phi } \right\| _{L^{2}(0, T;H^{1}(\Omega))}+ C_{3}\left\| { \phi } \right\| ^{2}_{L^{2}(0, T;H^{1}(\Omega))}+C_{4}\int_{0}^{t}\left\| { u(., s)} \right\| _{L^{3}(\Omega)}ds. \end{array} \end{equation} | (2.40) |
According to (2.28), we can deduce that
\begin{equation*} \begin{array}{lcr} \left\| { u(., t)} \right\| _{L^{3}(\Omega)}\leq C_{5}\Big(1+\left\| { u_{0}} \right\| _{L^{3}(\Omega)}+\left\| { \phi_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}+\left\| { u_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}\\ \quad + \Vert I_{i}\Vert_{L^2({\cal Q})}+ \Vert I\Vert_{L^2({\cal Q})}+\Vert \phi_{0}\Vert_\mathbb{H}\Big)+C_{6}\int_{0}^{t}\left\| { u(., s)} \right\| _{L^{3}(\Omega)}ds. \end{array} \end{equation*} |
Consequently (by using Gronwall lemma)
\begin{equation*} \begin{array}{lcr} \left\| { u(., t)} \right\| _{L^{3}(\Omega)}\leq C\Big(1+\left\| { u_{0}} \right\| _{L^{3}(\Omega)}+\left\| { \phi_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}+\left\| { u_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}\\ \quad + \Vert I_{i}\Vert_{L^2({\cal Q})}+ \Vert I\Vert_{L^2({\cal Q})}+\Vert \phi_{0}\Vert_\mathbb{H}\Big). \end{array} \end{equation*} |
Since
(u_{0}, \phi_{0}, I, I_{i}, u_{past}, \phi_{past})\in L^{3}(\Omega)\times L^{2}(\Omega)\times L^2({\cal Q})\times L^2({\cal Q})\times L^{2}(-\delta(0), 0;L^{3}(\Omega))\times L^{2}(-\delta(0), 0;L^{3}(\Omega)) , then, u(., t)\in L^{3}(\Omega) (for all t\in (0, T) ).
(ⅱ) From (2.33), we have (for all t )
\begin{equation} \begin{array}{lcr} u({\bf{x}}, t) \! = \! u_{0}-\int_{0}^{t}(\mathcal{I}_{2}({\bf{x}}, s;\phi_{1})-\mathcal{I}_{2}({\bf{x}}, s;\phi_{2}))ds-\int_{0}^{t}\hbar u({\bf{x}}, s)ds\\ \quad +\sum\limits_{k = 1}^{n_{1}}\! \int_{0}^{t}c_k({\bf{x}}, s)\phi({\bf{x}}, s-\xi_{k}(s))ds +\sum\limits_{l = 1}^{n_{2}}\! \int_{0}^{t}d_l({\bf{x}}, s) u({\bf{x}}, s-\eta_{l}(s))ds. \end{array} \end{equation} | (2.41) |
Consequently,
\begin{equation} \begin{array}{lcr} u({\bf{x}}, t) \! = \! u_{0}-\int_{0}^{t}\left(\mathcal{I}_{2}({\bf{x}}, s;\phi_{1})-\mathcal{I}_{2}({\bf{x}}, s;\phi_{2})\right)ds-\int_{0}^{t}\hbar u({\bf{x}}, s)ds\\ \quad +\sum\limits_{k = 1}^{n_{1}}\! \int_{-\xi_{k}(0)}^{t-\xi_{k}(t)}c_k({\bf{x}}, e_k(\theta))e'_k(\theta) \phi({\bf{x}}, \theta)d\theta +\sum\limits_{l = 1}^{n_{2}}\! \int_{-\eta_{l}(0)}^{t-\eta_{l}(t)}d_l({\bf{x}}, q_{l}(\theta))q'_{l}(\theta) u({\bf{x}}, \theta)d\theta. \end{array} \end{equation} | (2.42) |
Since -\delta(0) \leq -\xi_{k}(0) , -\delta(0) \leq -\eta_{k}(0) , t -\xi_{k}(t) \leq t and t -\eta_{k}(t) \leq t we can deduce that (according to the regularity of c_{k} , d_{l} and \hbar )
\begin{equation} \begin{array}{lcr} \mid u({\bf{x}}, t)\mid \leq \int_{0}^{t}\mid\mathcal{I}_{2}({\bf{x}}, s;\phi_{1})-\mathcal{I}_{2}({\bf{x}}, s;\phi_{2})\mid ds +\mid u_{0}\mid\\ \quad +C\Big(\int_{0}^{t}\mid u({\bf{x}}, s)\mid ds+ \int_{0}^{t}\mid \phi({\bf{x}}, \theta)\mid d\theta\\ \quad +\int_{-\delta(0)}^{0}\mid \phi({\bf{x}}, \theta)\mid d\theta + \int_{-\delta(0)}^{0}\mid u({\bf{x}}, \theta)\mid d\theta\Big) \end{array} \end{equation} | (2.43) |
and then (since from (H4) we have \mid \mathcal{I}_{2}(.; \phi_{1})-\mathcal{I}_{2}(.; \phi_{2}) \mid \leq C\mid \phi\mid (1+\mid \phi_{1} \mid +\mid \phi_{2} \mid) )
\begin{equation} \begin{array}{lcr} \mid u({\bf{x}}, t)\mid \leq C\Big(\int_{0}^{t}\mid \phi({\bf{x}}, s)\mid \mid \phi_{1}({\bf{x}}, s)\mid ds+\int_{0}^{t}\mid \phi({\bf{x}}, s)\mid \mid \phi_{2}({\bf{x}}, s)\mid ds\\ \quad +\int_{0}^{t}\mid u({\bf{x}}, s)\mid ds+ \int_{0}^{t}\mid \phi({\bf{x}}, s)\mid ds +\mid u_{0}\mid\\ \quad + \int_{-\delta(0)}^{0}\mid \phi_{past}({\bf{x}}, \theta)\mid d\theta+ \int_{-\delta(0)}^{0}\mid u_{past}({\bf{x}}, \theta)\mid d\theta\Big). \end{array} \end{equation} | (2.44) |
Consequently
\begin{equation} \begin{array}{lcr} \Big(\int_{\Omega}\mid u({\bf{x}}, t)\mid^{3} d{\bf{x}}\Big)^{1/3}\\ \quad \leq C\left(\left\| { u_{0}} \right\| _{L^{3}(\Omega)}+\Big[\int_{\Omega}\big(\int_{0}^{t}\mid u({\bf{x}}, s)\mid ds\big)^{3}d{\bf{x}}\Big]^{1/3}+ \Big[\int_{\Omega}\big(\int_{0}^{t}\mid \phi({\bf{x}}, s)\mid ds\big)^{3}d{\bf{x}}\Big]^{1/3} \right.\\ \quad \left.+\Big[\int_{\Omega}\big(\int_{0}^{t}\mid \phi({\bf{x}}, t)\mid\; \mid \phi_{2}({\bf{x}}, s)\mid ds\big)^{3}d{\bf{x}}\Big]^{1/3} +\Big[\int_{\Omega}\big(\int_{0}^{t}\mid \phi({\bf{x}}, s)\mid\; \mid \phi_{1}({\bf{x}}, s)\mid ds\big)^{3}d{\bf{x}}\Big]^{1/3}\right. \\ \quad \left.+ \Big[\int_{\Omega}\big(\int_{-\delta(0)}^{0}\mid \phi_{past}({\bf{x}}, \theta)\mid d\theta\big)^{3}d{\bf{x}}\Big]^{1/3}+ \Big[\int_{\Omega}\big(\int_{-\delta(0)}^{0}\mid u_{past}({\bf{x}}, \theta)\mid d\theta\big)^{3}d{\bf{x}}\Big]^{1/3}\right). \end{array} \end{equation} | (2.45) |
This implies (by using Hölder's and Minkowski inequalities)
\begin{equation} \begin{array}{lcr} \left\| { u(., t)} \right\| _{L^{3}(\Omega)}\leq C\left(\int_{0}^{t}\big(\int_{\Omega}\mid u({\bf{x}}, s)\mid^{3} d{\bf{x}}\big)^{1/3}ds +\int_{0}^{t}\big(\int_{\Omega}\mid \phi({\bf{x}}, \theta)\mid^{3} d{\bf{x}}\big)^{1/3}d\theta \right.\\ \quad \left.+\int_{0}^{t}\big(\int_{\Omega}\mid \phi({\bf{x}}, s)\mid^{6} d{\bf{x}}\big)^{1/6}\big(\int_{\Omega}\mid \phi_{1}({\bf{x}}, s)\mid^{6} d{\bf{x}}\big)^{1/6}ds\right.\\ \quad \left.+\int_{0}^{t}\big(\int_{\Omega}\mid \phi({\bf{x}}, s)\mid^{6} d{\bf{x}}\big)^{1/6}\big(\int_{\Omega}\mid \phi_{2}({\bf{x}}, s)\mid^{6} d{\bf{x}}\big)^{1/6}ds +\left\| { u_{0}} \right\| _{L^{3}(\Omega)}\right.\\ \quad \left.+ \int_{-\delta(0)}^{0}\big(\int_{\Omega}\mid \phi_{past}({\bf{x}}, \theta)\mid^{3} d{\bf{x}}\big)^{1/3}d\theta + \int_{-\delta(0)}^{0}\big(\int_{\Omega}\mid u_{past}({\bf{x}}, \theta)\mid^{3} d{\bf{x}}\big)^{1/3}d\theta\right). \end{array} \end{equation} | (2.46) |
Since \phi \in L^{2}(0, T, H^{1}(\Omega))\subset L^{2}(0, T, L^{r}(\Omega)) ( r\in [1, 6] ), then
\begin{equation} \begin{array}{lcr} \left\| { u(., t)} \right\| _{L^{3}(\Omega)}\leq C_{1}\left(\left\| { u_{0}} \right\| _{L^{3}(\Omega)}+\left\| { \phi_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}+\left\| { u_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}\right)\\ \quad + C_{2}\left\| { \phi } \right\| _{L^{2}(0, T;H^{1}(\Omega))}\left(1+\left\| { \phi_{1} } \right\| _{L^{2}(0, T;H^{1}(\Omega))}+\left\| { \phi_{2} } \right\| _{L^{2}(0, T;H^{1}(\Omega))}\right)\\ \quad +C_{3}\int_{0}^{t}\left\| { u(., s)} \right\| _{L^{3}(\Omega)}ds. \end{array} \end{equation} | (2.47) |
According to (2.29), we can deduce that
\begin{equation*} \begin{array}{lcr} \left\| { u(., t)} \right\| _{L^{3}(\Omega)}\leq C_{4}\left(\left\| { u_{0}} \right\| _{L^{3}(\Omega)}+\left\| { \phi_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}+\left\| { u_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}\right)\\ \quad + C_{5}\left(\Vert I_{i}\Vert_{L^2({\cal Q})}+ \Vert I\Vert_{L^2({\cal Q})}+\Vert \phi_{0}\Vert_\mathbb{H}\right)+C_{6}\int_{0}^{t}\left\| { u(., s)} \right\| _{L^{3}(\Omega)}ds. \end{array} \end{equation*} |
Consequently (by using Gronwall lemma)
\begin{equation} \begin{array}{lcr} \left\| { u(., t)} \right\| _{L^{3}(\Omega)}\leq C\Big(\left\| { u_{0}} \right\| _{L^{3}(\Omega)}+\left\| { \phi_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}+\left\| { u_{past}} \right\| _{L^{2}(-\delta(0), 0;L^{3}(\Omega))}\\ \quad + \Vert I_{i}\Vert_{L^2({\cal Q})}+ \Vert I\Vert_{L^2({\cal Q})}+\Vert \phi_{0}\Vert_\mathbb{H}\Big). \end{array} \end{equation} | (2.48) |
(ⅲ).a. Put S_{1} = \mathcal{H}(., \phi_\tau, u_\tau) , then from Lemma 2.6, we can deduce that
\begin{equation*} \begin{array}{lcr} \left\| { S_{1}} \right\| _{L^{2}(0, t;L^{2}(\Omega))} \leq C\Big(\Vert \phi \Vert_{L^{2}(0, t;L^{2}(\Omega))}+\Vert u\Vert_{L^{2}(0, t;L^{2}(\Omega))}\\ \quad +\Vert \phi_{past}\Vert_{L^{2}(-\delta(0), 0;L^{2}(\Omega))}+\Vert u_{past}\Vert_{L^{2}(-\delta(0), 0;L^{2}(\Omega))}\Big). \end{array} \end{equation*} |
Since (\phi, u)\in (L^{2}(0, T; L^{2}(\Omega)))^{2} and (\phi_{past}, u_{past})\in (L^{2}(-\delta(0), 0;L^{2}(\Omega)))^{2} , we can deduce that
S_{1}\in L^{2}({\cal Q}). |
From (2.15), we can deduce that (\phi, u) satisfies (for all v\in\mathbb{V}, \; v_e\in\mathbb{U} )
\begin{equation} \begin{array}{lcr} \big\langle \mathfrak{c}_{m}\frac{\partial \phi}{\partial t}, v\big\rangle_{\mathbb{V}', \mathbb{V}}+ \int_\Omega \mathcal{I}(.;\phi, u)v d{\bf{x}} + A_i(\phi+\varphi_e, v) = \int_\Omega (I_{i}+S_{1}) v d{\bf{x}}, \\ A_i(\phi + \varphi_e, v_e)+ A_e(\varphi_e, v_e) = \int_\Omega I v_{e} d{\bf{x}}. \end{array} \end{equation} | (2.49) |
In order to derive the result of (a), we will just sketch the proof based on suitable a priori estimates. From (2.49) with (v, v_{e}) = (\frac{\partial \phi}{\partial t}, \frac{\partial \varphi_{e}}{\partial t}) (since \mathfrak{c}_{m} = \mathfrak{b}_{m}^{2} )
\begin{equation} \begin{array}{lcr} \left\| { \mathfrak{b}_{m}\frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}+ \int_\Omega \mathcal{I}_{0}(\phi) \frac{\partial \phi}{\partial t}d{\bf{x}} + A_i(\phi+\varphi_e, \frac{\partial \phi}{\partial t}) +\int_{\Omega} \mathcal{I}_{1}(\phi) u\frac{\partial \phi}{\partial t}d{\bf{x}}\\ = \int_\Omega I_{i} \frac{\partial \phi}{\partial t} d{\bf{x}}+ \int_\Omega S_{1} \frac{\partial \phi}{\partial t}d{\bf{x}}, \\ A_i(\phi + \varphi_e, \frac{\partial \varphi_{e}}{\partial t})+ A_e(\varphi_e, \frac{\partial \varphi_{e}}{\partial t}) = \int_\Omega I \frac{\partial \varphi_{e}}{\partial t} d{\bf{x}}. \end{array} \end{equation} | (2.50) |
Then
\begin{equation} \begin{array}{lcr} \left\| { \mathfrak{b}_{m} \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}+ \frac{d}{dt}(\int_\Omega \tilde{\mathcal{I}}_{0}(\phi) d{\bf{x}}) +\int_{\Omega} \frac{\partial \tilde{\mathcal{I}}_{0}}{\partial t}(\phi)d{\bf{x}} + \int_{\Omega} \mathcal{I}_{1}(\phi) u\frac{\partial \phi}{\partial t}d{\bf{x}}\\ + \frac{d}{2dt}A_i(\phi+\varphi_e, \phi+\varphi_e) +\frac{d}{2dt}A_i(\varphi_e, \varphi_e)\\ = -\int_{\Omega} \varphi_{e}\frac{\partial I}{\partial t} d{\bf{x}}+\frac{d}{dt}(\int_{\Omega} I \varphi_{e} d{\bf{x}})+ \int_\Omega I_{i} \frac{\partial \phi}{\partial t} d{\bf{x}}+ \int_\Omega S_{1} \frac{\partial \phi}{\partial t}d{\bf{x}}. \end{array} \end{equation} | (2.51) |
Since I\in U_{c}\subset C^{0}([0, T]; L^{2}(\Omega)) , then I(0)\in L^{2}(\Omega) and we have \left\| { I(0)} \right\| ^{2}_{L^{2}(\Omega)}\leq C \left\| { I } \right\| ^{2}_{U_{c}} . Consequently, by integrating (2.51) by time we can deduce (from (2.4), (2.5), (H5), the boundedness of \mathfrak{b}_{m} and, the coercivity and continuity of A_{i} and A_{e} )
\begin{equation} \begin{array}{lcr} 2\underline{c}_{m}\int_{0}^{t}\left\| { \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+ \alpha_{i}\left\| { \phi(t)} \right\| ^{2}_{H^{1}(\Omega)}+(\alpha_{e}+\alpha_{i})\left\| { \varphi_e(t)} \right\| ^{2}_{H^{1}(\Omega)}\\ \quad +2\beta_{12} \int_{0}^{t}\left\| { \phi } \right\| ^{4}_{L^{4}(\Omega)}ds +2\beta_{10}\left\| { \phi(t) } \right\| ^{4}_{L^{4}(\Omega)}\leq \delta_{0}\int_{0}^{t}\left\| { \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+\delta_{1}\left\| { \varphi_{e}(t)} \right\| ^{2}_{H^{1}(\Omega)}\\ \quad +C_{0}\int_{0}^{t} \left\| { u } \right\| ^{2}_{L^{3}(\Omega)}\left\| { \phi } \right\| ^{2}_{H^{1}(\Omega)}ds +C_{1}\left(\left\| { I_{i} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { S_{1} } \right\| ^{2}_{L^{2}({\cal Q})} +\left\| { I } \right\| ^{2}_{U_{c}}\right)\\ \quad + C_{2}\left(1+\left\| { \phi_{0}} \right\| ^{2}_{H^{1}(\Omega)}+\left\| { \varphi_e^{(0)}} \right\| ^{2}_{H^{1}(\Omega)}+\left\| { \phi_{0} } \right\| ^{2}_{L^{2}(\Omega)}\right)+\int_\Omega \tilde{\mathcal{I}}_{0}(\phi_{0}) d{\bf{x}}\\ \quad +C_{3}\left(\left\| { \phi } \right\| ^{4}_{L^{4}({\cal Q})}+\left\| { \phi } \right\| ^{2}_{L^{\infty}(0, T;L^{2}(\Omega))}+\left\| { \varphi_{e} } \right\| ^{2}_{L^{2}({\cal Q})}\right). \end{array} \end{equation} | (2.52) |
From (H5) we can deduce that \int_\Omega \tilde{\mathcal{I}}_{0}(\phi_{0}) d{\bf{x}}\leq C(1+\left\| { \phi_{0} } \right\| ^{4}_{L^{4}(\Omega)})\leq C(1+\left\| { \phi_{0} } \right\| ^{4}_{H^{1}(\Omega)}) and then (by choosing \delta_{0} = \underline{c}_{m} and \delta_{1} = (\alpha_{e}+\alpha_{i})/2 )
\begin{equation} \begin{array}{lcr} \underline{c}_{m}\int_{0}^{t}\left\| { \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+ \alpha_{i}\left\| { \phi(t)} \right\| ^{2}_{H^{1}(\Omega)}+(\alpha_{e}+\alpha_{i})/2\left\| { \varphi_e(t)} \right\| ^{2}_{H^{1}(\Omega)}\\ \quad +2\beta_{12} \int_{0}^{t}\left\| { \phi } \right\| ^{4}_{L^{4}(\Omega)}ds+2\beta_{10}\left\| { \phi(t) } \right\| ^{4}_{L^{4}(\Omega)}\leq C_{4}\int_{0}^{t} \left\| { u } \right\| ^{2}_{L^{3}(\Omega)}\left\| { \phi } \right\| ^{2}_{H^{1}(\Omega)}ds \\ \quad +C_{5}\left(\left\| { I_{i} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { S_{1} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { I } \right\| ^{2}_{U_{c}}\right)\\ \quad + C_{6}\left(1+\left\| { \phi_{0}} \right\| ^{2}_{H^{1}(\Omega)}+\left\| { \varphi_e^{(0)}} \right\| ^{2}_{H^{1}(\Omega)}+\left\| { \phi_{0} } \right\| ^{4}_{H^{1}(\Omega)}\right)\\ \quad +C_{7}\left(\left\| { \phi } \right\| ^{4}_{L^{4}({\cal Q})}+\left\| { \phi } \right\| ^{2}_{L^{\infty}(0, T;L^{2}(\Omega))}+\left\| { \varphi_{e} } \right\| ^{2}_{L^{2}({\cal Q})}\right). \end{array} \end{equation} | (2.53) |
Consequently (since I\in U_{c} , \phi\in L^{4}({\cal Q}) \cap L^{\infty}(0, T; L^{2}(\Omega))\cap L^{2}(0, T; H^{1}(\Omega)) , \varphi_{e}\in L^{2}({\cal Q}) , u\in L^{\infty}(0, T; L^{3}(\Omega)) , (\phi_{past}, u_{past})\in (L^{2}({\cal Q}_{0}))^{2} , (I_{i}, S_{1})\in (L^{2}({\cal Q}))^{2} and (\phi_{0}, \varphi_{e}^{(0)})\in (H^{1}(\Omega))^{2} ),
\begin{equation} \begin{array}{lcr} \int_{0}^{t}\left\| { \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+ \left\| { \phi(t)} \right\| ^{2}_{H^{1}(\Omega)}+\left\| { \varphi_e(t)} \right\| ^{2}_{H^{1}(\Omega)} \leq C \end{array} \end{equation} | (2.54) |
and then \frac{\partial \phi}{\partial t}\in L^{2}({\cal Q}) and (\phi, \varphi_{e})\in L^{\infty}(0, T; H^{1}(\Omega)) .
The proof of (a) can be completed by implementing the classical Faedo-Galerkin method and by taking advantage of the above estimates. So we omit the details.
Prove now that u\in C^{0}([0, T]; L^{3}(\Omega)) . Let S = -\mathcal{I}_{2}(.; \phi)-\hbar u+\mathcal{E}(\phi_{\tau}, u_{\tau}) be the right hand side of (2.33). According to the expression of \mathcal{I}_{2} , we can deduce that (since \phi \in L^{\infty}(0, T; H^{1}(\Omega)) and u\in L^{\infty}(0, T; L^{3}(\Omega)) )
\left\| { S} \right\| _{L^{3}(\Omega)}\leq C\left(1+\left\| { \phi } \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { u } \right\| _{L^{\infty}(0, T;L^{3}(\Omega))}+\left\| { \mathcal{E}(\phi_{\tau}, u_{\tau})} \right\| _{L^{3}(\Omega)}\right) |
and then (from Lemma 2.6)
\begin{equation} \begin{array}{lcr} \left\| { S} \right\| _{L^{2}(0, T;L^{3}(\Omega))}\leq C\Big(1+\left\| { \phi } \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { u} \right\| _{L^{\infty}(0, T;L^{3}(\Omega))}+\Vert \phi \Vert_{L^{2}(0, T;L^{3}(\Omega))} \\ \quad +\Vert u \Vert_{L^{2}(0, T;L^{3}(\Omega))}+\Vert \phi_{past}\Vert_{L^{2}(-\delta(0), 0;L^{3}(\Omega))}+\Vert u_{past}\Vert_{L^{2}(-\delta(0), 0;L^{3}(\Omega))}\Big). \end{array} \end{equation} | (2.55) |
Consequently, since (\phi_{past}, u_{past})\in (L^{2}(-\delta(0), 0;L^{3}(\Omega)))^{2} , we can deduce that S\in L^{2}(0, T; L^{3}(\Omega)) and then \frac{\partial u}{\partial t}\in L^{2}(0, T; L^{3}(\Omega)) . Since u\in L^{2}(0, T; L^{3}(\Omega)) and \frac{\partial u}{\partial t}\in L^{2}(0, T; L^{3}(\Omega)) then, from Remark 2.3, u\in C^{0}([0, T]; L^{3}(\Omega)) .
(ⅲ).b. Prove now the estimate relation. Let (\phi_{j}, \varphi_{e, j}, u_{j}) (for j = 1, 2 ) be two solutions corresponding to (I_{i}^{(j)}, I^{(j)})\in L^{2}({\cal Q})\times U_{c} with (\phi_{1}-\phi_{2}, \varphi_{e, 1}-\varphi_{e, 2}, u_{1}-u_{2})(t = 0) = (0, 0, 0) and (\phi_{past}, u_{past}) = (\phi_{past, 1}-\phi_{past, 2}, u_{past, 1}-u_{past, 2}) = (0, 0) . Then (\phi, \varphi_{e}, u) = (\phi_{1}-\phi_{2}, \varphi_{e, 1}-\varphi_{e, 2}, u_{1}-u_{2}) satisfies, from (2.15) with (v, v_{e}) = (\frac{\partial \phi}{\partial t}, \frac{\partial \varphi_{e}}{\partial t})
\begin{equation} \begin{array}{lcr} \left\| { \mathfrak{b}_{m}\frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}+ \int_\Omega \left(\mathcal{I}_{0}(.;\phi_{1}) -\mathcal{I}_{0}(.;\phi_{2})\right)\frac{\partial \phi}{\partial t}d{\bf{x}} + A_i(\phi+\varphi_e, \frac{\partial \phi}{\partial t}) \\ \quad + \int_\Omega \left(u_{1}\mathcal{I}_{1}(.;\phi_{1}) -u_{2}\mathcal{I}_{2}(.;\phi_{2})\right)\frac{\partial \phi}{\partial t}d{\bf{x}} = \int_\Omega I_{i} \frac{\partial \phi}{\partial t} d{\bf{x}}+ \int_\Omega \mathcal{H}(., \phi_\tau, u_\tau) \frac{\partial \phi}{\partial t}d{\bf{x}}, \\ A_i(\phi + \varphi_e, \frac{\partial \varphi_{e}}{\partial t})+ A_e(\varphi_e, \frac{\partial \varphi_{e}}{\partial t}) = \int_\Omega I \frac{\partial \varphi_{e}}{\partial t} d{\bf{x}}, \\ \frac{\partial u}{\partial t} = -\left(\mathcal{I}_{2}(.;\phi_{1})-\mathcal{I}_{2}(.;\phi_{2})\right)-\hbar u+\mathcal{E}(\phi_{\tau}, u_{\tau}). \end{array} \end{equation} | (2.56) |
Then
\begin{equation} \begin{array}{lcr} \left\| { \mathfrak{b}_{m}\frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}+ \frac{d}{2dt}A_i(\phi+\varphi_e, \phi+\varphi_e) +\frac{d}{2dt}A_i(\varphi_e, \varphi_e)\\ \quad = -\int_\Omega \left(\mathcal{I}_{0}(.;\phi_{1}) -\mathcal{I}_{0}(.;\phi_{2})\right)\frac{\partial \phi}{\partial t}d{\bf{x}} - \int_\Omega u\mathcal{I}_{1}(.;\phi_{1}) \frac{\partial \phi}{\partial t}d{\bf{x}} \\ \quad - \int_\Omega u_{2}\left(\mathcal{I}_{1}(.;\phi_{1}) -\mathcal{I}_{1}(.;\phi_{2})\right) \frac{\partial \phi}{\partial t}d{\bf{x}} -\int_{\Omega} \varphi_{e}\frac{\partial I}{\partial t} d{\bf{x}}\\ \quad +\frac{d}{dt}\left(\int_{\Omega} I \varphi_{e} d{\bf{x}}\right)+ \int_\Omega I_{i} \frac{\partial \phi}{\partial t} d{\bf{x}}+ \int_\Omega \mathcal{H}(., \phi_\tau, u_\tau) \frac{\partial \phi}{\partial t}d{\bf{x}}, \\ \frac{\partial u}{\partial t} = -\left(\mathcal{I}_{2}(.;\phi_{1})-\mathcal{I}_{2}(.;\phi_{2})\right)-\hbar u+\mathcal{E}(\phi_{\tau}, u_{\tau}). \end{array} \end{equation} | (2.57) |
According to assumptions (H2), (H4) and (H6), we can deduce that (from the boundedness of \mathfrak{b}_{m} )
\begin{equation} \begin{array}{lcr} \underline{c}_{m}\left\| { \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}+ \frac{d}{2dt}A_i(\phi+\varphi_e, \phi+\varphi_e) +\frac{d}{2dt}A_i(\varphi_e, \varphi_e)\\ \quad \leq C_{1}\left(\int_\Omega \mid \phi\mid (\mid \phi_{1}\mid^{2}+\mid \phi_{2}\mid^{2} +1)\mid \frac{\partial \phi}{\partial t}\mid d{\bf{x}} +\int_\Omega \mid u\mid (1+\mid \phi_{1}\mid) \mid \frac{\partial \phi}{\partial t}\mid d{\bf{x}}\right) \\ \quad +C_{2}\int_\Omega \mid \phi \mid\mid u_{2}\mid \mid \frac{\partial \phi}{\partial t}\mid d{\bf{x}} +\int_{\Omega} \mid \varphi_{e}\mid \mid \frac{\partial I}{\partial t}\mid d{\bf{x}}\\ \quad +\frac{d}{dt}\left(\int_{\Omega} I \varphi_{e} d{\bf{x}}\right)+\int_\Omega \mid I_{i} \mid \mid \frac{\partial \phi}{\partial t}\mid d{\bf{x}}+ \int_\Omega \mid \mathcal{H}(., \phi_\tau, u_\tau) \mid \mid \frac{\partial \phi}{\partial t}\mid d{\bf{x}}, \\ \left\| { \frac{\partial u}{\partial t}} \right\| _{L^{3}(\Omega)}\leq C_{3}\left(\left\| { \phi } \right\| _{L^{6}(\Omega)}\left\| { \phi_{1} } \right\| _{L^{6}(\Omega)}+ \left\| { \phi } \right\| _{L^{6}(\Omega)}\left\| { \phi_{2} } \right\| _{L^{6}(\Omega)}+\left\| { \phi } \right\| _{L^{3}(\Omega)}+\left\| { u} \right\| _{L^{3}(\Omega)}\right)\\ \quad +C_{4}\left\| { \mathcal{E}(\phi_{\tau}, u_{\tau})} \right\| _{L^{3}(\Omega)}. \end{array} \end{equation} | (2.58) |
Integrating by time we obtain (according to the coercivity and continuity of A_{i} and A_{e} , and to the estimate of \mathcal{H}(.; \phi_\tau, u_\tau) and \mathcal{E}(.; \phi_\tau, u_\tau) given by Lemma 2.6)
\begin{equation*} \begin{array}{lcr} 2\underline{c}_{m}\int_{0}^{t}\left\| { \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+ \alpha_{i}\left\| { \phi(t)} \right\| ^{2}_{H^{1}(\Omega)}+(\alpha_{e}+\alpha_{i})\left\| { \varphi_e(t)} \right\| ^{2}_{H^{1}(\Omega)}\\ \quad \leq C_{5}\left\| { \phi } \right\| ^{2}_{L^{2}(0, T;H^{1}(\Omega)}\left(\left\| { \phi_{1} } \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { \phi_{2} } \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { u_{2} } \right\| ^{2}_{L^{\infty}(0, T;L^{3}(\Omega))}\right)\\ \quad +C_{6}\left\| { u} \right\| ^{2}_{L^{\infty}(0, T;L^{3}(\Omega))}\left\| { \phi_{1} } \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))} +\delta_{0}\int_{0}^{t}\left\| { \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+\delta_{1}\left\| { \varphi_{e}(t)} \right\| ^{2}_{H^{1}(\Omega)}\\ \quad +C_{7}\left(\left\| { \phi } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { u } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { I_{i} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { I } \right\| ^{2}_{U_{c}}+\left\| { \varphi_{e} } \right\| ^{2}_{L^{2}({\cal Q})}\right), \\ \left\| { \frac{\partial u}{\partial t}} \right\| _{L^{2}(0, T;L^{3}(\Omega))}\leq C_{8}\left\| { \phi } \right\| _{L^{\infty}(0, T;H^{1}(\Omega))}\left(\left\| { \phi_{1} } \right\| _{L^{\infty}(0, T;H^{1}(\Omega))}+ \left\| { \phi_{2} } \right\| _{L^{\infty}(0, T;H^{1}(\Omega))}\right)\\ \quad +C_{9}\left(\left\| { \phi } \right\| _{L^{2}(0, T;L^{3}(\Omega))}+\left\| { u} \right\| _{L^{2}(0, T;L^{3}(\Omega))}\right). \end{array} \end{equation*} |
Since (\phi_{j}, u_{j})\in L^{\infty}(0, T; H^{1}(\Omega))\times L^{\infty}(0, T; L^{3}(\Omega)) (for j = 1, 2 ) we can deduce that (by choosing \delta_{0} = \underline{c}_{m} and \delta_{1} = (\alpha_{e}+\alpha_{i})/2 )
\begin{equation} \begin{array}{lcr} \underline{c}_{m}\int_{0}^{t}\left\| { \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+ \alpha_{i}\left\| { \phi(t)} \right\| ^{2}_{H^{1}(\Omega)}+\frac{\alpha_{e}+\alpha_{i}}{2}\left\| { \varphi_e(t)} \right\| ^{2}_{H^{1}(\Omega)} \\ \quad \leq C_{10}\left(\left\| { u} \right\| ^{2}_{L^{\infty}(0, T;L^{3}(\Omega))}+\left\| { \phi } \right\| ^{2}_{L^{2}(0, T;H^{1}(\Omega))}+\left\| { I_{i} } \right\| ^{2}_{L^{2}({\cal Q})} +\left\| { \varphi_{e} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { I } \right\| ^{2}_{U_{c}}\right)\\ \left\| { \frac{\partial u}{\partial t}} \right\| ^{2}_{L^{2}(0, T;L^{3}(\Omega))}\leq C_{11}\left(\left\| { \phi } \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { \phi } \right\| ^{2}_{L^{2}(0, T;L^{3}(\Omega))}+\left\| { u} \right\| ^{2}_{L^{2}(0, T;L^{3}(\Omega))}\right). \end{array} \end{equation} | (2.59) |
Consequently (from the estimates (2.29)–(2.48))
\begin{equation} \begin{array}{lcr} \left\| { \frac{\partial \phi}{\partial t}} \right\| ^{2}_{L^{2}({\cal Q})}+ \left\| { \frac{\partial u}{\partial t}} \right\| ^{2}_{L^{2}(0, T;L^{3}(\Omega))}+ \left\| { \phi} \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { \varphi_e} \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}\\ \quad \leq C\left(\left\| { I_{i} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { I } \right\| ^{2}_{U_{c}}\right). \end{array} \end{equation} | (2.60) |
This completes the proof.
According to previous Theorems, we can derive the following results.
Theorem 2.4. Consider the case of p = 4 . Let assumptions (H1)-(H6) and (RC) be fulfilled. For (u_{0}, u_{past}, \phi_{past}, \phi_{0}) and (I, I_{i}) given such that (\phi_{past}, u_{past})\in (L^{2}(-\delta(0), 0;L^{3}(\Omega)))^{2} , u_{0}\in L^{3}(\Omega) , \varphi_{i}^{(0)}-\varphi_{e}^{(0)}\! = \! \phi_{0} with (\phi_{0}, \varphi_{e}^{(0)})\in (H^{1}(\Omega))^{2} , and (I_{i}, I)\in L^2({\cal Q})\times U_{c} , there exists a unique solution (\phi, \varphi_{e}, u) of problem (2.15) verifying (\phi, \varphi_{e}, u) \in \mathbb{D} , with
\begin{equation} \begin{array}{lcr} \mathbb{D} = \mathbb{A}\times L^{\infty}(0, T;\mathbb{U})\times\mathbb{B} \mathit{\; \text{such that}}\; \; \mathbb{A} = L^{\infty}(0, T;\mathbb{V})\cap H^{1}(0, T;\mathbb{H})\cap C^{0}([0, T]; \mathbb{H})\\ \mathit{\; \text{and}}\; \mathbb{B} = C^{0}([0, T]; L^{3}(\Omega))\cap H^{1}(0, T;L^{3}(\Omega)), \; \; \mathit{\text{equipped with }}\; \; (\forall (v, v_{e}, \rho)\in \mathbb{D})\\ \quad \left\| { (v, v_{e}, \rho)} \right\| ^{2}_{\mathbb{D}} = \left\| { v} \right\| ^{2}_{\mathbb{A}}+\left\| { v_{e}} \right\| ^{2}_{L^{\infty}(0, T;\mathbb{U})}+\left\| { \rho} \right\| ^{2}_{\mathbb{B}}, \end{array} \end{equation} | (2.61) |
and it holds that
\begin{equation} \begin{array}{lcr} \left\| { (\phi, \varphi_{e}, u)} \right\| ^{2}_{\mathbb{D}}\leq C\left(1+\Vert I_{i}\Vert_{L^2({\cal Q})}^2 + \Vert I\Vert_{U_{c}}^2 \right). \end{array} \end{equation} | (2.62) |
Moreover the Lipschitz continuity relation is satisfied, i.e., for any element (I^{(j)}_{i}, I^{(j)})\in L^2({\cal Q})\times U_{c} for j = 1, 2 , we have
\begin{equation} \begin{array}{lcr} \left\| { (\phi_{1}-\phi_{2}, \varphi_{e, 1}-\varphi_{e, 2}, u_{1}-u_{2})} \right\| ^{2}_{\mathbb{D}} \leq C\left(\Vert I^{(1)}_{i}-I^{(2)}_{i}\Vert_{L^2({\cal Q})}^2 + \Vert I^{(1)}-I^{(2)}\Vert_{U_{c}}^2 \right), \end{array} \end{equation} | (2.63) |
where (\phi_{j}, u_{j}, \varphi_{e, j}) is the solution of (2.15), which corresponds to data (\phi_{0}, u_{0}) , (\phi_{past}, u_{past}) and (I^{(j)}_{i}, I^{(j)}) (for j = 1, 2 ).
In this section, we formulate the minimax control problem and discuss the existence and necessary optimality conditions for an optimal solution.
Our problem in this section is to find the best admissible source function \xi in presence of the admissible disturbance in the function \pi . In order to illustrate our minimax control problem, we assume here that the control \xi is in I and the disturbance \pi is in I_{i} (which act on control domain \Omega_{c} and disturbance domain \Omega_{d} , respectively), i.e., I = {\cal B}_{1}\xi and I_{i} = {\cal B}_{2}\pi+f , where f\in L^{2}({\cal Q}) , support(\xi) \subset \Omega_{c}\times (0, T) and support(\pi) \subset \Omega_{d}\times (0, T) , and the operators {\cal B}_{i}\in {\mathfrak L}(L^{2}({\Omega}); L^{2}({\Omega})) , with \left\| { {\cal B}_{i}\theta} \right\| _{L^{2}(\Omega)} \leq C\left\| { \theta } \right\| _{L^{2}(\Omega)} , for all \theta (for i = 1, 2 ). Therefore, the function (\phi, \varphi_{e}, u) is assumed to be related to the disturbance \pi and control \xi through the problem (under conditions (1.3) and (1.2) for \varphi_{e} and {\cal B}_{1}\xi , respectively)
\begin{equation} \begin{array}{lcr} \mathfrak{c}_{m}\frac{\partial \phi}{\partial t}+{\cal I}(\phi, u)-div(\mathcal{K}_{i}\nabla (\phi+\varphi_{e})) = \mathcal{H}(\phi_{\tau}, u_{\tau})+{\cal B}_{2}\pi+f, \; \; \text{in} \; \; {\cal Q}\\ -div(\mathcal{K}_{i}\nabla \phi)-div((\mathcal{K}_{i}+\mathcal{K}_{e})\varphi_{e}) = {\cal B}_{1}\xi, \; \; \text{in}\; \; {\cal Q}\\ \frac{\partial u}{\partial t}+{\cal G}(\phi, u) = \mathcal{E}(\phi_{\tau}, u_{\tau}), \; \; \text{in} \; \; {\cal Q}\\ \big(\mathcal{K}_{i}\nabla \phi\big).{\bf{n}} + \big(K_{i}\nabla \varphi_{e}\big).{\bf{n}} = 0, \; \; \text{on}\; \; \Sigma\\ \big(\mathcal{K}_{e}\nabla \varphi_{e}\big).{\bf{n}} = 0, \; \; \text{on}\; \; \Sigma\\ \phi (., t = 0) = \phi_{0}, \; \; \; u(., t = 0) = u_{0}, \; \; \text{in}\; \; \Omega\\ \phi = \phi_{past}, \; \; \; u = u_{past}, \; \; \text{in}\; \; {\cal Q}_{0} \end{array} \end{equation} | (3.1) |
under pointwise constraints
\begin{equation} \begin{array}{lcr} \mid \xi\mid \leq \tau_{1}\; \; a.e. \; \text{in}\; {\cal Q}, \\ \mid \pi\mid \leq \tau_{2}\; \; a.e. \; \text{in}\; {\cal Q}. \end{array} \end{equation} | (3.2) |
Let {\cal U}_{ad} and {\cal V}_{ad} be convex, closed, non-empty and bounded subsets of U_{c} and L^{2}({\cal Q}) , respectively, and describing constraints (3.2) and compatibility condition (1.2) such that
\begin{equation*} \begin{array}{lcr} {\cal U}_{ad}\! = \!\left\{\xi\in U_{c}:\; support(\xi) \subset \Omega_{c}\times (0, T), \; \int_{\Omega}\!\!\!{\cal B}_{1}\xi d{\bf{x}} = 0, \; \mid \xi\mid \leq \tau_{1} \; a.e. \; in\; {\cal Q}\right\}, \\ {\cal V}_{ad}\! = \!\Big\{\pi\in L^{2}({\cal Q}):\; support(\pi) \subset \Omega_{d}\times (0, T), \; \mid \pi\mid \leq \tau_{2} \; \; a.e. \; in\; {\cal Q} \Big\}. \end{array} \end{equation*} |
Although {\cal U}_{ad} and {\cal V}_{ad} are subsets of L^{\infty}({\cal Q}) , we prefer to use standard norms of space L^{2}({\cal Q}) . the reason is that we would like to take advantages of differentiability of the latter norms away from the origin to perform our variational analysis. Moreover the spaces {\cal U}_{ad} and {\cal V}_{ad} form closed, convex, weak-star sequentially compact subsets of spaces L^{\infty}(0, T; L^{2}(\Omega)) (for weak-star sequential compactness see e.g. [59] and for similar result see e.g. [45]). For operator {\cal B}_{1} , we can consider, for example, the operator
\begin{equation} \theta\in L^{2}(\Omega) \longrightarrow {\cal B}_{1}\theta = \theta-\left(\frac{1}{mes(\Omega_{c})}\int_{\Omega}\theta d{\bf{x}}\right)\chi_{\Omega_{c}}. \end{equation} | (3.3) |
Then \int_{\Omega}\!\!{\cal B}_{1}\theta d{\bf{x}} = 0 and if support(\theta) \subset \Omega_{c} we have \int_{\Omega_{c}}\!\!{\cal B}_{1}\theta d{\bf{x}} = 0 . Moreover the operator {\cal B}_{1} is autoadjoint on the domain \big\{\theta\in L^{2}(\Omega):\; \; support(\theta)\subset \Omega_{c}\big\} . The studied control problem is to find a saddle point of cost function {\cal J} which measures the distance between known observations (a nominal desired states) and the prognostic variables \phi . We assume that we have observations on some domain \Omega_{obs} at certain times t_k , k = 1;...; N_{obs} or at final time T . Precisely we will study the following control problem ({\bf SP}) .
Find an admissible control-disturbance (\xi^{*}, \pi^{*})\in {\cal U}_{ad}\times {\cal V}_{ad} such that cost functional (in the reduced form)
\begin{equation} \begin{array}{lcr} {\cal J}(\xi , \pi) = \frac{m_{1}}{2}\int_{0}^{T}\Upsilon(t)\left\| { \phi-\phi_{obs}} \right\| ^{2}_{L^{2}({\Omega_{obs}})}dt+\frac{m_{2}}{2}\left\| { \phi(T)-\psi_{obs}} \right\| ^{2}_{L^{2}({\Omega_{obs}})}\\ \quad +\frac{\alpha}{2}\left\| { \xi} \right\| ^{2}_{U_{c}}-\frac{\beta}{2}\int_{0}^{T}\left\| { \pi } \right\| ^{2}_{L^{2}({\Omega_{d}})}dt, \end{array} \end{equation} | (3.4) |
is minimized with respect to \xi and maximized with respect to \pi subject to problem (3.1), where the weight function \Upsilon is given by (with \varpi > 0 large enough)
\begin{equation} \Upsilon(t) = \sum\limits_{k = 1}^{N_{obs}} exp(-\varpi(t-t_{k})^{2}), \end{equation} | (3.5) |
\alpha, \beta > 0 and m_{i}\geq 0 ( i = 1, 2 ) are fixed such that m_{1}+m_{2} > 0 , the functions \phi_{obs}\in L^{2}(0, T; \Omega_{obs}) and \psi_{obs}\in L^{2}(\Omega_{obs}) are the observations (given), \Omega_{obs}\subset \Omega is the observation domain, \Omega_{c}\subset \Omega is the control domain and \Omega_{d}\subset \Omega is the disturbance domain. Clearly, find (\xi^{*}, \pi^{*})\in {\cal U}_{ad}\times {\cal V}_{ad} (a saddle point for the functional \mathcal{J} ) such that ( \forall (\xi, \pi)\in {\cal U}_{ad}\times {\cal V}_{ad} )
\begin{equation} {\cal J}(\xi^{*}, \pi)\leq {\cal J}(\xi^{*}, \pi^{*})\leq {\cal J}(\xi, \pi^{*}). \end{equation} | (3.6) |
Remark 3.1. (i) The coefficient \alpha > 0 can be interpreted as the measure of price of control (that the engineer can afford) and the coefficient \beta > 0 can be interpreted as the measure of price of disturbance (that the environment can afford).
(ii) Operators {\cal B}_{i} , i = 1, 2 , also include the quantification of source profiles, inside the considered area, which results of change in disturbance and control variables.
In the sequel of this paper, we restrict our analysis to a generalized form of the three models mentioned at the beginning of article, namely : Rogers-McCulloch model, Fitzhugh-Nagumo model and Aliev-Panfilov model. More precisely:
(HMC) p = 4 and the operators \mathcal{I} and \mathcal{G} are supposed to be of the form given in (2.21).
Remark 3.2. According to expression (2.21) of operators \mathcal{I} and \mathcal{G} , we verify easily that hypotheses (H5)-(H6) are satisfied.
In this section, we study the Fréchet differentiability of the nonlinear operator solution and derive some necessary estimates. For a given f\in L^{2}({\cal Q}) , initial condition (\phi_{0}, \varphi_{e}^{(0)}, u_{0}) in (H^{1}(\Omega))^{2}\times L^{3}(\Omega) and past condition (\phi_{past}, u_{past})\in (L^{2}(-\delta_{0}, 0; L^{3}(\Omega)))^{2} , let us introduce the following mapping {\cal F}: {\cal U}_{ad}\times {\cal V}_{ad} \to \ \mathbb{D} , which maps the source term (\xi, \pi) \in {\cal U}_{ad}\times {\cal V}_{ad} of (3.1) into the corresponding solution X = (\phi, \varphi_{e}, u) in \mathbb{D} , where \mathbb{D} is defined by (2.61).
Before proceeding with investigation of Fréchet differentiability of operator {\cal F} , we study the following linear parabolic problem, for (h_{1}, h_{2})\in U_{c}\times L^{2}(\Omega_{d}) ,
\begin{equation} \begin{array}{lcr} \mathfrak{c}_{m}\frac{\partial \Psi}{\partial t}+\frac{\partial {\cal I}}{\partial \phi}(\phi, u).\Psi+\frac{\partial {\cal I}}{\partial u}(\phi, u).w-div(\mathcal{K}_{i}\nabla (\Psi+\psi_{e})) = \mathcal{H}(\Psi_{\tau}, w_{\tau})+{\cal B}_{2}h_{2}, \; \; \text{in} \; \; {\cal Q}\\ -div(\mathcal{K}_{i}\nabla \Psi)-div((\mathcal{K}_{i}+\mathcal{K}_{e})\psi_{e}) = {\cal B}_{1}h_{1}, \; \; \text{in} \; \; {\cal Q}\\ \frac{\partial w}{\partial t}+\frac{\partial {\cal G}}{\partial \phi}(\phi, u).\Psi+\frac{\partial {\cal G}}{\partial u}(\phi, u).w = \mathcal{E}(\Psi_{\tau}, w_{\tau}), \; \; \text{in} \; \; {\cal Q}\\ \big(\mathcal{K}_{i}\nabla \Psi\big).{\bf{n}} + \big(\mathcal{K}_{i}\nabla \psi_{e}\big).{\bf{n}} = 0, \; \; \text{on}\; \Sigma\\ \big(\mathcal{K}_{e}\nabla \psi_{e}\big).{\bf{n}} = 0, \; \text{on}\; \Sigma\\ \Psi(., t = 0) = 0, \; \; \; w(., t = 0) = 0, \; \; \text{in}\; \; \Omega\\ \Psi = 0, \; \; \; w = 0, \; \; \text{in}\; \; {\cal Q}_{0} \end{array} \end{equation} | (3.7) |
and under conditions (1.3) and (1.2) for \psi_{e} and {\cal B}_{1}h_{1} , respectively.
The weak formulation of Problem (3.7) can be written as follows ( \forall (v, v_{e}, \rho)\in \mathbb{V}\times\mathbb{U}\times \mathbb{H} and a.e. in (0, T) )
\begin{equation} \begin{array}{lcr} \big\langle \mathfrak{c}_{m}\frac{\partial \Psi}{\partial t}, v \big\rangle_{\mathbb{V}', \mathbb{V}} +\int_{\Omega}\mathcal{K}_{i}\nabla \Psi\nabla v d{\bf{x}} +\int_{\Omega}\mathcal{K}_{i}\nabla \psi_{e}\nabla v d{\bf{x}} \quad +\int_{\Omega}\frac{\partial {\cal I}}{\partial \phi}\Psi v d{\bf{x}}\\ \quad +\int_{\Omega}\frac{\partial {\cal I}}{\partial u}w vd{\bf{x}} = \int_{\Omega}\mathcal{H}(\Psi_{\tau}, w_{\tau}) v d{\bf{x}}+\int_{\Omega}{\cal B}_{2}h_{2} vd{\bf{x}}, \\ \int_{\Omega}(\mathcal{K}_{i}+\mathcal{K}_{e})\nabla\psi_{e}\nabla v_{e} d{\bf{x}} +\int_{\Omega}\mathcal{K}_{i}\nabla \Psi\nabla v_{e} d{\bf{x}} = \int_{\Omega}{\cal B}_{1}h_{1}v_{e} d{\bf{x}}, \\ \big(\frac{\partial w}{\partial t}, \rho \big)_{\mathbb{H}}+\int_{\Omega}\frac{\partial {\cal G}}{\partial \phi}.\Psi \rho d{\bf{x}} +\int_{\Omega}\hbar w \rho d{\bf{x}} = \int_{\Omega}\mathcal{E}(\Psi_{\tau}, w_{\tau})\rho d{\bf{x}}. \end{array} \end{equation} | (3.8) |
We are now going to prove the existence and uniqueness result of problem (3.8). To this end, we begin by proving the following necessary Lemma, which correspond to the existence, uniqueness and some regularity properties for the following nondelayed problem on {\cal Q}_{S} = \Omega\times(S_{0}, S_{f}) (with 0\leq S_{0} < S_{f}\leq T )
\begin{equation} \begin{array}{lcr} \mathfrak{c}_{m}\frac{\partial \Pi }{\partial t}+\frac{\partial {\cal I}}{\partial \phi}(\phi, u).\Pi+\frac{\partial {\cal I}}{\partial u}(\phi, u).V-div(\mathcal{K}_{i}\nabla (\Pi+\pi_{e})) = g_{1}, \\ -div\left((\mathcal{K}_{e}+\mathcal{K}_{i})\nabla \pi_{e}\right)-div(\mathcal{K}_{i}\nabla \Pi) = {\cal B}_{1}h_{1}\in U_{c}, \\ \frac{\partial V}{\partial t}+\frac{\partial {\cal G}}{\partial \phi}(\phi, u).\Pi+\frac{\partial {\cal G}}{\partial u}(\phi, u).V = g_{2}, \\ \text{with initial and boundary conditions}\\ \big(\mathcal{K}_{i}\nabla (\Pi + \pi_{e})\big) \cdot{\bf{n}} = 0, \; \text{on}\; \Sigma_{S} = \partial\Omega\times(S_{0}, S_{f})\\ \big(\mathcal{K}_{e}\nabla \pi_{e}\big)\cdot{\bf{n}} = 0, \; \text{on}\; \Sigma_{S} \\ \Pi(t = S_{0}) = \Pi_{0}, V(t = S_{0}) = V_{0}, \; \text{on}\; \Omega. \end{array} \end{equation} | (3.9) |
Lemma 3.1. Let assumptions (H1)-(H4) and (HMC) be fulfilled. Suppose that (\phi, \varphi_{e}, u) \in \mathbb{D} and h_{1}\in U_{c} , then the following results hold.
1. For (V_{0}, \Pi_{0})\in (L^{2}(\Omega))^{2} and (g_{1}, g_{2})\in (L^2({\cal Q}_{S}))^{2} given, there exists a unique solution (\Pi, \pi_{e}, V) of problem (3.9) verifying the following regularity
(\Pi, \pi_{e}, V)\in \mathcal{W}({\cal Q}_{S}), \big(\frac{\partial \Pi}{\partial t}, \frac{\partial V}{\partial t}\big)\in L^{4/3}(S_0, S_f;\mathbb{V}')\times L^{2}({\cal Q}_{S}), (\Pi, V)\in (C^{0}([S_0, S_f];L^{2}(\Omega)))^{2}, |
where \mathcal{W}({\cal Q}_{S}) = \big(L^{2}(S_0, S_f; \mathbb{V})\cap L^{\infty}(S_0, S_f; \mathbb{H})\big)\times L^{2}(S_0, S_f; \mathbb{U}) \times L^{\infty}(S_0, S_f; L^{2}(\Omega)) .
2. For (V_{0}, \Pi_{0}) and (g_{1}, g_{2}) given such that V_{0}\in L^{3}(\Omega) , \pi_{i}^{(0)}-\pi_{e}^{(0)}\! = \! \Pi_{0} with (\Pi_{0}, \pi_{e}^{(0)})\in (H^{1}(\Omega))^{2} and (g_{1}, g_{2})\in L^2({\cal Q}_{S})\times L^{2}(S_{0}, S_{f}; L^{3}(\Omega)) , the solution (\Pi, \pi_{e}, V) of (3.9) is in \mathbb{D}({\cal Q}_{S}) , where
\mathbb{D}({\cal Q}_{S}) = \mathbb{A}_{S}\times L^{\infty}(S_{0}, S_{f}; \mathbb{U})\times \mathbb{B}_{S} , with \mathbb{A}_{S} = L^{\infty}(S_{0}, S_{f}; \mathbb{V})\cap H^{1}(S_{0}, S_{f}; \mathbb{H})\cap C^{0}([S_{0}, S_{f}]; \mathbb{H}) and \mathbb{B}_{S} = C^{0}([S_{0}, S_{f}]; L^{3}(\Omega))\cap H^{1}(S_{0}, S_{f}; L^{3}(\Omega)) .
Proof. We will just sketch the proof based on suitable a priori estimates. The weak formulation of Problem (3.9) can be written as follows ( \forall (v, v_{e}, \rho)\in \mathbb{V}\times\mathbb{U}\times \mathbb{H} and a.e. in (S_0, S_f) )
\begin{equation} \begin{array}{lcr} \big\langle \mathfrak{c}_{m}\frac{\partial \Pi}{\partial t}, v \big\rangle_{\mathbb{V}', \mathbb{V}} +\int_{\Omega}\mathcal{K}_{i}\nabla \Pi\nabla v d{\bf{x}} +\int_{\Omega}\mathcal{K}_{i}\nabla \pi_{e}\nabla v d{\bf{x}} \quad +\int_{\Omega}\frac{\partial {\cal I}}{\partial \phi}\Pi v d{\bf{x}}\\ \quad +\int_{\Omega}\frac{\partial {\cal I}}{\partial u}V vd{\bf{x}} = \int_{\Omega}g_{1} vd{\bf{x}}, \\ \int_{\Omega}(\mathcal{K}_{i}+\mathcal{K}_{e})\nabla\pi_{e}\nabla v_{e} d{\bf{x}} +\int_{\Omega}\mathcal{K}_{i}\nabla \Pi\nabla v_{e} d{\bf{x}} = \int_{\Omega}B_{1}h_{1}v_{e} d{\bf{x}}, \\ \big( \frac{\partial V}{\partial t}, \rho \big)_{\mathbb{H}}+\int_{\Omega}\frac{\partial {\cal G}}{\partial \phi}.\Pi \rho d{\bf{x}} +\int_{\Omega}\hbar V \rho d{\bf{x}} = \int_{\Omega}g_{2}\rho d{\bf{x}}. \end{array} \end{equation} | (3.10) |
According to Theorem 2.1, we have then
\begin{equation} \begin{array}{lcr} \big\langle \mathfrak{c}_{m}\frac{\partial \Pi}{\partial t}, v \big\rangle_{\mathbb{V}', \mathbb{V}} +\underline{A}_{i}(\Pi, v) \quad +\int_{\Omega}\frac{\partial {\cal I}}{\partial \phi}\Pi v d{\bf{x}}\\ \quad +\int_{\Omega}\frac{\partial {\cal I}}{\partial u}V vd{\bf{x}} = \int_{\Omega}g_{1} vd{\bf{x}}-\int_{\Omega}\mathcal{K}_{i}\nabla \overline{\pi}_{e}\nabla v d{\bf{x}}, \\ \big(\frac{\partial V}{\partial t}, \rho \big)_{\mathbb{H}}+\int_{\Omega}\frac{\partial {\cal G}}{\partial \phi}.\Pi \rho d{\bf{x}} +\int_{\Omega}\hbar V \rho d{\bf{x}} = \int_{\Omega}g_{2}\rho d{\bf{x}}, \end{array} \end{equation} | (3.11) |
where \underline{\pi}_e and \overline{\pi}_e are the unique solutions of
\begin{equation} (A_i + A_e)(\underline{\pi}_e, v_e) +A_i(\Pi, v_e) = 0, \; \; \forall v_e \in \mathbb{U} \end{equation} | (3.12) |
and
\begin{equation} (A_i + A_e)(\overline{\pi}_e, v_e) = \langle {\cal B}_{1}h_{1}, v_{e}\rangle_{\mathbb{V}', \mathbb{V}}, \; \; \forall v_e \in \mathbb{U}. \end{equation} | (3.13) |
From (3.13), we can deduce that (by taking v_e = \overline{\pi}_e )
\begin{equation} \left\| { \overline{\pi}_e } \right\| _{\mathbb{V}}\leq C \left\| { h_{1} } \right\| _{L^{2}(\Omega)} \end{equation} | (3.14) |
and then
\begin{equation} \begin{array}{lcr} \mid \int_{\Omega}g_{1} vd{\bf{x}}-\int_{\Omega}\mathcal{K}_{i}\nabla \overline{\pi}_{e}\nabla v d{\bf{x}}\mid \leq C_{1}\left\| { h_{1} } \right\| _{L^{2}(\Omega)}\left\| { v} \right\| _{\mathbb{V}}+ C_{2}\left\| { g_{1} } \right\| _{L^{2}(\Omega)}\left\| { v} \right\| _{\mathbb{V}}\\ \quad \leq C_{3}(\left\| { h_{1} } \right\| ^{2}_{L^{2}(\Omega)}+\left\| { g_{1} } \right\| ^{2}_{L^{2}(\Omega)})+ \epsilon_{0}\left\| { v} \right\| ^{2}_{\mathbb{V}} \end{array} \end{equation} | (3.15) |
where \epsilon_{0} can be chosen. For (v, \rho) = (\Pi, V) , we have
\begin{equation} \begin{array}{lcr} \frac{1}{2}\frac{d}{dt} \left\| { \mathfrak{b}_{m}\Pi} \right\| _{L^{2}(\Omega)}^{2}+\underline{A}_{i}(\Pi, \Pi) \quad +\int_{\Omega}\frac{\partial {\cal I}}{\partial \phi}\Pi^{2} d{\bf{x}}\\ \quad +\int_{\Omega}\frac{\partial {\cal I}}{\partial u}V \Pi d{\bf{x}} = \int_{\Omega}g_{1} \Pi d{\bf{x}}-\int_{\Omega}\mathcal{K}_{i}\nabla \overline{\pi}_{e}\nabla \Pi d{\bf{x}}, \\ \frac{1}{2}\frac{d}{dt} \left\| { V} \right\| _{L^{2}(\Omega)}^{2} +\int_{\Omega}\frac{\partial {\cal G}}{\partial \phi}.\Pi V d{\bf{x}}+\int_{\Omega}\hbar V^{2} d{\bf{x}} = \int_{\Omega}g_{2} V d{\bf{x}}. \end{array} \end{equation} | (3.16) |
According to the partial derivatives of {\cal I} and {\cal G} given by (2.21), we can deduce from (3.16) that (with \hbar_{11} > 0 )
\begin{equation} \begin{array}{lcr} \frac{1}{2}\frac{d}{dt} \left\| { \mathfrak{b}_{m}\Pi} \right\| _{L^{2}(\Omega)}^{2}+\underline{\alpha}_{i}\left\| { \Pi} \right\| _{\mathbb{V}}^{2}+\int_{\Omega}3\hbar_{11}\Pi^{2}\phi^{2}d{\bf{x}}\\ \quad \leq C_{4}\big(\left\| { h_{1} } \right\| ^{2}_{L^{2}(\Omega)}+\left\| { g_{1} } \right\| ^{2}_{L^{2}(\Omega)}\big)+ \epsilon_{0}\left\| { \Pi} \right\| ^{2}_{\mathbb{V}} \\ \quad +C_{5}\left(\int_{\Omega}\Pi^{2}\mid \phi\mid d{\bf{x}}+\int_{\Omega}\Pi^{2}\mid u\mid d{\bf{x}}+\int_{\Omega}\Pi^{2}d{\bf{x}} +\int_{\Omega}V^{2}d{\bf{x}}+\int_{\Omega} \mid u\mid \mid V\mid \mid \Pi\mid d{\bf{x}}\right), \\ \frac{1}{2}\frac{d}{dt} \left\| { V} \right\| _{L^{2}(\Omega)}^{2} \leq C_{6}\left(\int_{\Omega}\mid\Pi\mid \mid \phi\mid \mid V\mid d{\bf{x}}+\int_{\Omega}\mid\Pi\mid \mid V\mid d{\bf{x}}\right)\\ \quad +C_{7}\big(\left\| { V} \right\| _{L^{2}(\Omega)}^{2}+\left\| { g_{2}} \right\| _{L^{2}(\Omega)}^{2}\big). \end{array} \end{equation} | (3.17) |
Then
\begin{equation} \begin{array}{lcr} \frac{1}{2}\frac{d}{dt} \left\| { \mathfrak{b}_{m}\Pi} \right\| _{L^{2}(\Omega)}^{2}+\underline{\alpha}_{i}\left\| { \Pi} \right\| _{\mathbb{V}}^{2}+\int_{\Omega}3\hbar_{11}\Pi^{2}\phi^{2}d{\bf{x}}\\ \quad \leq C_{4}\big(\left\| { h_{1} } \right\| ^{2}_{L^{2}(\Omega)}+\left\| { g_{1} } \right\| ^{2}_{L^{2}(\Omega)}\big)+ \epsilon_{0}\left\| { \Pi} \right\| ^{2}_{\mathbb{V}}\\ \quad +C_{8}\left\| { \Pi} \right\| _{L^{2}(\Omega)}\big(\left\| { \phi} \right\| _{L^{4}(\Omega)}\left\| { \Pi} \right\| _{L^{4}(\Omega)}+\left\| { u} \right\| _{L^{3}(\Omega)}\left\| { \Pi} \right\| _{L^{6}(\Omega)}\big)\\ \quad +C_{9}\big(\left\| { \Pi } \right\| ^{2}_{L^{2}(\Omega)}+\left\| { u } \right\| _{L^{3}(\Omega)}\left\| { \Pi} \right\| _{L^{6}(\Omega)} \left\| { V} \right\| _{L^{2}(\Omega)}+\left\| { V} \right\| ^{2}_{L^{2}(\Omega)}\big), \\ \frac{1}{2}\frac{d}{dt} \left\| { V} \right\| _{L^{2}(\Omega)}^{2} \leq C_{10}\left\| { \phi } \right\| _{L^{4}(\Omega)}\left\| { \Pi} \right\| _{L^{4}(\Omega)} \left\| { V} \right\| _{L^{2}(\Omega)}\\ \quad +C_{11}\big(\left\| { \Pi} \right\| _{L^{2}(\Omega)}^{2}+\left\| { V} \right\| _{L^{2}(\Omega)}^{2}+\left\| { g_{2}} \right\| _{L^{2}(\Omega)}^{2}\big). \end{array} \end{equation} | (3.18) |
Since L^{r}(\Omega)\subset H^{1}(\Omega) for r\leq 6 , then
\begin{equation} \begin{array}{lcr} \frac{1}{2}\frac{d}{dt} \left\| { \mathfrak{b}_{m}\Pi} \right\| _{L^{2}(\Omega)}^{2}+\underline{\alpha}_{i}\left\| { \Pi} \right\| _{\mathbb{V}}^{2} \quad +\int_{\Omega}3\hbar_{11}\Pi^{2}\phi^{2}d{\bf{x}}\\ \quad \leq C_{4}(\left\| { h_{1} } \right\| ^{2}_{\mathbb{V}'}+\left\| { g_{1} } \right\| ^{2}_{L^{2}(\Omega)})+ (\epsilon_{0}+\epsilon_{1})\left\| { \Pi} \right\| ^{2}_{\mathbb{V}} \\ \quad +C_{12}\left\| { \Pi} \right\| ^{2}_{L^{2}(\Omega)}\big(1+\left\| { \phi} \right\| ^{2}_{L^{4}(\Omega)}+\left\| { u} \right\| ^{2}_{L^{3}(\Omega)}\big) +C_{13}\big(1+\left\| { u } \right\| ^{2}_{L^{3}(\Omega)}\big)\left\| { V} \right\| ^{2}_{L^{2}(\Omega)}, \\ \frac{1}{2}\frac{d}{dt} \left\| { V} \right\| _{L^{2}(\Omega)}^{2} \leq \epsilon_{2}\left\| { \Pi} \right\| ^{2}_{\mathbb{V}}+C_{14}\left\| { V} \right\| ^{2}_{L^{2}(\Omega)}\big(1+\left\| { \phi } \right\| ^{2}_{L^{4}(\Omega)}\big)\\ \quad +C_{15}\left\| { \Pi} \right\| _{L^{2}(\Omega)}^{2}+C_{16}\left\| { g_{2}} \right\| _{L^{2}(\Omega)}^{2}. \end{array} \end{equation} | (3.19) |
By adding the first and second equation of previous system we obtain
\begin{equation} \begin{array}{lcr} \frac{1}{2}\left(\frac{d}{dt} \left\| { \mathfrak{b}_{m}\Pi} \right\| _{L^{2}(\Omega)}^{2}+\frac{d}{dt} \left\| { V} \right\| _{L^{2}(\Omega)}^{2}\right) +\underline{\alpha}_{i}\left\| { \Pi} \right\| _{\mathbb{V}}^{2} +\int_{\Omega}3\hbar_{11}\Pi^{2}\phi^{2}d{\bf{x}}\\ \quad \leq (\epsilon_{0}+\epsilon_{1}+\epsilon_{2})\left\| { \Pi} \right\| ^{2}_{\mathbb{V}} +C_{17}\big(\left\| { h_{1} } \right\| ^{2}_{L^{2}(\Omega)}+\left\| { g_{1} } \right\| ^{2}_{L^{2}(\Omega)}+\left\| { g_{2}} \right\| _{L^{2}(\Omega)}^{2}\big) \\ \quad +C_{18}\big(\left\| { V} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { \Pi} \right\| ^{2}_{L^{2}(\Omega)}\big)\big(1+\left\| { \phi } \right\| ^{2}_{L^{4}(\Omega)}+\left\| { u } \right\| ^{2}_{L^{3}(\Omega)}\big). \end{array} \end{equation} | (3.20) |
By choosing \epsilon_{i} such that (\epsilon_{0}+\epsilon_{1}+\epsilon_{2}) = \frac{1}{2}\underline{\alpha}_{i} , we can deduce that
\begin{equation} \begin{array}{lcr} \left(\frac{d}{dt} \left\| { \mathfrak{b}_{m}\Pi} \right\| _{L^{2}(\Omega)}^{2}+\frac{d}{dt} \left\| { V} \right\| _{L^{2}(\Omega)}^{2}\right) +\left\| { \Pi} \right\| _{\mathbb{V}}^{2}+\int_{\Omega}\Pi^{2}\phi^{2}d{\bf{x}}\\ \quad \leq C_{19}\big(\left\| { h_{1} } \right\| ^{2}_{L^{2}(\Omega)}+\left\| { g_{1} } \right\| ^{2}_{L^{2}(\Omega)}+\left\| { g_{2}} \right\| _{L^{2}(\Omega)}^{2}\big)\\ \quad +C_{20}\big(\left\| { V} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { \mathfrak{b}_{m} \Pi} \right\| ^{2}_{L^{2}(\Omega)}\big)\big(1+\left\| { \phi } \right\| ^{2}_{L^{4}(\Omega)}+\left\| { u } \right\| ^{2}_{L^{3}(\Omega)}\big). \end{array} \end{equation} | (3.21) |
Then, Gronwall's lemma yields (for all t\in (S_0, S_f) )
\left\| { V(t)} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { \mathfrak{b}_{m}\Pi(t)} \right\| ^{2}_{L^{2}(\Omega)} \leq \Big(\left\| { V_{0}} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { \Pi_{0}} \right\| ^{2}_{L^{2}(\Omega)}\Big)e^{G(t)}+e^{G(t)}\int_{S_0}^{S_f}e^{-G(s)}F(s)ds, |
where the functions
\begin{equation*} \begin{array}{lcr} G(s) = \int_{0}^{s}C_{20}(1+\left\| { \phi } \right\| ^{2}_{L^{4}(\Omega)}+\left\| { u } \right\| ^{2}_{L^{3}(\Omega)})d\tau, \\ F(s) = C_{19}\big(\left\| { h_{1} } \right\| ^{2}_{L^{2}(\Omega)}+\left\| { g_{1} } \right\| ^{2}_{L^{2}(\Omega)}+\left\| { g_{2}} \right\| _{L^{2}(\Omega)}^{2}\big), \end{array} \end{equation*} |
exist (since (h_{1}, g_{1}, g_{2}, u, \phi)\in L^{2}({\cal Q}_{S})\times L^{2}({\cal Q}_{S})\times L^{2}({\cal Q}_{S})\times L^{2}(S_0, S_f; L^{3}(\Omega))\times L^{4}({\cal Q}_{S})) . Consequently, according to the boundedness of \mathfrak{b}_{m} (for all t\in (S_0, S_f) )
\begin{equation} \begin{array}{lcr} \left\| { V(t)} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { \Pi(t)} \right\| ^{2}_{L^{2}(\Omega)} \leq C\Big(\left\| { V_{0}} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { \Pi_{0}} \right\| ^{2}_{L^{2}(\Omega)}\\ \quad +\left\| { h_{1} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { g_{1} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { g} \right\| _{L^{2}({\cal Q}_{S})}^{2}\Big). \end{array} \end{equation} | (3.22) |
Then according to (3.21), we can deduce that
\begin{equation} \begin{array}{lcr} \left\| { \Pi} \right\| ^{2}_{L^{2}(S_0, S_f;\mathbb{V})} \leq C\Big(\left\| { V_{0}} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { \Pi_{0}} \right\| ^{2}_{L^{2}(\Omega)}\\ \quad +\left\| { h_{1} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { g_{1} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { g_{2}} \right\| _{L^{2}({\cal Q}_{S})}^{2}\Big), \\ \int_{S_0}^{S_f}\int_{\Omega}\Pi^{2}\phi^{2}d{\bf{x}}dt = \left\| { \Pi \phi} \right\| ^{2}_{L^{2}({\cal Q}_{S})} \leq C\Big(\left\| { V_{0}} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { \Pi_{0}} \right\| ^{2}_{L^{2}(\Omega)}\\ \quad +\left\| { h_{1} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { g_{1} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { g_{2}} \right\| _{L^{2}({\cal Q}_{S})}^{2}\Big). \end{array} \end{equation} | (3.23) |
From the second equation of (3.10) and relation (3.23), we can deduce that
\begin{equation} \left\| { \pi_{e}} \right\| ^{2}_{L^{2}(S_0, S_f;\mathbb{V})} \leq C\Big(\!\left\| { V_{0}} \right\| ^{2}_{L^{2}(\Omega)}\!+\!\left\| { \Pi_{0}} \right\| ^{2}_{L^{2}(\Omega)}\!+\!\left\| { h_{1} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}\!+\! \left\| { g_{1} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}\!+\!\left\| { g_{2}} \right\| _{L^{2}({\cal Q}_{S})}^{2}\! \Big). \end{equation} | (3.24) |
According to (3.22), (3.23) and (3.24), we can conclude that
\begin{equation} (\Pi, V) \in (L^{\infty}(S_{0}, S_{f};L^{2}(\Omega)))^{2}, \; \; (\Pi, \pi_{e}) \in (L^{2}(S_0, S_f;\mathbb{V}))^{2}\; \; \text{and}\; \; \Pi \phi\in L^{2}({\cal Q}_{S}). \end{equation} | (3.25) |
From (3.10), we can deduce that
\begin{equation} \begin{array}{lcr} \mid \big\langle \mathfrak{c}_{m}\frac{\partial \Pi}{\partial t}, v \big\rangle_{\mathbb{V}', \mathbb{V}}\mid \leq C_{1} \left\| { v } \right\| _{\mathbb{V}}\big(\left\| { \Pi } \right\| _{\mathbb{V}}+\left\| { \pi_{e} } \right\| _{\mathbb{V}}+\left\| { g_{1} } \right\| _{L^{2}(\Omega)}\big)\\ \quad +C_{2}\int_{\Omega}\mid \phi\mid \mid \phi \Pi\mid \mid v\mid d{\bf{x}}\\ \quad +C_{3}\big(\int_{\Omega} \mid \phi\mid \mid \Pi\mid \mid v\mid d{\bf{x}}+ \int_{\Omega} \mid u\mid\mid \Pi\mid \mid v\mid d{\bf{x}}+\int_{\Omega} \mid \phi\mid \mid V\mid \mid v\mid d{\bf{x}}\big), \\ \mid \big(\frac{\partial V}{\partial t}, \rho \big)_{\mathbb{H}}\mid\leq C_{4}\left\| { \rho } \right\| _{L^{2}(\Omega)}\big(\left\| { \Pi } \right\| _{L^{2}(\Omega)}+\left\| { V } \right\| _{L^{2}(\Omega)}+\left\| { g_{2} } \right\| _{L^{2}(\Omega)}\big)\\ \quad +C_{5}\int_{\Omega} \mid \phi\mid \mid \Pi\mid \mid \rho\mid d{\bf{x}}, \end{array} \end{equation} | (3.26) |
and then
\begin{equation} \begin{array}{lcr} \left\| { \frac{\partial \Pi}{\partial t}} \right\| _{\mathbb{V}'}\leq C_{6}\big(\left\| { \Pi } \right\| _{\mathbb{V}}+\left\| { \pi_{e} } \right\| _{\mathbb{V}}+\left\| { g_{1} } \right\| _{L^{2}(\Omega)}\big)+C_{7}\left\| { V } \right\| _{L^{2}(\Omega)}\left\| { \phi } \right\| _{L^{4}(\Omega)}\\ \quad +C_{8}\big(\left\| { \phi \Pi} \right\| _{L^{2}(\Omega)}\left\| { \phi } \right\| _{L^{4}(\Omega)}+ \left\| { \phi } \right\| _{L^{2}(\Omega)}\left\| { \Pi } \right\| _{\mathbb{V}}+ \left\| { u } \right\| _{L^{2}(\Omega)}\left\| { \Pi } \right\| _{\mathbb{V}}\big), \\ \left\| { \frac{\partial V}{\partial t}} \right\| _{L^{2}(\Omega)}\leq C_{9}\big(\left\| { V} \right\| _{L^{2}(\Omega)}+\left\| { \Pi } \right\| _{L^{2}(\Omega)}+\left\| { \phi \Pi} \right\| _{L^{2}(\Omega)}+\left\| { g_{2} } \right\| _{L^{2}(\Omega)}\big). \end{array} \end{equation} | (3.27) |
Consequently
\begin{equation*} \begin{array}{lcr} \int_{S_0}^{S_f}\left\| { \frac{\partial \Pi}{\partial t}} \right\| ^{4/3}_{\mathbb{V}'} dt \leq C_{10}\big(\left\| { \Pi } \right\| ^{4/3}_{L^{2}(S_0, S_f;\mathbb{V})}+\left\| { \pi_{2} } \right\| ^{4/3}_{L^{2}(S_0, S_f;\mathbb{V})}+ \left\| { g_{1} } \right\| ^{4/3}_{L^{2}({\cal Q}_{S})}\big)\\ \quad +C_{11}\big(\left\| { \phi \Pi} \right\| ^{4/3}_{L^{2}({\cal Q}_{S})}\left\| { \phi } \right\| ^{4/3}_{L^{4}(S_0, S_f;L^{4}(\Omega))}+\left\| { V } \right\| ^{4/3}_{L^{2}({\cal Q}_{S})}\left\| { \phi } \right\| ^{4/3}_{L^{4}(S_0, S_f;L^{4}(\Omega))}\big)\\ \quad +C_{12}\big(\left\| { \Pi} \right\| ^{4/3}_{L^{2}(S_0, S_f;\mathbb{V})}\left\| { u } \right\| ^{4/3}_{L^{4}(S_0, S_f;L^{2}(\Omega))}+\left\| { \Pi} \right\| ^{4/3}_{L^{2}(S_0, S_f;\mathbb{V})}\left\| { \phi} \right\| ^{4/3}_{L^{4}(S_0, S_f;L^{2}(\Omega))}\big), \\ \int_{S_0}^{S_f}\left\| { \frac{\partial V}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}dt\leq C_{13}\big(\left\| { \Pi } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { V } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+ \left\| { \phi \Pi } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { g } \right\| ^{2}_{L^{2}({\cal Q}_{S})}\big) \end{array} \end{equation*} |
and then (according to (3.22), (3.23) and (3.24))
\begin{equation} \begin{array}{lcr} \left\| { \frac{\partial \Pi}{\partial t}} \right\| _{L^{4/3}(S_0, S_f;\mathbb{V}')}+\left\| { \frac{\partial V}{\partial t}} \right\| _{L^{2}(S_0, S_f;L^{2}(\Omega))}\\ \quad \leq C\left(\left\| { V_{0}} \right\| _{L^{2}(\Omega)}+\left\| { \Pi_{0}} \right\| _{L^{2}(\Omega)}+\left\| { h_{1} } \right\| _{L^{2}({\cal Q}_{S})}+\left\| { g_{1} } \right\| _{L^{2}({\cal Q}_{S})}+\left\| { g_{2}} \right\| _{L^{2}({\cal Q}_{S})}\right). \end{array} \end{equation} | (3.28) |
We can conclude that \frac{\partial \Pi}{\partial t}\in L^{4/3}(S_0, S_f; \mathbb{V}') \; \; \text{and}\; \; \frac{\partial V}{\partial t} \in L^{2}({\cal Q}_{S}). The proof of Theorem can be completed by implementing the Galerkin method and by taking advantage of the above estimates, (3.22)–(3.24) and (3.28), and Lemma 2.4 and Remark 2.1. So we omit the details. Since the problem is linear, then from the estimates (3.22)–(3.24) and (3.28) we can deduce the Lipschitz continuity result and then the uniqueness of solution.
(ⅱ) Prove first that V\in L^{2}(S_0, S_f; L^{3}(\Omega)) . Since V satisfies the equation
\begin{equation} \frac{\partial V }{\partial t}\! = \! -\frac{\partial \mathcal{I}_{2}}{\partial \phi}(.;\phi)\Pi-\hbar V +g_{2}. \end{equation} | (3.29) |
Since (3.29) is similar as (3.39), then by using similar argument to derive (3.43) we obtain
\begin{equation} \begin{array}{lcr} \left\| { V(., t)} \right\| _{L^{3}(\Omega)} \leq C_{1}\int_{S_0}^{t} \left\| { V(., s)} \right\| _{L^{3}(\Omega)} ds\\ \quad +C_{2}\left(\left\| { g_{2}} \right\| _{L^{2}(S_0, S_f;L^{3}(\Omega))} +\left\| { \phi} \right\| _{L^{2}(S_0, S_f;L^{6}(\Omega))}\left\| { \Pi} \right\| _{L^{2}(S_0, S_f;L^{6}(\Omega))} \right). \end{array} \end{equation} | (3.30) |
Since \Pi and \phi are in L^{2}(0, T, H^{1})\subset L^{2}(0, T, L^{r}) ( r\in [1, 6] ) and g_{2}\in L^{2}(S_0, S_f; L^{3}(\Omega)) , then
\begin{equation} \begin{array}{lcr} \left\| { V(., t)} \right\| _{L^{3}(\Omega)}\leq C_{3}+C_{4}\int_{S_0}^{t}\left\| { V(., s)} \right\| _{L^{3}(\Omega)}ds. \end{array} \end{equation} | (3.31) |
Consequently, by using Gronwall lemma, \left\| { V(., t)} \right\| _{L^{3}(\Omega)}\leq C and then V\in L^{\infty}(S_0, S_f; L^{3}(\Omega)) .
Prove now that (\Pi, \pi_{e})\in (L^{\infty}(S_0, S_f; H^{1}(\Omega)))^{2} . For this we will just sketch the proof based on suitable a priori estimates. From (3.10) with (v, v_{e}) = (\frac{\partial \Pi}{\partial t}, \frac{\partial \pi_{e}}{\partial t}) (put \tilde{h}_{1} = B_{1} h_{1} )
\begin{equation} \begin{array}{lcr} \left\| { \mathfrak{b}_{m}\frac{\partial \Pi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}+A_i(\Pi+\pi_e, \frac{\partial \Pi}{\partial t}) +\int_{\Omega}\frac{\partial {\cal I}}{\partial \phi}\Pi \frac{\partial \Pi}{\partial t} d{\bf{x}} +\int_{\Omega}\frac{\partial {\cal I}}{\partial u}V \frac{\partial \Pi}{\partial t} d{\bf{x}} = \int_{\Omega} {g}_{1}\frac{\partial \Pi}{\partial t}d{\bf{x}}, \\ A_i(\Pi + \pi_e, \frac{\partial \pi_{e}}{\partial t})+ A_e(\pi_e, \frac{\partial \pi_{e}}{\partial t}) = \int_{\Omega}\tilde{h}_{1} \frac{\partial \pi_{e}}{\partial t}d{\bf{x}}. \end{array} \end{equation} | (3.32) |
Then
\begin{equation} \begin{array}{lcr} \left\| { \mathfrak{b}_{m}\frac{\partial \Pi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)} + \frac{d}{2dt}A_i(\Pi+\pi_e, \Pi+\pi_e) +\frac{d}{2dt}A_i(\pi_e, \pi_e)\\ \quad = -\int_{\Omega}\frac{\partial {\cal I}}{\partial \phi}\Pi \frac{\partial \Pi}{\partial t} d{\bf{x}} -\int_{\Omega}\frac{\partial {\cal I}}{\partial u}V\frac{\partial \Pi}{\partial t} d{\bf{x}}\\ \quad +\int_{\Omega} g_{1}\frac{\partial \Pi}{\partial t}d{\bf{x}}+\int_{\Omega} \pi_{e}\frac{\partial \tilde{h}_{1}}{\partial t} d{\bf{x}}+\frac{d}{dt}\big(\int_{\Omega} \tilde{h}_{1} \pi_{e} d{\bf{x}}\big). \end{array} \end{equation} | (3.33) |
Since {h}_{1}\in U_{c} then \tilde{h}_{1}\in U_{c}\subset C^{0}([S_0, S_f]; L^{2}(\Omega)) . So, \tilde{h}_{1}(S_0)\in L^{2}(\Omega) and we have \left\| { \tilde{h}_{1}(S_0)} \right\| ^{2}_{L^{2}(\Omega)}\leq C_{1} \left\| { \tilde{h}_{1}} \right\| ^{2}_{U_{c}} . Consequently, by integrating (3.33) by time, we can deduce (from (2.21), the boundedness of \mathfrak{b}_{m} , coercivity and continuity of A_{i} and A_{e} , and regularity H^{1} of (\Pi_{0}, \pi_{e}^{(0)}) )
\begin{equation} \begin{array}{lcr} 2\underline{c}_{m}\int_{S_0}^{t}\left\| { \frac{\partial \Pi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+ \alpha_{i}\left\| { \Pi(t)} \right\| ^{2}_{H^{1}(\Omega)}+(\alpha_{e}+\alpha_{i})\left\| { \pi_e(t)} \right\| ^{2}_{H^{1}(\Omega)}\\ \quad \leq C_{0}\Big( \int_{S_0}^{t}\int_{\Omega}\mid \phi \mid ^{4}\mid \Pi \mid^{2} d{\bf{x}} ds+\int_{S_0}^{t}\int_{\Omega}\mid \phi \mid ^{2}\mid \Pi \mid^{2} d{\bf{x}} ds\Big)\\ \quad +C_{1}\Big(\int_{S_0}^{t}\int_{\Omega}\mid u \mid ^{2}\mid \Pi \mid^{2} d{\bf{x}} ds+\int_{S_0}^{t}\int_{\Omega}\mid \phi \mid ^{2}\mid V \mid^{2} d{\bf{x}} ds\Big)\\ \quad +C_{2} \Big(\int_{S_0}^{t}\int_{\Omega}\mid \Pi \mid^{2} d{\bf{x}} ds+\int_{S_0}^{t}\int_{\Omega}\mid V\mid^{2} d{\bf{x}} ds+\int_{S_0}^{t}\int_{\Omega}\mid \pi_{e}\mid^{2} d{\bf{x}} ds\Big)\\ \quad +\delta_{0}\int_{S_0}^{t}\left\| { \frac{\partial \Pi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+\delta_{1}\left\| { \pi_{e}(t)} \right\| ^{2}_{H^{1}(\Omega)} +C_{3}\big(1+\left\| { g_{1} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { \tilde{h}_{1}} \right\| ^{2}_{U_{c}}\big). \end{array} \end{equation} | (3.34) |
Since \phi is in L^{\infty}(0, T, H^{1})\subset L^{\infty}(0, T, L^{r}) ( r\in [1, 6] ), then (by choosing \delta_{0} = \underline{c}_{m} , \delta_{1} = (\alpha_{e}+\alpha_{i})/2 )
\begin{equation} \begin{array}{lcr} \int_{S_0}^{t}\left\| { \frac{\partial \Pi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+ \left\| { \Pi(t)} \right\| ^{2}_{H^{1}(\Omega)}+\left\| { \pi_e(t)} \right\| ^{2}_{H^{1}(\Omega)}\\ \quad \leq C_{4}\big(\left\| { \phi } \right\| ^{4}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { \phi } \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))} +\left\| { u } \right\| ^{2}_{L^{\infty}(0, T;L^{3}(\Omega))}\big)\int_{S_0}^{t}\left\| { \Pi } \right\| ^{2}_{H^{1}(\Omega)}ds\\ \quad +C_{5} \big(\left\| { \phi } \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}\left\| { V } \right\| ^{2}_{L^{\infty}(S_0, S_f;L^{3}(\Omega))} +\left\| { \Pi } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { \pi_{e} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { V } \right\| ^{2}_{L^{2}({\cal Q}_{S})}\big)\\ \quad +C_{6}\big(1+\left\| { g_{1} } \right\| ^{2}_{L^{2}({\cal Q}_{S})}+\left\| { \tilde{h}_{1}} \right\| ^{2}_{U_{c}}\big). \end{array} \end{equation} | (3.35) |
From the regularity of (\Pi, \pi_{e}, V, g_{1}, \tilde{h}_{1}) , estimation of (\Pi, \pi_{e}) in L^{\infty}(S_0, S_f; H^{1}(\Omega)) and and estimation of V in L^{\infty}(S_0, S_f; L^{3}(\Omega)) , respectively, we can deduce that
\begin{equation} \begin{array}{lcr} \int_{S_0}^{t}\left\| { \frac{\partial \Pi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+ \left\| { \Pi(t)} \right\| ^{2}_{H^{1}(\Omega)}+\left\| { \pi_e(t)} \right\| ^{2}_{H^{1}(\Omega)} \leq C \end{array} \end{equation} | (3.36) |
and then \frac{\partial \Pi}{\partial t}\in L^{2}({\cal Q}_{S}) and (\Pi, \pi_{e})\in L^{\infty}(S_0, S_f; H^{1}(\Omega)) .
The proof of the result can be completed by implementing the classical Faedo-Galerkin method and by taking advantage of above estimates. So we omit the details.
Finally we prove that V\in C^{0}([S_0, S_f], L^{3}(\Omega)) . Let R = -\frac{\partial \mathcal{I}_{2}}{\partial \phi}-\hbar V+g_{2} be the right hand side of equation (3.29). Since \frac{\partial \mathcal{I}_{2}}{\partial \phi} is a polynomial of degree 1 on \phi , we can deduce that (since (\phi, \Pi)\in (L^{\infty}(S_{0}, S_{f}; H^{1}(\Omega)))^{2} and V\in L^{\infty}(S_{0}, S_{f}; L^{3}(\Omega)) )
\begin{equation*} \begin{array}{lcr} \left\| { R } \right\| _{L^{3}(\Omega)}\leq C\left(\left\| { \Pi} \right\| _{L^{\infty}(S_{0}, S_{f}; H^{1}(\Omega))}\big(1+\left\| { \phi} \right\| _{L^{\infty}(S_{0}, S_{f}; H^{1}(\Omega))}\big)+\left\| { V} \right\| _{L^{\infty}(S_{0}, S_{f}; L^{3}(\Omega))}+ \left\| { g_{2}} \right\| _{L^{3}(\Omega)}\right). \end{array} \end{equation*} |
Since g_{2}\in L^{2}(S_{0}, S_{f}; L^{3}(\Omega)) , then \left\| { R } \right\| _{L^{2}(S_{0}, S_{f}; L^{3}(\Omega))}\leq C . Consequently, R\in L^{2}(S_{0}, S_{f}; L^{3}(\Omega)) and then, from (3.29), \frac{\partial V }{\partial t}\in L^{2}(S_{0}, S_{f}; L^{3}(\Omega)) . As V\in L^{2}(S_{0}, S_{f}; L^{3}(\Omega)) and \frac{\partial V}{\partial t}\in L^{2}(S_{0}, S_{f}; L^{3}(\Omega)) we can conclude, from Remark 2.3, that V\in C^{0}([S_{0}, S_{f}]; L^{3}(\Omega)) . This completes the proof.
We can now prove the well-posedness of problem (3.7).
Theorem 3.1. Let assumptions (H1)-(H4), (HMC) and (RC) be fulfilled. Suppose that (\phi, \varphi_{e}, u) \in \mathbb{D} , then the following results hold.
(i) For any (h_{1}, h_{2})\in U_{c}\times L^2({\cal Q}) under the compatibility condition (1.2), there exists a unique weak solution (\Psi, \psi_{e}, w)\in \mathbb{D} , of the linear problem (3.7).
(ii) Let (h_{1, i}, h_{2, i}) , i = 1, 2 be given in U_{c}\times L^2({\cal Q}) . If (\Psi_{i}, \psi_{e, i}, w_{i}) is the solution of (3.7) corresponding to data (h_{1, i}, h_{2, i}) , for i = 1, 2 , then
\begin{equation} \begin{array}{lcr} \left\| { (\Psi, \psi_{e}, w)} \right\| ^{2}_{\mathbb{D}} \leq C\left(\Vert h_{1}\Vert_{U_{c}}^2 + \Vert h_{2}\Vert_{L^2({\cal Q})}^2 \right), \end{array} \end{equation} | (3.37) |
where (\Psi, \psi_{e}, w) = (\Psi_{1}-\Psi_{2}, \psi_{e, 1}-\psi_{e, 2}, w_{1}-w_{2}) and (h_{1}, h_{2}) = (h_{1, 1}-h_{1, 2}, h_{2, 1}-h_{2, 2}) .
Proof. To prove the existence of a unique solution on {\cal Q} , we first establish the existence of a unique solution on \tilde{\cal Q}_{j} = \Omega\times (s_{0}, s_{j}), j\geq1 and obtain some estimations.
We solve the problem on \tilde{\cal Q}_{1} and obtain the existence of a unique solution on \tilde{\cal Q}_{1} . Then, the existence of a unique solution on \tilde{\cal Q}_{2} is proved by using the solution on \tilde{\cal Q}_{1} to generate the initial data at s_{1} . This advancing process is repeated for \tilde{\cal Q}_{3}, \tilde{\cal Q}_{4}, ... until the final set is reached. Hereafter, the solution on \tilde{\cal Q}_{j} will be denoted by (\Psi_j, \psi_{e, j}, w_j) for j = 1, ... .
Now we introduce the following problems ({\cal P}_{j}) for j\in \mbox{IN}^{{}}-\{0\} (for ({\bf x}, t)\in \Omega\times (s_{j-1}, s_{j}) )
\begin{equation} \begin{array}{lcr} \mathfrak{c}_{m}\frac{\partial \Pi_j }{\partial t}+\frac{\partial {\cal I}}{\partial \phi}(\phi, u).\pi_j+\frac{\partial {\cal I}}{\partial u}(\phi, u).V_j-div(\mathcal{K}_{i}\nabla (\Pi_j+\pi_{e, j})) = g_{1, j}, \\ -div\left((\mathcal{K}_{e}+\mathcal{K}_{i})\nabla \pi_{e, j}\right)-div(\mathcal{K}_{i}\nabla \Pi_j) = {\cal B}_{1}h_{1}, \\ \frac{\partial V_j }{\partial t}+\frac{\partial {\cal G}}{\partial \phi}(\phi, u).\Pi_j+\frac{\partial {\cal G}}{\partial u}(\phi, u).V_j = g_{2, j}, \\ \big(\mathcal{K}_{i}\nabla (\Pi_j + \pi_{e, j})\big) \cdot{\bf{n}} = 0, \; \text{on}\; \partial\Omega\times(s_{j-1}, s_{j}) \\ \big(\mathcal{K}_{e}\nabla \pi_{e, j}\big)\cdot{\bf{n}} = 0, \; \text{on}\; \partial\Omega\times(s_{j-1}, s_{j}) \\ \Pi_j({\bf{x}}, s_{j-1}) = \Psi_{j-1}({\bf{x}}, s_{j-1})\in L^{2}(\Omega), \\ V_j({\bf{x}}, s_{j-1}) = w_{j-1}({\bf{x}}, s_{j-1})\in L^{2}(\Omega) \end{array} \end{equation} | (3.38) |
where (\Psi_{j-1}, w_{j-1})\in (L^{2}(\tilde{\cal Q}_{j-1}))^{2} and
\begin{eqnarray*} g_{1, j}({\bf x}, t) = {\cal B}_{2}h_{2}+\sum\limits_{k = 1}^{n_{1}} a_k({\bf{x}}, t) \Psi_{j-1}({\bf{x}}, t-\xi_{k}(t))+\sum\limits_{l = 1}^{n_{2}}b_l({\bf{x}}, t) w_{j-1}({\bf{x}}, t-\eta_{l}(t)), \\ g_{2, j}({\bf x}, t) = \sum\limits_{k = 1}^{n_{1}} c_k({\bf{x}}, t) \Psi_{j-1}({\bf{x}}, t-\xi_{k}(t))+\sum\limits_{l = 1}^{n_{2}}d_l({\bf{x}}, t) w_{j-1}({\bf{x}}, t-\eta_{l}(t)). \end{eqnarray*} |
Since B_{2}h_{2} \in L^{2}({\cal Q}) , (\Psi_{j-1}, w_{j-1})\in L^{2}(\tilde{\cal Q}_{j-1}) and a_{k}, c_{k}, b_l, d_l (for 1\leq k\leq n_{1} and 1\leq l\leq n_{2} ) are in {\cal C}^{\infty}(\overline{\cal Q}) , then, according to Lemma 2.6, we have that g_{1, j} and g_{2, j} are in L^{2}(\tilde{\cal Q}_{j}) . Then using Lemma 3.1 we have that the problem ({\cal P}_{j}) admits a unique solution (\Pi_{j}, \pi_{e, j}, V_{j}) \in {\cal W}(\Omega\times (s_{j-1}, s_{j})) and verifying (\frac{\partial \Pi_{j}}{\partial t}, \frac{\partial V_{j}}{\partial t})\in L^{4/3}(s_{j-1}, s_{j}; \mathbb{V}')\times L^{2}(s_{j-1}, s_{j}; L^{2}(\Omega))) and (\Pi_{j}, V_{j})\in (C^{0}([s_{j-1}, s_{j}]; L^{2}(\Omega)))^{2} . Then, we can extend the result to the cylinder set \tilde{\cal Q}_{j+1} by taking (\Psi_{j}, \psi_{e, j}, w_{j}) = (\Psi_{j-1}, \psi_{e, j-1}, w_{j-1}) on \tilde{\cal Q}_{j-1} and (\Psi_{j}, \psi_{e, j}, w_{j}) = (\Pi_{j}, \pi_{e, j}, V_{j}) on \Omega\times(s_{j-1}, s_{j}) .
We observe that for j = 1, we have (according to the initial condition in (3.7))
\begin{eqnarray*} g_{1, 1}({\bf x}, t) = {\cal B}_{2}h_{2}, \; \; \text{and}\; \; g_{2, 1}({\bf x}, t) = 0 \end{eqnarray*} |
and then g_{1, 1}, g_{2, 1}\in L^{2}(\tilde{\cal Q}_{1}) . By using the previous result, the problem ({\cal P}_{1}) admits a unique solution (\Pi_{1}, \pi_{e, 1}, V_{1}) and then the solution (\Psi_{1}, \psi_{e, 1}, w_{1}) . We inject now (\Psi_{1}, \psi_{e, 1}, w_{1}) in the problem ({\cal P}_{2}) and by using the same approach, we obtain the existence and uniqueness of (\Pi_{2}, \pi_{e, 2}, V_{2}) (solution of ({\cal P}_{2}) ).
We can now iterate the process for any domain \tilde{\cal Q}_{j}, for \; j\geq 1 and we obtain the existence and uniqueness of (\Pi_{j}, \pi_{e, j}, V_{j}) solution of ({\cal P}_{j}) .
We deduce then the existence and uniqueness of the solution (\Psi, \psi_{e}, w)\in{\cal W}({\cal Q}) of (3.8) (verifying (\frac{\partial \Psi}{\partial t}, \frac{\partial w}{\partial t})\in L^{4/3}(0, T; \mathbb{V}')\times L^{2}(0, T; L^{2}(\Omega))) and (\Psi, w)\in (C^{0}([0, T]; L^{2}(\Omega)))^{2} ) such that (\Psi, \psi_{e}, w)|_{\tilde{\cal Q}_{j}} = (\Psi_{j}, \psi_{e, j}, w_{j}), j\geq 1 .
Prove now that the unique solution (\Psi, \psi_{e}, w) satisfies the following regularity: (\Psi, \psi_{e}, w)\in (L^{\infty}(0, T, ; H^{1}(\Omega)))^{2}\times C^{0}([0, T]; L^{3}(\Omega)) and (\frac{\partial \Psi}{\partial t}, \frac{\partial w}{\partial t})\in L^{2}({\cal Q})\times L^{2}(0, T; L^{3}(\Omega))) .
Step1. Prove first that w\in L^{2}(0, T; L^{3}(\Omega)) . Since w satisfies the equation
\begin{equation} \frac{\partial w }{\partial t}\! = \! -\frac{\partial \mathcal{I}_{2}}{\partial \phi}(.;\phi)\Psi-\hbar w +\mathcal{E}(\Psi_{\tau}, w_{\tau}). \end{equation} | (3.39) |
Then for all t we have (since w(t = 0) = 0 )
\begin{equation} \begin{array}{lcr} w({\bf{x}}, t) \! = \! -\int_{0}^{t} \frac{\partial \mathcal{I}_{2}}{\partial \phi}({\bf{x}}, s;\phi)\Psi ds-\int_{0}^{t}\hbar w({\bf{x}}, s)ds+\int_{0}^{t}\mathcal{E}(\Psi_{\tau}, w_{\tau}) ds. \end{array} \end{equation} | (3.40) |
Consequently,
\begin{equation} \begin{array}{lcr} \mid w({\bf{x}}, t)\mid \leq C\left(\int_{0}^{t}\mid \frac{\partial \mathcal{I}_{2}}{\partial \phi}({\bf{x}}, s;\phi)\mid \mid \Psi\mid ds+\int_{0}^{t}\mid w({\bf{x}}, s)\mid ds\right)+\int_{0}^{t}\mid \mathcal{E}(\Psi_{\tau}, w_{\tau})\mid ds \end{array} \end{equation} | (3.41) |
and then (since \frac{\partial \mathcal{I}_{2}}{\partial \phi}({\bf{x}}, s; \phi) is linear operator)
\begin{equation} \begin{array}{lcr} \mid w({\bf{x}}, t)\mid \leq C\left(\int_{0}^{t}\mid w({\bf{x}}, s)\mid ds+\int_{0}^{t}\mid \Psi({\bf{x}}, s)\mid \mid \phi({\bf{x}}, s)\mid ds\right)+\int_{0}^{t}\mid \mathcal{E}(\Psi_{\tau}, w_{\tau})\mid ds. \end{array}\ \end{equation} | (3.42) |
This implies
\begin{equation*} \begin{array}{lcr} \left(\int_{\Omega}\mid w({\bf{x}}, t)\mid^{3} d{\bf{x}}\right)^{1/3} \leq C\left(\big(\int_{\Omega}\big(\int_{0}^{t}\mid w({\bf{x}}, s)\mid ds\big)^{3}d{\bf{x}}\big)^{1/3}\right.\\ \quad \left.+\big(\int_{\Omega}\big(\int_{0}^{t}\mid \mathcal{E}(\Psi_{\tau}, w_{\tau})\mid ds\big)^{3}d{\bf{x}}\big)^{1/3} \right.\\ \quad \left.+\big(\int_{\Omega}\big(\int_{0}^{t}\mid \Psi({\bf{x}}, s)\mid \mid \phi({\bf{x}}, s)\mid ds\big)^{3}d{\bf{x}}\big)^{1/3}\right) \end{array} \end{equation*} |
and then (using Minkowski inequality)
\begin{equation} \begin{array}{lcr} \left(\int_{\Omega}\mid w({\bf{x}}, t)\mid^{3} d{\bf{x}}\right)^{1/3} \leq C\Big(\int_{0}^{t}\!\!\!\big(\int_{\Omega}\mid w({\bf{x}}, s)\mid^{3} d{\bf{x}}\big)^{1/3}ds\\ \quad +\int_{0}^{t}\!\!\!\big(\int_{\Omega}\mid \mathcal{E}(\Psi_{\tau}, w_{\tau})\mid^{3} d{\bf{x}}\big)^{1/3}ds \\ \quad +\int_{0}^{t}\!\!\!\big(\int_{\Omega}\mid \Psi({\bf{x}}, s)\mid^{3}\mid \phi({\bf{x}}, s)\mid^{3} d{\bf{x}}\big)^{1/3}ds \Big). \end{array} \end{equation} | (3.43) |
According to Lemma 2.6 and the fact that (\Psi_{past}, w_{past}) = (0, 0) , we can deduce that
\begin{equation} \begin{array}{lcr} \left\| { w(., t)} \right\| _{L^{3}(\Omega)} \leq C_{1}\int_{0}^{t} \left\| { w(., s)} \right\| _{L^{3}(\Omega)} ds\\ \quad +C_{2}\left(\left\| { \Psi} \right\| _{L^{2}(0, T;L^{3}(\Omega))} +\left\| { \phi} \right\| _{L^{2}(0, T;L^{6}(\Omega)}\left\| { \Psi} \right\| _{L^{2}(0, T;L^{6}(\Omega)} \right). \end{array} \end{equation} | (3.44) |
Since \Psi and \phi are in L^{2}(0, T, H^{1}(\Omega))\subset L^{2}(0, T, L^{r}(\Omega)) ( r\in [1, 6] ), then
\begin{equation} \begin{array}{lcr} \left\| { w(., t)} \right\| _{L^{3}(\Omega)}\leq C_{3}+C_{4}\int_{0}^{t}\left\| { w(., s)} \right\| _{L^{3}(\Omega)}ds. \end{array} \end{equation} | (3.45) |
Consequently, by using Gronwall lemma, \left\| { w(., t)} \right\| _{L^{3}(\Omega)}\leq C and then w\in L^{\infty}(0, T; L^{3}(\Omega)) .
Step2. Prove now that (\Psi, \psi_{e})\in (L^{\infty}(0, T; H^{1}(\Omega)))^{2} and w\in C^{0}([0, T]; L^{3}(\Omega)) .
Put g_{1} = \mathcal{H}(\Psi_{\tau}, w_{\tau})+{\cal B}_{2}h_{2} and g_{2} = \mathcal{E}(\Psi_{\tau}, w_{\tau}) . Since w\in L^{\infty}(0, T; L^{3}(\Omega)) and \Psi \in L^{2}(0, T; H^{1}(\Omega)) then, according to Lemma 2.6, g_{1}\in L^{2}({\cal Q}) and g_{2}\in L^{2}(0, T, L^{3}(\Omega)) . Since (\Psi, \psi_{e}, w) is a solution of (3.9) which correspond to data (g_{1}, g_{2}, h_{1}) and (\Psi, w)(t = 0) = (0, 0) with (S_0, S_{f}) = (0, T) , we can deduce from Lemma 3.1 that (\Psi, \psi_{e})\in (L^{\infty}(0, T; H^{1}(\Omega)))^{2} and w\in C^{0}([0, T]; L^{3}(\Omega)) .
We are now going to prove the estimation given in theorem.
Let (\Psi_{i}, \psi_{e, i}, w_{i})_{i = 1, 2} be two solutions of (3.8), corresponding to data (h_{i, 1}, h_{i, 2})_{i = 1, 2} , respectively. We denote \Psi = \Psi_{1}-\Psi_{2} , w = w_{1}-w_{2} , \psi_{e} = \psi_{e, 1}-\psi_{e, 2} , h_{1} = h_{1, 1}-h_{2, 1} , h_{2} = h_{1, 2}-h_{2, 2} . Then according to (3.8) and setting (v, u, v_{e}) = (\Psi, w, \psi_{e}) we can deduce that (since delay operators are linear)
\begin{equation} \begin{array}{lcr} \frac{d}{2dt}\Vert \mathfrak{b}_{m}\Psi\Vert^{2}_{L^{2}(\Omega)} +\int_{\Omega}\mathcal{K}_{i}\nabla \Psi\nabla \Psi d{\bf{x}} +\int_{\Omega}\mathcal{K}_{i}\nabla \psi_{e}\nabla \Psi d{\bf{x}} +\int_{\Omega}\frac{\partial {\cal I}}{\partial \phi}(\phi, u)\Psi^{2} d{\bf{x}}\\ \quad +\int_{\Omega}\frac{\partial {\cal I}}{\partial u}(\phi, u)w \Psi d{\bf{x}} = \int_{\Omega}\mathcal{H}(\Psi_{\tau}, w_{\tau}) \Psi d{\bf{x}}+\int_{\Omega}{\cal B}_{2}h_{2} \Psi d{\bf{x}}, \\ \int_{\Omega}(\mathcal{K}_{i}+\mathcal{K}_{e})\nabla\psi_{e}\nabla \psi_{e} d{\bf{x}} +\int_{\Omega}\mathcal{K}_{i}\nabla \Psi\nabla \psi_{e} d{\bf{x}} = \int_{\Omega}{\cal B}_{1}h_{1}\psi_{e} d{\bf{x}}, \\ \frac{d}{2dt}\Vert w\Vert^{2}_{L^{2}(\Omega)}+\int_{\Omega}\frac{\partial {\cal G}}{\partial \phi}(\phi, u)\Psi w d{\bf{x}} +\int_{\Omega}\hbar w^{2} d{\bf{x}} = \int_{\Omega}\mathcal{E}(\Psi_{\tau}, w_{\tau})w d{\bf{x}}\\ \Psi(., t = 0) = 0, \; \; \; w(., t = 0) = 0, \; \; \text{in}\; \; \Omega\\ \Psi = 0, \; \; \; w = 0, \; \; \text{in}\; \; {\cal Q}_{0}. \end{array} \end{equation} | (3.46) |
Consequently (from the expression of the derivatives of {\cal I} and {\cal G} )
\begin{equation} \begin{array}{lcr} \frac{d}{2dt}\big(\Vert \mathfrak{b}_{m} \Psi\Vert^{2}_{L^{2}(\Omega)}+\Vert w\Vert^{2}_{L^{2}(\Omega)}\big) +A_{i}(\Psi+\psi_{e}, \Psi+\psi_{e})+A_{e}(\psi_{e}, \psi_{e})\\ \quad +\int_{\Omega}(\hbar_{11} \phi^{2}-\hbar_{12} \phi + \hbar_{13} + \epsilon_{1}\hbar_{21} u)\Psi^{2} d{\bf{x}}+\int_{\Omega}\hbar w^{2} d{\bf{x}}\\ +\int_{\Omega}(\epsilon_{1}\hbar_{21}\phi+h_{22})w \Psi d{\bf{x}}+\int_{\Omega}(-\epsilon_{2}\hbar_{31}\phi+(2\epsilon_{2}-1)h_{32})\Psi w d{\bf{x}} \\ = \int_{\Omega}\mathcal{H}(\Psi_{\tau}, w_{\tau}) \Psi d{\bf{x}}+\int_{\Omega}\mathcal{E}(\Psi_{\tau}, w_{\tau})w d{\bf{x}}+\int_{\Omega}{\cal B}_{1}h_{1}\psi_{e} d{\bf{x}}+\int_{\Omega}{\cal B}_{2}h_{2} \Psi d{\bf{x}}, \\ \Psi(., t = 0) = 0, \; \; \; w(., t = 0) = 0, \; \; \text{in}\; \; \Omega\\ \Psi = 0, \; \; \; w = 0, \; \; \text{in}\; \; {\cal Q}_{0}. \end{array} \end{equation} | (3.47) |
Using the boundedness of the function \hbar_{ij} and \hbar , coercivity of A_i and A_e , and the assumption concerning the operators {\cal B}_{i} , i = 1, 2 ), we obtain
\begin{equation} \begin{array}{lcr} \frac{d}{2dt}\big(\Vert \mathfrak{b}_{m}\Psi\Vert^{2}_{L^{2}(\Omega)}+\Vert w\Vert^{2}_{L^{2}(\Omega)}\big) +\alpha_{i}\Vert \Psi+\psi_{e}\Vert^{2}_{\mathbb{V}}+\alpha_{e}\Vert \psi_{e}\Vert^{2}_{\mathbb{V}} +\int_{\Omega}\hbar_{11} \phi^{2}\Psi^{2} d{\bf{x}}\\ \quad \leq C_{1}\left(\int_{\Omega}(\mid \phi \mid + 1+ \mid u\mid )\Psi^{2} d{\bf{x}}+\int_{\Omega}w^{2} d{\bf{x}}+\int_{\Omega}(\mid \phi \mid + 1)\mid w\mid \mid \Psi\mid d{\bf{x}}\right)\\ \quad + \left\| { \mathcal{H}(\Psi_{\tau}, w_{\tau})} \right\| _{L^{2}(\Omega)} \left\| { \Psi} \right\| _{L^{2}(\Omega)}+\left\| { \mathcal{E}(\Psi_{\tau}, w_{\tau})} \right\| _{L^{2}(\Omega)} \left\| { w} \right\| _{L^{2}(\Omega)}\\ \quad +C_{0}\left(\left\| { h_{2}} \right\| _{L^{2}(\Omega)} \left\| { \Psi} \right\| _{L^{2}(\Omega)}+\left\| { h_{1}} \right\| _{L^{2}(\Omega)} \left\| { \psi_{e}} \right\| _{L^{2}(\Omega)}\right) \end{array} \end{equation} | (3.48) |
and then
\begin{equation*} \begin{array}{lcr} \frac{d}{2dt}\big(\Vert \mathfrak{b}_{m} \Psi\Vert^{2}_{L^{2}(\Omega)}+\Vert w\Vert^{2}_{L^{2}(\Omega)}\big) +\alpha_{i}\Vert \Psi+\psi_{e}\Vert^{2}_{\mathbb{V}}+\alpha_{e}\Vert \psi_{e}\Vert^{2}_{\mathbb{V}} +\int_{\Omega}\hbar_{11} \phi^{2}\Psi^{2} d{\bf{x}}\\ \quad \leq C_{2}\left\| { \Psi} \right\| _{L^{2}(\Omega)}\big(\left\| { \phi} \right\| _{L^{4}(\Omega)}\left\| { \Psi} \right\| _{L^{4}(\Omega)}+\left\| { u} \right\| _{L^{3}(\Omega)}\left\| { \Psi} \right\| _{L^{6}(\Omega)}\big)\\ \quad +C_{3}\left\| { \phi } \right\| _{L^{4}(\Omega)}\left\| { \Psi} \right\| _{L^{4}(\Omega)} \left\| { w} \right\| _{L^{2}(\Omega)}\\ \quad + C_{4}(\left\| { \mathcal{H}(\Psi_{\tau}, w_{\tau})} \right\| ^{2}_{L^{2}(\Omega)} +\left\| { \mathcal{E}(\Psi_{\tau}, w_{\tau})} \right\| ^{2}_{L^{2}(\Omega)})\\ \quad +C_{5}\big(\left\| { h_{2}} \right\| ^{2}_{L^{2}(\Omega)} +\left\| { h_{1}} \right\| ^{2}_{L^{2}(\Omega)}\big)+C_{6}\big(\left\| { \Psi} \right\| ^{2}_{L^{2}(\Omega)} +\left\| { w} \right\| ^{2}_{L^{2}(\Omega)}\big). \end{array} \end{equation*} |
Consequently (since \Vert \Psi\Vert_{\mathbb{V}}\leq \Vert \Psi+\psi_{e}\Vert_{\mathbb{V}}+\Vert \psi_{e}\Vert_{\mathbb{V}} )
\begin{equation} \begin{array}{lcr} \frac{d}{2dt}\big(\Vert \mathfrak{b}_{m} \Psi\Vert^{2}_{L^{2}(\Omega)}+\Vert w\Vert^{2}_{L^{2}(\Omega)}\big) +\min(\alpha_{i}, \alpha_{e})\big(\Vert \Psi+\psi_{e}\Vert^{2}_{\mathbb{V}}+\Vert \psi_{e}\Vert^{2}_{\mathbb{V}}\big)\\ \quad +\int_{\Omega}\hbar_{11} \phi^{2}\Psi^{2} d{\bf{x}} \leq \frac{1}{2}\min(\alpha_{i}, \alpha_{e})\big(\Vert \Psi+\psi_{e}\Vert^{2}_{\mathbb{V}}+\Vert \psi_{e}\Vert^{2}_{\mathbb{V}}\big)\\ \quad +C_{6}\big(\left\| { \mathcal{H}(\Psi_{\tau}, w_{\tau})} \right\| ^{2}_{L^{2}(\Omega)} +\left\| { \mathcal{E}(\Psi_{\tau}, w_{\tau})} \right\| ^{2}_{L^{2}(\Omega)})\\ \quad +C_{7}\big(\left\| { h_{2}} \right\| ^{2}_{L^{2}(\Omega)} +\left\| { h_{1}} \right\| ^{2}_{L^{2}(\Omega)})\\ \quad +C_{8}\big(\left\| { \Psi} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { w} \right\| ^{2}_{L^{2}(\Omega)}\big)\big(1+\left\| { \phi } \right\| ^{2}_{L^{4}(\Omega)}+\left\| { u} \right\| ^{2}_{L^{3}(\Omega)}\big). \end{array} \end{equation} | (3.49) |
By integrating in time between 0 and t (since \Psi(0) = w(0) = 0 ), we can deduce that (according to the boundedness of \mathfrak{b}_{m} )
\begin{equation} \begin{array}{lcr} \Vert \Psi(., t)\Vert^{2}_{L^{2}(\Omega)}+\Vert w(., t)\Vert^{2}_{L^{2}(\Omega)} +\int_{0}^{t}\big(\Vert \Psi+\psi_{e}\Vert^{2}_{\mathbb{V}}+\Vert \psi_{e}\Vert^{2}_{\mathbb{V}}\big)ds +\int_{0}^{T}\int_{\Omega}\phi^{2}\Psi^{2} d{\bf{x}} ds\\ \quad \leq C_{9}\left(\int_{0}^{t}\left\| { \mathcal{H}(\Psi_{\tau}, w_{\tau})} \right\| ^{2}_{L^{2}(\Omega)}ds +\int_{0}^{t}\left\| { \mathcal{E}(\Psi_{\tau}, w_{\tau})} \right\| ^{2}_{L^{2}(\Omega)}ds \right)\\ \quad +C_{10}\left(\int_{0}^{T}\left\| { h_{2}} \right\| ^{2}_{L^{2}(\Omega)}ds +\int_{0}^{T}\left\| { h_{1}} \right\| ^{2}_{L^{2}(\Omega)}ds\right)\\ \quad +C_{11}\int_{0}^{t}\big(\left\| { \Psi} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { w} \right\| ^{2}_{L^{2}(\Omega)}\big)\big(1+\left\| { \phi } \right\| ^{2}_{L^{4}(\Omega)}+\left\| { u} \right\| ^{2}_{L^{3}(\Omega)}\big)ds. \end{array} \end{equation} | (3.50) |
According to Lemma 2.6, we can deduce that
\begin{equation} \begin{array}{lcr} \int_{0}^{t}\left(\left\| { \mathcal{H}(\Psi_{\tau}, w_{\tau})} \right\| ^{2}_{L^{2}(\Omega)} +\left\| { \mathcal{E}(\Psi_{\tau}, w_{\tau})} \right\| ^{2}_{L^{2}(\Omega)}\right)ds\leq C\int_{0}^{t}\left(\left\| { \Psi} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { w} \right\| ^{2}_{L^{2}(\Omega)}\right)ds. \end{array} \end{equation} | (3.51) |
From (3.51), (3.50) becomes (since \Vert \Psi\Vert_{\mathbb{V}}\leq \Vert \Psi+\psi_{e}\Vert_{\mathbb{V}}+\Vert \psi_{e}\Vert_{\mathbb{V}} )
\begin{equation} \begin{array}{lcr} (\Vert \Psi(., t)\Vert^{2}_{L^{2}(\Omega)}+\Vert w(., t)\Vert^{2}_{L^{2}(\Omega)}) +\int_{0}^{t}(\Vert \Psi\Vert^{2}_{\mathbb{V}}+\Vert \psi_{e}\Vert^{2}_{\mathbb{V}})ds +\int_{0}^{T}\int_{\Omega}\phi^{2}\Psi^{2} d{\bf{x}} ds\\ \quad \leq C_{12}(\int_{0}^{T}\left\| { h_{2}} \right\| ^{2}_{L^{2}(\Omega)}ds +\int_{0}^{T}\left\| { h_{1}} \right\| ^{2}_{L^{2}(\Omega)}ds)\\ \quad +C_{13}\int_{0}^{t}(\left\| { \Psi} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { w} \right\| ^{2}_{L^{2}(\Omega)})(1+\left\| { \phi } \right\| ^{2}_{L^{4}(\Omega)}+\left\| { u} \right\| ^{2}_{L^{3}(\Omega)})ds \end{array} \end{equation} | (3.52) |
and then
\begin{equation} \begin{array}{lcr} (\Vert \Psi(., t)\Vert^{2}_{L^{2}(\Omega)}+\Vert w(., t)\Vert^{2}_{L^{2}(\Omega)}) \leq C_{12}(\int_{0}^{T}\left\| { h_{2}} \right\| ^{2}_{L^{2}(\Omega)}ds +\int_{0}^{T}\left\| { h_{1}} \right\| ^{2}_{L^{2}(\Omega)}ds)\\ \quad +C_{13}\int_{0}^{t}(\left\| { \Psi} \right\| ^{2}_{L^{2}(\Omega)}+\left\| { w} \right\| ^{2}_{L^{2}(\Omega)})(1+\left\| { \phi } \right\| ^{2}_{L^{4}(\Omega)}+\left\| { u} \right\| ^{2}_{L^{3}(\Omega)})ds. \end{array} \end{equation} | (3.53) |
By first using Gronwall lemma in (3.53), we can deduce that
\begin{equation*} \begin{array}{lcr} \Vert \Psi(., t)\Vert^{2}_{L^{2}(\Omega)}+\Vert w(., t)\Vert^{2}_{L^{2}(\Omega)}\\ \leq C(\int_{0}^{T}\left\| { h_{2}} \right\| ^{2}_{L^{2}(\Omega)}ds +\int_{0}^{T}\left\| { h_{1}} \right\| ^{2}_{L^{2}(\Omega)}ds) \exp(\left\| { \phi } \right\| ^{2}_{L^{2}(0, T;L^{4}(\Omega))}+\left\| { u} \right\| ^{2}_{L^{2}(0, T;L^{3}(\Omega))}), \end{array} \end{equation*} |
and then from (3.52), we easily deduce the following estimates
\begin{array}{l} \left\| { \Psi} \right\|^{2}_{L^{\infty}(0, T;L^{2}(\Omega))}+\left\| { w} \right\|^{2}_{L^{\infty}(0, T;L^{2}(\Omega))} \leq C\big(\left\| { h_{2}} \right\| ^{2}_{L^{2}({\cal Q})} +\left\| { h_{1}} \right\| ^{2}_{L^{2}({\cal Q})}\big), \\ \left\| { \Psi} \right\|^{2}_{L^{2}(0, T;\mathbb{V})}+\left\| { \psi_{e}} \right\|^{2}_{L^{2}(0, T;\mathbb{V})}\leq C\big(\left\| { h_{2}} \right\| ^{2}_{L^{2}({\cal Q})} +\left\| { h_{1}} \right\| ^{2}_{L^{2}({\cal Q})}\big), \\ \left\| {\phi \Psi } \right\|_{{L^2}({\cal Q})}^2 \le C(\left\| {{h_2}} \right\|_{{L^2}({\cal Q})}^2 + \left\| {{h_1}} \right\|_{{L^2}({\cal Q})}^2). \end{array} | (3.54) |
Derive now the estimate of results, for the solution (\Psi, \psi_{e}, w) , in the following space: (L^{\infty}(0, T; H^{1}(\Omega))\cap H^{1}(0, T; L^{2}(\Omega)))\times L^{\infty}(0, T; H^{1}(\Omega))\times (L^{\infty}(0, T; H^{1}(\Omega))\cap H^{1}(0, T; L^{3}(\Omega))) .
Since w satisfies: \frac{\partial w }{\partial t}\! = \! -\frac{\partial \mathcal{I}_{2}}{\partial \phi}(.; \phi)\Psi-\hbar w +\mathcal{E}(\Psi_{\tau}, w_{\tau}) , then for all t we have (since w(t = 0) = 0 )
\begin{equation} \begin{array}{lcr} w({\bf{x}}, t) \! = \! -\int_{0}^{t} \frac{\partial \mathcal{I}_{2}}{\partial \phi}({\bf{x}}, s;\phi)\Psi ds-\int_{0}^{t}\hbar w({\bf{x}}, s)ds+\int_{0}^{t}\mathcal{E}(\Psi_{\tau}, w_{\tau})ds. \end{array} \end{equation} | (3.55) |
Since relation (3.55) is similar to (3.40), then we can use similar argument as to derive (3.44) and we obtain (since \psi and \phi are in L^{2}(0, T, H^{1})\subset L^{2}(0, T, L^{r}) , for r\in [1, 6] )
\begin{equation} \begin{array}{lcr} \left\| { w(., t)} \right\| _{L^{3}(\Omega)} \leq C_{1}\int_{0}^{t} \left\| { w(., s)} \right\| _{L^{3}(\Omega)} ds\\ \quad +C_{2}\left(\left\| { \Psi} \right\| _{L^{2}(0, T;L^{3}(\Omega))}+\left\| { \phi} \right\| _{L^{2}(0, T;H^{1}(\Omega))}\left\| { \Psi} \right\| _{L^{2}(0, T;H^{1}(\Omega))} \right). \end{array} \end{equation} | (3.56) |
According to (3.54), we can deduce that
\begin{equation*} \begin{array}{lcr} \left\| { w(., t)} \right\| _{L^{3}(\Omega)}\leq C_{3}\int_{0}^{t} \left\| { w(., s)} \right\| _{L^{3}(\Omega)} ds+ C_{4}\big(\Vert h_{1}\Vert_{L^2({\cal Q})}+ \Vert h_{2}\Vert_{L^2({\cal Q})}\big). \end{array} \end{equation*} |
Consequently (by using Gronwall lemma)
\begin{equation} \begin{array}{lcr} \left\| { w(., t)} \right\| _{L^{3}(\Omega)}\leq C\left(\Vert h_{1}\Vert_{L^2({\cal Q})}+ \Vert h_{2}\Vert_{L^2({\cal Q})}\right). \end{array} \end{equation} | (3.57) |
Finally, from (3.10) with (v, v_{e}) = (\frac{\partial \Psi}{\partial t}, \frac{\partial \psi_{e}}{\partial t}) (put \tilde{h}_{i} = {\cal B}_{i} h_{i} )
\begin{equation} \begin{array}{lcr} \left\| { \mathfrak{b}_{m}\frac{\partial \Psi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)} +A_i(\Psi+\psi_e, \frac{\partial \Psi}{\partial t}) +\int_{\Omega}\frac{\partial {\cal I}}{\partial \phi}\Psi \frac{\partial \Psi}{\partial t} d{\bf{x}}\\ \quad +\int_{\Omega}\frac{\partial {\cal I}}{\partial u}w \frac{\partial \Psi}{\partial t} d{\bf{x}} = \int_\Omega \mathcal{H}(., \phi_\tau, u_\tau) \frac{\partial \phi}{\partial t}d{\bf{x}}+\int_{\Omega} \tilde{h}_{2}\frac{\partial \Psi}{\partial t}d{\bf{x}}, \\ A_i(\Psi + \psi_e, \frac{\partial \psi_{e}}{\partial t})+ A_e(\psi_e, \frac{\partial \psi_{e}}{\partial t}) = \int_{\Omega}\tilde{h}_{1} \frac{\partial \psi_{e}}{\partial t}d{\bf{x}}, \\ \frac{\partial w }{\partial t}\! = \! -\frac{\partial \mathcal{I}_{2}}{\partial \phi}(.;\phi)\Psi-\hbar w +\mathcal{E}(\Psi_{\tau}, w_{\tau}). \end{array} \end{equation} | (3.58) |
Then (since \frac{\partial \mathcal{I}_{2}}{\partial \phi}(.; \phi) is a polynomial of degree 1 on \phi )
\begin{equation} \begin{array}{lcr} \left\| { \mathfrak{b}_{m}\frac{\partial \Psi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)} + \frac{d}{2dt}A_i(\phi+\varphi_e, \phi+\varphi_e) +\frac{d}{2dt}A_i(\varphi_e, \varphi_e) = \\ \quad -\int_{\Omega}\frac{\partial {\cal I}}{\partial \phi}\Psi \frac{\partial \Psi}{\partial t} d{\bf{x}} -\int_{\Omega}\frac{\partial {\cal I}}{\partial u}w \frac{\partial \Psi}{\partial t} d{\bf{x}}\\ \quad +\int_{\Omega} \tilde{h}_{2}\frac{\partial \Psi}{\partial t}d{\bf{x}}+\int_{\Omega} \psi_{e}\frac{\partial \tilde{h}_{1}}{\partial t} d{\bf{x}}+\frac{d}{dt}(\int_{\Omega} \tilde{h}_{1} \psi_{e} d{\bf{x}})+\int_\Omega \mathcal{H}(., \Psi_\tau, w_\tau) \frac{\partial \Psi}{\partial t}d{\bf{x}}, \\ \left\| { \frac{\partial w }{\partial t}} \right\| _{L^{3}(\Omega)}\leq C\big(\left\| { \Psi } \right\| _{L^{\infty}(0, T;H^{1}(\Omega))}(1+\left\| { \phi } \right\| _{L^{\infty}(0, T;H^{1}(\Omega))})+ \left\| { w} \right\| _{L^{\infty}(0, T;L^{3}(\Omega))}\big)\\ \quad +\left\| { \mathcal{E}(\Psi_{\tau}, w_{\tau})} \right\| _{L^{3}(\Omega)}. \end{array} \end{equation} | (3.59) |
Since \tilde{h}_{1}\in U_{c}\subset C^{0}([0, T]; L^{2}(\Omega)) , then \tilde{h}_{1}(0)\in L^{2}(\Omega) and we have \left\| { \tilde{h}_{1}(0)} \right\| ^{2}_{L^{2}(\Omega)}\leq C \left\| { \tilde{h}_{1} } \right\| ^{2}_{U_{c}} . Consequently, by integrating (3.59) by time, we can deduce (from (2.21), the boundedness of \mathfrak{b}_{m} , the coercivity and continuity of A_{i} and A_{e} , and the estimate of \mathcal{H} and \mathcal{E} given by Lemma 2.6)
\begin{equation} \begin{array}{lcr} 2\underline{c}_{m}\int_{0}^{t}\left\| { \frac{\partial \Psi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+ \alpha_{i}\left\| { \Psi(t)} \right\| ^{2}_{H^{1}(\Omega)}+(\alpha_{e}+\alpha_{i})\left\| { \psi_e(t)} \right\| ^{2}_{H^{1}(\Omega)}\\ \leq C_{1}\left( \int_{0}^{t}\int_{\Omega}\mid \phi \mid ^{4}\mid \Psi \mid^{2} d{\bf{x}} ds+\int_{0}^{t}\int_{\Omega}\mid \phi \mid ^{2}\mid \Psi \mid^{2} d{\bf{x}} ds+\int_{0}^{t}\int_{\Omega}\mid u \mid ^{2}\mid \Psi \mid^{2} d{\bf{x}} ds\right)\\ \quad +C_{2} \left(\int_{0}^{t}\int_{\Omega}\mid \Psi \mid^{2} d{\bf{x}} ds+\int_{0}^{t}\int_{\Omega}\mid w\mid^{2} d{\bf{x}} ds+\int_{0}^{t}\int_{\Omega}\mid \psi_{e}\mid^{2} d{\bf{x}} ds\right)\\ \quad +C_{3}\int_{0}^{t}\int_{\Omega}\mid \phi \mid ^{2}\mid w \mid^{2} d{\bf{x}} ds+\delta_{0}\int_{0}^{t}\left\| { \frac{\partial \Psi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+\delta_{1}\left\| { \psi_{e}(t)} \right\| ^{2}_{H^{1}(\Omega)}\\ \quad +C_{4}\left(\left\| { \tilde{h}_{2} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { \tilde{h}_{1}} \right\| ^{2}_{U_{c}}\right), \\ \left\| { \frac{\partial w }{\partial t}} \right\| _{L^{2}(0, T;L^{3}(\Omega))}\!\leq \! C_{5}\left(\left\| { \Psi } \right\| _{L^{\infty}(0, T;H^{1}(\Omega))}(1+\left\| { \phi } \right\| _{L^{\infty}(0, T;H^{1}(\Omega))})+ \left\| { w} \right\| _{L^{\infty}(0, T;L^{3}(\Omega))}\right). \end{array} \end{equation} | (3.60) |
Since \phi is in L^{\infty}(0, T, H^{1})\subset L^{\infty}(0, T, L^{r}) ( r\in [1, 6] ), then (by choosing \delta_{0} = \underline{c}_{m} , \delta_{1} = (\alpha_{e}+\alpha_{i})/2 )
\begin{equation} \begin{array}{lcr} \underline{c}_{m}\int_{0}^{t}\left\| { \frac{\partial \Psi}{\partial t}} \right\| ^{2}_{L^{2}(\Omega)}ds+ \alpha_{i}\left\| { \Psi(t)} \right\| ^{2}_{H^{1}(\Omega)}+(\alpha_{e}+\alpha_{i})/2\left\| { \psi_e(t)} \right\| ^{2}_{H^{1}(\Omega)}\\ \quad \leq C_{6}\left(\left\| { \phi } \right\| ^{4}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { \phi } \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))} +\left\| { u } \right\| ^{2}_{L^{\infty}(0, T;L^{3}(\Omega))}\right)\int_{0}^{t}\left\| { \Psi } \right\| ^{2}_{H^{1}(\Omega)}ds\\ \quad +C_{7} \left(\left\| { \phi } \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}\left\| { w } \right\| ^{2}_{L^{\infty}(0, T;L^{3}(\Omega))} +\left\| { \Psi } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { \psi_{e} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { w } \right\| ^{2}_{L^{2}({\cal Q})}\right)\\ \quad +C_{8}\left(\left\| { \tilde{h}_{2} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { \tilde{h}_{1}} \right\| ^{2}_{U_{c}}\right), \\ \left\| { \frac{\partial w }{\partial t}} \right\| _{L^{2}(0, T;L^{3}(\Omega))}\leq \! C_{5}\left(\left\| { \Psi } \right\| _{L^{\infty}(0, T;H^{1}(\Omega))}(1+\left\| { \phi } \right\| _{L^{\infty}(0, T;H^{1}(\Omega))})+ \left\| { w} \right\| _{L^{\infty}(0, T;L^{3}(\Omega))}\right). \end{array} \end{equation} | (3.61) |
From the regularity of (\phi, \varphi_{e}, u) , estimation of (\Psi, \psi_{e}) in L^{2}({\cal Q}) and estimation of w in L^{\infty}(0, T; L^{3}(\Omega)) (see the estimates (3.54)–(3.57)), we can deduce that (according to the assumption concerning the operators {\cal B}_{i} , i = 1, 2 )
\begin{equation} \begin{array}{lcr} \left\| { \frac{\partial \Psi}{\partial t}} \right\| ^{2}_{L^{2}({\cal Q})}+ \left\| { \Psi(t)} \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { \psi_e(t)} \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { \frac{\partial w }{\partial t}} \right\| _{L^{2}(0, T;L^{3}(\Omega))}\\ \quad \leq C\left(\left\| { {h}_{2} } \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { {h}_{1}} \right\| ^{2}_{U_{c}}\right). \end{array} \end{equation} | (3.62) |
This completes the proof.
We are now going to study the Fréchet differentiability of {\cal F} .
Theorem 3.2. Let assumptions (H1)-(H4), (HMC) and (RC) be fulfilled. Suppose that (\phi, \varphi_{e}, u) \in \mathbb{D} , then the following results hold (for all (\xi, \pi)\in {\cal U}_{ad}\times {\cal V}_{ad} ).
(i) Let (\xi, \pi+h_{2})\in {\cal U}_{ad}\times {\cal V}_{ad} , with h_{2}\in L^{\infty}({\cal Q}) such that \pi+h_{2}\in {\cal V}_{ad} and, {\cal F}(\xi, \pi) and {\cal F}(\xi, \pi+h_{2}) being the corresponding solutions of (3.1). Then
\begin{equation} \begin{array}{lcr} \left\| { {\cal F}(\xi, \pi+h_{2})-{\cal F}(\xi, \pi)- {\cal F}'_{\pi}(\xi, \pi)h_{2}} \right\| _{\mathbb{D}}\leq C\left\| { h_{2}} \right\| ^{2}_{L^{2}({\cal Q})}, \end{array} \end{equation} | (3.63) |
where {\cal F}'_{\pi}(\xi, \pi):L^{\infty}({\cal Q}) \to \ \mathbb{D} is a linear operator, and (\Psi, \psi_{e}, w) = {\cal F}'_{\pi}(\xi, \pi)h is the solution of problem (3.7) with h_{1} = 0 (we denote this problem by ({\bf{\cal P}}_{FP}) ). Moreover \forall (\xi_{i}, \pi_{i})\in {\cal U}_{ad}\times {\cal V}_{ad} ( i = 1, 2 ), we have the following estimate
\begin{equation} \begin{array}{lcr} \left\| { {\cal F}'_{\pi}(\xi_{1}, \pi_{1})h_{2}-{\cal F}'_{\pi}(\xi_{2}, \pi_{2})h_{2}} \right\| _{\mathbb{D}} \leq C_{e}\left\| { h_{2}} \right\| _{L^{2}({\cal Q})}\left\| { X_{h}} \right\| _{U_{c}\times L^{2}({\cal Q})}, \end{array} \end{equation} | (3.64) |
where X_{h} = (\xi, \pi) = (\xi_{1}-\xi_{2}, \pi_{1}-\pi_{2}) .
(ii) Let (\xi+h_{1}, \pi)\in {\cal U}_{ad}\times {\cal V}_{ad} , with h_{1}\in L^{\infty}({\cal Q})\cap U_{c} such that \xi+h_{1}\in {\cal U}_{ad} and, {\cal F}(\xi, \pi) and {\cal F}(\xi+h_{1}, \pi) being the corresponding solutions of (3.1). Then
\begin{equation} \begin{array}{lcr} \left\| { {\cal F}(\xi+h_{1}, \pi)-{\cal F}(\xi, \pi)- {\cal F}'_{\xi}(\xi, \pi)h_{1}} \right\| _{\mathbb{D}}\leq C\left\| { h_{1}} \right\| ^{2}_{U_{c}}, \end{array} \end{equation} | (3.65) |
where {\cal F}'_{\xi}(\xi, \pi):L^{\infty}({\cal Q})\cap U_{c} \to \ \mathbb{D} is a linear operator, and (\Psi, \psi_{e}, w) = {\cal F}'_{\xi}(\xi, \pi)h is the solution of the problem (3.7) with h_{2} = 0 (we denote this problem by ({\bf{\cal P}}_{FX}) ). Moreover \forall (\xi_{i}, \pi_{i})\in {\cal U}_{ad}\times {\cal V}_{ad} ( i = 1, 2 ), we have the following estimate
\begin{equation} \begin{array}{lcr} \left\| { {\cal F}'_{\xi}(\xi_{1}, \pi_{1})h_{1}-{\cal F}'_{\xi}(\xi_{2}, \pi_{2})h_{1}} \right\| _{\mathbb{D}} \leq C_{e}\left\| { h_{1}} \right\| _{U_{c}}\left\| { X_{h}} \right\| _{U_{c}\times L^{2}({\cal Q})}, \end{array} \end{equation} | (3.66) |
where X_{h} = (\xi, \pi) = (\xi_{1}-\xi_{2}, \pi_{1}-\pi_{2}) .
Proof. According to Theorem 3.1 the problems ({\cal P}_{FP}) , ({\cal P}_{FX}) have a unique solution in \mathbb{D} .
(ⅰ) Let Y = (\phi, \varphi_{e}, u) = {\cal F}(\xi, \pi) and Y_{h} = (\phi_{h}, \varphi_{e, h}, u_{h}) = {\cal F}(\xi, \pi+h_{2}) . From the stability estimate in Theorem 2.4, we know that
\begin{equation} \left\| { Y-Y_{h}} \right\| ^{2}_{\mathbb{D}}\leq C\left\| { h_{2}} \right\| ^{2}_{L^{2}({\cal Q})}. \end{equation} | (3.67) |
Denote by \Phi_{h} = \phi_{h}-\phi , \Pi_{e, h} = \varphi_{e, h}-\varphi_{e} , U_{h} = u_{h}-u , \phi^{*} = \Phi_{h}-\Psi , \varphi^{*}_{e} = \Pi_{e, h}-\psi_{e} and u^{*} = U_{h}-w . It is easy to see that (\phi^{*}, \varphi^{*}_{e}, u^{*}) satisfies the linear problem
\begin{equation} \begin{array}{lcr} \mathfrak{c}_{m}\frac{\partial \phi^{*}}{\partial t}-div(\mathcal{K}_{i}\nabla (\phi^{*}+\varphi^{*}_{e})) +\frac{\partial \mathcal{I}}{\partial u}u^{*}+\frac{\partial \mathcal{I}}{\partial \phi}\phi^{*} = S_{1}+\mathcal{H}(\phi^{*}_{\tau}, u^{*}_{\tau}), \; \; \text{in}\; \; {\cal Q}\\ -div(\mathcal{K}_{i}\nabla \phi^{*})-div((\mathcal{K}_{i}+\mathcal{K}_{e})\varphi^{*}_{e}) = 0, \; \; \text{in}\; \; {\cal Q}\\ \frac{\partial u^{*}}{\partial t}+\hbar u^{*}+\frac{\partial \mathcal{G}}{\partial \phi}\phi^{*} = S_{2}+\mathcal{E}(\phi^{*}_{\tau}, u^{*}_{\tau}), \; \; \text{in}\; \; {\cal Q}\\ \big(\mathcal{K}_{i}\nabla \phi^{*}\big).{\bf{n}} + \big(\mathcal{K}_{i}\nabla \varphi^{*}_{e}\big).{\bf{n}} = 0, \; \; \text{on}\; \; \Sigma\\ \big(\mathcal{K}_{e}\nabla \varphi^{*}_{e}\big).{\bf{n}} = 0, \; \; \text{on}\; \; \Sigma\\ \phi^{*} (., t = 0) = 0, \; \; \; u^{*}(., t = 0) = 0, \; \; \text{in}\; \; \Omega\\ \phi^{*} = 0, \; \; \; u^{*} = 0, \; \; \text{in}\; \; {\cal Q}_{0} \end{array} \end{equation} | (3.68) |
where the condition (1.3) for \varphi^{*}_{e} holds and
\begin{equation} \begin{array}{lcr} S_{1} = -(g_{0}+ug_{1})-U_{h}\big(\mathcal{I}_{1}(\phi_{h})-\mathcal{I}_{1}(\phi)\big), \; \; \; \; S_{2} = -g_{2}, \end{array} \end{equation} | (3.69) |
with g_{i} = \big(\mathcal{I}_{i}(\phi_{h})-\mathcal{I}_{i}(\phi)-\mathcal{I}_{i}'(\phi).\Phi_{h}\big) and \mathcal{I}_{i}'(\phi) = \frac{\partial \mathcal{I}_{i}}{\partial \phi} , for i = 0, 2. Now we have to derive some estimates necessary to prove the result of theorem. By using a simple manipulation we obtain that
g_{i} = \int_{0}^{1}\big(\mathcal{I}_{i}'(\phi)-\mathcal{I}_{i}'(\phi+s\Phi_{h})\big)\Phi_{h} ds\; \; \text{(for $i = 0, 1, 2$)}. |
Since \mathcal{I}_{0}'(\phi) = \hbar_{11}\phi^{2}-\hbar_{12}\phi+\hbar_{13} , \mathcal{I}_{1}'(\phi) = \epsilon_{1}\hbar_{21} , \mathcal{I}_{2}'(\phi) = -\epsilon_{2}\hbar_{31}\phi+(\epsilon_{2}-1)\hbar_{32} and \mathcal{I}_{1}(\phi_{h})-\mathcal{I}_{1}(\phi) = \epsilon_{1}\hbar_{21}\Phi_{h} then
\begin{equation} \begin{array}{lcr} g_{0} = \int_{0}^{1}-s^{2}\Phi_{h}^{2} (\hbar_{11}(2\phi+s\Phi_{h})-\hbar_{12})ds = \frac{-2\hbar_{11}}{3}\Phi_{h}^{2}\phi-\frac{\hbar_{11}}{4}\Phi_{h}^{3}+\frac{\hbar_{12}}{3}\Phi_{h}^{2}, \\ g_{1} = 0, \\ g_{2} = \int_{0}^{1}s\Phi_{h}^{2} \epsilon_{2}\hbar_{32} ds = \frac{\epsilon_{2}\hbar_{32}}{2}\Phi_{h}^{2}. \end{array} \end{equation} | (3.70) |
According to the regularity of \hbar_{ij} we can deduce that
\begin{equation} \begin{array}{lcr} \mid S_{1}\mid^{2} \leq C\left(\mid \Phi_{h}\mid^{6}+\mid \phi \mid^{2}\mid \Phi_{h}\mid^{4}+\mid \Phi_{h}\mid^{4}+ \mid U_{h}\mid^{2} \mid \Phi_{h}\mid^{2}\right), \\ \mid S_{2}\mid^{3} \leq C\mid \Phi_{h}\mid^{6} \end{array} \end{equation} | (3.71) |
Integrating by space and using Young's formula, we can deduce
\begin{equation*} \begin{array}{lcr} \int_{\Omega}\mid S_{1}\mid^{2} d{\bf{x}} \leq C\left(\left\| { \Phi_{h}} \right\| ^{6}_{L^{6}(\Omega)}\!+\!\left\| { \Phi_{h}} \right\| ^{4}_{L^{6}(\Omega)}\left\| { \phi} \right\| ^{2}_{L^{6}(\Omega)} \!+\!\left\| { \Phi_{h}} \right\| ^{4}_{L^{4}(\Omega)}\!+\!\left\| { \Phi_{h}} \right\| ^{2}_{L^{6}(\Omega)}\left\| { U_{h}} \right\| ^{2}_{L^{3}(\Omega)}\right), \\ \int_{\Omega}\mid S_{2}\mid^{3} d{\bf{x}}\leq C\left\| { \Phi_{h}} \right\| ^{6}_{L^{6}(\Omega)}. \end{array} \end{equation*} |
From Gagliardo-Nirenberg inequality, we have, for all v\in H^{1}(\Omega) , \left\| { v} \right\| _{L^{r}(\Omega)}\leq C\left\| { v} \right\| _{H^{1}(\Omega)} , for r\in [2, 6] , we have (since \phi, U_{h}, \Phi_{h}\in L^{\infty}(0, T; H^{1}(\Omega)) )
\begin{equation} \begin{array}{lcr} \int_{\Omega}\mid S_{1}\mid^{2}d{\bf{x}} \leq C\left(\left\| { \Phi_{h}} \right\| ^{4}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { U_{h}} \right\| ^{4}_{L^{\infty}(0, T;L^{3}(\Omega))}\right), \\ \int_{\Omega}\mid S_{2}\mid^{3} d{\bf{x}}\leq C\left\| { \Phi_{h}} \right\| ^{6}_{L^{\infty}(0, T;H^{1}(\Omega))}. \end{array} \end{equation} | (3.72) |
We can deduce that (according to the estimate (3.67))
\begin{equation} \begin{array}{lcr} \left\| { S_{1}} \right\| _{L^{\infty}(0, T; L^{2}(\Omega))}\leq C\left\| { h_{2}} \right\| ^{2}_{L^{2}({\cal Q})}, \\ \left\| { S_{2}} \right\| _{L^{\infty}(0, T; L^{3}(\Omega))}\leq C\left\| { h_{2}} \right\| ^{2}_{L^{2}({\cal Q})}. \end{array} \end{equation} | (3.73) |
We can conclude that source term (S_{1}, S_{2}) in (3.68) is in L^{2}(0, T; L^{2}(\Omega))\times L^{2}(0, T; L^{3}(\Omega)) and satisfies estimates (3.73). Since (3.68) is similar as system (3.7) with source term (S_{1}, S_{2}) then from Theorem 3.1 we can deduce that
\begin{equation*} \begin{array}{lcr} \left\| { (\phi^{*}, \varphi_{e}^{*}, u^{*})} \right\| ^{2}_{\mathbb{D}} \leq C\left(\left\| { S_{1}} \right\| ^{2}_{L^{2}({\cal Q})}+\left\| { S_{2}} \right\| ^{2}_{L^{2}(0, T; L^{3}(\Omega))}\right) \end{array} \end{equation*} |
and then according to estimates (3.73), we can deduce estimate (3.63).
Prove now the second part of (ⅰ). Let (\xi_{i}, \pi_{i})\in {\cal U}_{ad}\times {\cal V}_{ad} , i = 1, 2 be given and {\bf{w}} _{i} = (\Psi_{i}, \psi_{e, i}, w_{i}) = {\cal F}_{\pi}'(\xi_{i}, \pi_{i}).h_{2} solution of ({\cal P}_{FP}) (we denote by Y_{i} = (\phi_{i}, \varphi_{e, i}, u_{i}) = {\cal F}(\xi_{i}, \pi_{i}) and by (\phi, \varphi_{e}, u) = Y_{1}-Y_{2} ). Set {\bf{w}} = (\Psi, \psi_{e}, w) = {\bf{w}} _{1}-{\bf{w}} _{2} , (\xi, \pi) = (\xi_{1}-\xi_{2}, \pi_{1}-\pi_{2}) . According to equations satisfied by {\bf{w}} _{1} and {\bf{w}} _{2} we have
\begin{equation} \begin{array}{lcr} \mathfrak{c}_{m}\frac{\partial \Psi}{\partial t}-div(\mathcal{K}_{i}\nabla (\Psi+\psi_{e})) +\frac{\partial {\cal I}}{\partial \phi}(\phi_{1}, u_{1})\Psi +\frac{\partial {\cal I}}{\partial u}(\phi_{1}, u_{1})w = S_{1}+\mathcal{H}(\Psi_{\tau}, w_{\tau}), \\ -div(K_{i}\nabla \Psi)-div((\mathcal{K}_{i}+\mathcal{K}_{e})\psi_{e}) = 0, \\ \frac{\partial w}{\partial t} +\frac{\partial {\cal G}}{\partial \phi}(\phi_{1}, u_{1})\Psi +\hbar w = S_{2}+\mathcal{E}(\Psi_{\tau}, w_{\tau}), \\ \big(\mathcal{K}_{i}\nabla \Psi\big).{\bf{n}} + \big(\mathcal{K}_{i}\nabla \psi_{e}\big).{\bf{n}} = 0, \; \; \text{on}\; \; \Sigma\\ \big(\mathcal{K}_{e}\nabla \psi_{e}\big).{\bf{n}} = 0, \; \; \text{on}\; \; \Sigma\\ \Psi(., t = 0) = 0, \; \; \; w(., t = 0) = 0, \; \; \text{in}\; \; \Omega\\ \Psi = 0, \; \; \; w = 0, \; \; \text{in}\; \; {\cal Q}_{0} \end{array} \end{equation} | (3.74) |
where the condition (1.3) for \psi_{e} holds and
\begin{equation} \begin{array}{lcr} S_{1} = (\frac{\partial {\cal I}}{\partial \phi}(\phi_{1}, u_{1})-\frac{\partial {\cal I}}{\partial \phi}(\phi_{2}, u_{2}))\Psi_{2} \quad +(\frac{\partial {\cal I}}{\partial u}(\phi_{1}, u_{1})-\frac{\partial {\cal I}}{\partial u}(\phi_{2}, u_{2}))w_{2}, \\ S_{2} = (\frac{\partial {\cal G}}{\partial \phi}(\phi_{1}, u_{1})-\frac{\partial {\cal G}}{\partial \phi}(\phi_{2}, u_{2}))\Psi_{2}. \end{array} \end{equation} | (3.75) |
From the expression (2.21) of partial derivatives of \mathcal{I} and \mathcal{G} and Hölder's inequality, we can have (according to regularity of \hbar_{ij} )
\begin{equation} \begin{array}{lcr} \mid S_{1}\mid \leq C(\mid \phi\mid (1+\mid \phi_{1}\mid +\mid \phi_{2}\mid) +\mid u\mid)) \mid \Psi_{2}\mid+\mid \phi\mid \mid w_{2}\mid, \\ \mid S_{2}\mid \leq C\mid \phi\mid \mid \Psi_{2}\mid. \end{array} \end{equation} | (3.76) |
So
\begin{equation} \begin{array}{lcr} \int_{\Omega}\mid S_{1}\mid^{2} d{\bf{x}}\leq C_{1}(1+\left\| { \phi_{1}} \right\| ^{2}_{L^{6}(\Omega)}+\left\| { \phi_{2}} \right\| ^{2}_{L^{6}(\Omega)})\left\| { \phi} \right\| ^{2}_{L^{6}(\Omega)}\left\| { \Psi_{2}} \right\| ^{2}_{L^{6}(\Omega)}\\ \quad +C_{2}\left\| { u} \right\| ^{2}_{L^{3}(\Omega)}\left\| { \Psi_{2}} \right\| ^{2}_{L^{6}(\Omega)}+C_{3}\left\| { \phi} \right\| ^{2}_{L^{6}(\Omega)}\left\| { w_{2}} \right\| ^{2}_{L^{3}(\Omega)}, \\ \int_{\Omega}\mid S_{2}\mid^{3} d{\bf{x}} \leq C_{4}\left\| { \phi} \right\| ^{3}_{L^{6}(\Omega)}\left\| { \Psi_{2}} \right\| ^{3}_{L^{6}(\Omega)}. \end{array} \end{equation} | (3.77) |
According to the regularity of \phi_{i} ( i = 1, 2 ) in L^{\infty}(0, T; L^{6}(\Omega))\subset L^{\infty}(0, T; H^{1}(\Omega)) , we can deduce that
\begin{equation*} \begin{array}{lcr} \int_{\Omega}\mid S_{1}\mid^{2} d{\bf{x}}\leq C_{5}(\left\| { \phi} \right\| ^{2}_{L^{6}(\Omega)}+\left\| { u} \right\| ^{2}_{L^{3}(\Omega)})\left\| { \Psi_{2}} \right\| ^{2}_{L^{6}(\Omega)} +C_{6}\left\| { \phi} \right\| ^{2}_{L^{6}(\Omega)}\left\| { w_{2}} \right\| ^{2}_{L^{3}(\Omega)}, \\ \int_{\Omega}\mid S_{2}\mid^{3} d{\bf{x}} \leq C_{4}\left\| { \phi} \right\| ^{3}_{L^{6}(\Omega)}\left\| { \Psi_{2}} \right\| ^{3}_{L^{6}(\Omega)} \end{array} \end{equation*} |
and then (from Gagliardo-Nirenberg inequalities)
\begin{equation} \begin{array}{lcr} \int_{\Omega}\mid S_{1}\mid^{2} d{\bf{x}}\leq C_{5}(\left\| { \phi} \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}+\left\| { u} \right\| ^{2}_{L^{\infty}(0, T;L^{3}(\Omega))})\left\| { \Psi_{2}} \right\| ^{2}_{H^{1}(\Omega)}\\ \quad +C_{6}\left\| { \phi} \right\| ^{2}_{L^{\infty}(0, T;H^{1}(\Omega))}\left\| { w_{2}} \right\| ^{2}_{L^{3}(\Omega)}, \\ \int_{\Omega}\mid S_{2}\mid^{3} d{\bf{x}} \leq C_{4}\left\| { \phi} \right\| ^{3}_{L^{\infty}(0, T;H^{1}(\Omega))}\left\| { \Psi_{2}} \right\| ^{3}_{H^{1}(\Omega)}. \end{array} \end{equation} | (3.78) |
According to Theorems 2.4 and 3.1, we can deduce that
\begin{equation} \begin{array}{lcr} \left\| { S_{1}} \right\| _{L^{2}(0, T;L^{2}(\Omega))}\leq C\left\| { h_{2}} \right\| _{L^{2}({\cal Q})}\left\| { (\xi, \pi)} \right\| _{U_{c}\times L^{2}({\cal Q})}, \\ \left\| { S_{2}} \right\| _{L^{2}(0, T;L^{3}(\Omega))} \leq C\left\| { h_{2}} \right\| _{L^{2}({\cal Q})}\left\| { (\xi, \pi)} \right\| _{U_{c}\times L^{2}({\cal Q})}. \end{array} \end{equation} | (3.79) |
We can conclude that source term (S_{1}, S_{2}) , in (3.74), is in L^{2}(0, T; L^{2}(\Omega))\times L^{2}(0, T; L^{3}(\Omega)) and satisfies estimates (3.79). Since (3.74) is similar as system (3.7) with source term (S_{1}, S_{2}) then from Theorem 3.1 we can deduce that the result (3.64) of the theorem holds.
(ⅱ) By using the same technique as in the proof of results of (ⅰ), we have results of (ⅱ). Therefore, we omit the details.
Theorem 3.3. Assume that assumptions of Theorem 3.2 are satisfied. Then, for \alpha and \beta sufficiently large (i.e. there exist (\alpha_{l}, \beta_{l}) such that \alpha\geq \alpha_{l} and \beta\geq\beta_{l} ) there exist (\xi^{*}, \pi^{*})\in {\cal U}_{ad}\times {\cal V}_{ad} and (\phi^{*}, \varphi_{e}^{*}, u^{*})\in \mathbb{D} such that (\xi^{*}, \pi^{*}) is a saddle point of {\cal J} and (\phi^{*}, \varphi_{e}^{*}, u^{*}) = {\cal F}(\xi^{*}, \pi^{*}) is the solution of (3.1).
Proof. Let P_{\pi} be the map: \xi \to \ {\cal J}(\xi, \pi) and Q_{\xi} be the map: \pi \to \ {\cal J}(\xi, \pi) . To obtain the existence of minimax control problem we prove that P_{\pi} is convex and lower semicontinuous for all \pi \in {\cal V}_{ad} , Q_{\xi} is concave and upper semicontinuous for all \xi\in {\cal U}_{ad} , and we use the classical minimax theorem in infinite dimensions (see, e.g. [10,35]).
First we prove, for \alpha and \beta sufficiently large, the convexity of the map P_{\pi} and the concavity of the map Q_{\xi} . In order to prove the convexity, it is sufficient to show that for all (\xi_{1}, \xi_{2})\in {\cal U}_{ad} we have:
(P_{\pi}'(\xi_{1})-P_{\pi}'(\xi_{2})).\xi\geq 0, |
where \xi = \xi_{1}-\xi_{2} (because P_{\pi} is Fréchet differentiable).
According to definition of {\cal J} , we have that
\begin{equation} \begin{array}{lcr} (P_{\pi}'(\xi_{1})-P_{\pi}'(\xi_{2})).\xi = \alpha \left\| { \xi } \right\| ^{2}_{U_{c}}\\ \quad +m_{1}\int_{0}^{T}\!\!\!\int_{\Omega_{obs}}\!\Upsilon(t)(\phi_{1}-\phi_{2})\Psi_{2} d{\bf{x}}dt+m_{2}\int_{\Omega_{obs}}\!(\phi_{1}(T)-\phi_{2}(T))\Psi_{2}(T) d{\bf{x}}\\ \quad +m_{1}\int_{0}^{T}\!\!\!\int_{\Omega_{obs}}\!\Upsilon(t)(\phi_{1}-\phi_{obs})(\Psi_{1}-\Psi_{2}) d{\bf{x}}dt+m_{2}\int_{\Omega_{obs}}\!(\phi_{1}(T)-\psi_{obs})(\Psi_{1}(T)-\Psi_{2}(T)) d{\bf{x}}, \end{array} \end{equation} | (3.80) |
where (\phi_{i}, \varphi_{e, i}, u_{i}) = {\cal F}(\xi_{i}, \pi) and function (\Psi_{i}, \psi_{e, i}, w_{i}) = {\cal F}'(\xi_{i}, \pi).(\xi, 0) is the solution of problem (3.7), for i = 1, 2 . According to Theorems 2.4 and 3.1 we can deduce that
\begin{equation*} \begin{array}{lcr} \mid m_{1}\int_{0}^{T}\!\!\!\int_{\Omega_{obs}}\!\Upsilon(t)(\phi_{1}-\phi_{2})\Psi_{2} d{\bf{x}}dt+m_{2}\int_{\Omega_{obs}}(\phi_{1}(T)-\phi_{2}(T))\Psi_{2}(T) d{\bf{x}}\mid \\ \quad \leq C_{1}\Big(\left\| { \phi_{1}-\phi_{2}} \right\| _{L^{2}(0, T;\Omega_{obs})}\left\| { \psi_{2}} \right\| _{L^{2}(0, T;\Omega_{obs})}\\ \quad +\left\| { \phi_{1}(T)-\phi_{2}(T)} \right\| _{L^{2}(\Omega_{obs})}\left\| { \psi_{2}(T)} \right\| _{L^{2}(\Omega_{obs})}\Big)\\ \quad \leq M_{0}\left\| { \xi } \right\| ^{2}_{U_{c}} \end{array} \end{equation*} |
and
\begin{equation*} \begin{array}{lcr} \mid m_{1}\int_{0}^{T}\!\!\!\int_{\Omega_{obs}}\!\Upsilon(t)(\phi_{1}-\phi_{obs})(\Psi_{1}-\Psi_{2}) d{\bf{x}}dt+m_{2}\int_{\Omega_{obs}}\!(\phi_{1}(T)-\psi_{obs})(\Psi_{1}(T)-\Psi_{2}(T)) d{\bf{x}}\mid \\ \quad \leq C_{2}\Big(\left\| { \phi_{1}-\phi_{obs}} \right\| _{L^{2}(0, T;\Omega_{obs})}\left\| { \psi_{1}-\psi_{2}} \right\| _{L^{2}(0, T;\Omega_{obs})}\\ \quad +\left\| { \phi_{1}(T)-\psi_{obs}} \right\| _{L^{2}(\Omega_{obs})}\left\| { \psi_{1}(T)-\psi_{2}(T)} \right\| _{L^{2}(\Omega_{obs})}\Big)\\ \quad \leq M_{1}\left\| { \xi } \right\| ^{2}_{U_{c}}. \end{array} \end{equation*} |
From (3.80) and the previous relations we deduce that for \alpha\geq \alpha_{l} such that \alpha_{l}\geq M_{0}+M_{1} we have (P_{\pi}'(\xi_{1})-P_{\pi}'(\xi_{2})).\xi\geq 0 and then the convexity of P_{\pi} . In the same way, we can find \beta_{l} such that for \beta\geq \beta_{l} we have the concavity of Q_{\xi} .
We prove now that P_{\pi} is lower semicontinuous for all \pi \in {\cal V}_{ad} , and Q_{\xi} is upper semicontinuous for all \xi \in {\cal U}_{ad} .
Let \xi^{(k)} be a minimizing sequence of {\cal J} i.e. \mathop {\lim \inf }\limits_k{\cal J}(\xi^{(k)}, \pi) = \mathop {\min }\limits_{\xi \in {{\cal U}_{ad}}}{\cal J}(\xi, \pi) \; (\forall \pi\in {\cal V}_{ad} ). Then \xi^{(k)} is uniformly bounded in {\cal U}_{ad} . Set (\phi_{\pi}, \varphi_{e, \pi}, u_{\pi}) = {\cal F}(\xi, \pi) , and (\phi^{(k)}, \varphi^{(k)}_{e}, u^{(k)}) = {\cal F}(\xi^{(k)}, \pi) . In view of Theorem 2.2 and the nature of the operator {\cal B}_{1} , we can deduce that the sequence (\phi^{(k)}, \varphi^{(k)}_{e}, u^{(k)}) is uniformly bounded in \mathcal{W}({\cal Q}) with respect to k . Therefore, we can extract from (\xi^{(k)}, \phi^{(k)}, \varphi^{(k)}_{e}, u^{(k)}) a subsequence also denoted by (\xi^{(k)}, \phi^{(k)}, \varphi^{(k)}_{e}, u^{(k)}) and such that
\begin{equation*} \begin{array}{lcr} {\cal B}_{1}\xi^{(k)}\rightharpoonup {\cal B}_{1}\xi_{\pi} \; weakly \; in \; U_{c}, \\ (\phi^{(k)}, \varphi^{(k)}_{e}, u^{(k)}) \rightharpoonup (\underline{\phi}_{\pi}, \underline{\varphi}_{e, \pi}, \underline{u}_{\pi})\; weakly\; in \; \mathcal{W}({\cal Q}), \\ (\phi^{(k)}, \varphi^{(k)}_{e}, u^{(k)}) \to \ (\underline{\phi}_{\pi}, \underline{\varphi}_{e, \pi}, \underline{u}_{\pi})\; \; strongly \; in \; L^{2}({\cal Q}). \end{array} \end{equation*} |
Passing to limit in the corresponding system satisfied by (\xi^{(k)}, \phi^{(k)}, \varphi^{(k)}_{e}, u^{(k)}) , we can conclude that (\underline{\phi}_{\pi}, \underline{\varphi}_{e, \pi}, \underline{u}_{\pi}) = {\cal F}(\xi, \pi) and according to uniqueness of solution of (3.1), we have then (\underline{\phi}_{\pi}, \underline{\varphi}_{e, \pi}, \underline{u}_{\pi}) = ({\phi}_{\pi}, {\varphi}_{e, \pi}, {u}_{\pi}) . Therefore, using the sequential weak lower semicontinuous of {\cal J} with respect to convergence and uniform boundedness of sequence \xi^{(k)} (according to structure of {\cal J} ), we conclude that the map P_{\pi}:\xi \to \ {\cal J}(\xi, \pi) is lower semicontinuous for all \pi \in {\cal V}_{ad} . By using the same technique we obtain then Q_{\xi} is upper semicontinuous for all \xi \in {\cal K}_{1} .
We now turn to necessary optimality conditions which have be satisfied by each solution of the control problem. In order to simplify the presentation we assume that the functions (p_l)_{1\leq l\leq n_{2}} and (r_k)_{1\leq k\leq n_{1}} at final time T satisfy
\begin{equation} \begin{array}{lcr} p_1(T)\geq ... \geq p_{n_2}(T)\; \; \text{and}\; \; r_1(T)\geq ... \geq r_{n_1}(T), \\ \text{with the convention $r_{n_{1}+1}(T) = 0$ and $p_{n_{2}+1}(T) = 0$}. \end{array} \end{equation} | (3.81) |
Since the cost {\cal J} is a composition of F-differentiable maps then {\cal J} is F-differentiable and we have
\begin{equation} \begin{array}{lcr} {\cal J}'(\xi, \pi).{\bf{h}} = m_{1}\int_{0}^{T}\int_{\Omega_{obs}}\Upsilon(t)(\phi-\phi_{obs})\psi d{\bf{x}}dt+m_{2}\int_{\Omega_{obs}}(\phi(T)-\psi_{obs})\psi(T) d{\bf{x}}\\ \quad +\alpha (\xi, h_{1})_{U_{c}}-\beta \int_{0}^{T}\int_{\Omega_{d}}\pi h_{2} d{\bf{x}}dt \; \text{($\forall {\bf{h}} = (h_{1}, h_{2})\in {\cal U}_{ad}\times {\cal V}_{ad}$)}\; , \end{array} \end{equation} | (3.82) |
where (\Psi, \psi_{e}, w) = {\cal F}'(\xi, \pi).{\bf{h}} is solution of problem (3.7).
Theorem 3.4. Assume that assumptions of Theorem 3.3 are satisfied and \alpha and \beta are sufficiently large. Let (\xi^{*}, \pi^{*})\in {\cal U}_{ad}\times {\cal V}_{ad} be an optimal solution of (3.6) and (\phi^{*}, \varphi_{e}^{*}, u^{*}) = {\cal F}(\xi^{*}, \pi^{*}) \in \mathbb{D} be its corresponding solution. Then ( \forall (\xi, \pi)\in {\cal U}_{ad}\times {\cal V}_{ad} )
\begin{equation} \begin{array}{lcr} \frac{\partial {\cal J}}{\partial \xi}(\xi^{*} , \pi^{*}).(\xi-\xi^{*}) = \left(\alpha\xi^{*}+ \chi_{\Omega_{c}}\aleph_{1}^{*}\tilde{\varphi}^{*}_{e}, \xi-\xi^{*} \right)_{U_{c}}\geq 0, \\ \frac{\partial {\cal J}}{\partial \phi}(\xi^{*} , \pi^{*}).(\pi -\pi^{*}) = \int_{0}^{T}\!\!\int_{\Omega_{d}}(-\beta\pi^{*}+ {\cal B}_{2}^{*}\tilde{\phi}^{*}) (\pi-\pi^{*})d{\bf{x}}dt\leq 0, \end{array} \end{equation} | (3.83) |
where (\tilde{\phi}^{*}, \tilde{\varphi}^{*}_{e}, \tilde{u}^{*}): = {\cal F}^{\bot}(\xi^{*}, \pi^{*}) is the solution of the so-called adjoint problem (3.86) where the condition (1.3) for \tilde{\varphi}_{e} holds (given below), corresponding to (\phi^{*}, \varphi^{*}_{e}, u^{*}) , and with \aleph_{1}^{*} = \Lambda^{-1}{\cal B}_{1}^{*} where \Lambda is the canonical isomorphism : U_{c} \to \ U_{c}' such that
\begin{equation} \langle g, v\rangle _{U_{c}, U_{c}'} = (\Lambda g, v)_{U_{c}'} = (g, \Lambda^{-1}v)_{U_{c}}, \; \; \forall (g, v)\in U_{c}\times U_{c}'. \end{equation} | (3.84) |
Moreover the gradients of {\cal J} at any point (\xi, \pi) , in the weak sense, are given by
\begin{equation} \begin{array}{lcr} \frac{\partial {\cal J}}{\partial \xi}(\xi , \pi) = \alpha\xi+ \chi_{\Omega_{c}}\aleph_{1}^{*}\tilde{\varphi}_{e}, \; \; \; \; \frac{\partial {\cal J}}{\partial \phi}(\xi , \pi) = -\beta\pi+ \chi_{\Omega_{d}}{\cal B}_{2}^{*}\tilde{\phi}, \end{array} \end{equation} | (3.85) |
where (\tilde{\phi}, \tilde{\varphi}_{e}, \tilde{u}) = {\cal F}^{\bot}(\xi, \pi) is the solution of adjoint problem (corresponding to (\phi, \varphi_{e}, u) = {\cal F}(\xi, \pi) )
\begin{equation} \begin{array}{lcr} -\mathfrak{c}_{m}\frac{\partial \tilde{\phi} }{\partial t}+ \frac{\partial \mathcal{I}}{\partial \phi}(\phi, u)\tilde{\phi} + \frac{\partial \mathcal{G} }{\partial \phi}(\phi, u) \tilde{u}\\ \quad -div(\mathcal{K}_{i}\nabla (\tilde{\phi}+\tilde{\varphi}_e)) = m_1(\phi - \phi_{obs})\Upsilon(t)\chi_{\Omega_{obs}}, \; \; \mathit{\text{in}}\; \; \Omega\times ({r_1(T)}, T)\\ -\mathfrak{c}_{m}\frac{\partial \tilde{\phi} }{\partial t} + \frac{\partial \mathcal{I}}{\partial \phi}(\phi, u)\tilde{\phi}+ \frac{\partial \mathcal{G} }{\partial \phi}(\phi, u) \tilde{u} -div(\mathcal{K}_{i}\nabla (\tilde{\phi}+\tilde{\varphi}_e)) \\ \quad + \sum\limits_{i = 1}^k \left( a_i(., e_i(t)) \tilde{\phi}(., e_i(t)) + c_i(., e_i(t)) \tilde{u}(., e_i(t)) \right) e_i^\prime(t)\\ \quad = m_1(\phi- \phi^{obs})\Upsilon(t)\chi_{\Omega_{obs}}, \; \; \mathit{\text{in $\Omega\times ({r_{k+1}(T)}, {r_{k}(T)})$, $k = n_1, \dots, 1$}}\\ -div(\mathcal{K}_{i}\nabla \tilde{\phi}) - div((\mathcal{K}_{i} + \mathcal{K}_{e})\nabla \tilde{\varphi}_e) = 0, \; \; \mathit{\text{in ${\cal Q}$}}\\ -\frac{\partial \tilde{u}}{\partial t}+ \frac{\partial \mathcal{G} }{\partial u}(\phi, u)\tilde{u}+\frac{\partial \mathcal{I}}{\partial u}(\phi, u) \tilde{\phi} = 0, \; \; \mathit{\text{in $\Omega\times(p_{1}(T), T)$}}\\ -\frac{\partial \tilde{u}}{\partial t}+ \frac{\partial \mathcal{G} }{\partial u}(\phi, u)\tilde{u}+\frac{\partial \mathcal{I}}{\partial u}(\phi, u) \tilde{\phi}\\ \quad +\sum\limits_{i = 1}^{l}\left( b_i(., q_i(t)) \tilde{\phi}(., q_i(t)) + d_i(., q_i(t)) \tilde{u}(., q_i(t))\right) q_i^\prime(t)\\ \quad = 0, \; \; \mathit{\text{in $\Omega\times ({p_{l+1}(T)}, {p_{l}(T)})$, $l = n_2, \dots, 1$, }}\\ \mathit{\text{subject to boundary and final conditions}}\\ (\mathcal{K}_{i}\nabla (\tilde{\phi}+\tilde{\varphi}_{e})).{\bf n} = 0, \; \; (\mathcal{K}_{e}\nabla \tilde{\varphi}_{e}).{\bf n} = 0, \mathit{\; \; \text{on}}\; \; \Sigma, \\ \tilde{\phi}(., T) = \frac{m_2}{\mathfrak{c}_{m}}(\phi(T) - \psi_{obs})\chi_{\Omega_{obs}}, \; \; \tilde{u}(., T) = 0, \mathit{\; \; \text{in }}\; \Omega. \end{array} \end{equation} | (3.86) |
Proof. Let \widetilde{\textbf{X}} = (\tilde{\phi}, \tilde{\varphi}_{e}, \tilde{u}) be sufficiently regular such that (\tilde{\phi}, \tilde{u})(T) = (\frac{m_2}{\mathfrak{c}_{m}}(\phi(T)-\psi_{obs})\chi_{\Omega_{obs}}, 0) .
Now multiplying the system (3.7) by \widetilde{\textbf{X}} and integrating over \mathcal{Q} , we obtain
\begin{equation} \begin{array}{lrc} \int_0^T\!\!\!\int_\Omega \left(\mathfrak{c}_{m}\frac{\partial \Psi }{\partial t}+\frac{\partial \mathcal{I}}{\partial \phi}(\phi, u)\Psi + \frac{\partial \mathcal{I}}{\partial u}(\phi, u)w\right)\tilde\phi d{\bf{x}} dt- \int_0^T\!\!\!\int_\Omega div(\mathcal{K}_{i}\nabla (\Psi + \psi_e )) \tilde\phi d{\bf{x}} dt\\ = \int_0^T\!\!\!\int_\Omega \left( \sum\limits_{k = 1}^{n_1} a_k({\bf{x}}, t) \Psi({\bf{x}}, r_k(t)) + \sum\limits_{l = 1}^{n_2} b_l({\bf{x}}, t)w({\bf{x}}, p_l(t)) \right)\tilde\phi d{\bf{x}} dt +\int_0^T\!\!\!\int_\Omega \tilde\phi{\cal B}_{2}h_{2} d{\bf{x}} dt, \\ - \int_0^T\!\!\!\int_\Omega div(\mathcal{K}_{i}\nabla \Psi + \left( \mathcal{K}_{i} + \mathcal{K}_{e}\right)\nabla \psi_e)\tilde{\varphi}_e d{\bf{x}} dt = \int_0^T\!\!\!\int_\Omega \tilde{\varphi}_{e}{\cal B}_{1}h_{1} d{\bf{x}} dt, \\ \int_0^T\!\!\!\int_\Omega \left(\frac{\partial w}{\partial t} +\frac{\partial \mathcal{G}}{\partial \phi}(\phi, u) \Psi+ \frac{\partial \mathcal{G} }{\partial u}(\phi, u) w\right)\tilde{u} d{\bf{x}} dt \\ = \int_0^T\!\!\!\int_\Omega \left(\sum\limits_{k = 1}^{n_1} c_k({\bf{x}}, t) \Psi({\bf{x}}, r_k(t)) + \sum\limits_{l = 1}^{n_2} d_l({\bf{x}}, t)w({\bf{x}}, p_l(t))\right)\tilde{u} d{\bf{x}} dt. \end{array} \end{equation} | (3.87) |
Using Green's theorem and integrating by part in time, the above system takes the following form (since (\Psi, w)(0) = 0) and according to boundary conditions satisfied by (\Psi, \psi_{e}) )
\begin{equation} \begin{array}{lcr} \int_0^T\!\!\!\int_\Omega \left(-\mathfrak{c}_{m}\frac{\partial \tilde{\phi}}{\partial t} +\frac{\partial \mathcal{I} }{\partial \phi}(\phi, u)\tilde{\phi}-div(\mathcal{K}_{i}\nabla \tilde{\phi}) \right)\Psi d{\bf{x}} dt+\int_{\Omega_{obs}} m_{2}(\phi(T)-\psi_{obs})\Psi(T) d{\bf{x}}\\ \quad - \int_0^T\!\!\!\int_\Omega div(\mathcal{K}_{i}\nabla \tilde{\phi})\psi_e d{\bf{x}} dt -\int_0^T\!\!\!\int_\Omega \Big(\sum\limits_{k = 1}^{n_1} a_k({\bf{x}}, t)\tilde{\phi} ({\bf{x}}, t) \Psi({\bf{x}}, r_k(t))\Big) d{\bf{x}} dt\\ \quad +\int_0^T\!\!\!\int_\Omega (\frac{\partial \mathcal{I} }{\partial u}(\phi, u)\tilde{\phi})w d{\bf{x}} dt - \int_0^T\!\!\!\int_\Omega \Big(\sum\limits_{l = 1}^{n_2} b_l({\bf{x}}, t)\tilde{\phi}({\bf{x}}, t) w({\bf{x}}, p_l(t))\Big)d{\bf{x}} dt\\ \quad = \int_0^T\!\!\!\int_\Omega \tilde{\phi} {\cal B}_{2}h_{2}d{\bf{x}} dt -\int_0^T\!\int_\Gamma ((\mathcal{K}_{i}\nabla \tilde{\phi}).{\bf n})(\Psi +\psi_e )d\gamma dt, \\ - \int_0^T\!\!\!\int_\Omega div(\mathcal{K}_{i}\nabla\tilde{\varphi}_e)\Psi d{\bf{x}} dt - \int_0^T\!\!\!\int_\Omega div((\mathcal{K}_{i} + \mathcal{K}_{e})\nabla \tilde{\varphi}_e)\psi_e d{\bf{x}} dt\\ \quad = -\int_0^T\!\int_\Gamma \left\{\Big(\big((\mathcal{K}_{i}+\mathcal{K}_{e})\nabla \tilde{\varphi}_{e}\big).{\bf n}\Big)\psi_e+\Big(\big(\mathcal{K}_{i}\nabla \tilde{\varphi}_{e} \big).{\bf n}\Big)\Psi \right\}d\gamma dt +\int_0^T\!\!\!\int_\Omega \tilde{\varphi}_{e} {\cal B}_{1}h_{1}d{\bf{x}} dt, \\ \int_0^T\!\!\!\int_\Omega \left(-\frac{\partial \tilde{u}}{\partial t} + \frac{\partial \mathcal{G} }{\partial u}(\phi, u) \tilde{u}\right)w d{\bf{x}} dt- \int_0^T\!\!\!\int_\Omega \Big(\sum\limits_{l = 1}^{n_2} d_l({\bf{x}}, t)\tilde{u}({\bf{x}}, t) w({\bf{x}}, p_l(t))\Big) d{\bf{x}} dt \\ \quad +\int_0^T\!\!\!\int_\Omega\frac{\partial \mathcal{G}}{\partial \phi}(\phi, u) \tilde{u} \Psi d{\bf{x}} dt - \int_0^T\!\!\!\int_\Omega \Big(\sum\limits_{k = 1}^{n_1} c_k({\bf{x}}, t) \tilde{u}({\bf{x}}, t)\Psi({\bf{x}}, r_k(t))\Big)d{\bf{x}} dt = 0. \end{array} \end{equation} | (3.88) |
By summing the three relations of the system (3.88), one obtains
\begin{equation} \begin{array}{lcr} \int_0^T\!\!\!\int_\Omega \left(-\mathfrak{c}_{m}\frac{\partial \tilde{\phi}}{\partial t} +\frac{\partial \mathcal{I} }{\partial \phi}(\phi, u)\tilde{\phi}+\frac{\partial \mathcal{G} }{\partial \phi}(\phi, u)\tilde{u}-div(\mathcal{K}_{i}\nabla (\tilde{\phi}+\tilde{\varphi}_{e})) \right)\Psi d{\bf{x}} dt\\ \quad -\int_0^T\!\!\!\int_\Omega \Big(\sum\limits_{k = 1}^{n_1} \big(a_k({\bf{x}}, t)\tilde{\phi} ({\bf{x}}, t) +c_k({\bf{x}}, t) \tilde{u}({\bf{x}}, t)\big)\Psi({\bf{x}}, r_k(t))\Big) d{\bf{x}} dt\\ - \int_0^T\!\!\!\int_\Omega \left(div((\mathcal{K}_{i}+ \mathcal{K}_{e})\nabla \tilde{\varphi}_e)+div(\mathcal{K}_{i}\nabla \tilde{\phi})\right)\psi_e d{\bf{x}} dt\\ +\int_0^T\!\!\!\int_\Omega \left(-\frac{\partial \tilde{u}}{\partial t} + \frac{\partial \mathcal{G} }{\partial u}(\phi, u) \tilde{u}+\frac{\partial \mathcal{I} }{\partial u}(\phi, u)\tilde{\phi}\right)w d{\bf{x}} dt+\int_{\Omega_{obs}} m_{2}(\phi(T)-\psi_{obs})\Psi(T) d{\bf{x}}\\ - \int_0^T\!\!\!\int_\Omega \Big(\sum\limits_{l = 1}^{n_2} \big(b_l({\bf{x}}, t)\tilde{\phi}({\bf{x}}, t)+ d_l({\bf{x}}, t)\tilde{u}({\bf{x}}, t) \big)w({\bf{x}}, p_l(t))\Big) d{\bf{x}} dt \\ = \int_0^T\!\!\!\int_\Omega \tilde{\varphi}_{e} {\cal B}_{1}h_{1} d{\bf{x}} dt+\int_0^T\!\!\!\int_\Omega \tilde{\phi} {\cal B}_{2}h_{2} d{\bf{x}} dt -\int_0^T\!\int_\Gamma ((\mathcal{K}_{i}\nabla \tilde{\phi}).{\bf n})(\Psi +\psi_e )d\gamma dt\\ -\int_0^T\!\int_\Gamma \left\{\Big(\big((\mathcal{K}_{i}+\mathcal{K}_{e})\nabla \tilde{\varphi}_{e}\big).{\bf n}\Big)\psi_e+\Big(\big(\mathcal{K}_{i}\nabla \tilde{\varphi}_{e}\big).{\bf n}\Big )\Psi \right\}d\gamma dt. \end{array} \end{equation} | (3.89) |
Now we calculate the terms corresponding to delays operators. For this let s = r_{k}(t) (respectively s = p_{l}(t) ), then t = r^{-1}_{k}(s) = e_{k}(s) (respectively t = p^{-1}_{l}(s) = q_{l}(s) ) and dt = e'_{k}(s)ds (respectively dt = q'_{l}(s)ds ). So (for \tilde{a}_{k} = a_{k} or c_{k} , and \tilde{b}_{l} = b_{l} or d_{l} )
\begin{equation*} \begin{array}{lcr} \int_0^T\! \tilde{a}_k({\bf{x}}, t)\Psi({\bf{x}}, r_k(t))\tilde{\phi}({\bf{x}}, t)dt & = \int_{r_{k}(0)}^{r_{k}(T)} \tilde{a}_k({\bf{x}}, e_k(s)) \Psi({\bf{x}}, s) \tilde{\phi}({\bf{x}}, e_k(s)) e_k^\prime(s)ds, \\ \int_0^T\! \tilde{b}_l({\bf{x}}, t)w({\bf{x}}, p_l(t))\tilde{\phi}({\bf{x}}, t)dt & = \int_{p_{l}(0)}^{p_{l}(T)} \tilde{b}_l({\bf{x}}, q_l(s)) w({\bf{x}}, s) \tilde{\phi}({\bf{x}}, q_l(s)) q_l^\prime(s)ds. \end{array} \end{equation*} |
Since \Psi and w are zero on {\cal Q}_{0} , we can conclude that
\begin{equation} \begin{array}{lcr} \int_0^T\! \sum\limits_{k = 1}^{n_1}\tilde{a}_k({\bf{x}}, t)\theta({\bf{x}}, r_k(t))\tilde\phi({\bf{x}}, t)dt \! = \! \sum\limits_{i = n_{1}}^{1}\int_{r_{i+1}(T)}^{r_i(T)} \sum\limits_{k = 1}^{i}\tilde{a}_k({\bf{x}}, e_k(s))e_k^\prime(s) \tilde\phi({\bf{x}}, e_k(s)) \Psi({\bf{x}}, s)ds, \\ \int_0^T\! \sum\limits_{l = 1}^{n_2} \tilde{b}_l({\bf{x}}, t)w({\bf{x}}, p_l(t))\tilde\phi({\bf{x}}, t)dt \! = \! \sum\limits_{j = n_{2}}^{1}\int_{p_{j+1}(T)}^{p_j(T)} \sum\limits_{l = 1}^{i} \tilde{b}_l({\bf{x}}, q_l(s))q_l^\prime(s) \tilde\phi({\bf{x}}, q_l(s))w({\bf{x}}, s)ds. \end{array} \end{equation} | (3.90) |
According to (3.90), system (3.89) becomes
\begin{equation*} \begin{array}{lcr} \int_0^T\!\!\!\int_\Omega \left(-\mathfrak{c}_{m}\frac{\partial \tilde{\phi}}{\partial t} +\frac{\partial \mathcal{I} }{\partial \phi}(\phi, u)\tilde{\phi}+\frac{\partial \mathcal{G} }{\partial \phi}(\phi, u)\tilde{u}-div(\mathcal{K}_{i}\nabla (\tilde{\phi}+\tilde{\varphi}_{e})) \right)\Psi d{\bf{x}} dt\\ -\sum\limits_{i = n_{1}}^{1}\int_\Omega\int_{r_{i+1}(T)}^{r_i(T)}\sum\limits_{k = 1}^{i}\Big(a_k({\bf{x}}, e_k(t))\tilde\phi({\bf{x}}, e_k(t)) +c_k({\bf{x}}, e_k(t)) \tilde{u}({\bf{x}}, e_k(t)) \Big)e_k^\prime(t) \Psi({\bf{x}}, t)d{\bf{x}}dt \end{array} \end{equation*} |
\begin{equation*} \begin{array}{lcr} - \int_0^T\!\!\!\int_\Omega \left(div((\mathcal{K}_{i}+ \mathcal{K}_{e})\nabla \tilde{\varphi}_e)+div(\mathcal{K}_{i}\nabla \tilde{\phi})\right)\psi_e d{\bf{x}} dt\\ +\int_0^T\!\!\!\int_\Omega \left(-\frac{\partial \tilde{u}}{\partial t} + \frac{\partial \mathcal{G} }{\partial u}(\phi, u) \tilde{u}+\frac{\partial \mathcal{I} }{\partial u}(\phi, u)\tilde{\phi}\right)wd{\bf{x}} dt\\ - \sum\limits_{j = n_{2}}^{1}\int_\Omega\int_{p_{j+1}(T)}^{p_j(T)}\sum\limits_{l = 1}^{j}\Big(b_l({\bf{x}}, q_l(t))\tilde{\phi}({\bf{x}}, q_l(t)) +d_l({\bf{x}}, q_l(t)) \tilde{u}({\bf{x}}, q_l(t)) \Big)q_l^\prime(t) w({\bf{x}}, t)d{\bf{x}} dt \\ \quad = \int_0^T\!\!\!\int_{\Omega_{c}} \tilde{\varphi}_{e} {\cal B}_{1}h_{1} d{\bf{x}} dt+\int_0^T\!\!\!\int_{\Omega_{d}} \tilde{\phi} {\cal B}_{2}h_{2} d{\bf{x}} dt-\int_{\Omega_{obs}} m_{2}(\phi(T)-\psi_{obs})\Psi(T) d{\bf{x}}\\ \quad -\int_0^T\!\int_\Gamma \big((\mathcal{K}_{i}\nabla \tilde{\phi}).{\bf n}\big)(\Psi +\psi_e )d\gamma dt\\ \quad -\int_0^T\!\int_\Gamma \left\{\Big(\big((\mathcal{K}_{i}+\mathcal{K}_{e})\nabla \tilde{\varphi}_{e}\big).{\bf n}\Big)\psi_e+\Big((\mathcal{K}_{i}\nabla \tilde{\varphi}_{e}).{\bf n}\Big)\Psi \right\}d\gamma dt. \end{array} \end{equation*} |
In order to simplify the previous system, we suppose that (\tilde{\phi}, \tilde{\varphi}_{e}, \tilde{u}) satisfies the following "adjoint" system
\begin{equation} \begin{array}{lcr} -\mathfrak{c}_{m}\frac{\partial \tilde{\phi} }{\partial t}+ \frac{\partial \mathcal{I}}{\partial \phi}(.;\phi, u)\tilde{\phi} + \frac{\partial \mathcal{G} }{\partial \phi}(.;\phi, u) \tilde{u} \\ \quad -div(\mathcal{K}_{i}\nabla (\tilde{\phi}+\tilde{\varphi}_e)) = m_1(\phi - \phi_{obs})\Upsilon(t)\chi_{\Omega_{obs}}, \; \; \text{in}\; \; \Omega\times ({r_1(T)}, T)\\ -\mathfrak{c}_{m}\frac{\partial \tilde{\phi} }{\partial t} + \frac{\partial \mathcal{I}}{\partial \phi}(.;\phi, U)\tilde{\phi}+ \frac{\partial \mathcal{G} }{\partial \phi}(.;\phi, u) \tilde{u}({\bf{x}}, t) -div(\mathcal{K}_{i}\nabla (\tilde{\phi}+\tilde{\varphi}_e)) \\ \quad + \sum\limits_{i = 1}^k \left( a_i(., e_i(t)) \tilde{\phi}(., e_i(t)) + c_i(., e_i(t)) \tilde{u}(., e_i(t)) \right) e_i^\prime(., t)\\ \quad = m_1(\phi- \phi^{obs})\Upsilon(t)\chi_{\Omega_{obs}}, \; \; \text{in $\Omega\times ({r_{k+1}(T)}, {r_{k}(T)})$, $k = n_1, \dots, 1$}\\ -div\big(\mathcal{K}_{i}\nabla \tilde{\phi}({\bf{x}}, t)\big) - div\big((\mathcal{K}_{i} + \mathcal{K}_{e})\nabla \tilde{\varphi}_e({\bf{x}}, t)\big) = 0, \; \; \text{in ${\cal Q}$}\\ -\frac{\partial \tilde{u}}{\partial t} ({\bf{x}}, t)+ \frac{\partial \mathcal{G} }{\partial u}(.;\phi, u)\tilde{u}({\bf{x}}, t) + \frac{\partial \mathcal{I}}{\partial u}(.;\phi, u) \tilde{\phi}({\bf{x}}, t) = 0, \; \; \text{in $\Omega\times(p_{1}(T), T)$}\\ -\frac{\partial \tilde{u}}{\partial t}+ \frac{\partial \mathcal{G} }{\partial u}(.;\phi, u)\tilde{u}+\frac{\partial \mathcal{I}}{\partial u}(.;\phi, u) \tilde{\phi}\\ \quad +\sum\limits_{i = 1}^{l}\left( b_i(., q_i(t)) \tilde{\phi}(., q_i(t)) + d_i(., q_i(t)) \tilde{u}(., q_i(t))\right) q_i^\prime\\ \quad = 0, \; \; \text{in $\Omega\times ({p_{l+1}(T)}, {p_{l}(T)})$, $l = n_2, \dots, 1$, }\\ (\mathcal{K}_{i}\nabla (\tilde{\phi}+\tilde{\varphi}_{e})).{\bf n} = 0, \; \; (\mathcal{K}_{e}\nabla \tilde{\varphi}_{e}).{\bf n} = 0, \; \; \text{on}\; \; \Sigma, \\ \tilde{\phi}(., T) = \frac{m_2}{\mathfrak{c}_{m}}(\phi(., T) - \psi_{obs})\chi_{\Omega_{obs}}, \; \; \tilde{u}(., T) = 0, \; \; \text{in}\; \; \Omega. \end{array} \end{equation} | (3.91) |
Since \int_0^T\!\! = \!\!\sum\limits_{i = n_{1}}^{1}\int_{r_{i+1}(T)}^{r_i(T)}\!+\!\int_{r_{1}(T)}^{T}\! = \!\sum\limits_{j = n_{2}}^{1}\int_{p_{j+1}(T)}^{p_j(T)}\!+\!\int_{p_{1}(T)}^{T} we can then deduce from the two previous systems that
\begin{equation} \begin{array}{lcr} \int_0^T\!\!\!\int_{\Omega_{obs}} m_1(\phi - \phi_{obs})\Upsilon(t) \Psi d{\bf{x}} dt+\int_{\Omega_{obs}} m_2(\phi(., T) - \phi_{obs})\Psi(., T) d{\bf{x}}\\ \quad = \int_0^T\!\!\!\int_{\Omega_{c}} h_{1} {\cal B}^{*}_{1}\tilde{\varphi}_{e} d{\bf{x}} dt+\int_0^T\!\!\!\int_{\Omega_{d}} h_{2} {\cal B}^{*}_{2}\tilde{\phi} d{\bf{x}} dt. \end{array} \end{equation} | (3.92) |
According to (3.92) and (3.84), the expression (3.82) of {\cal J}^\prime takes the form
\begin{equation} \begin{array}{c} {\cal J}^\prime(\xi, \pi)\cdot (h_{1}, h_{2}) = (h_{1}, \chi_{\Omega_{c}}\aleph_{1}^{*}\tilde{\varphi}_{e} +\alpha \xi)_{U_{c}} +\int_0^T\!\!\!\int_{\Omega_{d}} h_{2} ({\cal B}^{*}_{2}\tilde{\phi}-\beta \pi )d{\bf{x}} dt. \end{array} \end{equation} | (3.93) |
Since (\xi^{*}, \pi^{*}) is an optimal solution we have ( \forall (\xi, \pi)\in {\cal U}_{ad}\times {\cal V}_{ad} )
\begin{equation} \begin{array}{lcr} \frac{\partial {\cal J}}{\partial \xi}(\xi^{*} , \pi^{*}).(\xi-\xi^{*}) = (\chi_{\Omega_{c}}\aleph_{1}^{*}\tilde{\varphi}^{*}_{e} +\alpha \xi^{*}, \xi-\xi^{*})_{U_{c}}\geq 0, \\ \frac{\partial {\cal J}}{\partial \phi}(\xi^{*} , \pi^{*}).(\pi -\pi^{*}) = \int_0^T\!\!\!\int_{\Omega_{d}}(-\beta\pi^{*}+ {\cal B}_{2}^{*}\tilde{\phi}^{*}) (\pi -\pi^{*})d{\bf{x}}dt\leq 0, \\ \end{array} \end{equation} | (3.94) |
where (\tilde{\phi}^{*}, \tilde{\varphi}^{*}_{e}, \tilde{u}^{*}) is the solution of (3.91) corresponding to {\cal F}(\xi^{*}, \pi^{*}) . This completes the proof.
Remark 3.3. By using a standard control argument (see e.g. [10], page 207) concerning the sign of the variations (h_{1}, h_{2}) (depending on the size of (\xi, \pi) ), we obtain that
\begin{equation} \begin{array}{lcr} \xi^{*} = \max\Big(-\tau_{1}, \min\big(\frac{-\chi_{\Omega_{c}}\aleph_{1}^{*}\tilde{\varphi}_{e}^{*}}{\alpha}, \tau_{1}\big)\Big), \pi^{*} = \max\Big(-\tau_{2}, \min\big(\frac{\chi_{\Omega_{d}}{\cal B}_{2}^{*}\tilde{\phi}^{*}}{\beta}, \tau_{2}\big)\Big). \end{array} \end{equation} | (3.95) |
Let us now give the well-posedness of adjoint system (3.86).
Proposition 3.1. Assume that assumptions of Theorem 3.2 hold and that (\phi, {\varphi}_{e}, u) is in \mathbb{D} . Adjoint problem (3.86) admits one unique solution (\tilde{\phi}, \tilde{\varphi}_{e}, \tilde{u})\in \mathcal{W}({\cal Q}) with (\frac{\partial \tilde{\phi}}{\partial t}, \frac{\partial \tilde{u}}{\partial t})\in L^{4/3}(0, T; \mathbb{V}')\times L^{2}({\cal Q}) and (\tilde{\phi}, \tilde{u})\in (C^{0}([0, T]; L^{2}(\Omega)))^{2} .
Proof. To prove the existence of a unique solution (\tilde{\phi}, \tilde{\varphi}_{e}, \tilde{u}) of linear problem (3.86), which is backward in time, we transform problem (3.86) into an initial-boundary value problem by reversing the sense of time i.e., t: = T-t . By using the results of Lemma 7 on each time interval, we obtain the existence and uniqueness of the solution.
Remark 3.4. In this section, our main results investigate Fréchet differentiability properties of solution operator and minimax control problems related to the nonlinear delayed dynamic system (3.1) with an abstract class of ionic models, including some classical models as Rogers-McCulloch, Fitz-Hugh-Nagumo and Aliev-Panfilov. We can consider other ionic model type including Mitchell-Schaeffer model (see [52]). This two-variable model can be defined with operators \mathcal{I} and \mathcal{G} as (for example)
\begin{equation} \begin{array}{lcr} \mathcal{I}(\phi, u) = -\frac{u}{\tau_{in}}\frac{(\phi-\phi_{min})^2(\phi_{max}-\phi)}{(\phi_{max}-\phi_{min})} - \frac{1}{\tau_{out}}\frac{\phi-\phi_{min}}{(\phi_{max}-\phi_{min})}, \\ \mathcal{G}(\phi, u) = \left\{\begin{array}{ll} \frac{u}{\tau_{open}}-\frac{1}{\tau_{open}(\phi_{max}-\phi_{min})^2} &if \ \phi \lt \phi_{gate}, \\ \frac{u}{\tau_{close}}& if\ \phi\geq\phi_{gate}. \end{array} \right.\\ \end{array} \end{equation} | (3.96) |
These operators depend on the change-over voltage \phi_{gate} , the resting potential \phi_{min} , the maximum potential \phi_{max} , and on times constants \tau_{in} , \tau_{out} , \tau_{open} and \tau_{close} . The two times \tau_{open} and \tau_{close} , respectively controlling the durations of the action potential and of the recovery phase, and the two times \tau_{in} and \tau_{out} , respectively controlling the length of depolarization and repolarization phases. This model is well-known to be valid under the assumption \tau_{in}\ll\tau_{out}\ll\min (\tau_{open}, \tau_{close}).
In order to guarantee the well-posedness of system (3.1) with Mitchell-Schaeffer ionical model, we can use the following regularized version of ionic operator \mathcal{G}
\begin{equation} \begin{array}{r} \mathcal{G}_{\zeta}(\phi, u) = \left( \frac{1}{\tau_{close}} - \big(\frac{1}{\tau_{close}}-\frac{1}{\tau_{open}}\big) h_\zeta(\phi) \right)\left(u-h_\zeta(\phi) \right)\\ \quad +\frac{1}{\tau_{open}}\left(1-\frac{1}{(\phi_{max}-\phi_{min})^2 }\right)h_\zeta(\phi), \end{array} \end{equation} | (3.97) |
where the differentiable function 0\leq h_\zeta \leq 1 is given by
\begin{equation*} h_\zeta(\phi) = \frac{1}{2}\left(1- \tanh \left(\frac{\phi-\phi_{gate}}{\zeta} \right) \right), \end{equation*} |
with \zeta a positive parameter.
The operator \mathcal{G}_{\zeta}(\phi, u) can be written as \mathcal{G}_{\zeta}(\phi, u) = \mathcal{I}_{2, \zeta}(\phi)+\hbar_{\zeta}(\phi) u where
\begin{equation} \begin{array}{ll} \mathcal{I}_{2, \zeta}(\phi) = -\left( \frac{1}{\tau_{close}} - \big(\frac{1}{\tau_{close}}-\frac{1}{\tau_{open}}\big) h_\zeta(\phi) \right)h_\zeta(\phi) +\frac{1}{\tau_{open}}\left(1-\frac{1}{(\phi_{max}-\phi_{min})^2 }\right)h_\zeta(\phi), \\ \hbar_{\zeta}(\phi) = \frac{1}{\tau_{close}} - \Big(\frac{1}{\tau_{close}}-\frac{1}{\tau_{open}}\Big) h_\zeta(\phi) . \end{array} \end{equation} | (3.98) |
According to the definition of \tanh , we can deduce that \lim\limits_{\zeta\rightarrow 0} h_{\zeta}(\phi) = \left\{ \begin{array}{ll} 1 \ if\ \phi < \phi_{gate}, \\ 0\ if\ \phi > \phi_{gate} \end{array} \right. and then \lim\limits_{\zeta\rightarrow 0} \mathcal{G}_\zeta(\phi, u) = \mathcal{G}(\phi, u) . The regularized Mitchell-Schaeffer model has a slightly different structure compared to models in (2.3) because in this model, \hbar_{\zeta} depend on \phi through the function h_\zeta . Since h_\zeta is sufficiently regular, the arguments of this paper can be adapted with some necessary modifications to analyse minimax control problems with the regularized Mitchell-Schaeffer ionical model.
More general, the study developed in this paper remains valid if we consider the operator \mathcal{G} in the form of \mathcal{G}({\bf{x}}, t; \phi, u) = \mathcal{I}_{2}({\bf{x}}, t; \phi)+\hbar({\bf{x}}, t; \phi) u (i.e. a general form of Hodgkin-Huxley model including Beeler-Reuter and Luo-Rudy ionic models described by continuous or regularized discontinuous functions, see [5,49,50]) with \mathcal{G} Carathéodory function from (\Omega \times \mathit{IR}^{{}}) \times \mathit{IR}^{{2}} into \mathit{IR}^{{}} and locally Lipschitz continuous function on (\phi, u) and, \mathcal{I}_{2} and \hbar sufficiently regulars.
We end this section by a description of a gradient algorithm to solve the minimax control problem, by using adjoint model. The method is formulated in terms of continuous variables which are independent of a specific numerical discretization. For more details concerning some optimization strategies in order to solve minimax control problem, by using the adjoint model, in term of the continuous variables and in terms of discrete variables (based on the discretization of continuous direct, adjoint and sensitive models) see Chapter 9 of [10].
We present algorithms where the descent direction is calculated by using the adjoint variables, particularly by choosing an admissible step size. For a given observation (\phi_{obs}, \psi_{obs}) , initial states (\phi_{0}, u_{0}) and past states (\phi_{past}, u_{past}) , the resolution of the nonlinear minimax control problem (3.6), with cost functional given by (3.4), by gradient methods requires, at each iteration of the optimization algorithm, the resolution of direct problem (3.1) and its corresponding adjoint problem (3.86).
The gradient algorithm for the resolution of treated saddle point problems is given by:
for k\geq 1 , (iteration index) we denote by (\xi_{k}, \pi_{k}) the numerical approximation of the control-disturbance at the kth iteration of the algorithm.
(1) Initialization: k = 0 and (\xi_{0}, \pi_{0}) (given initial guess).
(2) Resolution of direct problem (3.1) with source term (\xi_{k}, \pi_{k}) , gives (\phi^{(k)}, \varphi^{(k)}_{e}, u^{(k)}) = {\cal F}(\xi_{k}, \pi_{k}) .
(3) Resolution of adjoint problem (3.86) (based on (\xi_{k}, \pi_{k}, {\cal F}(\xi_{k}, \pi_{k})) , gives (\tilde{\phi}^{(k)}, \tilde{\varphi}^{(k)}_{e}, \tilde{u}^{(k)}) .
(4) Local expression of the gradient of {\cal J} at point (\xi_{k}, \pi_{k}) :
\begin{equation*} ({\bf{GJ}}), \left\{ \begin{array}{lcr} \upsilon_{k}\stackrel{def}{ = }\frac{\partial {\cal J}}{\partial \xi}(\xi_{k}, \pi_{k}) = \alpha\xi_{k}+ \chi_{\Omega_{c}}\aleph_{1}^{*}\tilde{\varphi}_{e, k}, \\ \varpi_{k}\stackrel{def}{ = }\frac{\partial {\cal J}}{\partial \pi}(\xi_{k}, \pi_{k}) = -\beta\pi_{k}+ \chi_{\Omega_{d}}{\cal B}^{*}_{2}\tilde{\phi}_{k}, \\ G_{k} = (\upsilon_{k}, \varpi_{k}). \end{array} \right. \end{equation*} |
(5) Determine (\xi_{k+1}, \pi_{k+1}) :
\begin{equation*} \left\{ \begin{array}{lcr} \xi_{k+1}: = \xi_{k}-\varsigma_{k}\upsilon_{k}, \\ \pi_{k+1}: = \pi_{k}+\delta_{k}\varpi_{k}, \end{array} \right. \end{equation*} |
where 0 < m\leq \varsigma_{k}, \delta_{k}\leq M are the sequences of step lengths.
(6) IF the gradient is sufficiently small (convergence) THEN end; ELSE set k: = k+1 and REPEAT from (2) UNTIL convergence.
The approximation of optimal Solution is: (\xi^{*}, \pi^{*}; \phi^{*}, \varphi^{*}_{e}, u^{*}): = (\xi_{k}, \pi_{k}; \phi^{(k)}, \varphi^{(k)}_{e}, u^{(k)}) . The convergence of the algorithm depends on the second Fréchet derivative of {\cal J} (i.e. m, M depend on the second Fréchet derivative of {\cal J} ).
In order to obtain an algorithm which is numerically efficient, the best choice of \varsigma_{k}, \delta_{k} will be the result of a line minimization and maximization algorithm, respectively. Otherwise, at each iteration step k of previous algorithm, we solve the one-dimensional optimization problem of parameters \varsigma_{k} and \delta_{k} :
\begin{equation} \begin{array}{lcr} \varsigma_{k} = \min\limits_{\lambda \gt 0}{\cal J}(\xi_{k}-\lambda \upsilon_{k}, \pi_{k}), \\ \delta_{k} = \min\limits_{\lambda \gt 0}{\cal J}(\xi_{k}, \pi_{k}+\lambda \varpi_{k}). \end{array} \end{equation} | (3.99) |
From the numerical computation viewpoint, it is most efficient to compute (\varsigma_{k}, \delta_{k}) only approximately, in order to reduce computational cost. To derive an approximation for a pair (\varsigma_{k}, \delta_{k}) we can use a purely heuristic approach, for example, by taking \varsigma_{k} = min(1, \left\| { \upsilon_{k}} \right\| ^{-1}_{\infty}) and \delta_{k} = min(1, \left\| { \varpi_{k}} \right\| ^{-1}_{\infty}) or by using the linearization of {\cal F}(\xi_{k}-\lambda \upsilon_{k}, \pi_{k}) at \xi_{k} and {\cal F}(\xi_{k}, \pi_{k}+\lambda \varpi_{k}) at \pi_{k} by
\begin{equation*} \begin{array}{lcr} {\cal F}(\xi_{k}-\lambda \upsilon_{k}, \pi_{k})\approx {\cal F}(\xi_{k}, \pi_{k})-\lambda \frac{\partial {\cal F}}{\partial \xi}(\xi_{k}, \pi_{k}).\upsilon_{k}, \\ {\cal F}(\xi_{k}, \pi_{k}+\lambda \varpi_{k})\approx {\cal F}(\xi_{k}, \pi_{k})+\lambda \frac{\partial {\cal F}}{\partial \pi}(\xi_{k}, \pi_{k}).\varpi_{k}, \end{array} \end{equation*} |
where
\begin{equation*} \begin{array}{lcr} (\Psi^{(k)}_{c}, \psi^{(k)}_{e, c}, w^{(k)}_{c}) = \frac{\partial {\cal F}}{\partial \xi}(\xi_{k}, \pi_{k}).\upsilon_{k}, \\ (\Psi^{(k)}_{d}, \psi^{(k)}_{e, d}, w^{(k)}_{d}) = \frac{\partial {\cal F}}{\partial \pi}(\xi_{k}, \pi_{k}).\varpi_{k}, \end{array} \end{equation*} |
are solutions of the sensitivity problem (3.7).
According to the previous approximation, we can approximate the problem (3.99) by
\begin{equation} \begin{array}{lcr} \varsigma_{k} = \min\limits_{\lambda \gt 0}H(\lambda) \; \text{and}\; \delta_{k} = \max\limits_{\lambda \gt 0}R(\lambda), \end{array} \end{equation} | (3.100) |
where H(\lambda) = {\cal J}(\xi_{k}-\lambda \upsilon_{k}, \pi_{k}) and R(\lambda) = {\cal J}(\xi_{k}, \pi_{k}+\lambda \varpi_{k}) . Since H and R are polynomial functions of degree 2 (since the functional {\cal J} is quadratic), then problem (3.100) can be solved exactly. Consequently, we obtain explicitly the value of the parameters \varsigma_{k} and \delta_{k} .
Remark 3.5. 1. After derived the gradient of functional {\cal J} , by using the adjoint model corresponding to sensitivity state (which corresponds to the direct problem), we can use any other classical optimization strategies (as conjugate gradient method, Lagrange-Newton method) to solve control problem considered in this paper.
2. In the numerical treatment of minimax control problem, the direct system, adjoint system, sensitivity system and objective functional must be discretized (reduction of infinite-dimensional dynamics to finite-dimensional problems). The discretized formulation for direct, sensitivity and adjoint systems can be performed by combining Galerkin and finite element methods to the variational formulations associated to these coupled problems, for space discretization and semi-implicit backward differentiation schemes with an explicit treatment of ionic current, for time discretization, or by using lattice Boltzmann methods. In objective functional, the integrals with respect to time can be approximated by composition trapezoidal rules (see e.g. [7]).
3. Despite its apparent complexity, the proposed gradient algorithm is quite easy to implement. The main difficulty, in practical applications, is due to enormous storage requirements of state solution (and control-disturbance variables) for evaluating the adjoint equation over the whole time interval [0, T] , for large time horizons T or fine space-time meshes (because the computation of the discrete gradient by discrete adjoint methods requires one forward solve of the discrete state system and one backward solve of adjoint system in which state trajectory is an input). Fortunately, these storage requirements can be lowered by using e.g. the so-called ``checkpointing'' techniques (see e.g. [39]).
Modeling and control of electrical cardiac activity represent nowadays a very valuable tool to maximize the efficiency and safety of treatment for cardiac disease. For predicting and acting on phenomena and processes occurring inside and surrounding cardiac medium, we have discussed stabilization and regulation processes in order to determine, from some observations (desired target), the best optimal prognostic values of sources, in presence of disturbance and fluctuations. Coupling the proposed method with technical improvements in Magnetic Resonance Imaging (MRI) measurements and genetic, and ionic measurements, will be very beneficial and great help for diagnostics and treatments in medical practices.
The well-posedness and regularity of the governing nonlinear systems are discussed. The Fréchet differentiability and some properties of nonlinear operator solution are derived. Afterwards, minimax control problems have been formulated. Under suitable hypotheses, it is shown that one has existence of an optimal solution, and the appropriate necessary optimality conditions for an optimal solution, by introducing adjoint problems, are derived. These conditions (obtained in a Lagrangian form) correspond to identify the gradient of the cost functional that is necessary to develop numerical optimization methods (gradient methods, Newton methods, etc.). Some numerical methods, combining the obtained optimal necessary conditions and gradient-iterative algorithms, are presented in order to solve the minimax control problems.
It is clear that, in accordance with practical applications and available experimental observations, we can consider other observations, controls and/or disturbances (which can appear in boundary conditions, in initial conditions, in parameters of ionic models or in time-delay functions) in order to take into account at best the influence of uncertainty on the main phenomena and their mutual interactions that take place during the bioelectrical cardiac activity, and we obtain similar results by using similar approach as used in this work (for more details see [10]).
The author thank the referees for their valuable comments and suggestions to improve the quality of the manuscript.
The author declares there is no conflicts of interest in this paper.
[1] | Espín Susana & Samaniego Iván (2016) Manual para el análisis de parámetros químicos asociados a la calidad del cacao. Quito—Ecuador: Estación Experimental Santa Catalina- INIAP. |
[2] | Arvelo M, González D, Steven A, et al. (2017) Manual Técnico del Cultivo de Cacao Prácticas Latinoamericanas. Available from: https://iica.int/es. |
[3] |
Barišić V, Kopjar M, Jozinović A, et al.(2019) The chemistry behind chocolate production. Molecules 24: 3163. https://doi.org/10.3390/molecules24173163 doi: 10.3390/molecules24173163
![]() |
[4] | Beckett ST, Fowler MS, Ziegler GR (2017) Beckett's Industrial Chocolate Manufacture and Use. https://doi.org/10.1002/9781118923597 |
[5] | López Guerrero A (2017) Producción y Comercialización de Cacao Fino de Aroma en el Ecuador-Año 2012–2014. Available from: https://www.scpm.gob.ec/sitio/wp-content/uploads/2019/03/ESTUDIO-DEL-CACAO-IZ7-version-publica-ultima.pdf. |
[6] | ISO 2451 (2017) Cocoa Beans Specification and Quality Requirements. Switzerland. |
[7] |
Becerra LD, Quintanilla-Carvajal MX, Escobar S, et al. (2023) Correlation between color parameters and bioactive compound content during cocoa seed transformation under controlled process conditions. Food Biosci 53: 102526. https://doi.org/10.1016/j.fbio.2023.102526 doi: 10.1016/j.fbio.2023.102526
![]() |
[8] |
Pallares Pallares A, Perea Villamil JA, López Giraldo LJ (2017) Influence of fermentation and drying processes in the aroma precursor compounds of cocoa beans (Theobroma cacao L) CCN-51. Respuestas 21: 120–133. https://doi.org/10.22463/0122820X.726 doi: 10.22463/0122820X.726
![]() |
[9] |
Haruna L, Abano EE, Teye E, et al. (2024) Effect of partial pulp removal and fermentation duration on drying behavior, nib acidification, fermentation quality, and flavor attributes of Ghanaian cocoa beans. J Agric Food Res 17: 101211. https://doi.org/10.1016/j.jafr.2024.101211 doi: 10.1016/j.jafr.2024.101211
![]() |
[10] |
Llerena W, Samaniego I, Vallejo C, et al. (2023) Profile of bioactive components of cocoa (Theobroma cacao L.) by-products from Ecuador and evaluation of their antioxidant activity. Foods 12: 2583. https://doi.org/10.3390/foods12132583 doi: 10.3390/foods12132583
![]() |
[11] |
Cortez D, Quispe-Sanchez L, Mestanza M, et al. (2023) Changes in bioactive compounds during fermentation of cocoa (Theobroma cacao) harvested in Amazonas-Peru. Curr Res Food Sci. 6: 100494. https://doi.org/10.1016/j.crfs.2023.100494 doi: 10.1016/j.crfs.2023.100494
![]() |
[12] |
Wollgast J, Anklam E (2000) Review on polyphenols in Theobroma cacao: Changes in composition during the manufacture of chocolate and methodology for identification and quantitation. Food Res Int 33: 423–447. https://doi.org/10.1016/S0963-9969(00)00068-5 doi: 10.1016/S0963-9969(00)00068-5
![]() |
[13] |
Onomo PE, Niemenak N, Djocgoue PF, et al. (2015) Heritability of polyphenols, anthocyanins, and antioxidant capacity of Cameroonian cocoa (Theobroma cacao L.) beans. Afr J Biotechnol 14: 2672–2682. https://doi.org/10.5897/AJB2015.14715 doi: 10.5897/AJB2015.14715
![]() |
[14] |
Urbańska B, Kowalska J (2019) Comparison of the total polyphenol content and antioxidant activity of chocolate obtained from roasted and unroasted cocoa beans from different regions of the world. Antioxidants 8: 283. https://doi.org/10.3390/antiox8080283 doi: 10.3390/antiox8080283
![]() |
[15] |
Caporaso N, Whitworth MB, Fowler MS, et al. (2018) Hyperspectral imaging for non-destructive prediction of fermentation index, polyphenol content, and antioxidant activity in single cocoa beans. Food Chem 258: 343–351. https://doi.org/10.1016/j.foodchem.2018.03.039 doi: 10.1016/j.foodchem.2018.03.039
![]() |
[16] | Borja-Fajardo JG, Horta-Tellez HB, Peñaloza-Atuesta GC, et al. (2022) Antioxidant activity, total polyphenol content, and methylxantine ratio in four materials of Theobroma cacao L. from Tolima, Colombia. Heliyon 8: e09402. https://doi.org/10.1016/j.heliyon.2022.e09402 |
[17] | Arrazate C, Villarreal J, Campos E, et al. (2011) Diagnóstico del cacao en México SINAREFI Sistema Nacional de Recursos Fitogenéticos para la Alimentación y la Agricultura. |
[18] |
do Carmo Brito BDN, Campos Chisté R, da Silva Pena R, et al. (2017) Bioactive amines and phenolic compounds in cocoa beans are affected by fermentation. Food Chem 228: 484–490. https://doi.org/10.1016/j.foodchem.2017.02.004 doi: 10.1016/j.foodchem.2017.02.004
![]() |
[19] |
Calvo AM, Botina BL, García MC, et al. (2021) Dynamics of cocoa fermentation and its effect on quality. Sci Rep Dec 11: 16746. https://doi.org/10.1038/s41598-021-95703-2 doi: 10.1038/s41598-021-95703-2
![]() |
[20] |
Moretti LK, Ramos KK, Ávila PF, et al. (2023) Influence of cocoa varieties on carbohydrate composition and enzymatic activity of cocoa pulp. Food Res Int 173: 113393. https://doi.org/10.1016/j.foodres.2023.113393 doi: 10.1016/j.foodres.2023.113393
![]() |
[21] |
Samaniego I, Espín S, Quiroz J, et al. (2020) Effect of the growing area on the methylxanthines and flavan-3-ols content in cocoa beans from Ecuador. J Food Compos Anal 88: 103448. https://doi.org/10.1016/j.jfca.2020.103448 doi: 10.1016/j.jfca.2020.103448
![]() |
[22] | Quiroz-Vera J, Mestanza S, Parada-Vera N, et al. (2021) Catálogo de cultivares de cacao en Ecuador. 1era. Ed. 2021. Instituto Nacional de Investigaciones Agropecuarias. Boletín Técnico No. 449. |
[23] |
Melo TS, Pires TC, Engelmann JVP, et al. (2021) Evaluation of the content of bioactive compounds in cocoa beans during the fermentation process. J Food Sci Technol 58: 1947–1957. https://doi.org/10.1007/s13197-020-04706-w doi: 10.1007/s13197-020-04706-w
![]() |
[24] |
Menéndez-Cevallos LT, Burgos-Briones GA (2021) Efectos de la fermentación y secado en el contenido de polifenoles y alcaloides del cacao. Dominio De Las Ciencias 7: 1280–1304. http://dx.doi.org/10.23857/dc.v7i5.2310 doi: 10.23857/dc.v7i5.2310
![]() |
[25] |
Fang Y, Li R, Chu Z, et al. (2020) Chemical and flavor profile changes of cocoa beans (Theobroma cacao L.) during primary fermentation. Food Sci Nutr 8: 4121–4133. https://doi.org/10.1002/fsn3.1701 doi: 10.1002/fsn3.1701
![]() |
[26] | Tello Alonso S, Avendaño Arrazate CH, Vásquez Murrieta MS, et al. (2020) Contenido de compuestos bioactivos en Theobroma cacao L. Investigación y Desarrollo en Ciencia y Tecnología de Alimentos. |
[27] | Velásquez Reyes D, Gschaedler A, Kirchmayr M, et al. (2021) Cocoa bean turning as a method for redirecting the aroma compound profile in artisanal cocoa fermentation. Heliyon 7: e07694. https://doi.org/10.1016/j.heliyon.2021.e07694 |
[28] |
Papalexandratou Z, Kaasik K, Kauffmann LV, et al. (2019) Linking cocoa varietals and microbial diversity of Nicaraguan fine cocoa bean fermentations and their impact on final cocoa quality appreciation. Int J Food Microbiol 304: 106–118. https://doi.org/10.1016/j.ijfoodmicro.2019.05.012 doi: 10.1016/j.ijfoodmicro.2019.05.012
![]() |
[29] | Instituto Ecuatoriano de Normalización. NTE INEN 176 (2018) Granos de cacao. requisitos cocoa beans requirements. |
[30] | López M, Hernández E (2018) El proceso de fermentación del CACAO (Theobroma cacao L.). Agro Productividad 4: 20–24. Available from: https://revista-agroproductividad.org/index.php/agroproductividad/article/view/572/441 |
[31] |
De Vuyst L, Leroy F (2020) Functional role of yeasts, lactic acid bacteria, and acetic acid bacteria in cocoa fermentation processes. FEMS Microbiol Rev 44: 432–453. https://doi.org/10.1093/femsre/fuaa014 doi: 10.1093/femsre/fuaa014
![]() |
[32] | Sarbu I, Csutak O (2019) The microbiology of cocoa fermentation. In: Grumezescu AM, Holban AM (Eds.), Caffeinated and Cocoa Based Beverages, Volume 8: The Science of Beverages, Woodhead Publishing, 423–446. https://doi.org/10.1016/B978-0-12-815864-7.00013-1 |
[33] |
Visintin S, Ramos L, Batista N, et al. (2017) Impact of Saccharomyces cerevisiae and Torulaspora delbrueckii starter cultures on cocoa beans fermentation. Int J Food Microbiol 257: 31–40. https://doi.org/10.1016/j.ijfoodmicro.2017.06.004 doi: 10.1016/j.ijfoodmicro.2017.06.004
![]() |
[34] |
Afoakwa EO, Quao J, Budu AS, et al. (2011) Effect of pulp preconditioning on acidification, proteolysis, sugars, and free fatty acids concentration during fermentation of cocoa (Theobroma cacao) beans. Int J Food Sci Nutr 62: 755–764. https://doi.org/10.3109/09637486.2011.581224 doi: 10.3109/09637486.2011.581224
![]() |
[35] |
Chire GC, Verona PA, Guzmán JH (2016) Color changes during post-harvest of Peruvian cocoa beans from Piura. Ciencia e Investigación 19: 29–34. https://doi.org/10.15381/ci.v19i1.13625 doi: 10.15381/ci.v19i1.13625
![]() |
[36] |
Efraim P, Pezoa-Garcı́a N, Jardim D, et al. (2010) Influence of cocoa beans fermentation and drying on the polyphenol content and sensory acceptance. Food Sci Technol 30: 142–150. https://doi.org/10.1590/S0101-20612010000500022 doi: 10.1590/S0101-20612010000500022
![]() |
1. | Aziz Belmiloudi, Cardiac memory phenomenon, time-fractional order nonlinear system and bidomain-torso type model in electrocardiology, 2021, 6, 2473-6988, 821, 10.3934/math.2021050 |