Next Article in Journal
On the Normalized Laplacian Spectrum of the Zero-Divisor Graph of the Commutative Ring Zp1T1p2T2
Next Article in Special Issue
Optimal Control Problems for Erlang Loss Systems
Previous Article in Journal
Parameters Determination via Fuzzy Inference Systems for the Logistic Populations Growth Model
Previous Article in Special Issue
About Stabilization of the Controlled Inverted Pendulum Under Stochastic Perturbations of the Type of Poisson’s Jumps
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bilinear Optimal Control for a Nonlinear Parabolic Equation Involving Nonlocal-in-Time Term

Laboratoire L.A.M.I.A., Département de Mathématiques et Informatique, Université des Antilles, Campus Fouillole, 97159 Pointe-à-Pitre, France
*
Author to whom correspondence should be addressed.
Axioms 2025, 14(1), 38; https://doi.org/10.3390/axioms14010038
Submission received: 20 November 2024 / Revised: 24 December 2024 / Accepted: 30 December 2024 / Published: 4 January 2025
(This article belongs to the Special Issue Advances in Mathematical Optimal Control and Applications)

Abstract

:
We study a bilinear optimal control problem for an evolution equation with a nonlinear term that depends on both the state and its time integral. First, we establish existence and uniqueness results for this evolution equation. Then, we derive weak maximum principle results to improve the regularity of the state equation. We proceed by formulating an optimal control problem aimed at steering the system’s state to a desired final state. Finally, we demonstrate that this optimal control problem admits a solution and derive the first- and second-order optimality conditions.

1. Introduction

We denote by Ω a bounded open set of R d , d = 2 , 3 with a Lipschitz boundary Γ . Also, let ω be an open subset of Ω . For any T > 0 , we set Q = Ω × ( 0 , T ) , Σ = Γ × ( 0 , T ) and ω T = ω × ( 0 , T ) . Then, we turn our attention to the optimal control problem:
min J ( y , v ) ,
subject to the constraints that y solves the nonlinear and nonlocal-in-time parabolic
y t Δ y + β 0 t g ( s ) y ( x , s ) d s y = v χ ω y in Q = Ω × ( 0 , T ) , y ν = 0 on Σ = Γ × ( 0 , T ) , y ( · , 0 ) = y 0 in Ω ,
where
J ( y , v ) = 1 2 y ( · , T ) y d L 2 ( Ω ) 2 + N 2 v L 2 ( ω T ) 2 ,
with N > 0 and y d L ( Ω ) are given, and the set of admissible controls
U a d = v L ( ω T ) : v a v v b ,
with v b , v b R , v b > v a . The function y 0 L ( Ω ) , g L ( 0 , T ) , β C 2 ( R ) L ( R ) and χ ω denote the characteristic function of the control set ω . We assume that there exists g > 0 such that
g L ( 0 , T ) g
and the function β C 2 ( R ) , and there exists M > 0 such that
1 β ( s ) M for all s R
and
| β ( s ) | + | β ( s ) | M for all s R ,
Equation (2) can be viewed as a model for population dynamics, where y ( x , t ) represents the population density at a specific location x and time t. The real function 0 t g ( s ) y ( x , s ) d s represents at any given time t and, at the location x, the aggregated history of a population’s density weighted by factors such as pathogen exposure, stress or other time-dependent influences. This dynamic approach allows a wide range of biological processes to be captured, including pathogen dynamics, resource competition and adaptive immunity or resistance. The real function β 0 t g ( s ) y ( x , s ) d s models at any given time t and at the location x the mortality rate as a response to cumulative environmental, physiological or population-related factors. As a rate function, β is positive and naturally bounded. We refer to [1], where the proof that β can be lower bounded, i.e., 1 β ( s ) , s R is given. Mathematically, the assumption (6)–(7) on β ensure the well-posedness of the system, the boundedness of solutions, and the characterization of the local optimal control.
In the literature, a system similar to (2), but with homogeneous Dirichlet boundary conditions and a second member for the state equation living in L 2 ( Q ) , has been studied. In [1], the author proved the existence and uniqueness of the weak solution and analyzed its asymptotic behavior. The establishment of the existence of a weak solution relies on the use of a fixed-point argument and Lebesgue’s dominated convergence theorem. For uniqueness, the standard approach of considering two solutions satisfying the same system and then proving that they are equal by using a subtraction technique that leads to a differential inequality, which forces the difference between the two solutions to be zero, has been applied. Another similar nonlocal-in-time parabolic problem, but with a different nonlinearity and homogeneous Dirichlet boundary conditions, has been studied in [2]. The nonlocal term involves an integral over the entire time interval, making the problem non-evolutionary in the traditional sense because the solution at any time depends on future values. The author proved the existence of a weak solution as well as its uniqueness. For the existence, the author built a weakly continuous map and applied Tikhonov’s fixed-point theorem. Using the standard method, the author proved the uniqueness of the solution under the conditions that the derivative of the nonlocal-in-time term is bounded, and the time interval T is sufficiently small. From the same author, Ref. [3] presents a similar system but with a slightly different nonlocal-in-time nonlinear term. The difference lies in the fact that the integral over the entire time interval is weighted. The author proved the existence of a weak solution without the use of any continuity properties of the solution in time. He has two existence results, one where the nonlinearity is a continuous bounded function and the other where the nonlinear term is a non-negative function that is not necessarily bounded. The demonstrations involve the use of estimates and weak convergences to show that a sequence of approximate solutions converges to a weak solution.
Optimal control of nonlinear partial differential Equations (PDEs) has been a subject of intensive works and applications in various fields. We refer, for instance, to [4,5,6,7,8,9,10,11] and the references therein. It is well established that the first-order optimality condition is both necessary and sufficient for studying optimal control problems associated with linear partial differential equations. However, in the case of nonlinear partial differential equations, the cost function is no longer convex. As a result, while the first-order optimality condition remains necessary, it is not sufficient. To address this, the second-order optimality condition is required, which involves analyzing the second derivative of the cost function with respect to the control. This ensures the local uniqueness of the optimal control. In this paper, we study an optimal control problem for a time-nonlocal parabolic equation. We start by establishing some regularity results for both the state equation and the cost function. Then, we prove that our optimal control problem as the local optimal solution that we characterize by means of the necessary and sufficient first-order and second-order optimality conditions.
The rest of this paper is organized as follows. Section 2 is devoted to the general framework and preliminary results. In Section 3, we establish a maximum principle and study the existence and uniqueness of a generalized nonlocal-in-time problem. Then, we deduce similar properties for system (2). In Section 4, we study the Lipschitz continuity and the differentiability of the solution to (2), then we prove existence of a local optimal control to problem (1) and characterize it using first- and second-order optimality conditions. Some concluding remarks are also given in Section 5.

2. General Framework and Preliminary Results

We give the general setting and some results that will be useful for the study of the control problem (1)–(2).
We recall that the Sobolev Space H 1 ( Ω ) ) is defined as
H 1 ( Ω ) ) = ρ L 2 ( Ω ) , ρ x i L 2 ( Ω ) , i = 1 , , d .
Equipped with the norm
ρ H 1 ( Ω ) = Ω ρ 2 ( x ) d x + i = 1 d Ω ρ x i ( x ) 2 d x 1 / 2 ,
the space H 1 ( Ω ) ) is a Hilbert space. We denote by ( H 1 ( Ω ) ) the dual of H 1 ( Ω ) ) . Next, we introduce the space
W ( 0 , T ; H 1 ( Ω ) ) = ρ L 2 ( 0 , T ; H 1 ( Ω ) ) ; ρ t L 2 0 , T ; ( H 1 ( Ω ) ) .
Endowed with the norm
ψ W ( 0 , T ; H 1 ( Ω ) ) = ψ L 2 ( 0 , T ; H 1 ( Ω ) ) 2 + ψ t L 2 ( 0 , T ; ( H 1 ( Ω ) ) ) 2 1 / 2 ,
the space W ( 0 , T ; H 1 ( Ω ) ) is a Hilbert space. Furthermore, we have from ([12], Theorem 1.1, p. 102) that the following embedding is continuous:
W ( 0 , T ; H 1 ( Ω ) ) C ( [ 0 , T ] ; L 2 ( Ω ) ) .
Also, using ([13] Theorem 16.1, p. 110) and ([14], Theorem 5.1, p. 58), we obtain the following compact embedding:
W ( 0 , T ; H 1 ( Ω ) ) L 2 ( 0 , T ; H 1 2 ( Ω ) ) .
The following results are essential for proving the existence of a solution to the nonlinear Equation (2), as well as the existence of a solution to the optimization problem discussed in Section 4.
Lemma 1.
Let the sequence ( y n ) be such that
y n W ( 0 , T ; H 1 ( Ω ) ) C 2
for some constant C 2 > 0 independent of n . Then, there exists y W ( 0 , T ; H 1 ( Ω ) ) such that
y n y weakly in W ( 0 , T ; H 1 ( Ω ) ) ,
y n y weakly in L 2 ( 0 , T ; H 1 ( Ω ) ) ,
y n y strongly in L 2 ( Q ) .
Proof. 
We then deduce from (11) that there exists y W ( 0 , T ; H 1 ( Ω ) ) such that
y n y weakly in W ( 0 , T ; H 1 ( Ω ) ) .
Hence, from the compact embedding (10), we obtain that
y n y strongly in L 2 ( 0 , T ; H 1 2 ( Ω ) ) ,
from which, thanks to the continuous embedding of L 2 ( 0 , T ; H 1 2 ( Ω ) ) into L 2 ( Q ) , we have that
y n y strongly in L 2 ( Q ) .
From now on, we adopt the following notation for any appropriate function ϕ :
h ϕ g ( x , t ) = 0 t g ( s ) ϕ ( x , s ) d s for a . e . ( x , t ) Q , h ϕ ϑ ( x , t ) = T t T ϑ ( x , s ) ϕ ( x , s ) d s for a . e . ( x , t ) Q ,
where g L ( 0 , T ) and ϑ L ( Q ) . We assume that there exists ϑ > 0 such that
ϑ L ( Q ) ϑ .
Lemma 2.
Let g L ( 0 , T ) and ϑ L ( Q ) be such that (5) and (16) hold. Also, let ϕ L ( Q ) . Then,
h ϕ g L ( Q ) g T ϕ L ( Q )
and
h ϕ ϑ L ( Q ) ϑ T ϕ L ( Q ) ,
where h ϕ g and h ϕ ϑ are defined as in (15).
Proof. 
Using (5), (15), the Cauchy–Schwarz inequality and the fact that ϕ L ( Q ) , we easily have (17) and (18). □
Lemma 3.
Let the functions g L 2 ( 0 , T ) , β C 2 ( R ) and ϑ L ( Q ) be such that (5), (6), (7) and (16) hold. Also, let ϕ L 2 ( Q ) and ( z n ) be sequence of L 2 ( Q ) such that
z n z strongly in L 2 ( Q ) .
Then, we have the following estimation
h ϕ g L 2 ( Q ) g T ϕ L 2 ( Q )
and
h ϕ ϑ L 2 ( Q ) ϑ T ϕ L 2 ( Q ) ,
where h ϕ g and h ϕ ϑ are defined as in (15). Moreover, as n ,
h z n g h z g strongly in L 2 ( Q ) ,
h z n ϑ h z ϑ strongly in L 2 ( Q ) ,
β ( l ) ( h z n g ) β ( l ) ( h z g ) strongly in L 2 ( Q ) , k = 0 , 1 , 2 ,
where β ( l ) stands for the derivative of order l = 0 , 1 , 2 of the function β with β ( 0 ) = β .
Proof. 
Using (15) and the Cauchy–Schwarz inequality, we have that
h ϕ g L 2 ( Q ) 2 = 0 T Ω 0 t g ( s ) ϕ ( x , s ) d s 2 d x d t T 2 g 2 ϕ L 2 ( Q ) 2 ,
from which we deduce (20). Proceeding exactly as above while using (16), we deduce (21).
From the linearity of the integral, (20) and (21), we can write
h z n g h z g L 2 ( Q ) 2 = 0 T Ω 0 t g ( s ) z n ( x , s ) z ( x , s ) d s 2 d x d t T 2 g 2 z n z L 2 ( Q ) 2
and
h z n ϑ h z ϑ L 2 ( Q ) 2 = 0 T Ω T t T ϑ ( x , s ) z n ( x , s ) z ( x , s ) d s 2 d x d t T 2 ϑ 2 z n z L 2 ( Q ) 2 .
Passing to the limit when n in these two inequalities while using (19), we obtain (22) and (23). To prove (24), we use the Lebesgue theorem. Indeed, in view of (22), we can extract a subsequence of h z n g n still denoted h z n g n such that
h z n g ( x , t ) h z g ( x , t ) for a . e . ( x , t ) Q , as n + ,
and since β C 2 ( R ) , we have that
β ( l ) h z n g ( x , t ) β ( l ) h z g ( x , t ) for a . e . ( x , t ) Q , as n + .
Therefore, using (6) and (7), we have that there exists M > 0 such that
β ( l ) h z n g ( x , t ) M for a . e . ( x , t ) Q .
Since Q is bounded domain, it follows from the Lebesgue dominated convergence theorem that
β ( l ) h z n g β ( l ) h z g strongly in L 2 ( Q ) , l = 0 , 1 , 2 as n + .

3. Analysis of the State Equation

In this section, we establish maximum principle and study the existence and uniqueness of a generalized nonlocal-in-time problem. Actually, the study of this auxiliary system allow us obtain similar properties for system (2) which also includes related systems such as its adjoint model. More precisely, we consider the following system:
y t Δ y + b 0 h y g + g ( t ) h y ϑ + μ β ( h y g ) y + a 0 y = v χ ω y + f in Q , y ν = 0 on Σ , y ( · , 0 ) = y 0 in Ω ,
where μ 0 , v χ ω L ( Q ) , f L 2 ( Q ) and y 0 L 2 ( Ω ) . The functions g L ( 0 , T ) , b 0 L ( Q ) , ϑ L ( Q ) and a 0 L ( Q ) . The function h y ϑ and h y g are given by notation (15). We assume that (16) holds and that there exist b > 0 and a > 0 such that
b 0 L ( Q ) b ,
and
a 0 L ( Q ) a .
Before going further, we give the definition of a weak solution to system (25).
Definition 1.
Assume that β C 2 ( R ) satisfies (6)–(7). Let μ 0 , v L ( ω T ) , g L ( Q ) ,   f L 2 ( Q ) , b 0 L ( Q ) ,   ϑ L ( Q ) ,   a 0 L ( Q ) , and y 0 L 2 ( Ω ) . We say that a function y L 2 ( 0 , T ; H 1 ( Ω ) ) is a weak solution to (25) if
Ω y 0 ϕ ( · , 0 ) d x + Q f ϕ d x d t = 0 T ϕ t , y ( H 1 ( Ω ) ) , H 1 ( Ω ) d t + Q a 0 y ϕ d x d t + Q ϕ · y d x d t + μ Q β ( h y g ) y ϕ d x d t ω T v y ϕ d x d t + Q b 0 h y g ϕ d x d t + Q g ( t ) h y ϑ ϕ d x d t ,
holds for every ϕ H ( Q ) : = { φ W ( 0 , T ; H 1 ( Ω ) ) such that ϕ ( · , T ) = 0 } .

3.1. Maximum Principle

In this subsection, we will prove some estimates in the space L ( Q ) . One necessary result is the one directly obtained from the Gagliardo–Niremberg Theorem (see, e.g., [15,16]) which we recall here.
Lemma 4.
Let y W ( 0 , T ; H 1 ( Ω ) ) . Then, y L μ ( Q ) , with μ = 2 ( d + 2 ) d . In addition, there exists a constant C = C ( d ) such that the following estimate
Q | y | μ d x d t C y L ( 0 , T ; L 2 ( Ω ) ) 4 d | y L 2 ( 0 , T ; H 1 ( Ω ) ) 2
holds.
Another preliminary result in proving the maximum principle is the following theorem.
Theorem 1
([17], Lemma 4.1.1). We consider ρ a non-negative and non-increasing function on [ k 0 , + ) such that
ρ ( m ) C m k θ ( ρ ( k ) ) η , m > k > k 0 ,
where C, θ ,   η are positive constants with η > 1 . Then, ρ ( m ) = 0 for all m k 0 + ϖ , where ϖ : = C 2 η η 1 ( ρ ( k 0 ) ) η 1 θ .
Theorem 2.
Let μ 0 ,   v L ( ω T ) , f L ( Q ) and y 0 L ( Ω ) . Also, let g L ( 0 , T ) , b 0 L ( Q ) ,   ϑ L ( Q ) ,   a 0 L ( Q ) and β C 2 ( R ) be such that (5), (26), (16), (27), and (6) hold. Assume that y W ( 0 , T , H 1 ( Ω ) ) is a weak solution to (25). Then, we have that y L ( Q ) . Moreover, there exists C ( d , T , Ω , a , b , v L ( ω T ) , ϑ , g ) > 0 such that
y L ( Q ) C f L ( Q ) + y 0 L ( Ω ) .
Proof. 
First we need to make the following change for variable: p = e r t y where y W ( 0 , T ; H 1 ( Ω ) ) is a solution to (25) and r > 0 . It follows that p W ( 0 , T ; H 1 ( Ω ) ) is a weak solution to
p t Δ p = ( r + a 0 ) p μ β ( h y g ) p e r t b 0 h e r ( · ) p g e r t g ( t ) h e r ( · ) p ϑ + v χ ω p + e r t f in Q , p ν = 0 on Σ , p ( · , 0 ) = y 0 in Ω ,
where in view of (15), h e r ( · ) p g ( x , t ) = 0 t e r s g ( s ) p ( x , s ) d s and h e r ( · ) p ϑ ( x , t ) = T t T e r s b 0 ( x , s ) p ( x , s ) d s , for a.e. ( x , t ) Q .
Next, we introduce a real number κ such that κ > y 0 L ( Ω ) . We easily deduce that for almost every t ( 0 , T ) , p ( t ) κ H 1 ( Ω ) , and then we set p + ( t ) : = max { p ( t ) κ , 0 } which also resides in H 1 ( Ω ) . Multiplying the first equation of system (25) by φ H 1 ( Ω ) and then integrating by parts over Ω , we obtain the following:
Ω e r t f ( t ) φ ( t ) d x = p t ( t ) , φ ( t ) ( H 1 ( Ω ) ) , H 1 ( Ω ) + Ω p ( t ) · φ ( t ) d x + Ω ( a 0 + r ) p ( t ) φ ( t ) d x + μ Ω β h y g ( t ) p ( t ) φ ( t ) d x ω v ( t ) p ( t ) φ ( t ) d x + Ω e r t b 0 0 t e r s g ( s ) p ( x , s ) d s φ d x + Ω e r t g ( t ) T t T e r s ϑ ( x , s ) p ( x , s ) d s φ d x .
Hence,
Ω e r t f ( t ) φ ( t ) d x = t ( p κ ) ( t ) , φ ( t ) ( H 1 ( Ω ) ) , H 1 ( Ω ) + Ω ( p κ ) ( t ) · φ ( t ) d x + μ Ω β h y g ( t ) ( p κ ) ( t ) φ d x + κ μ Ω β h y g ( t ) φ ( t ) d x + Ω ( a 0 + r ) ( p κ ) ( t ) φ ( t ) d x + κ Ω e r t g ( t ) T t T e r s ϑ ( x , s ) d s φ d x + Ω e r t b 0 0 t e r s g ( s ) ( p κ ) ( x , s ) d s φ ( t ) d x + κ Ω e r t b 0 0 t e r s g ( s ) d s φ ( t ) d x + κ Ω ( a 0 + r ) φ ( t ) d x ω v ( t ) ( p κ ) ( t ) φ ( t ) d x κ ω v ( t ) φ ( t ) d x + Ω e r t b 0 T t T e r s g ( s ) ( p κ ) ( x , s ) d s φ d x .
Taking φ = p + in this latter identity yields
Ω e r t f ( t ) p + ( t ) d x = 1 2 d d t p + ( t ) L 2 ( Ω ) 2 + μ Ω β h y g ( t ) ( p + ( t ) ) 2 d x + μ κ Ω β h y g ( t ) p + ( t ) d x + Ω | p + ( t ) | 2 d x + r p + ( t ) L 2 ( Ω ) 2 + Ω a 0 ( p + ( t ) ) 2 d x + Ω e r t b 0 0 t e r s g ( s ) ( p κ ) ( x , s ) d s p + ( t ) d x + κ Ω e r t b 0 0 t e r s g ( s ) d s p + ( t ) d x + κ Ω ( a 0 + r ) p + ( t ) d x ω v ( t ) ( p + ( t ) ) 2 d x κ ω v ( t ) p + ( t ) d x + Ω e r t b 0 T t T e r s g ( s ) ( p κ ) ( x , s ) d s p + ( t ) d x + κ Ω e r t g ( t ) T t T e r s ϑ ( x , s ) d s p + ( t ) d x ,
and from (5), (6), (26), and (27) we obtain that for a.e. t ( 0 , T ) ,
1 2 d d t p + ( t ) L 2 ( Ω ) 2 + Ω | p + ( t ) | 2 d x + r p + ( t ) L 2 ( Ω ) 2 a p + ( t ) L 2 ( Ω ) 2 + b g T 1 / 2 p + ( t ) L 2 ( Ω ) Ω 0 t ( p + ( s ) ) 2 d s d x 1 / 2 + ϑ g T 1 / 2 p + ( t ) L 2 ( Ω ) Ω T t T ( p + ( s ) ) 2 d s d x 1 / 2 + Ω | f ( t ) | | p + ( t ) | d x + a κ Ω p + ( t ) d x + v L ( ω T ) p + ( t ) L 2 ( Ω ) 2 + κ v L ( ω T ) Ω | p + ( t ) | d x + κ ( ϑ + b ) g T Ω p + ( t ) d x r κ Ω p + ( t ) d x .
Now, observing that p + ( 0 ) : = max ( y ( 0 ) k , 0 ) = 0 and integrating (34) on ( 0 , s ) , with s [ 0 , T ] and then using Cauchy–Schwarz’s inequality, we obtain that
1 2 p + ( s ) L 2 ( Ω ) 2 + 0 s Ω | p + ( t ) | 2 d x d t + r 0 s p + ( t ) L 2 ( Ω ) 2 d t r κ 0 s Ω p + ( t ) d x d t + b g T 0 s p + ( t ) L 2 ( Ω ) 2 d t + ϑ g T 0 s p + ( t ) L 2 ( Ω ) 2 d t + a 0 s p + ( t ) L 2 ( Ω ) 2 d t + 0 s Ω | f ( t ) | | p + ( t ) | d x d t + v L ( ω T ) 0 s p + ( t ) L 2 ( Ω ) 2 d t + κ v L ( ω T ) 0 s Ω | p + ( t ) | d x d t + κ ( ϑ + b ) g T 0 s Ω p + ( t ) d x d t + a κ 0 s Ω p + ( t ) d x d t +
and then choosing r = 1 + ( ϑ + b ) g T + a + v L ( ω T ) , we finally obtain
1 2 p + ( s ) L 2 ( Ω ) 2 + 0 s p + ( t ) H 1 ( Ω ) 2 d t Q | f | | p + | d x d t .
Hence,
sup t [ 0 , T ] p + ( t ) L 2 ( Ω ) 2 2 Q | f | | p + | d x d t ,
p + L 2 ( 0 , T ; H 1 ( Ω ) ) 2 Q | f | | p + | d x d t ,
and it follows from (36) that
p + L ( 0 , T ; L 2 ( Ω ) ) 2 2 Q | f | | p + | d x d t .
Now, using Lemma 4, we obtain that
Q | p + | 2 ( d + 2 ) d d x d t d d + 2 C ( d ) p + L ( 0 , T ; L 2 ( Ω ) ) 4 d p + L 2 ( 0 , T ; H 1 ( Ω ) ) 2 d d + 2
for some C ( d ) > 0 . Observing that
p + L ( 0 , T ; L 2 ( Ω ) ) 4 d p + L 2 ( ( 0 , T ; H 1 ( Ω ) ) 2 d d + 2 = p + L ( 0 , T ; L 2 ( Ω ) ) 4 d + 2 p + L 2 ( 0 , T ; H 1 ( Ω ) ) 2 d d + 2
and using the Young’s inequality with p = d + 2 2 and q = d + 2 d and then (39), we have that
Q | p + | 2 ( d + 2 ) d d x d t d d + 2 C ( d ) p + L ( 0 , T ; L 2 ( Ω ) ) 2 + p + L 2 ( 0 , T ; H 1 ( Ω ) ) 2 .
Therefore, it follows from (37) and (38) that
Q | p + | 2 ( d + 2 ) d d x d t d d + 2 3 C ( d ) Q | f | | p + | d x d t .
We set B κ : = { ( x , t ) Q , p ( x , t ) > κ } . Then, B κ is a measurable set. Moreover, we deduce from (41) that
B κ | p + | 2 ( d + 2 ) d d x d t d d + 2 3 C ( d ) B κ | f | | p + | d x d t
Using Hölder’s inequality with p = 2 ( d + 2 ) d + 4 and q = 2 ( d + 2 ) d in the right hand side of (42), we obtain
B κ | p + | 2 ( d + 2 ) d d x d t 3 C ( d ) d + 2 d B κ | f | 2 ( d + 2 ) d + 4 d x d t d + 4 2 d B κ | p + | 2 ( d + 2 ) d d x d t 1 2 ,
from which we deduce that
B κ | p + | 2 ( d + 2 ) d d x d t 1 2 3 C ( d ) d + 2 d B κ | f | 2 ( d + 2 ) d + 4 d x d t d + 4 2 d 3 C ( d ) d + 2 d f | L ( Q ) d + 2 d | B κ | d + 4 2 d ,
where | B κ | denotes the measure of the set B κ .
Let m > κ , then B m B κ . Consequently, for a.e. ( x , t ) B m , we have p ( x , t ) k > m κ and then, p κ = p + .
Thus,
B κ | p + | 2 ( d + 2 ) d d x d t B m | p + | 2 ( d + 2 ) d d x d t = B m | y k | 2 ( d + 2 ) d d x d t B m ( m k ) 2 ( d + 2 ) d d x d t = ( m k ) 2 ( d + 2 ) d | B m | .
Then, we obtain that
| B m | 1 2 1 ( m k ) ( n + 2 ) n B κ | p + | 2 ( d + 2 ) d d x d t 1 2 3 C ( d ) m k d + 2 d f L ( Q ) d + 2 d | B κ | d + 4 2 d
Hence,
| B m | 3 C ( d ) m k 2 ( d + 2 ) d f L ( Q ) 2 ( d + 2 ) d | B κ | d + 4 d .
Observing that B m B κ for any m > k , we have that | B m | | B κ | . Hence, applying Theorem 1 with φ ( k ) = | B κ | ,   k 0 = y 0 L ( Ω ) and ϖ = C ( d , T , Ω ) f L ( Q ) , we deduce that | B m | = 0 where m = C ( d , T , Ω ) f L ( Q ) + y 0 L ( Ω ) . Therefore, we have
p ( x , t ) C ( d , T , Ω ) f L ( Q ) + y 0 L ( Ω ) ( x , t ) Q .
We prove by setting p = min ( p + k , 0 ) and B ˜ κ = { ( x , t ) Q : p ( x , t ) < k } that the set B ˜ m , where p < C ( d , T , Ω ) f L ( Q ) + y 0 L ( Ω ) , is a measure zero set. Consequently,
p L ( Q ) C ( d , T , Ω ) f L ( Q ) + y 0 L ( Ω ) .
Now, since p = e ( a + ( ϑ + b ) g T + v L ( ω T ) + 1 ) t y , we deduce that
y L ( Q ) C ( d , T , Ω ) e ( a + ( ϑ + b ) g T + v L ( ω T ) + 1 ) T f L ( Q ) + y 0 L ( Ω ) .
From this Theorem based on the generalized model (25), we can derive similar properties for system (2).
Corollary 1.
Let v L ( ω T ) and y 0 L ( Ω ) . Also, let g L ( 0 , T ) and β C 2 ( R ) be such that (5) and (6) hold. Assume that system (2) has a weak solution y W ( 0 , T , H 1 ( Ω ) ) . Then, y L ( Q ) and there exists a constant C ( d , T , Ω ) > 0 such that
y L ( Q ) C ( d , T , Ω ) e ( v L ( Q ) + 1 ) T y 0 L ( Ω ) .
Proof. 
Proceeding as in Theorem 2 but with a 0 0 , ϑ 0 , b 0 0 and f 0 , we easily prove that y L ( Q ) and that estimate (45) holds. □
In what following, we need the following result.
Lemma 5.
Let ( ψ n ) n be sequence of L ( Q ) . Assume that there exists ψ > 0 and ψ L 2 ( Q ) such that
ψ n L ( Q ) ψ
and
ψ n ψ strongly in L 2 ( Q ) , as n + .
Then, for any ζ L 2 ( Q ) ,
ψ n ζ ψ ζ strongly in L 2 ( Q ) , as n + .
Proof. 
Since ψ n ψ strongly in L 2 ( Q ) , a s n + , we can extract a subsequence of ( ψ n ) n still denoted ( ψ n ) n such that
ψ n ( x , t ) ψ ( x , t ) for a . e . ( x , t ) Q , as n + .
Therefore, for any ζ L 2 ( Q ) , we obtain that
ψ n ( x , t ) ζ ( x , t ) ψ ( x , t ) ζ ( x , t ) for a . e . ( x , t ) Q , as n +
and as there exists ψ > 0 such that
ζ ( x , t ) ψ n ( x , t ) ψ ζ ( x , t ) for a . e . ( x , t ) Q .
Relation (46) follows from the Lebesgue’s dominated convergence theorem. □

3.2. Existence and Uniqueness Results

Theorem 3.
Let μ 0 , v L ( ω T ) , f L ( Q ) and y 0 L ( Ω ) . Also, let g L ( 0 , T ) , b 0 L ( Q ) ,   ϑ L ( Q ) ,   a 0 L ( Q ) , and β C 2 ( R ) be such that (5), (26), (16), (27), (6) and (7) hold. Then, there exists a unique weak solution y W ( 0 , T ; H 1 ( Ω ) ) to (25) in the sense of Definition 1.
Moreover, there exist C M , g , b , a , v L ( ω T ) , Γ , T > 0 and C ( a , b , g , v L ( ω T ) , T ) > 0 depending continuously on v L ( ω T ) such that
y L 2 ( 0 , T ; H 1 ( Ω ) ) C ( a , ϑ , b , g , v L ( ω T ) , T ) f L 2 ( Q ) + y 0 L 2 ( Ω ) ,
and
y W ( 0 , T ; H 1 ( Ω ) ) C M , g , b , a , v L ( ω T ) , ϑ , T f L 2 ( Q ) + y 0 L 2 ( Ω ) .
Proof. 
We proceed in three steps.
Step 1. We prove the existence results and estimations (48).
We use the Schauder’s fixed-point theorem. So, Let B be a non-empty, closed and convex subset of L 2 ( Q ) given by
B : = z L 2 ( Q ) , z L 2 ( Q ) C 1 ,
where
C 1 : = C ( a , b , g , v L ( ω T ) , T ) f L 2 ( Q ) + y 0 L 2 ( Ω ) .
Let z B . In view of Theorem A1, we have constructed the following nonlinear map
F : B B
where F ( z ) = y ( z ) . Here, y ( z ) W ( 0 , T ; H 1 ( Ω ) ) is the weak solution to (A1):
y t Δ y + b 0 h y g + g ( t ) h y ϑ + μ β h z y + a 0 y = v χ ω y + f in Q , y ν = 0 on Σ , y ( · , 0 ) = y 0 in Ω ,
Proving that F has a fixed point will be enough to prove that (25) has a solution in L 2 ( Q ) . To this end, we use the Schauder’s fixed-point theorem.
Let z B . Since F ( z ) = y ( z ) , where y ( z ) is the unique weak solution to (A1), using Corollary A1, we have that there exist C ( a , b , g , v L ( ω T ) , T ) > 0 and C M , g , b , a , v L ( ω T ) , Γ , T > 0 such that
y L 2 ( 0 , T ; H 1 ( Ω ) ) C ( a , ϑ , b , g , v L ( ω T ) , T ) f L 2 ( Q ) + y 0 L 2 ( Ω ) ,
and
y W ( 0 , T ; H 1 ( Ω ) ) C M , μ , g , ϑ , b , a , v L ( ω T ) , T f L 2 ( Q ) + y 0 L 2 ( Ω ) .
From (51), we deduce that F ( B ) B , and from (52), we have that F ( B ) is relatively compact in B because W ( 0 , T ; H 1 ( Ω ) ) is relatively compact in L 2 ( Q ) . It remains to be proven that F is continuous on B to obtain that F a fixed point. So, let z k be a sequence of B such that
z k z strongly in L 2 ( Q ) , as k + .
Let y k = F ( z k ) , the solution to (A1) with z = z k :
y k t Δ y k + b 0 h y k g + g ( t ) h y k ϑ + μ β h z k g y k + a 0 y k = v χ ω y k + f in Q , y k ν = 0 on Σ , y k ( · , 0 ) = y 0 in Ω ,
Then, from Corollary A1, we have that y k verifies
Ω y 0 ϕ ( · , 0 ) d x = 0 T ϕ t , y k ( H 1 ( Ω ) ) , H 1 ( Ω ) d t ω T v y k ϕ d x d t + Q ϕ · y k d x d t + μ Q β ( h z k g ) y k ϕ d x d t + Q a 0 y k ϕ d x d t + Q b 0 h y k g ϕ d x d t + Q g ( t ) h y k ϑ ϕ d x d t Q f ϕ d x d t ,
and there exist positive constants C 1 = C ( a , ϑ , b , g , v L ( ω T ) , T ) and C 2 = C M , μ , g , b , ϑ , a , v L ( ω T ) T such that
y k L 2 ( 0 , T ; H 1 ( Ω ) ) C 1 f L 2 ( Q ) + y 0 L 2 ( Ω ) ,
and
y k W ( 0 , T ; H 1 ( Ω ) ) C 2 f L 2 ( Q ) + y 0 L 2 ( Ω ) .
From (56), we have that there exists y W ( 0 , T ; H 1 ( Ω ) ) and a subsequence of ( y k ) k still denoted by ( y k ) k such that
y k y weakly in L 2 ( 0 , T ; H 1 ( Ω ) ) as k + ,
and
y k t y t weakly in L 2 0 , T ; H 1 ( Ω ) as k + .
Using Lemma 1 with y n = y k , we deduce from (56) that
y k y strongly in L 2 ( Q ) as k + .
Thanks to Lemma 3, we have from (22) and (23) that
h z k g h z g strongly in L 2 ( Q ) as k + ,
h y k g h y g strongly in L 2 ( Q ) as k +
and
h y k ϑ h y ϑ strongly in L 2 ( Q ) as k + .
Applying Lemma 3 again, we have from (24) with l = 0 that
β h z k g β h z g strongly in L 2 ( Q ) as k + .
From (6), (7) and Lemma 5, we obtain for any ϕ H ( Q ) that
ϕ β h z k g ϕ β h z g strongly in L 2 ( Q ) , as k + .
Passing to the limit in (54) when k + while using (5), (16), (27), (26), (57), (59), (61), (62), (63), and (64), we obtain that
Ω y 0 ϕ ( · , 0 ) d x + Q f ϕ d x d t = 0 T ϕ t , y ( H 1 ( Ω ) ) , H 1 ( Ω ) d t ω T v y ϕ d x d t + Q ϕ · y d x d t + μ Q β ( h z g ) y ϕ d x d t + Q a 0 y ϕ d x d t + Q b 0 h y g ϕ d x d t + Q g ( t ) h y ϑ ϕ d x d t .
Consequently, y is a weak solution to (A1). The uniqueness of the solution to (A1) allows us to conclude that F ( z k ) = y k converges strongly to F ( z ) = y in L 2 ( Q ) . This means that F is continuous on B.
Step 2. We prove estimate (48)
If we set z = e r t y with r = a + ( ϑ + b ) g T + v L ( ω T ) + 1 , where y is a solution to (25), we obtain that z W ( 0 , T ; H 1 ( Ω ) ) is a solution to
z t Δ z + μ β ( h y g ) z = e r t b 0 h e r ( · ) z g e r t g ( t ) h e r ( · ) z ϑ ( a 0 + r ) z + v χ ω z + e r t f in Q , z ν = 0 on Σ , z ( · , 0 ) = y 0 in Ω ,
where in view of notation (15), h e r ( · ) z g ( x , t ) = 0 t e r s g ( s ) z ( x , s ) d s and h e r ( · ) z ϑ ( x , t ) = T t T e r s ϑ ( x , s ) z ( x , s ) d s , for a.e. ( x , t ) Q .
Then, proceeding exactly as for the computation of (A11) and (A13), we prove that there there exist C ( a , ϑ , b , g , v L ( ω T ) , T ) > 0 and C M , μ , g , b , ϑ , a , v L ( ω T ) , T > 0 such that
y L 2 ( 0 , T ; H 1 ( Ω ) ) C ( a , ϑ , b , g , v L ( ω T ) , T ) f L 2 ( Q ) + y 0 L 2 ( Ω ) .
y W ( 0 , T ; H 1 ( Ω ) ) C M , μ , g , b , ϑ , a , v L ( ω T ) , T f L 2 ( Q ) + y 0 L 2 ( Ω ) ,
Step 3. We show the uniqueness of the solution.
Let y 1 , y 2 be two solutions to (25) with the same source term f and initial value y 0 . If we set z = e r t ( y 1 y 2 ) with r > 0 , we obtain that z W ( 0 , T ; H 1 ( Ω ) ) is a solution to
z t Δ z + η ( y 1 , y 2 , z ) + ( β ( h y 1 g ) + a 0 + r ) z = v χ ω z in Q , z ν = 0 on Σ , z ( · , 0 ) = 0 in Ω ,
where
η ( y 1 , y 2 , z ) = μ β ( h y 1 g ) z + μ e r t β ( h y 1 g ) β ( h y 2 g ) y 2 + e r t b 0 h e r 1 ( · ) z g + e r t g ( t ) h e r ( · ) z ϑ .
If we multiply the first equation of (67) by z and integrate over Q, we obtain
1 2 z ( · , T ) L 2 ( Ω ) 2 + z L 2 ( Q ) 2 + Q β ( h y 1 g ) z 2 d x d t + r z L 2 ( Q ) 2 = μ Q e r t β ( h y 1 g ) β ( h y 2 g ) y 2 z d x d t Q e r t b 0 0 t e r 1 s g ( s ) z ( x , s ) d s z d x d t Q e r t g ( t ) T t T e r 1 s ϑ ( x , s ) z ( x , s ) d s z d x d t Q a 0 z 2 d x d t + ω T v s . z 2 d x d t
which in view of (6) and the Mean Value Theorem implies that
z L 2 ( Q ) 2 + r z L 2 ( Q ) 2 M μ Q 0 t | g ( s ) | | z ( x , s ) | d s | y 2 | | z | d x d t + a z L 2 ( Q ) 2 + b Q 0 t | g ( s ) | | z ( x , s ) | d s | z | d x d t + v L ( ω T ) z L 2 ( Q ) 2 + g Q T t T | ϑ ( x , s ) | | z ( x , s ) | d s | z | d x d t .
Hence, thanks to Theorem 2, (20) and (21), we deduce that
z L 2 ( Q ) 2 + r z L 2 ( Q ) 2 M μ g T C f L ( ω T ) + y 0 L ( Ω ) z L 2 ( Q ) 2 + ( b g T + ϑ g T + a + v L ( ω T ) ) z L 2 ( Q ) 2 ,
where C = C ( d , T , Ω , a , b , g ) > 0 .
Choosing r M μ g T C f L ( ω T ) + y 0 L ( Ω ) + ( b g T + ϑ g T + a + v L ( ω T ) ) + 1 ,
z L 2 ( Q ) 2 + z L 2 ( Q ) 2 0 ,
from which we deduce that z = 0 in Q. This means that y 1 = y 2 in Q .
Corollary 2.
Let v L ( ω T ) and y 0 L ( Ω ) . Also, let g L ( 0 , T ) and β C 2 ( R ) be such that (5), (6) and (7) hold. Then, there exists a unique weak solution y W ( 0 , T ; H 1 ( Ω ) ) to (2) provided that
Ω y 0 ϕ ( · , 0 ) d x = 0 T ϕ t ( t ) , y ( t ) ( H 1 ( Ω ) ) , H 1 ( Ω ) d t + Q ϕ · y d x d t + Q β ( h y g ) y ϕ d x d t ω T v y ϕ d x d t ,
holds for every ϕ H ( Q ) = { φ W ( 0 , T ; H 1 ( Ω ) ) such that ϕ ( · , T ) = 0 } .
Moreover, there exist positive constants C ( g , v L ( ω T ) , T ) and C M , g , v L ( ω T ) , T depending continuously on v L ( ω T ) such that
y L 2 ( 0 , T ; H 1 ( Ω ) ) C ( g , v L ( ω T ) , T ) y 0 L 2 ( Ω )
and
y W ( 0 , T ; H 1 ( Ω ) ) C M , g , v L ( ω T ) , T y 0 L 2 ( Ω ) .
Proof. 
Proceeding as in Theorem 3 while taking μ = 1 , f 0 , ϑ 0 ,   a 0 0 and b 0 0 , we obtain the existence of a weak solution to (2) as well as its uniqueness and the estimates (69) and (70). □
Lemma 6.
Let g L ( 0 , T ) , v L ( ω T ) , y 0 L ( Ω ) and β C 2 ( R ) be such that (6) and (7) hold. Also, let y = y ( v ) W ( 0 , T ; H 1 ( Ω ) ) L ( Q ) be the solution of (2). Then, there exist M > 0 and C ( d , T , Ω ) > 0 such that
y β ( h y g ) L ( Q ) M C d , T , Ω , v L ( ω T ) y 0 L ( Ω ) ,
and
β ( h y g ) L ( Q ) M .
Proof. 
Since y = y ( v ) W ( 0 , T ; H 1 ( Ω ) ) L ( Q ) is solution of (2), using (6), (7) and (45), we deduce (71) and (72) □
We now consider the system:
q t Δ q + β ( h y g ) q = v χ ω q g ( t ) t T y ( x , τ ) β ( h y g ) ( x , τ ) q ( x , τ ) d τ in Q , q ν = 0 in Σ , q ( v ; · , T ) = y ( v ; · , T ) y d in Ω .
Proposition 1.
Let v L ( ω T ) and y d L ( Ω ) and y = y ( v ) W ( 0 , T ; H 1 ( Ω ) ) L ( Q ) be solution to (2). Also, let g L ( 0 , T ) and β C 2 ( R ) be such that (5), (6) and (7) hold. Then, (73) has a unique weak solution q W ( 0 , T ; H 1 ( Ω ) ) L ( Q ) . Moreover, there exist C = C ( M , d , T , Ω , v L ( ω T ) , y 0 L ( Ω ) , g ) > 0 such that
q W ( 0 , T ; H 1 ( Ω ) ) C y 0 L 2 ( Ω ) + y d L ( Ω )
and
q L ( Q ) C y 0 L ( Ω ) + y d L ( Ω ) .
Proof. 
If we make the change for variable t T t in (73), we obtain that q ˜ ( x , t ) = q ( x , T t ) is a solution to
q ˜ t Δ q ˜ + β ( h ˜ y g ) q ˜ = v ˜ χ ω q ˜ g ( T t ) t T T y ( x , τ ) β ( h y g ) ( x , τ ) q ( x , τ ) d τ = 0 in Q , q ˜ ν = 0 in Σ , q ˜ ( v ; · , 0 ) = y ( v ; · , 0 ) y d in Ω ,
where h ˜ y g ( x , t ) = 0 T t g ( s ) y ( x , s ) d s , for almost every ( x , t ) Q , v ˜ ( x , t ) = v ( x , T t ) . Since y = y ( v ) W ( 0 , T ; H 1 ( Ω ) ) L ( Q ) is a solution of (2), we have on the one hand that y ( v ; · , 0 ) y d = y 0 y d L ( Ω ) , and on the other hand, using (71) and (72), we have that ϑ = y β ( h y g ) L ( Q ) and a 0 = β ( h y g ) L ( Q ) . Therefore, q ˜ ( x , t ) = q ( x , T t ) is a solution to (A1) with μ = 0 , b 0 = 0 ,   f = 0 , a 0 = β ( h y g ) and ϑ = y β ( h y g ) . Thanks to Corollary A1 and Theorem 2, we deduce that there exists a unique weak solution q ˜ W ( 0 , T ; H 1 ( Ω ) ) L ( Q ) to (76). Moreover, using (71) and (72), the continuous embedding of L ( Ω ) into L 2 ( Ω ) , the continuous embedding of L ( Q ) into L 2 ( Q ) , (A16) and (31) with μ = 0 , b 0 = 0 ,   f = 0 , a 0 = β ( h y g ) , ϑ = y β ( h y g ) , we have that there exists C = C ( M , d , T , Ω , v L ( ω T ) , y 0 L ( Ω ) , g ) > 0 such that
q ˜ W ( 0 , T ; H 1 ( Ω ) ) C 1 y d L ( Ω ) + y 0 L ( Ω ) .
Using (31), (71) and (72) again with μ = 0 , b 0 = 0 ,   f = 0 , a 0 = β ( h y g ) , ϑ = y β ( h y g ) , then (70), the embedding W ( 0 , T ; H 1 ( Ω ) ) into C ( [ 0 , T ] ; L 2 ( Ω ) ) , (71) and (72), we deduce that there exists C = C ( d , g , M , T , Ω , v L ( ω T ) , y 0 L ( Ω ) ) > 0 such that
q ˜ L ( Q ) C y d L ( Ω ) + y 0 L ( Ω ) .
Since q ˜ ( x , t ) = q ( x , T t ) , we deduce (74) and (75). □

4. Resolution of an Optimal Control Problem

From Corollaries 1 and 2, we know that the weak solution y of system (2) resides in X = W ( 0 , T , H 1 ( Ω ) ) L ( Q ) . We introduce the following space:
Y = ϕ X : ϕ t Δ ϕ L ( Q ) ,
and we observe that endowed with the graph norm
ϕ Y = ϕ X + ϕ t Δ ϕ L ( Q ) , ϕ Y ,
Y is a Banach space.
We also set the mapping S : Y × L ( ω T ) L ( Q ) × L ( Ω ) given by
( y , v ) S ( y , v ) : = y t Δ y + β 0 t g ( s ) y ( x , s ) d s y v χ ω y , y ( 0 ) y 0 ,
which is well defined, and the state equation as well as the initial data of system (2) can be viewed as
S ( y , v ) = 0 .
Moreover, taking into account the control-to-state mapping
S : L ( ω T ) X v y ( v ) ,
we can rewrite the optimal control problem (1) as:
inf v U a d J ( S ( v ) , v ) ,
where
J ( S ( v ) , v ) = J ( y ( v ) , v ) .

4.1. Sensitivity Analysis

Lemma 7.
Let β C 2 ( R ) be such that (6)–(7) hold and y = y ( v ) X be the solution of (2). Then, the mapping S is of class C 2 . Furthermore, the control-to-state mapping S is also of class C 2 .
Proof. 
It is clear that the second component of S is of class C with respect to y and v. Since β C 2 ( R ) , the first component of S is also of class C 2 . Therefore, we can conclude that S is of class C 2 . Moreover,
y S ( y , v ) φ = φ t Δ φ + y β ( h y g ) h φ g + β ( h y g ) φ v χ ω φ , φ ( 0 ) .
Let φ 0 L ( Ω ) and f L ( Q ) . We consider the linear problem
φ t Δ φ + y β ( h y g ) h φ g + β ( h y g ) φ = f + v χ ω φ in Q , φ ν = 0 in Σ , φ ( · , 0 ) = φ 0 in Ω .
Let b 0 = y β ( h y g ) and a 0 = β ( h y g ) . Then, from (71), (72), b 0 L ( Q ) , a 0 L ( Q ) and the system (81) is the same as (25) with μ = 0 , ϑ = 0 , y 0 = φ 0 . Consequently, we have from Theorems 2 and 3 that there exists a unique φ = φ ( φ 0 , f ) X which depends continuously on φ 0 L ( Ω ) and on f L ( Q ) . Since y L ( Q ) and β C 2 ( R ) satisfies (6) and (7), we have φ t Δ φ = f + v χ ω φ y β ( h y g ) h φ g β ( h y g ) φ L ( Q ) . Therefore, φ Y . Hence, y S ( y , v ) defines an isomorphism from Y to L ( Q ) × L ( Ω ) . Using the Implicit Function Theorem, we deduce that S ( y , v ) = 0 implicitly defines the control-to-state operator S : v y ( v ) which is itself of class C 2 . □
Proposition 2.
Let v , w , π L ( ω T ) . Under the assumptions of Lemma 7, the first derivative of the control-to-state operator S ( v ) is given by S ( v ) w = y w ( v ) , where y w ( v ) X is the unique weak solution to
y w ( v ) t Δ y w ( v ) = β ( h y g ) y w ( v ) + ( v y w ( v ) + w y ( v ) ) χ ω y ( v ) β ( h y g ) h y w ( v ) g i n Q , y w ( v ) ν = 0 i n Σ , y w ( v ; · , 0 ) = 0 i n Ω ,
 where according to (15) h y g ( x , t ) = 0 t g ( s ) y ( x , s ) d s for almost every ( x , t ) Q . The second derivative S ( v ) is given by S ( v ) ( w , π ) = y w π ( v ) , where y w π ( v ) X is the unique weak solution to
t y w π ( v ) Δ y w π ( v ) = β ( h y g ) y w π ( v ) y ( v ) β ( h y g ) h y w π ( v ) g + v χ ω y w π ( v ) + π χ ω y w ( v ) + w χ ω y π ( v ) + F i n Q , ν y w π ( v ) = 0 i n Σ , y w π ( v ; · , 0 ) = 0 i n Ω ,
where
F = β ( h y g ) y w h y π g β ( h y g ) y π h y w g β ( h y g ) y ( v ) h y w g h y π g ,
and y w = S ( v ) w , y π = S ( v ) π are respective solutions to (82) with w = w and w = π . Moreover, for every v , w L ( Ω ) , the linear mapping w y w = S ( v ) w can be extended to a linear continuous mapping on L 2 ( ω T ) and there exists C = C ( M , d , T , Ω , v L ( ω T ) , y 0 L ( Ω ) , g ) > 0 such that
y w L 2 ( 0 , T ; H 1 ( Ω ) ) C w L 2 ( ω T )
and
y w W ( 0 , T ; H 1 ( Ω ) ) C w L 2 ( ω T ) .
Proof. 
Let v , w L ( ω T ) . We have from Lemma 7 that the control-to-state mapping S : L ( ω T ) X , v y is of class C 2 . Therefore, S ( v ) w = y w ( v ) and S ( v ) ( w , π ) = y w π ( v ) exist. Using (6) and (7), we obtain after some computations that y w ( v ) is a solution to (82) and y w π ( v ) is a solution to (83).
Applying Theorems 2 and 3 to y w solution to the system (82) with f = w y ( v ) , b 0 = y β ( h y g ) ,   a 0 = β ( h y g ) , ϑ = 0 and μ = 0 , we deduce that y w X . Moreover, using (31), (47), (48), (71) and (72), we obtain that there exists C = C ( M , d , T , Ω , v L ( ω T ) , y 0 L ( Ω ) , g ) > 0 such that
y w L 2 ( 0 , T ; H 1 ( Ω ) ) C w L 2 ( ω T ) , y w W ( 0 , T ; H 1 ( Ω ) ) C w L 2 ( ω T )
and
y w L ( Q ) C ( M , d , T , Ω , v L ( ω T ) , y 0 L ( Ω ) , g ) w L ( ω T ) .
Proceeding as above, we also have that the y π solution to the system (82) with w = π belongs to X and satisfies the same estimation as y w . Since y w and y π belong to L ( Q ) , we have from Lemma 2 that h y π g and h y w g belong also to L ( Q ) . Consequently, using (6)–(7), the fact that the solution y to (2) belongs to X , we obtain that F defined in (84) belongs to L ( Q ) . Therefore, applying Theorem 3 and Theorem 2 to y w π solution to (83) with f = F + π y w ( v ) + w y π ( v ) ) χ ω L ( Q ) , μ = 0 , ϑ = 0 , b 0 = y β ( h y g ) L ( Q ) and a 0 = β ( h y g ) L ( Q ) , we deduce that the problem (83) has also a unique weak solution y w π X .
Proposition 3.
Let v L ( ω T ) and y 0 L ( Ω ) . Let g L ( 0 , T ) and β C 2 ( R ) be such that (5), (6) and (7) hold. Also, let y = y ( v ) X and q = q ( v ) X be the unique weak solution to (2) and to (73), respectively. Then, the mappings v S ( v ) = y ( v ) and v q ( v ) are Lipschitz continuous functions from L 2 ( ω T ) onto L 2 ( 0 , T ; H 1 ( Ω ) ) . More precisely, for all v 1 , v 2 L ( ω T ) , there is More precisely, for all v 1 , v 2 L ( ω T ) , there is a constant C 1 = C ( M , d , T , Ω , v 2 L ( Q ) , v 1 L ( Q ) , y 0 L ( Ω ) , g ) > 0 and C 2 = C d , T , Ω , v 2 L ( ω T ) , y 0 L ( Ω ) , g , M , v 1 L ( ω T ) , y d L ( Ω ) > 0 such that the following estimates hold
S ( v 1 ) S ( v 2 ) L 2 ( 0 , T ; H 1 ( Ω ) ) C 1 v 1 v 2 L 2 ( ω T )
and
q ( v 1 ) q ( v 2 ) L 2 ( 0 , T ; H 1 ( Ω ) ) C 2 v 1 v 2 L 2 ( ω T ) .
Proof. 
Let v 1 , v 2 L ( ω T ) and r > 0 . Let p = e r t ( y 1 y 2 ) where y 1 = y ( v 1 ) and y 2 = y ( v 2 ) are solutions of (2) with v = v 1 and v = v 2 , respectively. Then, p W ( 0 , T ; H 1 ( Ω ) ) is a weak solution to
p t + r p Δ p + β h y 1 g p = v 1 χ ω p + e r t ( v 1 v 2 ) χ ω y 2 β h y 1 g β h y 2 g y 2 in Q , p ν = 0 on Σ , p ( · , 0 ) = 0 in Ω .
We multiply the first equation in (90) by p and intergate by parts over Q, which gives us
1 2 p ( · , T ) L 2 ( Ω ) 2 + r p L 2 ( Q ) 2 + p L 2 ( Q ) 2 + Q β ( h y 1 g ) p 2 d x d t ω T v 1 p 2 d x d t + ω T e r t ( v 1 v 2 ) y 2 p d x d t + Q e r t β ( h y 1 g ) β ( h y 2 g ) y 2 p d x d t ,
which in view of (6) and (45), we obtain that
1 2 p ( · , T ) L 2 ( Ω ) 2 + r p L 2 ( Q ) 2 + p L 2 ( Q ) 2 1 2 p L 2 ( Q ) 2 + C ( d , T , Ω , v 2 L ( Q ) , y 0 L ( Ω ) ) v 1 v 2 L 2 ( ω T ) 2 + v 1 L ( ω T ) p L 2 ( Q ) 2 + C ( M , d , T , Ω , v 2 L ( Q ) , y 0 L ( Ω ) , g , T ) p L 2 ( Q ) 2 .
Choosing r = C ( M , d , T , Ω , v 2 L ( Q ) , y 0 L ( Ω ) , g , T ) + v 1 L ( ω T ) + 1 in this latter inequality, 1 2 p ( · , T ) L 2 ( Ω ) 2 + 1 2 p L 2 ( 0 , T ; H 1 ( Ω ) ) 2 C ( d , T , Ω , v 2 L ( Q ) , y 0 L ( Ω ) ) v 1 v 2 L 2 ( ω T ) 2 . Since p = e r t ( y 1 y 1 ) we deduce there exists C 1 = C ( M , d , T , Ω , v 2 L ( Q ) , v 1 L ( Q ) , y 0 L ( Ω ) , g ) > 0 such that (88) holds that
y ( v 1 ; · , T ) y ( v 2 ; · , T ) L 2 ( Ω ) = y 1 ( · , T ) y 2 ( · , T ) L 2 ( Ω ) C 1 v 1 v 2 L 2 ( ω T ) 2 .
Next, let v 1 , v 2 L ( ω T ) and r > 0 . Let z = e r t ( q 1 q 2 ) where q 1 = q ( v 1 ) and q 2 = q ( v 2 ) are solutions of (73) with v = v 1 and v = v 2 , respectively. Then, z W ( 0 , T ; H 1 ( Ω ) ) is a weak solution to
z t Δ z + r z + β ( h y 1 g ) z = v 1 χ ω z + e r t ( v 1 v 2 ) χ ω q 2 + η q in Q , z ν = 0 in Σ , z ( · , T ) = e r T ( y ( v 1 ; · , T ) y ( v 2 ; · , T ) ) in Ω
where
η q = e r t g ( t ) t T ( y 1 y 2 ) β ( h y 1 g ) q 2 d τ e r t g ( t ) t T y 1 e r τ β ( h y 1 g ) z d τ e r t g ( t ) t T y 2 q 2 β ( h y 1 g ) β ( h y 2 g ) d τ e r t q 2 ( β ( h y 1 g ) β ( h y 2 g ) ) .
Using the linearity of the integral, the notation (5), (6), (7), (15), (20) of Lemma 3, the Cauchy–Schwarz inequality and the fact that y 1 , y 2 , q 1 , q 2 L ( Q ) , we have that
Q η q z d x d t e r T M g q 2 L ( Q ) T y 1 y 2 L 2 ( Q ) z L 2 ( Q ) + g M y 1 L ( Q ) T z L 2 ( Q ) 2 + M q 2 L ( Q ) e r T g T y 1 y 2 L 2 ( Q ) z L 2 ( Q ) + g 2 e r T q 2 L ( Q ) y 2 L ( Q ) M T 2 y 1 y 2 L 2 ( Q ) z L 2 ( Q ) ,
which in view of Young’s inequality implies that
Q η q z d x d t 1 2 e 2 r T M 2 g 2 q 2 L ( Q ) 2 1 + T 2 + T 2 g 2 y 2 L ( Q ) 2 y 1 y 2 L 2 ( Q ) 2 + z L 2 ( Q ) 2 3 2 + g M y 1 L ( Q ) T
If we multiply the first equation in (92) by z and integrate by parts over Q, we obtain
1 2 z ( · , 0 ) L 2 ( Ω ) 2 + r z L 2 ( Q ) 2 + z L 2 ( Q ) 2 + Q β ( h y 1 g ) z 2 d x d t v 1 L ( ω T ) z L 2 ( Q ) 2 + 1 2 e 2 r T q 2 L ( Q ) 2 v 1 v 2 L 2 ( ω T ) + 1 2 z L 2 ( Q ) 2 + Q | η q | | z | d x d t + 1 2 e r T ( y ( v 1 ; · , T ) y ( v 2 ; · , T ) ) L 2 ( Ω ) 2 .
Choosing r = 3 + g M y 1 L ( Q ) T + v 1 L ( ω T ) in this latter inequality and using (6) and (93), we deduce that
z L 2 ( 0 , T ; H 1 ( Ω ) ) 2 1 2 e 2 r T q 2 L ( Q ) 2 v 1 v 2 L 2 ( ω T ) + e 2 r T 2 y 1 y 2 L 2 ( Q ) 2 1 + M 2 g 2 q 2 L ( Q ) 2 ( 1 + T + y 2 L ( Q ) 2 ) + 1 2 e r T ( y ( v 1 ; · , T ) y ( v 2 ; · , T ) ) L 2 ( Ω ) 2 ,
which in view of (45), (75), (88) and (91) implies that
z L 2 ( 0 , T ; H 1 ( Ω ) ) 2 C 2 v 1 v 2 L 2 ( ω T ) 2 ,
for some C 2 = C ( d , T , Ω , v 2 L ( Q ) , v 1 L ( Q ) , y 0 L ( Ω ) , M , g , y d L ( Ω ) ) > 0 .
Since z = e r t ( q 1 q 2 ) , we have that
q 1 q 2 L 2 ( 0 , T ; H 1 ( Ω ) ) 2 C v 1 v 2 L 2 ( ω T ) 2 .

4.2. Optimization Problem

We introduce the following notion of local solutions.
Definition 2.
We say that v U a d is an L p -local solution of (1) if there exists ε > 0 such that J ( S ( v ) , v ) J ( S ( w ) , w ) for every w U a d B ε p ( v ) where B ε p ( v ) = { w L p ( Ω ) : w v L p ( Q ) ε } . We say that v is a strict local minimum of (1) if the above inequality is strict whenever v w .
Theorem 4.
Let N > 0 , v U a d , y 0 L ( Ω ) and y d L ( Ω ) . Then, there exists at least a solution v U a d of the optimal control problem (2) and (79).
Proof. 
Since J ( S ( v ) , v ) 0 for all v U a d we can have a minimizing sequence { ( S ( v n ) , v n ) } n W ( 0 , T ; H 1 ( Ω ) ) × U a d such that
lim n J ( S ( v n ) , v n ) min w U a d J ( S ( w ) , w ) .
From the structure of cost function J given by (3), there exists a constant C > 0 independent of n such that
v n L 2 ( ω T ) C .
Since y n = S ( v n ) is the solution of (2) associated with the control v n , we know that y n satisfies
0 T ϕ t , y n ( H 1 ( Ω ) ) , H 1 ( Ω ) d t + Q ϕ y n d x d t + Q β ( h y n g ) y n ϕ d x d t = Ω y 0 ϕ ( · , 0 ) d x + ω T v n y n ϕ d x d t for every ϕ W ( 0 , T ; H 1 ( Ω ) ) , ϕ ( · , T ) = 0 in Ω
and in view of (70) in Corollary 2, there exists a constant C > 0 independent of n such that
y n W ( 0 , T ; H 1 ( Ω ) ) C .
Using, on the one hand, the boundedness of U a d in L ( ω T ) and (94), we have that there exist v L ( ω T ) such that
v n v weakly in L 2 ( ω T ) ,
v n v weak - star in L ( ω T ) ,
and on the other hand, using (96) and Lemma 1, we deduce the existence of y W ( 0 , T ; H 1 ( Ω ) ) such that
y n y weakly in W ( 0 , T ; H 1 ( Ω ) ) ,
y n y weakly in L 2 ( 0 , T ; H 1 ( Ω ) )
and
y n y strongly in L 2 ( Q ) .
Thanks to (22) and (101),
h y n g h y g strongly in L 2 ( Q ) as n + .
From (6) and Lemma 5, we have that for any ϕ L 2 ( Q ) ,
ϕ β h y n g ϕ β h y g strongly in L 2 ( Q ) , as n + .
Passing to the limit in (95), while using the convergences (98), (100), (101) and (103), we obtain that
Ω y 0 ϕ ( · , 0 ) d x + ω T v y ϕ d x d t = 0 T ϕ t , y ( H 1 ( Ω ) ) , H 1 ( Ω ) d t + Q ϕ y d x d t + Q β ( h y g ) y , ϕ d x d t , for every ϕ W ( 0 , T ; H 1 ( Ω ) ) , ϕ ( · , T ) = 0 in Ω .
Thanks to Equation (68), y = y ( v ) is a weak solution of (2). In addition, since U a d is a closed convex subset of L ( ω T ) , we have that v U a d .
Using (97) and (101) and the lower semi-continuity of the cost functional J, it is deduced that
J ( S ( v ) , v ) lim inf n J ( S ( v n ) , v n ) = inf w U a d J ( S ( w ) , w ) .
Proposition 4
(Twice Fréchet differentiability of J). Let y = y ( v ) X be the solution of (2) and set J ( v ) = J ( S ( v ) , v ) , where J ( S ( v ) , v ) is given by (80). Then, following the hypothesis of Lemma 7, the functional J : L ( ω T ) R is twice continuously Fréchet-differentiable, and for every π , v , w L ( ω T ) , we have
J ( v ) w = ω T y q + N v w d x d t
and
J ( v ) [ w , π ] = ω T ( π y w + w y π ) q d x d t + Q F q d x d t + Ω y w ( v ; x , T ) y π ( v ; x , T ) d x + N ω T w π d x d t .
where y w = S ( v ) w X , y π = S ( u ) π X are solutions to (82) with v = v and v = w , respectively. F L ( Q ) is defined in (84), and q X is the unique weak solution to the adjoint Equation (73).
Proof. 
From Lemma 7, we deduce that J is twice continuously Fréchet-differentiable.
Let π , v , w L ( ω T ) . After some computations, we obtain
J ( v ) w = Ω y w ( v ; x , T ) ( y ( v ; x , T ) y d ) d x + N ω T v w d x d t .
If we multiply the first equation in (82) by ϱ C ( Q ¯ ) such that ϱ ν = 0 on Σ and integrate by parts over Q, we obtain
Q p t Δ p + β ( h y g ) p v p χ ω y w ( v ) d x d t = Q y ( v ) β ( h y g ) h y w ( v ) g p d x d t + ω T w y ( v ) p d x d t Ω y w ( v ; x , T ) p ( x , T ) d x .
Hence, using the fact that h y w ( v ) g ( x , t ) = 0 t g ( s ) y w ( v ) ( x , s ) d s , we have
Q y ( v ) β ( h y g ) h y w ( v ) g p d x d t = Ω 0 T p ( x , t ) y ( v ) ( x , t ) β ( h y g ) ( x , t ) 0 t g ( s ) y w ( v ) ( x , s ) d s d t d x = Ω 0 T y w ( v ) ( x , s ) g ( s ) s T p ( x , t ) y ( v ) ( x , t ) β ( h y g ) ( x , t ) d t d s d x = Q y w ( v ) ( x , s ) g ( s ) s T p ( x , t ) y ( v ) ( x , t ) β ( h y g ) ( x , t ) d t d s d x .
Therefore, (108) can be rewritten as
Q p t Δ p + β ( h y g ) p v χ ω p y w ( v ) d x d t + Q g ( t ) t T p ( x , s ) y ( v ) ( x , s ) β ( h y g ) ( x , s ) d s y w ( v ) d x d t = ω T w y ( v ) p d x d t Ω y w ( v ; x , T ) p ( x , T ) d x .
So, if p verifies
p t Δ p = β ( h y g ) p g ( t ) t T y ( x , s ) β ( h y g ) ( x , s ) p ( x , s ) d s + v χ ω p in Q , p ν = 0 in Σ , p ( v ; · , T ) = y ( v ; · , T ) y d in Ω ,
where y ( v ; · , T ) y d L ( Ω ) , then, in view of Proposition 1 and the uniqueness of solution to (73), p = q W ( 0 , T ; H 1 ( Ω ) ) L ( Q ) . Hence, (109) becomes
Ω y w ( v ; x , T ) ( y ( v ; x , T ) y d ) d x = ω T y q w d x d t .
Combining (107) and (110), we obtain (105).
Next, after some calculations, we prove that
J ( v ) [ w , π ] = Ω y w π ( v ; x , T ) ( y ( v ; x , T ) y d ) d x + Ω y w ( v ; x , T ) y π ( v ; x , T ) d x + N ω T w π d x d t ,
where y w π W ( 0 , T ; H 1 ( Ω ) ) , the solution to (83). Multiplying the first equation of (73) by the y w π solution to (83) and integrating by parts over Q, we obtain
Ω y w π ( v ; x , T ) ( y ( v ; x , T ) y d ) d x = ω T ( π y w + w y π ) q d x d t + Q F q d x d t .
Thus, if we combine (111) and (112), we deduce (106). □
Theorem 5.
Let v U a d be an L -local minimum for (1). Then,
J ( v ) ( w v ) 0 for every w U a d ,
equivalently
ω T ( y q + N v ) ( w v ) d x 0 for every w U a d ,
where q is the unique weak solution to (73).
Proof. 
Let w U a d be arbitrary and v be an L -local minimum. Since U a d is convex, we have that v + λ ( w v ) U a d for all λ ( 0 , 1 ] . Then, for all w U a d ,
0 lim λ 0 J ( v + λ ( ( w v ) ) J ( v ) λ = J ( v ) ( w v ) ,
which in view of (105) implies that
ω T ( y q + N v ) ( ( w v ) d x 0 for every w U a d .
Lemma 8.
Assume that Lemma 7 holds. Let v , w , π L ( ω T ) , J ( v ) w and J ( v ) [ w , π ] be given by (105) and (106), respectively. Then, the linear mapping w J ( v ) w can be extended to a linear continuous mapping J ( v ) : L 2 ( ω T ) R and the bilinear mapping ( w , π ) J ( v ) [ w , π ] can be extended to a bilinear continuous mapping on J ( v ) : L 2 ( ω T ) × L 2 ( ω T ) R given by (106).
Proof. 
Let v L ( ω T ) and w L 2 ( ω T ) . From (105), we have that
J ( v ) w = ω T y q + N v w d x d t ,
where y = y ( v ) and q = q ( v ) are the solution of (2) and (73), respectively.
Using the Cauchy–Schwarz inequality, (45) and (75), we have that
| J ( v ) w | C w L 2 ( ω T )
for some C = C ( g , M , T , d , Ω , v L ( ω T ) , y 0 L ( Ω ) , y d L ( Ω ) ) > 0 . Thus, the mapping w J ( v ) w is linear and continuous on L 2 ( ω T ) .
From (84),
F = β ( h y g ) y w h y π g β ( h y g ) y π h y w g β ( h y g ) y h y w g h y π g ,
where y = y ( v ) X , y π X and y w X are solutions to (2), (82) and (83), respectively.
Using (7) and the Cauchy–Schwarz inequality, (17) of Lemma 2, we deduce that
Q | F | d x d t M y w L 2 ( Q ) h y π g L 2 y ( Q ) + M y π L 2 ( Q ) h y w g L 2 ( Q ) + M y L ( Q ) h y w g L 2 ( Q ) h y π g L 2 ( Q ) ,
which in view of (20), (45) and (85), and of Lemma 3 implies that there exist two positive constants C 2 = C ( d , T , Ω , v L ( ω T ) , y 0 L ( Ω ) , ) > 0 and C 1 = C ( M , d , T , Ω , v L ( ω T ) , y 0 L ( Ω ) , g , ) > 0 such that
Q | F | d x d t M C 1 2 w L 2 ( ω T ) g T π L 2 ( ω T ) + M C 1 2 π L 2 ( ω T ) g T w L 2 ( ω T ) + M g 2 T 2 C 1 2 C 2 w L 2 ( ω T ) π L 2 ( ω T ) .
Hence, we have that
F L 1 ( Q ) C w L 2 ( ω T ) π L 2 ( ω T ) ,
for some C = C ( M , d , T , Ω , v L ( ω T ) , y 0 L ( Ω ) , g , ) > 0 .
Applying the Cauchy–Schwartz inequality to (106), we obtain
| J ( v ) [ w , π ] | π L 2 ( ω T ) y w L 2 ( Q ) + w L 2 ( ω T ) y π L 2 ( Q ) q L ( Q ) + F L 1 ( Q ) q L ( Q ) + y w L 2 ( Q ) y π L 2 ( Q ) + N π L 2 ( ω T ) w L 2 ( ω T ) ,
which in view of (75), (85) and (115) implies that
| J ( v ) [ w , π ] | C w L 2 ( ω T ) π L 2 ( ω T )
for some constant there exists C = C ( M , d , T , Ω , v L ( ω T ) , y 0 L ( Ω ) , g , ) > 0 .
Since the cost functional J given by (80) is non-convex, the first-order optimality conditions given in Theorem 5 are necessary but not sufficient for optimality. Sufficient second-order conditions are required. To this end, we first recall the definition of the feasible direction and the critical cone that can be found [18,19].
Definition 3.
1. 
The set of feasible directions is the set A ( v ) defined by
A ( v ) = { π L ( ω T ) : π = λ ( w v ) for some λ > 0 and w U a d } .
2. 
The critical cone is the set C ^ ( v ) defined by
C ^ ( v ) = cl L 2 ( ω T ) ( A ( v ) ) { w L 2 ( ω T ) : J ( v ) w = 0 } .
where cl L 2 ( ω T ) ( A ( v ) ) denotes the closure in L 2 ( ω T ) of the admissible set A ( v ) .
Next, proceeding as in ([18] (Page 18)) (see also [19] (page 273)) we prove the following result that gives a characterization of the critical cone C ^ ( v ) defined in (119).
Proposition 5.
Let v U a d and w be defined for a.e ( x , t ) ω T as follows:
w ( x , t ) 0 i f v ( x , t ) = v a , w ( x , t ) 0 i f v ( x , t ) = v b , w ( x , t ) = 0 i f N v ( x , t ) + q ( x , t ) 0 ,
where q is a solution to (73). The critical cone C ^ ( v ) defined in (119) can be rewritten as
C ^ ( v ) = { w L 2 ( ω T ) : w fulfills ( 120 ) } .
Proceeding as in ([20] (p. 246)) or [21,22], we prove the following result.
Proposition 6
(Necessary second-order optimality conditions). Let v U a d be a L -local solution of system (79). Then, J ( v ) w 2 0 , for all w C ^ ( v ) .
Theorem 6.
Let v U a d be a control satisfying the first-order optimality conditions (113). Then, the following hold:
1. 
The functional J : L ( ω T ) R is of class C 2 . Furthermore, if we denote by L ( L 2 ( ω T ) , R ) and by B ( L 2 ( ω T ) , R ) the spaces of linear continuous functional on L 2 ( ω T ) and bilinear continuous functional on L 2 ( ω T ) × L 2 ( ω T ) , then for every v U a d , the following continuous extensions exist:
J ( v ) L ( L 2 ( ω T ) , R ) and J ( v ) B ( L 2 ( ω T ) , R ) .
2. 
For any sequence ( v n , w n ) n = 1 U a d × L 2 ( ω T ) such that
v n v . strongly in L 2 ( ω T ) ,
w n w weakly in L 2 ( ω T ) ,
as n ,
J ( v ) w = lim n J ( v n ) w n
and
J ( v ) [ w , w ] lim inf n J ( v n ) [ w n , w n ] .
In addition, if w = 0 , then
N lim inf n w n L 2 ( ω T ) 2 lim inf n J ( v n ) [ w n , w n ] .
Proof. 
We have that v U a d is a control satisfying the first-order optimality conditions (113). Thus, from Proposition 4 and Lemma 8, we have (122).
To prove (125), (126) and (127), we proceed in two steps.
Step 1. We prove (125).
Let v U a d and the sequence ( v n , w n ) n = 1 U a d × L 2 ( ω T ) be such that (123) and (124) hold. Then, we have that there exist C 4 > 0 and C 5 > 0 independent of n such that
v n L ( ω T ) C 4 , w n L 2 ( ω T ) C 5 .
If we denote by y n = y ( v n ) and y = y ( v ) , the solutions of (2) with v = v n and v = v , respectively, and by q n = q ( v n ) and q = q ( v ) , the solutions to (73) with v = v n and v = v , respectively, then from (45), (70), (74) and (75), we have that y , y n , q and q n belong to X . Moreover, from (45), (85), (75) and (87) we obtain
y n L ( ω T ) C ( d , T , Ω , C 4 , y 0 L ( Ω ) ) ,
and
q n L ( ω T ) C ( d , T , Ω , C 4 , y 0 L ( Ω ) , y d L ( Ω ) , g , M ) ,
for some C ( d , T , Ω , C 4 , y 0 L ( Ω ) ) > 0 and C ( d , T , Ω , C 4 , y 0 L ( Ω ) , y d L ( Ω ) , g , M ) > 0 . Using the Lipschitz continuity of the mapping v y ( v ) and v q ( v ) given in Proposition 3, we have that
y n y strongly in L 2 ( 0 , T ; H 1 ( Ω ) ) as n
and
q n q strongly in L 2 ( 0 , T ; H 1 ( Ω ) ) as n .
Thanks to (130) and (132), we have that
q n q weak - star in L ( Q ) , as n .
On the other hand, in view of (129) and (130), the sequences ( y n q n ) are bounded in L ( Q ) and thus in L 2 ( Q ) since Q is bounded. Therefore, we can prove using (129), (130), (131) and (132) that
y n q n y q strongly in L 2 ( Q ) as n .
Observing that
J ( v n ) w n = ω T y n q n + N v n w n d x d t ,
we have using (123), (124) and (134) that
lim n J ( v n ) w n = ω T y q + N v w d x d t = J ( v ) w .
Step 2. We show (126) and (127).
Let z n = y w ( v n ) and z = y w ( v ) be the solutions to (82) with ( v , w ) = ( v n , w n ) and ( v , w ) = ( v , w ) , respectively. Then, z n , z W ( 0 , T ; H 1 ( Ω ) ) and from (86) and (128), we have that there exist C 4 > 0 and C 5 > 0 independent of n such that
z n W ( 0 , T ; H 1 ( Ω ) ) C ( d , T , Ω , C 4 , y 0 L ( Ω ) , g , M ) ,
for some C ( d , T , Ω , C 4 , y 0 L ( Ω ) , g , M ) > 0 independent of n. Thanks to (20) of Lemma 3,
h z g L 2 ( Q ) g T z L 2 ( Q ) ,
and
h z n g L 2 ( Q ) C ( d , T , Ω , C 4 , y 0 L ( Ω ) , g , M ) ,
for some C ( d , T , Ω , C 4 , y 0 L ( Ω ) , g , M ) > 0 .
In addition, in view of (135), the compact embedding of W ( 0 , T ; H 1 ( Ω ) ) into L 2 ( Q ) and the continuous embedding of W ( 0 , T ; H 1 ( Ω ) ) into C [ 0 , T ] ; L 2 ( Ω ) , we deduce that there exists z L 2 ( Q ) such that
z n z strongly in L 2 ( Q ) as n
and
z n ( · , T ) z ( · , T ) weakly in L 2 ( Ω ) as n .
Thanks to (22) and (138) of Lemma 3,
h z n g h z g strongly in L 2 ( Q ) , as n .
Now, from (84) and (106),
J ( v ) [ w , w ] = 2 ω T w z q d x d t + Ω z ( x , T ) 2 d x + N ω T w 2 d x d t 2 Q β ( h y g ) z h z g q d x d t Q β ( h y g ) y ( v ) h z g 2 q d x d t .
Thus,
J ( v n ) [ w n , w n ] = 2 ω T w n z n q n d x d t + Ω z n ( x , T ) 2 d x + N ω T w n 2 d x d t 2 Q β ( h y n g ) z n h z n g q n d x d t Q β ( h y n g ) y n h z n g 2 q n d x d t .
where we recall z n = y w ( v n ) and z = y w ( v ) as the solutions to (82) with ( v , w ) = ( v n , w n ) and ( v , w ) = ( v , w ) , respectively. Since β C 2 satisfies (6) and (7), v U a d and the sequence ( v n , w n ) n = 1 U a d × L 2 ( ω T ) satisfies (123) and (124), we have from (7), (129), (130), (135) and (136) that
β ( h y n g ) q n z n L 2 ( Q ) d x d t M q n L ( Q ) z n L 2 ( Q ) C
and
β ( h y n g ) q n y n h z n g L 2 ( Q ) d x d t M q n L ( Q ) y n L ( Q ) h z n g L ( Q ) C
where there exists C = C ( d , T , Ω , C 4 , y 0 L ( Ω ) , y d L ( Q ) , g , M ) > 0 . Consequently, for any ϕ L 2 ( Q ) ,
Q β ( h y n g ) q n z n ϕ d x d t Q β ( h y g ) q z ϕ d x d t Q β ( h y n g ) ϕ q n ( z n z ) d x d t + Q β ( h y n g ) ϕ ( q n z q z ) d x d t + Q β ( h y n g ) ϕ β ( h y g ) ϕ q z d x d t
and
Q β ( h y n g ) q n y n h z n g ϕ d x d t Q β ( h y g ) q y h z g ϕ d x d t Q β ( h y n g ) y n q n ϕ h z n g h z g d x d t + Q β ( h y n g ) h z g q n ( y n ϕ y ϕ ) d x d t + Q β ( h y n g ) h z g y ( q n ϕ q ϕ ) d x d t + Q h z g y q β ( h y n g ) ϕ β ( h y g ϕ ) d x d t ,
which in view of the Cauchy—Schwarz inequality, (7), (129), (130) and (135) implies that
Q β ( h y n g ) q n z n ϕ d x d t Q β ( h y g ) q z ϕ d x d t M q n L ( Q ) ϕ L 2 ( Q ) z n z L 2 ( Q ) + M z L 2 ( Q ) q n ϕ q ϕ L 2 ( Q ) + q L ( Q ) z L 2 ( Q ) β ( h y n g ) ϕ β ( h y g ) ϕ L 2 ( Q )
and
Q β ( h y n g ) q n y n h z n g ϕ d x d t Q β ( h y g ) q y h z g ϕ d x d t M y n L ( Q ) q n L ( Q ) ϕ L 2 ( Q ) h z n g h z g L 2 ( Q ) + M q n L ( Q ) h z g L 2 ( Q ) y n ϕ y ϕ L 2 ( Q ) + M h z g L 2 ( Q ) q n ϕ q ϕ L 2 ( Q ) + y L ( Q ) q L ( Q ) β ( h y n g ) ϕ β ( h y g ϕ L 2 ( Q ) h z g L 2 ( Q ) .
Thanks to (7), (24), (129), (130), (131), (132) and Lemma 5, for any ϕ L 2 ( Q ) ,
q n ϕ q ϕ strongly in L 2 ( Q ) , y n ϕ y ϕ strongly in L 2 ( Q ) , β ( h y n g ) ϕ β ( h y g ϕ strongly in L 2 ( Q ) , β ( h y n g ) ϕ β ( h y g ) ϕ strongly in L 2 ( Q ) ,
as n + . Therefore, passing to limit in (142) and in (143) while using the above convergence, (129), (130), (138) and (140), we deuce that
Q β ( h y n g ) q n z n ϕ d x d t Q β ( h y g ) q z ϕ d x d t , ϕ L 2 ( Q )
and
Q β ( h y n g ) q n y n h z n g ϕ d x d t Q β ( h y g ) q y h z g ϕ d x d t , ϕ L 2 ( Q ) .
Thus
β ( h y n g ) q n z n β ( h y g ) q z weakly in L 2 ( Q )
and
β ( h y n g ) q n y n h z n g β ( h y g ) q y h z g weakly in L 2 ( Q ) ,
as n . Since q n z n is bounded in L 2 ( Q ) because (130) and (135),
Q ( q n z n q z ) 2 d x d t 2 q n L ( Q ) Q ( z n z ) 2 d x d t + 2 Q z 2 ( q n q ) 2 d x d t .
Passing to the limi in this latter inequality while using (130), (133), (138) and the fact that z 2 L 1 ( Q ) , we deduce that
lim n Q ( q n z n q z ) 2 d x d t = 0 .
This means that
q n z n q z strongly in L 2 ( Q ) .
Then, using (141), (124), (144), (145), (138), (139) and (146), we obtain that
lim inf n J ( v n ) [ w n , w n ] = lim n 2 ω T w n z n q n d x d t + lim inf n Q z n ( · , T ) 2 d x d t + N ω T w n 2 d x d t + lim n 2 Q β ( h y n g ) z n h z n g q n d x d t Q β ( h y n g ) y n h z n g 2 q n d x d t 2 ω T w z q d x d t + Q z 2 ( 0 , T ) d x d t + N ω T w 2 d x d t 2 Q β ( h y g ) z h z g q d x d t Q β ( h y g ) y h z g 2 q d x d t = J ( v ) [ w , w ] .
Finally, if w = 0 , then z = y w , being the solution of (105), we have z = 0 . Hence, passing to the lower limit in (141) when n , we obtain (127). □
We have the following result giving sufficient second-order conditions for locally optimal solutions.
Theorem 7.
Let v U a d be a control satisfying the first-order optimality conditions (113) and
J ( v ) [ w , w ] > 0 w C ( v ) { 0 } .
Then, there exist γ > 0 and η > 0 such that the quadratic growth condition holds
J ( w ) J ( v ) + η w v L 2 ( ω T ) 2 , w U a d B γ 2 ( v ) .
Therefore, v is locally optimal in the sense of L 2 ( ω T ) .
Proof. 
Relation (148) is obtained by proceeding as ([19] Theorem 2.3, page 265) while using Proposition 6. □

5. Concluding Remarks

In this paper, we studied the optimal control of a nonlinear parabolic equation involving a nonlocal-in-time term. We proved some regularity results of systems and the existence of the optimal solution to our control problem. Since the system is nonlinear and thus the cost function non-convex, we obtain local uniqueness by means of the second optimality conditions.

Author Contributions

Conceptualization, G.M. and A.F.; methodology, G.M., A.F. and C.J.-A.; validation, G.M., A.F. and C.J.-A.; formal analysis, G.M.; investigation, A.F.; writing—original draft preparation, G.M.; writing—review and editing, G.M.; visualization, G.M., A.F. and C.J.-A.; supervision, G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

For any z L 2 ( Q ) , we consider the following system:
y t Δ y + b 0 h y g + μ β h z g y + a 0 y + g ( t ) h y ϑ = v χ ω y + f in Q , y ν = 0 on Σ , y ( · , 0 ) = y 0 in Ω ,
where z L 2 ( Q ) , μ 0 , f L 2 ( Q ) and y 0 L 2 ( Ω ) . The functions g L ( Q ) ,   b 0 L ( Q ) ,   ϑ L ( Q ) and a 0 L ( Q ) are such that (5), (16), (26) and (27) hold. The functions h z g and h y ϑ are defined as in (15).
Let r > 0 . If we make the change for variable p = e r t y , where y is a solution of (A1), then p satisfies:
p t Δ p = ( r + a 0 ) p μ β ( h z g ) p e r t b 0 h e r ( · ) p g e r t g ( t ) h e r ( · ) p ϑ + v χ ω p + e r t f in Q , p ν = 0 on Σ , p ( · , 0 ) = y 0 in Ω ,
where in view of notation (15), h e r ( · ) p g ( x , t ) = 0 t e r s g ( s ) p ( x , s ) d s and h e r ( · ) p ϑ ( x , t ) = T t T e r s ϑ ( x , s ) p ( x , s ) d s , for a.e. ( x , t ) Q .
Definition A1.
Let r > 0 , μ 0 ,   z L 2 ( Q ) , y 0 L 2 ( Ω ) , g L ( Q ) ,   f L 2 ( Q ) , b 0 L ( Q ) ,   ϑ L ( Q ) and a 0 L ( Q ) . Assume that β C 2 ( R ) satisfies (6) and (7). We say that a function p L 2 ( 0 , T ; H 1 ( Ω ) ) is a weak solution to (A2) if
Ω y 0 ϕ ( · , 0 ) d x = 0 T ϕ t , p ( H 1 ( Ω ) ) , H 1 ( Ω ) d t + Q ϕ p d x d t + μ Q β ( h z g ) p ϕ d x d t ω T v p ϕ d x d t Q e r t f ϕ d x d t + Q e r t b 0 0 t e r s g ( s ) p ( x , s ) d s ϕ d x d t + Q e r t g ( t ) T t T e r s ϑ ( x , s ) p ( x , s ) d s ϕ d x d t ,
holds for every ϕ H ( Q ) : = { φ W ( 0 , T ; H 1 ( Ω ) ) such that ϕ ( · , T ) = 0 } .
Theorem A1.
Assume that β C 2 ( R ) satisfies (6) and (7). Let r > 0 , μ 0 , z L 2 ( Q ) , f L 2 ( Q ) and y 0 L 2 ( Ω ) . Also, let g L ( 0 , T ) , b 0 L ( Q ) , ϑ L ( Q ) and a 0 L ( Q ) be such that (5), (16), (26) and (27) hold. Then, there exists a unique weak solution p W ( 0 , T ; H 1 ( Ω ) ) to (A2) in the sense of Definition A1. Moreover,
p L 2 ( 0 , T ; H 1 ( Ω ) ) f L 2 ( Q ) + y 0 L 2 ( Ω )
and there exists C = C M , μ , g , b , ϑ , a , T > 0 such that
p W ( 0 , T ; H 1 ( Ω ) ) C f L 2 ( Q ) + y 0 L 2 ( Ω ) .
Proof. 
We proceed in four steps.
Step 1. We prove that there exists the p L 2 ( 0 , T ; H 1 ( Ω ) ) solution to (A2).
We recall that the norm on L 2 ( 0 , T ; H 1 ( Ω ) ) is given by:
p L 2 ( 0 , T ; H 1 ( Ω ) ) = 0 T p ( · , t ) H 1 ( Ω ) 2 d t 1 / 2
and we define the norm on H ( Q ) by:
p H ( Q ) = p L 2 ( 0 , T ; H 1 ( Ω ) ) 2 + p ( · , 0 ) L 2 ( Ω ) 2 1 / 2 .
Then, we have the continuous embedding H ( Q ) L 2 ( 0 , T ; H 1 ( Ω ) ) .
Let E ( · , · ) : L 2 ( 0 , T ; H 1 ( Ω ) ) × H ( Q ) R be the bilinear functional given for all ( p , ϕ ) L 2 ( 0 , T ; H 1 ( Ω ) ) × H ( Q ) by:
E ( p , ϕ ) = 0 T ϕ t , p ( H 1 ( Ω ) ) , H 1 ( Ω ) d t ω T v p ϕ d x d t + Q ϕ p d x d t + μ Q β ( h z g ) p ϕ d x d t + r Q p ϕ d x d t + Q e r t b 0 0 t e r s g ( s ) p ( x , s ) d s ϕ d x d t + Q a 0 p ϕ d x d t + Q e r t g ( t ) T t T e r s ϑ ( x , s ) p ( x , s ) d s ϕ d x d t
Using Cauchy–Schwarz’s inequality, (5), (6), (16), (26) and (27), we obtain that
| E ( p , ϕ ) | C p L 2 ( 0 , T ; H 1 ( Ω ) ) .
where C > 0 is given by
C = ϕ t L 2 ( 0 , T ; ( H 1 ( Ω ) ) ) + ϕ L 2 ( Q ) + μ M + T b g + T ϑ g + a + r + v L ( Q ) ϕ L 2 ( Q ) .
Consequently, for every fixed ϕ H ( Q ) , the functional p E ( p , ϕ ) is continuous on L 2 ( 0 , T ; H 1 ( Ω ) ) .
For every ϕ H ( Q ) ,
E ( ϕ , ϕ ) = 1 2 ϕ ( · , 0 ) L 2 ( Ω ) 2 + ϕ L 2 ( Q ) 2 ω T v ϕ 2 d x d t + μ Q β ( h z g ) ϕ 2 d x d t + r Q ϕ 2 d x d t + Q e r t b 0 0 t e r s g ( s ) ϕ ( x , s ) d s ϕ d x d t + Q a 0 ϕ 2 d x d t + Q e r t g ( t ) T t T e r s ϑ ( x , s ) ϕ ( x , s ) d s ϕ d x d t .
In view of (5), (16), (26) and (27), we have for any ϕ H ( Q ) ,
Q e r t b 0 0 t e r s g ( s ) ϕ ( x , s ) d s ϕ d x d t b g T ϕ L 2 ( Q ) 2 , Q e r t g ( t ) T t T e r s ϑ ( x , s ) ϕ ( x , s ) d s ϕ d x d t ϑ g T ϕ L 2 ( Q ) 2 , Q a 0 ϕ 2 d x d t a ϕ L 2 ( Q ) 2 .
Hence, using (6) and the definition of the norm on H ( Q ) given by (A6), we obtain that
E ( ϕ , ϕ ) 1 2 ϕ ( · , 0 ) L 2 ( Ω ) 2 + ϕ L 2 ( Q ) 2   + r a b g T ϑ g T v L ( ω T ) ϕ L 2 ( Q ) 2 .
Choosing in this latter inequality r = a + b g T + v L ( ω T ) + ϑ g T + 1 , we obtain that
E ( ϕ , ϕ ) 1 2 ϕ ( · , 0 ) L 2 ( Ω ) 2 + ϕ L 2 ( 0 , T ; H 1 ( Ω ) ) 2 = 1 2 ϕ H ( Q ) 2 .
Thus E ( · , · ) is coercive on H ( Q ) .
Finally, we consider the functional L : H ( Q ) R defined by
L ( ϕ ) : = Ω y 0 ϕ ( · , 0 ) d x + Q e r t f ϕ d x d t .
Using Cauchy–Schwarz’s inequality,
| L ( ϕ ) | y 0 L 2 ( Ω ) ϕ ( · , 0 ) L 2 ( Ω ) + f L 2 ( Q ) ϕ L 2 ( Q ) y 0 L 2 ( Ω ) 2 + f L 2 ( Q ) 2 1 / 2 ϕ ( · , 0 ) L 2 ( Ω ) 2 + ϕ L 2 ( 0 , T ; H 1 ( Ω ) ) 2 1 / 2 ,
which in view of the definition of he norm on H ( Q ) given by (A6) implies that L ( · ) is continuous on H ( Q ) . Since we have proved that the bilinear functional E ( · , · ) is coercive on H ( Q ) and continuous on L 2 ( 0 , T ; H 1 ( Ω ) ) , for every fixed ϕ H ( Q ) , Theorem 1.1 ([23] Page 37) allows us to say that there exists p L 2 ( 0 , T ; H 1 ( Ω ) ) such that E ( p , ϕ ) = L ( ϕ ) , ϕ H ( Q ) .
Step 2. We show that p t L 2 ( 0 , T ; ( H 1 ( Ω ) ) ) .
Since p L 2 ( 0 , T ; H 1 ( Ω ) ) , using (6) and the fact that a 0 , b 0 L ( Q ) , v L ( ω T ) and g L ( 0 , T ) , we have that for almost every t ( 0 , T ) ,
p t ( t ) = Δ p ( t ) ( r + a 0 ) p ( t ) μ β ( h z g ) ( t ) p ( t ) e r t b 0 h e r ( · ) p g + e r t g ( t ) h e r ( · ) p ϑ + v ( t ) χ ω p ( t ) + e r t f ( t ) ( H 1 ( Ω ) ) .
If we take the duality map between the first equation in (A2) and ϕ L 2 ( ( 0 , T ) ; H 1 ( Ω ) ) , we obtain
p t ( t ) , ϕ ( t ) ( H 1 ( Ω ) ) , H 1 ( Ω ) = ω v ( t ) p ( t ) ϕ ( t ) d x Ω ϕ ( t ) p ( t ) d x μ Ω β ( h z g ) ( t ) p ( t ) ϕ ( t ) d x Ω a 0 p ( t ) ϕ ( t ) d x Ω e r t b 0 0 t e r s g ( s ) p ( x , s ) d s ( t ) ϕ ( t ) d x r Ω p ( t ) ϕ ( t ) d x Ω e r t g ( t ) T t T e r s ϑ ( x , s ) p ( x , s ) d s ϕ ( t ) d x + Ω e r t f ( t ) ϕ ( t ) d x ,
which in view of (5), (6), (16), (26), (27) and the Cauchy–Schwarz inequality implies that
0 T p t ( t ) , ϕ ( t ) ( H 1 ( Ω ) ) , H 1 ( Ω ) d t f L 2 ( Q ) ϕ L 2 ( 0 , T ; H 1 ( Ω ) ) + ( 1 + M μ ) p L 2 ( 0 , T ; H 1 ( Ω ) ) ϕ L 2 ( 0 , T ; H 1 ( Ω ) ) + ( b + ϑ ) g T p L 2 ( 0 , T ; H 1 ( Ω ) ) ϕ L 2 ( 0 , T ; H 1 ( Ω ) ) + v L ( ω T ) + a + r p L 2 ( 0 , T ; H 1 ( Ω ) ) ϕ L 2 ( 0 , T ; H 1 ( Ω ) ) .
Observing that r = a + b g T + v L ( ω T ) + ϑ g T + 1 , we have that
p t L 2 ( 0 , T ; ( H 1 ( Ω ) ) ) f L 2 ( Q ) + C p L 2 ( 0 , T ; H 1 ( Ω ) ) ,
for some C = v L ( ω T ) , ϑ , M , μ , b , g , T , a , r > 0 .
Step 3. We establish the estimates (A4) and (A5).
If we multiply the first equation in (A2) by p L 2 ( ( 0 , T ) ; H 1 ( Ω ) ) and integrate by parts over Q, we have that
1 2 p ( · , T ) L 2 ( Ω ) + p L 2 ( Q ) 2 + μ Q β ( h z g ) p 2 d x d t + r p L 2 ( Q ) 2 = ω T v p 2 d x d t + Q e r t f p d x d t Q e r t b 0 0 t e r s g ( s ) p ( x , s ) d s p d x d t Q e r t g ( t ) T t T e r s ϑ ( x , s ) p ( x , s ) d s p d x d t Q a 0 p 2 d x d t + 1 2 y 0 L 2 ( Ω ) ,
Hence, using (5), (6), (16), (27), (26) and the Young’s inequality, we deduce that
p L 2 ( Q ) 2 + r p L 2 ( Q ) 2 v L ( ω T ) p L 2 ( Q ) 2 + 1 2 f L 2 ( Q ) 2 + 1 2 p L 2 ( Q ) 2 + ( b + ϑ ) g T p L 2 ( Q ) 2 + a p L 2 ( Q ) 2 + 1 2 y 0 L 2 ( Ω ) ,
which, because r = a + ( b + ϑ ) g T + v L ( ω T ) + 1 , gives
p L 2 ( 0 , T ; H 1 ( Ω ) ) f L 2 ( Q ) + y 0 L 2 ( Ω ) .
Hence, (A8) becomes
p t L 2 ( 0 , T ; ( H 1 ( Ω ) ) ) f L 2 ( Q ) + C ( M , μ , g , b , ϑ , a , v L ( ω T ) , T ) f L 2 ( Q ) + y 0 L 2 ( Ω ) .
From (A10), (A11) and the definition of the norm on W ( 0 , T ; H 1 ( Ω ) ) given by (8), we deduce that
p W ( 0 , T ; H 1 ( Ω ) ) C f L 2 ( Q ) + y 0 L 2 ( Ω ) ,
where C = C M , μ , g , b , ϑ , a , v L ( ω T ) , T > 0 depends continuously on v L ( ω T ) .
Step 4. We prove the uniqueness of p solution to (A2).
Assume that there exist p 1 and p 2 solutions to (A2) with the same right hand side f and initial value y 0 . Let q = p 1 p 2 . Then, q is a solution to (A2) with f = 0 and y 0 = 0 :
q t Δ q = ( r + a 0 ) q μ β ( h z g ) q e r t b 0 h e r ( · ) q g e r t g ( t ) h e r ( · ) q ϑ + v χ ω q in Q , q ν = 0 on Σ , q ( · , 0 ) = 0 in Ω ,
Multiplying the first equation in (A12) by q, integrating by parts over Ω , then using (A9), with p = q , f = 0 and y 0 = 0 , we obtain that
q L 2 ( Q ) 2 + r q L 2 ( Q ) 2 v L ( ω T ) q L 2 ( Q ) 2 + ( b + ϑ ) g T q L 2 ( Q ) 2 + a q L 2 ( Q ) 2 ,
which, because r = a + ( b + ϑ ) g T + v L ( ω T ) + 1 , gives
q L 2 ( 0 , T ; H 1 ( Ω ) ) 0 .
This implies that q = 0 in Q and therefore p 1 = p 2 in Q. □
The following results for system (A1) follow from Theorem A1 and the change of variable p = e r t y , with r = a + ( b + ϑ ) g T + v L ( ω T ) + 1 .
Corollary A1.
Assume that β C 2 ( R ) satisfies (6) and (7). Let μ 0 , z L 2 ( Q ) , f L 2 ( Q ) and y 0 L 2 ( Ω ) . Also, let g L ( 0 , T ) , b 0 L ( Q ) , ϑ L ( Q ) and a 0 L ( Q ) be such that (5), (16), (26) and (27) hold. Then, there exists a unique weak solution y W ( 0 , T ; H 1 ( Ω ) ) to (A1) provided
Ω y 0 ϕ ( · , 0 ) d x + Q f ϕ d x d t = 0 T ϕ t , y ( H 1 ( Ω ) ) , H 1 ( Ω ) d t + Q ϕ y d x d t + μ Q β ( h z g ) y ϕ d x d t ω T v y ϕ d x d t + Q b 0 0 t g ( s ) y ( x , s ) d s ϕ d x d t + Q a 0 y ϕ d x d t + Q g ( t ) T t T ϑ ( x , s ) y ( x , s ) d s ϕ d x d t
holds for every ϕ H ( Q ) : = { φ W ( 0 , T ; H 1 ( Ω ) ) such that ϕ ( · , T ) = 0 } . Moreover, there exists C = C ( a , b , ϑ , g , v L ( ω T ) , T ) > 0 such that
y L 2 ( 0 , T ; H 1 ( Ω ) ) C f L 2 ( Q ) + y 0 L 2 ( Ω )
and there exists C = C M , μ , g , b , ϑ , a , v L ( ω T ) , T > 0 such that
y W ( 0 , T ; H 1 ( Ω ) ) C f L 2 ( Q ) + y 0 L 2 ( Ω ) .

References

  1. Chipot, M. On a nonlocal-in-time parabolic problem. Adv. Math. Sci. Appl. 2022, 31, 1–10. [Google Scholar]
  2. Starovoitov, V.N. Boundary value problem for a global-in-time parabolic equation. Math. Methods Appl. Sci. 2021, 44, 1118–1126. [Google Scholar] [CrossRef]
  3. Starovoitov, V.N. Weak solvability of a boundary value problem for a parabolic equation with a global-in-time term that contains a weighted integral. J. Elliptic Parabol. Equ. 2021, 7, 623–634. [Google Scholar] [CrossRef]
  4. Casas, E.; Mateos, M. Critical cones for sufficient second order conditions in PDE constrained optimization. SIAM J. Optim. 2020, 30, 585–603. [Google Scholar] [CrossRef]
  5. Ebenbeck, M.; Knopf, P. Optimal medication for tumors modeled by a Cahn–Hilliard–Brinkman equation. Calc. Var. Partial. Differ. Equ. 2019, 58, 1–31. [Google Scholar] [CrossRef]
  6. Frankowska, H.; Marchini, E.; Mazzola, M. Second-order sufficient conditions in optimal control of evolution systems. J. Evol. Equ. 2024, 24, 40. [Google Scholar] [CrossRef]
  7. Frankowska, H.; Marchini, E.; Mazzola, M. Necessary optimality conditions for infinite dimensional state constrained control problems. J. Differ. Equ. 2018, 264, 7294–7327. [Google Scholar] [CrossRef]
  8. Garcke, H.; Lam, K.F.; Rocca, E. Optimal control of treatment time in a diffuse interface model of tumor growth. Appl. Math. Optim. 2018, 78, 495–544. [Google Scholar] [CrossRef]
  9. Kenne, C.; Djomegne, L.; Mophou, G. Optimal control of a parabolic equation with a nonlocal nonlinearity. J. Differ. Equ. 2024, 378, 234–263. [Google Scholar] [CrossRef]
  10. Matei, A.; Micu, S. Boundary optimal control for nonlinear antiplane problems. Nonlinear Anal. Theory Methods Appl. 2011, 74, 1641–1652. [Google Scholar] [CrossRef]
  11. Tang, H.; Yuan, Y. Optimal control for a chemotaxis–haptotaxis model in two space dimensions. Bound. Value Probl. 2022, 2022, 79. [Google Scholar] [CrossRef]
  12. Lions, J.L. Optimal Control of Systems Governed by Partial Differential Equations; Springer: Berlin/Heidelberg, Germany; New York, NY, USA, 1971. [Google Scholar]
  13. Lions, J.L.; Magenes, E.; Magenes, E. Problèmes aux Limites non Homogènes et Applications; Dunod Paris: Paris, France, 1968; Volume 1. [Google Scholar]
  14. Lions, J.L. Quelques Méthodes de Résolution des Problèmes aux Limites Non-Linéaires. Dunod 1969, 1. [Google Scholar]
  15. Nirenberg, L. On elliptic partial differential equations. In Il Principio di Minimo e sue Applicazioni alle Equazioni Funzionali; Springer: Berlin/Heidelberg, Germany, 2011; pp. 1–48. [Google Scholar]
  16. Breitenbach, T.; Borzì, A. A sequential quadratic Hamiltonian method for solving parabolic optimal control problems with discontinuous cost functionals. J. Dyn. Control. Syst. 2019, 25, 403–435. [Google Scholar] [CrossRef]
  17. Wu, Z.; Yin, J.; Wang, C. Elliptic & Parabolic Equations; World Scientific: Singapore, 2006. [Google Scholar]
  18. Aronna, M.S.; Tröltzsch, F. First and second order optimality conditions for the control of Fokker–Planck equations. ESAIM Control. Optim. Calc. Var. 2021, 27, 15. [Google Scholar] [CrossRef]
  19. Casas, E.; Tröltzsch, F. Second order analysis for optimal control problems: Improving results expected from abstract theory. SIAM J. Optim. 2012, 22, 261–279. [Google Scholar] [CrossRef]
  20. Tröltzsch, F. Optimal Control of Partial Differential Equations: Theory, Methods, and Applications; American Mathematical Soc.: Providence, RI, USA, 2010; Volume 112. [Google Scholar]
  21. Kenne, C.; Djomegne, L.; Zongo, P. Second-order optimality conditions for the bilinear optimal control of a degenerate parabolic equation. arXiv 2022, arXiv:2212.11046. [Google Scholar]
  22. Kenne, C.; Mophou, G.; Warma, M. Bilinear optimal control for a fractional diffusive equation. arXiv 2022, arXiv:2210.17494. [Google Scholar]
  23. Lions, J.L. Equations Différentielles Opérationnelles: Et Problèmes aux Limites; Springer: Berlin/Heidelberg, Germany, 2013; Volume 111. [Google Scholar]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mophou, G.; Fournier, A.; Jean-Alexis, C. Bilinear Optimal Control for a Nonlinear Parabolic Equation Involving Nonlocal-in-Time Term. Axioms 2025, 14, 38. https://doi.org/10.3390/axioms14010038

AMA Style

Mophou G, Fournier A, Jean-Alexis C. Bilinear Optimal Control for a Nonlinear Parabolic Equation Involving Nonlocal-in-Time Term. Axioms. 2025; 14(1):38. https://doi.org/10.3390/axioms14010038

Chicago/Turabian Style

Mophou, Gisèle, Arnaud Fournier, and Célia Jean-Alexis. 2025. "Bilinear Optimal Control for a Nonlinear Parabolic Equation Involving Nonlocal-in-Time Term" Axioms 14, no. 1: 38. https://doi.org/10.3390/axioms14010038

APA Style

Mophou, G., Fournier, A., & Jean-Alexis, C. (2025). Bilinear Optimal Control for a Nonlinear Parabolic Equation Involving Nonlocal-in-Time Term. Axioms, 14(1), 38. https://doi.org/10.3390/axioms14010038

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop