Open-access Hierarchical Control for the Oldroyd Equation in Memoriam to Professor Luiz Adauto Medeiros

Abstract

This manuscript deals with a hierarchical control problem for Oldroyd equation under the Stackelberg-Nash strategy. The Oldroyd equation model is defined by non-regular coefficients, that is, they are bounded measurable functions. We assume that we can act in the dynamic of the system by a hierarchy of controls, where one main control (the leader) and several additional secondary control (the followers) act in order to accomplish their given tasks: controllability for the leader and optimization for followers. We obtain the existence and uniqueness of Nash equilibrium and its characterization, the approximate controllability with respect to the leader control, and the optimality system for leader control.

Key words Hierarchical controllability; Stackelberg-Nash strategies; Oldroyd fluid; optimality system

Introduction

Bibliographical comments

With origin in game theory, and mainly motivated by economics, there exists several equilibrium concepts for multi-objective controllability of PDE. Each of them determines a strategy. For example, we mention the noncooperative optimization strategy proposed by Nash (1951), the Pareto cooperative strategy (1896), and the Von Stackelberg hierarchical-cooperative strategy (1934). The process in the problems above is a combination of strategies and is called Stackelberg-Nash strategy. The concept of hierarchical control was introduced by Lions (1994), where a simplified structure involving a single leader and a single follower control was considered to solve a problem of controllability for a hyperbolic equation.

For other equations, the hierarchical controllability has been considerably investigated. In the context of approximate controllability, we can cite some works involving Stackelberg-Nash strategy. In fact, in the paper by Díaz & Lions (2005), the approximate distributed controllability of a parabolic system has been established following a Stackelberg-Nash strategy, and in Límaco et al. (2009), this same strategy was developed to obtain the approximate controllability for the linear heat equation with moving boundaries. In the context of linear fluid models, some investigations into approximate controllability using Stackelberg-Nash strategies began with the results of Guillén-González et al. (2013) for the Stokes system. These were later expanded upon by Araruna et al. (2014) for linearized micropolar fluids. Subsequently, Jesus & Menezes (2015) extended the results of Araruna et al. (2014) to include moving domains. In both cases, the objective of the leader control is an approximate controllability result.

In recent years, much attention has been given to the investigation of new classes of problems in differential equations of hydrodynamics. Control problems for the Navier-Stokes equations and other models of fluid mechanics are examples of these . A considerable number of papers and books (see, for instance, Abergel & Temam (1990), Betts (2001), Blieger (2007), Fursikov (2000), Gunsburger (2003), Lions & Zuazua (1998), Stavre (2002), and references therein) deal with the theoretical and numerical study of the above mentioned problems. More precisely, in the context of fluid mechanics, significant controllability results are associated to the Burgers, Stokes, Euler, and Navier-Stokes equations. For instance, the local null controllability of the Burgers equation with distributed controls was investigated in Fernández-Cara & Guerrero (2006), and recent work by Araruna et al. (2024) explores null control for Burgers equations within hierarchical controllability using the Stackelberg-Nash strategy. For Stokes equations, the approximate and null controllability with distributed controls have been established in Fabre (1996) and Imanuvilov (2001), respectively. Additionally, non-null controllability of Stokes equations with memory was analyzed by Fernández-Cara et al. (2020). Global controllability results for Euler equations were proven by Coron (1996) and Glass (2000). Regarding Navier-Stokes equations, while only local exact controllability results are available for equations with initial and Dirichlet boundary conditions, as documented in Fursikov & Imanuvilov (1999), Imanuvilov (2001), Fernández-Cara et al. (2004, 2006), Coron et al. (2020) established a global exact controllability result for equations with Navier-slip (friction) boundary conditions.

The main novelties of this paper lie in the formulation and resolution of Nash and Stackelberg-Nash control problems within the framework of partial differential equations (PDEs) governing a non-conventional fluid model, specifically the equations governing Oldroyd fluids . The results obtained here can generate several interesting problems generalizing or improving the results to other similar models, for instance, this can be viewed as a first step in the path to understand similar questions in the context of the Navier-Stokes system and their variants.

The Oldroyd model

This model corresponds to an incompressible fluid which is described by the following system of partial differential equations:

ut+u.u+p=div(τ )+F(x,t),xΩ ,t>0, div(u)=0,xΩ ,t>0,(1)
with appropriate initial and boundary conditions. Here, Ω is an open bounded connected set of n with smooth boundary Γ, τ=(τik) denotes the stress tensor with trτ=0, u represents the velocity vector, p is the pressure of the fluid and F is an external force. The stress tensor τ plays a special role because the introduction of τ in (1) has the purpose of letting us consider reactions arising in the fluid during its motion. By establishing (Hooke’s Law) the connection between τ, the tensor of deformation velocities D=(Dik)=12(uixk+ukxi) and their derivatives, we thus establish the type of fluid. Such relation between τ and D is called a defining or rheological equation or an equation of state (see Serrin 1959). From Newton’s law, we set:
τ=2νD,(2)
where ν is the kinematic coefficient of viscosity. In this case, the fluid is called a Newtonian Fluid. Substituting (2) into (1) we obtain the equations of motion of Newtonian fluids, which is called Navier-Stokes equations.

Over the last century and half, the model of a Newtonian fluid has been the basic model of a viscous incompressible fluid. It describes flows of moderate velocities of the majority of viscous incompressible fluids encountered in practice. However, even earlier in the mid-nineteenth century it was known that there exists viscous incompressible fluid not subject to the Newtonian equation (2) . That is, it has a complex microstructure such as biological fluids, suspensions and liquid crystals, which are used in the current industrial process and shows (nonlinear) viscoelastic behavior that cannot be described by the classical linear viscous Newtonian models. The first models of such fluids, were proposed in the nineteenth century by Maxwell (1859, 1868), Kelvin(1875), Voigt (1889, 1892). In the mid-twentieth century, Oldroyd extended such models (see Oldroyd 1950, 1953, 1959, 1964).

The model for Oldroyd fluid (see Astarita & Marruci 1976, Wilkinson 1960) can predict the stress relaxation as well as the retardation of deformation. Due to this, it has become popular for describing polymer suspension. To model the behavior of a dilute polymer solution in a Newtonian solvent, the extra stress tensor is often split into two components: a viscoelastic one and a purely viscous one. So the Oldroyd fluids of order one as it is known in the Russian literature (see Oskolkov 1989, Oskolkov & Akhmatov 1974, Oskolkov & Kotsiolis 1986, Oskolkov et al. 1987) are described by defining relation:

(1+λt)τ=2ν(1+kν1t)D,(3)
where λ,ν,k are positive constants with νkλ1>0. Here, ν denotes the kinematic viscosity, λ is the relaxation time, and k represents the retardation time.

We observe that (3) can be rewritten in the form of an integral equation as follows:

τ(x,t)=2kλ1D(x,t)+2λ1(νkλ1)0te(tσ)λD(x,σ)dσ,(4)
where τ(x,0)=0.

Thus, the equation of motion of the Oldroyd fluid of first order can be described most naturally by the system of integro-differential equations:

ut+(u.)uμ Δ u0tg(tσ )Δ u(x,σ )dσ +p=F(x,t),xΩ ,t>0,(5)
and the incompressible condition:
div(u)=0,xΩ,t>0,(6)
with initial and boundary conditions:
u(x,0)=u0,xΩ,andu(x,t)=0xΓ,t0.(7)
Here μ=kλ1 and kernel g(t)=γeδt, where γ=λ1(νkλ1), with δ=λ1. For details of the physical background and its mathematical modeling, er refer to Oldroyd (1953), Astarita & Marruci (1976), Wilkinson (1960), and Oskolkov (1989).

Remark 0.1.. As the theory of viscoelastic fluids describes flows with moderate velocities, the equation (5) admits a reasonable simplification, i.e., the convective term (u.)u is neglected, as usual in mechanics.

Notations

As in Temam (1979) let us denote Hm(Ω) as the standard Hilbert-Sobolev space and by .m the norm defined on it. When m=0, we call H0(Ω) as the space of square-integrable functions L2(Ω) with the usual norm |.| and inner product (.,.). Further, let H01(Ω) be the completion of C0(Ω) concerning H1(Ω)-norm.

Let us consider the spaces:

𝒱:={ϕ(C0(Ω))n:div(ϕ)=0inΩ},H:=the closure of𝒱in(L2(Ω))nspace,
and
V:=the closure of𝒱in(H01(Ω))nspace.
The spaces of vector functions are indicated by boldface, for instance, 𝐇01(Ω)=(H01(Ω))n, 𝐋2(Ω)=(L2(Ω))n. The inner products on 𝐇01(Ω) and 𝐋2(Ω) are defined by:
((ϕ,w)):=i=1n(ϕi,wi)and(ϕ,w):=i=1n(ϕi,wi),
respectively. Similarly , we define the norms:
ϕ:=(i=1n|ϕi|2)12and|ϕ|:=(i=1n|ϕi|2)12.

Let us notice that under some smoothness assumptions on the boundary Γ, it is possible to characterize H and V:

H={u𝐋2(Ω);div(u)=0,andu.η|Γ=0}andV={u𝐇01(Ω);div(u)=0},
where η=η(x) is the outward unit normal vector at xΓ. By Poincaré’s inequality, it can be shown that the norm 𝐇01(Ω) is equivalent to 𝐇1(Ω)=(H1(Ω))nnorm. By V we denote the dual of V.

Main Result

Let T>0 be a real number. We consider the cylindrical domain Q:=Ω×(0,T) of n+1 with lateral boundary :=Γ×(0,T). We denote by 𝒪,𝒪1,𝒪2,,𝒪N non-empty disjoint open subsets of Ω. By χ𝒪, χ𝒪1, χ𝒪2, …, χ𝒪N we represent the characteristic functions of 𝒪, 𝒪1, 𝒪2, …, 𝒪N, respectively.

In this paper, we investigate the approximate controllability of the following system:

|utμΔu0tg(tσ)Δu(σ)dσ+p=vχ𝒪+i=1Nwiχ𝒪iinQ,div(u)=0inQ,u=0onΣ,u(x,0)=u0(x)inΩ,(8)
where u(x,t)=(u1(x,t),,un(x,t)) is the velocity vector (or state of the system) of moderate fluid evaluated at the point (x,t),x=(x1,...,xn)n, p=p(x,t) is the pressure of the fluid evaluated at the point (x,t), μ represents a constant, and u0(x) is the initial velocity.

The system (8) can be interpreted as a variant of the classical Oldroyd equations (5)-(7) , in which the nonlinearity (u.)u has been omitted; see Remark 0.1.

In (8) the subset 𝒪Ω is the main control domain (which is supposed to be as small as desired), 𝒪1, 𝒪2, …, 𝒪N are the secondary control domains, the function v is called leader control, and wi, (i=1,2,,N), are the followers controls.

Remark 2. By linearity of the system (8) without loss of generality, we may assume that u0=0.

As the solution u of (8) depends on v,w1,,wN then we denote it by u=u(x,t,v,𝐰), where 𝐰=(w1,,wN), or sometimes by u=u(x,t,v,w1,,wN).

To localize the action of the controls wi, (i=1,2,,N), we introduce the functions ρi(x), defined in Ω with real values, satisfying:

ρiL(Ω),ρi0,ρi=1inGiΩ,(9)
where Gi is a region where wi works.

We assume the leader objective to be of controllability type. On the other hand, the main objective of the followers is to hold u(x,t,v,𝐰), solution of the state equation (8) at the time T, near to a desired state uT(x), without a big cost for the controls w1,,wN, with cost functionals defined by:

Ji(v,𝐰):=12|wi|(L2(𝒪i×(0,T)))n2+αi2|ρiu(.,T,v,𝐰)ρiuT(.)|H2,(10)
where αi is a positive constant, 𝐰=(w1,....,wN), v(L2(𝒪×(0,T)))n, and wi varies in (L2(𝒪i×(0,T)))n, with i=1,2,,N.

The Stackelberg-Nash strategy is described as follows: for each choice of the leader v, we search for a Nash equilibrium for the cost functionals J1,,JN, that is, we look for controls w1,...,wN, depending on v, satisfying:

Ji(v,w1,,wN)Ji(v,w1,,ŵi,,wN)for allŵi(L2(𝒪i×(0,T)))n.(11)

The controls w1,,wN, solutions of the system of N inequalities (11), are called Nash equilibrium for the costs J1,,JN and they depend on v (cf. Aubin 1984).

For each i=1,2,,N, assuming that ρiL(Ω), and αi is small enough then there exists a unique Nash equilibrium w1,,wN, depending on v, given by the inequalities (11) Later, we will make explicit this assumption with more details.

The main problems to be answered in this paper can be read as follows:

  • Problem 1 The existence of solutions w1,,wN for the inequalities (11) that is, the existence of the Nash equilibrium for the functionals J1,,JN;

  • Problem 2 Assuming that the existence of the Nash equilibrium w1(v),...,wn(v) was proved, then when v varies in (L2(𝒪×(0,T)))n, to prove that the solutions u(x,t,v,𝐰(v)) of the state equation (8), evaluated at t=T, that is, u(x,T,v,𝐰(v)), generate a dense subset of H. This permits to approximate uT.

Let us consider the following set:

R(T):={u(x,T,v,w(v)),v(L2(O× (0,T)))nwhereu(x,t,v,w(v))is astrong solution of (8)withf=vχ O+i=1Nwi(v)χ Oiandu0=0.}

Thus, the main result of this paper is the following:

Theorem 1. Let T>0. Let us assume that for every v(L2(𝒪×(0,T)))n, there exists a unique Nash equilibrium w1,,wN, depending on v, given by the inequalities (11) that is, for each i=1,2,,N, assume that ρiL(Ω), and αi is small enough. Then R(T) is dense in H.

Organization of the paper

This paper is organized as follows. Initially, we present the optimality system for the followers controls. After this, we investigate the approximate controllability proving the density Theorem 0.1. The following section is devoted to establish the existence and uniqueness of the Nash equilibrium by using the Lax-Milgram’s Lemma. Posteriorly, we deal with the optimality system for the leader control. After this, we add some comments and point out open problems related to this article. Finally, in the last section, we present the Appendix, where we prove the existence and uniqueness of solutions.

Optimal system for the followers controls

The main objective here is to express the followers controls wi, (i=1,2,,N), as weak solutions of an appropriate system. For this, we suppose that there exists a Nash equilibrium 𝐰=(w1,,wN) for the cost functionals Ji,(i=1,2,....,N), defined in (10) This means that 𝐰=(w1,,wN) satisfies the Euler-Lagrange equation given by:

Ji(v,𝐰(v)).wî=0for allŵi(L2(𝒪i×(0,T)))n,(12)
where
Ji(v,𝐰(v)).wî=ddλJi(v,w1,...,wi+λŵi,...,wN)|λ=0.

The equation (12) will allow us to reach to the objective of this section. In fact, from (10) we have that:

Ji(v,w1,..,wi+λwî,,wN)=12(wi+λwî,wi+λwî)(L2(𝒪i×(0,T)))n

+αi2(ρiu(.,T,v,w1,,wi+λwî,,wN)ρiuT(.),ρiu(.,T,v,w1,,wi+λwî,,wN)ρiuT(.))H.
As the state equation (8) is linear, for any choice of the controls v,wi, its unique solution at the time T can be written as u(T)=L0v+i=1NLiwi, where Li are linear and continuous operators. Later, we will make explicit this assumption with more details. Thus, the equality above can be rewritten as follows: Ji(v,w1,..,wi+λwî,...,wN)=12(wi+λwî,wi+λwî)(L2(𝒪i×(0,T)))n
+αi2(ρiL(wi+λwî)ρiuT(.),ρiL(wi+λwî)ρiuT(.))H,(13)
where L(wi+λwî):=L0v+L1w1+...+Li(wi+λŵi)+...+LNwN.

Now, differentiating the equation (13) with respect to λ and evaluate at λ=0, we obtain: ddλJi(v,w1,,wi+λwî,,wN)|λ=0=(wi,wî)(L2(𝒪i×(0,T)))n

+αi(ρi2[u(.,T,v,w1,,wi,,wN)uT(.)],Liŵi)H,
for all ŵiL2(𝒪i×(0,T)), where Liwî=ûi(.,T,wî) and ûi(x,t,ŵi) is the unique strong solution of system:
|ûitμΔûi0tg(tσ)Δûi(σ)dσ+1Np=wîχ𝒪iinQ,div(ûi)=0inQ,ûi=0onΣ,ûi(x,0)=0inΩ,(14)
associate to the wî with N>0 given.

Then, admitting (11) for the cost functionals Ji,(i=1,2,....,N), we obtain the following characterization:

(wi,wî)(L2(𝒪i×(0,T)))n+αi(ρi2[u(.,T,v,w1,,wi,,wN)uT(.)],ûi(.,T))H=0,(15)
for all ŵi(L2(𝒪i×(0,T)))n.

To find an optimality system for the followers, we represent by p̂=p̂(x,t)L2(0,T;L2(Ω)) and ψi(x,t)=(ψi1(x,t),,ψin(x,t)), (x,t)Q the weak solution of system:

|ψitμΔψitTg(ηt)Δψi(η)dη+p̂=0inQ,div(ψi)=0inQ,ψi=0onΣ,ψi(x,T)=ρi2[u(x,T,v,𝐰(v))uT(x)]inΩ,(16)
where the condition ψi(x,T) is motivated by (15).

Observe that if we make the change of variables τ=Tt in (16) and set φi(x,τ)=ψi(x,t), we transform (16) into an equivalent system in the unknown φi but with φi(0)=ρi2[u(.,T,v,𝐰(v))uT(.)], where ρi2[u(.,T,v,𝐰(v))uT(.)]H. Furthermore, since the right hand side from (16)1 is zero, then (16) admits a unique weak solution ψi such that:

ψitμΔψitTg(ηt)Δψi(η)dη+p̂=0,
in the sense of L2(0,T;V).

If ûi(x,t,ŵi) is the strong solution of (14), we have that:

ûi(.,.,ŵi)C0([0,T];V)L2(0,T;V𝐇2(Ω)).
Therefore, it makes sense to take the duality between
(ψi)=ψitμΔψitTg(ηt)Δψi(η)dη+p̂
and ûi:=ûi(x,t,ŵi) in L2(0,T;V)×L2(0,T;V).

Formally, we multiply both sides of (16) by the strong solution ûi of (14), and integrate in Q. Let us recall that we had assumed u0=0 in (8) . Then:

(ψi(.,T),ûi(T))H+Qψi{ûitμΔûi}dxdt+0Tp̂(t),ûi(t)(H1(Ω))n×(H01(Ω))ndt0TtTg(ηt)Δψi(η)dη,ûi(t)V×Vdtdt=0.
From (14) it follows that:
ûitμΔûi=ŵiχ𝒪i1Np+0tg(tσ)Δûi(σ)dσa.e. inQ.
Hence,
(ρi2[ûi(.,T,v,𝐰(v))uT(.)],ûi(.,T))H+Qψiŵiχ𝒪idxdt+0Tp̂(t),ûi(t)(H1(Ω))n×(H01(Ω))ndt+0T(0tg(tσ)Δûi(σ)dσ,ψi(t))dt0TtTg(ηt)Δψi(η)dη,ûi(t)V×Vdt1N0T(p(t),ψi(t))dt=0.(17)

Remark 3. Let us notice that:

p(t),u(t)(H1(Ω))n×(H01(Ω))n=(p(t),divu(t))L2(Ω)=0,
since u(t)V and thus div(u(t))=0.

Remark 0.4. By Fubini’s formula, we have that:

0T(0tg(tσ)Δu(σ)dσ,ψ(t))dt0TtTg(ηt)Δψ(η)dη,u(t)V×Vdt=0.

From Remarks 0.3 and 0.4, we have that the equation (17) becomes:

(ρi2[u(.,T,v,𝐰(v))uT(.)],ûi(.,T))H=Qψiŵiχ𝒪idxdt, for all ŵi(L2(𝒪i×(0,T)))n.
Hence, from (15), we deduce that:
(wi+αiψi,wî)(L2(𝒪i×(0,T)))n=0, for all ŵi(L2(𝒪i×(0,T)))n,
i.e.,
wi=αiψia.e.in𝒪i×(0,T).(18)

Remark 0.5. The relation (18) is very important. It corresponds to finding the control functions as a weak solution of system (16) The main reason to express the followers controls in such way is to find an optimality system. Using this fact, one can derive numerical approximation algorithms.

Thus, the best wi satisfying (11) are given by wi=αiψi, where ψi is the unique solution of the following optimality system:

|utμ Δ u0tg(tσ )Δ u(σ )dσ +p+i=1Nα iψ iχ Oi=vχ O in Q,ψ itμ Δ ψ itTg(η t)Δ ψ i(η )dη +p^ =0 in Q,div(u)=0,div(ψ i)=0 in Q,u=0,ψ i=0 on Σ ,u(x,0)=0,ψ i(x,T)=ρ i2[u(x,T,v,w(v))uT(x)] in Ω .(19)

On approximate controllability

Our goal in this section is to show the approximate controllability for the state equation (8) assuming that v(L2(𝒪×(0,T)))n, i.e., we will prove that the solution u(x,t,v,𝐰(v)) of the state problem (8) evaluated at t=T, generate a dense subset of H. Observing the cost functionals Ji,(i=1,2,....,N), defined in (10) we will have the approximate controllability with uTH as in Díaz & Lions (2005).

Now, we present the proof of Theorem 0.1.

By linearity of optimality system (19) without loss of generality, let us assume that uT0 (it suffices to use a translation argument). We will prove that if fH satisfies (u(.,T,v,𝐰(v)),f)H=0 for all v(L2(𝒪×(0,T)))n, then f is the null vector of H, i.e., the orthogonal complement of R(T) in H is null.

In fact, multiplying (19)1 by φ, (19)2 by ξi, respectively, and integrating in Q, we obtain:

(u(.,T,v,w(v)),φ (T))H+Qu{φ tμ Δ φ }dxdt+0T(p(t),φ (t))dt0T(0tg(tσ )Δ u(σ )dσ ,φ (t))dt+Qφ i=1Nα iψ iχ Oidxdt=Qφ vχ Odxdt,(20)
and
Qψ i{ξ itμ Δ ξ i}dxdt+(ψ i(0),ξ i(0))H+0Tp^ (t),ξ idt0TtTg(η t)Δ ψ i(η )dη ,ξ i(t)dt(ψ i(T),ξ i(T))H=0,(21)
where i=1,2,,N. Fixing the conditions:
|φ tμ Δ φ tTg(η t)Δ φ (η )dη +p^ =0 in Q,ξ itμ Δ ξ i0tg(tσ )Δ ξ i(σ )dσ +p=α iφ χ Oi in Q,div(φ )=0,div(ξ i)=0 in Q,φ =0,ξ i=0 on Σ ,ξ i(x,0)=0,φ (x,T)=f(x)+i=1Nρ i2[ξ i(x,T)] in Ω ,(22)
we conclude that:

  • From (22), the equation (20) can be rewritten as follows:

    (u(.,T,v,w(v)),φ (T))H0T(0tg(tσ )Δ u(σ )dσ ,φ (t))dt+0T(p(t),φ (t))dt0Tu(t),p^ (t)dt+Qi=1Nφ α iψ iχ Oidxdt+0TtTg(η t)Δ φ (η )dη ,u(t)dt=Qφ vχ Odxdt.(23)

  • From (22)2, the expression in (21) becomes:

( ψ i ( T ) , ξ i ( T ) ) H Q ψ i α i φ χ O i d x d t + 0 T p ^ ( t ) , ξ i d t 0 T t T g ( η t ) Δ ψ i ( η ) d η , ξ i ( t ) d t 0 T ( ψ i , p ( t ) ) d t + 0 T ( 0 t g ( t σ ) Δ ξ i ( σ ) d σ , ψ i ( t ) ) d t = 0. (24)

Remark 0.6. The system (22) admits a unique solution {φ,ξi,p̂,p} in C0([0,T];H)×C0([0,T];V)×L2(0,T;L2(Ω))×L2(0,T;H1(Ω)); see Appendix. Therefore, it makes sense the calculus in (20) and (21).

From (19)3, (22)3, (23), (24) and Remarks 0.3 and 0.4 we have that:

  • Equation (23) is reduced to the expression:

( u ( . , T , v , 𝐰 ( v ) ) , φ ( T ) ) H + Q i = 1 N φ α i ψ i χ 𝒪 i d x d t = Q φ v χ 𝒪 d x d t . (25)
  • Expression in (24) becomes:

( ψ i ( T ) , ξ i ( T ) ) H Q ψ i α i φ χ 𝒪 i d x d t = 0 . (26)

Now, summing (26) from 1 at N, we obtain that:

i=1N(ψi(T),ξi(T))H=Qi=1Nψiαiχ𝒪iφdxdt.(27)

Hence, substituting (27) in (25), we get:

(u(.,T,v,𝐰(v)),φ(T))H+i=1N(ψi(T),ξi(T))H=Qφvχ𝒪dxdt.(28)

From (19)5 and (22)5 we find that:

φ(T)=f+i=1Nρi2ξi(T),ψi(T)=ρi2u(T).(29)

Thus, combining (28) and (29), we deduce that:

(u(T),f)H+(u(T),i=1Nρi2ξi(T))Hi=1N(ρi2u(T),ξi(T))H=Qφvχ𝒪dxdt,

i.e,

(u(T),f)H=Qφvχ𝒪dxdt,for allv(L2(𝒪×(0,T)))n.(30)

From assumption (u(T),f)H=0, then from (30), it follows that:

Qφvχ𝒪dxdt=0,for allv(L2(𝒪×(0,T)))n,
and therefore,
φ=0in𝒪×(0,T)Q.(31)
It follows by the unique continuation (cf. Doubova & Fernández-Cara 2012, Lemma 2.3, p. 576), that φ=0inQ. Thus, coming back to (22), we obtain that ξi=0inQ, i=1,2,,N. Therefore, as φC0([0,T];H) and ξiC0([0,T];V), we have that ξi(x,T)=0 and φ(x,T)=0, and from (22)5 implies that f0 in Ω. This ends the proof of Theorem 0.1.

Existence and uniqueness of Nash equilibrium

Our goal in this section is to prove the existence of Nash equilibrium for the cost functionals Ji,(i=1,2,,N), defined in (10), corresponding to the state equation (8). For this, let Hi=(L2(𝒪i×(0,T)))n and =i=1NHi be Hilbert spaces. Moreover, we consider for each i=1,...,N, the cost functionals Ji, given by (10) and the operator Li:(L2(𝒪i×(0,T)))nV such that Liŵi=ûi(T), where ûi is a unique strong solution of the linear system (14) The functionals Li are linear and are well defined. Indeed, since ûi is the strong solution of the system (14) with ŵi, it follows that ûi belongs to C0([0,T];V). Hence, ûi(T)V. Moreover, ûiC0([0,T];V)CTŵiHi, and thus, ûi(T)VCTŵiHi, that is, LiŵiVCTŵiHi, for all i=1,,N. Therefore Li(Hi,V). Since VH with continuous and dense immersion, we have that |Liŵi|HC0LiŵiVC̃0ŵiHi, where C̃0=C0CT.

Let u be a strong solution of the problem (8) with u0=0. From (14), we have:

|i=1Nuiti=1Nμ Δ ui0tg(tσ )i=1NΔ ui(σ )dσ +i=1N1Np=i=1Nwi(v)χ Oi in Q,i=1Ndiv(ui)=0 in Q,i=1Nui=0 on Σ ,i=1Nui(x,0)=0 in Ω .(32)
Since v(L2(𝒪×(0,T)))n, we compare (8) with (32) and conclude that u=z+i=1Nui, where ui is the strong solution of (14) with wi=ŵi and z is fixed depending only of v.

Since uC0([0,T];V), we have that u(x,T,v,𝐰(v))=zT+i=1NLiwi, for zT fixed.

This notation allows us to rewrite the cost functionals Ji,(i=1,2,,N), defined in (10) in the form:

Ji(v,𝐰)=12|wi|Hi2+αi2|ρi(j=1NLjwjηT(.))|H2,(33)
where ηT=zTuT.

This means that 𝐰=(w1,,wN) will be a Nash equilibrium for the convex functionals Ji given by (10), if their Gateaux derivatives are null in any direction wî, i.e., we must show that 𝐰 satisfies:

(wi,wî)Hi+αi(ρi[j=1NLjwjηT(.)],ρiLiŵi)H=0,for allŵiHi,
or
(wi,wî)Hi+αi(Li*[ρi2j=1NLjwj],ŵi)Hαi(Li*(ρi2ηT),ŵi)H=0
for all ŵiHi, where Li*(H,Hi) is the adjoint operator of the Li(Hi,H). Hence,
wi+Li*[ρi2j=1NLjwj]=αiLi*(ρi2ηT)inHi,for alli=1,2,,N.(34)

Since αiLi*(ρi2ηT)Hi, we can aim to find a vector

(α1L1*(ρ12ηT),,αNLN*(ρN2ηT))
in . After that, we define the functional A: such that (Aw,ŵ)=(f,ŵ),f,ŵ, where A(,) is given by (Aw)i=wi+Li*[ρi2j=1NLjwj]. Thus, the problem can be formulated as the following question:

Question : Given f, is there a unique w such that (Aw,ŵ)=(f,ŵ),ŵ ?

Therefore, we must prove that the linear equation (Aw,ŵ)=(f,ŵ) admits a solution w=(w1,,wN) in for each f=(f1,....,fN) in . To guarantee its solvability, we will apply the Lax-Milgram’s Lemma, with certain restrictions on αi and ρi. More precisely, the following result holds:

Proposition 0.1. Let us assume that

ρ iL(Ω ),α i=α for i = 1,.., N,andα issmallenough
such that β0:=C̃02αmaxi,j=1,,NρiρjL(Ω)maxi=1,,NρiL(Ω) <1. Then there exists a Nash equilibrium 𝐰=(w1,,wN) for the functionals Ji defined in (10).

Let us observe that:

(Aw,w)=i=1N|wi|Hi2+i=1Nαi(j=1NρjLjwj,ρiLiwi)H+i,j=1Nαi((ρiρj)Ljwj,ρiLiwi)H.
Therefore, according to the hypothesis α=αi,for alli=1,2,,N, we obtain:
(Aw,w)=α|i=1NρiLiwi|H2+i=1N|wi|Hi2+αi,j=1N((ρiρj)Ljwj,ρiLiwi)H.(35)
Next, let us that:
α|i,j=1N((ρiρj)Ljwj,ρiLiwi)H|C̃02αmaxi,j=1,,NρiρjL(Ω)maxi=1,..,NρiL(Ω)|w|2.
Hence, from (35), we obtain:
(Aw,w)(1β0)|w|2,(36)
where β0=C̃02αmaxi,j=1,,NρiρjL(Ω)maxi=1,..,NρiL(Ω). Since α is small enough, we have γ0=(1β0)>0. Therefore, from (36), it follows that (Aw,w)γ0|w|2.

Finally, since A(,) and (Aw,w)γ0|w|2, we have from Lax-Milgram’s Theorem that, for a given f, there exists a unique w such that Aw=f. In particular, for f=(α1L1*(ρ12ηT),,αNLN*(ρN2ηT)), there exists a unique solution w, that is, a Nash equilibrium for the cost functionals Ji, satisfying (Aw)i=αiLi*(ρi2ηT),for alli=1,2,,N, i.e.,

wi+Li*[ρi2j=1NLjwj]=αiLi*(ρi2ηT)inHi,for alli=1,2,,N.

Optimal system for the leader control

In the previous sections we have seen that no matter what strategy the leader assumes, the followers make their choices w1,....,wN satisfying the Nash equilibrium. Moreover, this choices can be made through of an optimality system. The goal of this section is to obtain an optimality system for the leader control. For this, we consider the functional:

J(v)=12𝒪×(0,T)v2dxdt,
and the minimization problem
{infJ(v)subject tou(T,v)uT+εB,(37)
where ε>0 is a given real number, BB(0,1) is the unitary ball of H, and u(v) is the unique solution of the optimality system defined in (19)

We introduce two convex proper functions as follows:

  • The first one is defined in (L2(𝒪×(0,T)))n by:

    F1(v):=12𝒪×(0,T)v2dxdt.(38)

  • The second one is defined in H by:

    F2(f):={0iffuT+εB,+out.(39)

With these notations, problem (37) is equivalent to

infv(L2(𝒪×(0,T)))n{F1(v)+F2(L(v))},(40)
where L(v)=u(.,T,v), with L([L2(𝒪×(0,T))]n,H).

Observe that there exists v(L2(𝒪×(0,T)))n such that F1 and F2 are finite, F1 is continuous in v, and F2 is continuous in L(v). By the Duality Theorem of Fenchel and Rockafellar (1969) (see also Brezis 2010, Ekeland & Temam 1974), we have:

infv(L2(𝒪×(0,T)))n(F1(v)+F2(Lv))=inffH(F1*(L*f)+F2*(f)),(41)
where L* denotes the adjoint of L and Fi* is the conjugate function of Fi(i=1,2). This means that the primal problem (37) is equivalent to its dual.

Now, our next goal is to get the best control function of the leader. In fact, using (22) then for fH, it follows that:

(L(v),f)H=Qφvχ𝒪dxdt,for allv(L2(𝒪×(0,T)))n.
Thus, we conclude that:
L*f=φχ𝒪,(42)
where φ is a solution of (22)

We see easily that:

F1*(v)=F1(v),(43)
and
F2*(f)=(f,uT)H+εf,(44)
where . denotes the norm of H.

So, from (42) – (44), the expression in (41) becomes:

infv(L2(𝒪×(0,T)))n(F1(v)+F2(Lv))=inffHF(f),(45)

where the functional F:H is defined by:

F(f):=12𝒪×(0,T)(L*f)2dxdt+εf(f,uT)H,(46)
with L*f given by (42).

Let us notice that the problem (37) has a unique solution (see Rockafellar 1967). Consequently, the dual problem has also a unique solution.

After a quick computation of the Gateaux derivative of the functional (46), then for φ̂H, we obtain the following variational inequality (cf. Ekeland & Temam (1974)):

𝒪×(0,T)φ(φ̂φ)dxdt+εf̂εf(f̂f,uT)H0,f̂H.(47)
Now, for each fH, we consider the unique solution of (22) φ and introduce (u,ψi), i=1,2,,N, as the unique solutions of the system:
|utμ Δ u0tg(tσ )Δ u(σ )dσ +p+i=1Nα iψ iχ Oi=φ χ O in Q,ψ itμ Δ ψ itTg(η t)Δ ψ i(η )dη +p^ =0 in Q,div(u)=0,div(ψ i)=0 in Q,u=0,ψ i=0 on Σ ,u(x,0)=0,ψ i(x,T)=ρ i2[u(x,T,v)] in Ω .(48)
Multiplying (48)1 by φ̂φ, and (48)2 by ξ̂iξi, respectively, and integrating in Q, we obtain:
(u(.,T),φ̂(T)φ(T))H+0T(p(t),(φ̂φ))dt+Qu{(φ̂φ)tμΔ(φ̂φ)}dxdt0Tu(t),tTg(ηt)Δ(φ̂φ)(η)dηdt+Q(φ̂φ)i=1Nαiψiχ𝒪idxdt=Qφ(φ̂φ)χ𝒪dxdt,(49)
and
(ψi(0),(ξ̂i(0)ξi(0)))H+0Tp̂(t),(ξ̂iξi)dt+Qψi{(ξ̂iξi)tμΔ(ξ̂iξi)}dxdt0T(ψi(t),0tg(tσ)Δ(ξ̂iξi)(σ)dσ)dt+(ψi(T),ξ̂i(T)ξi(T))H=0,(50)
where i=1,2,,N.

Setting φ=φ̂φ, ξi=ξ̂iξ, and fixing the conditions:

|φ ¯ tμ Δ φ ¯ tTg(η t)Δ φ ¯ (η )dη =0 in Q,ξ ¯ itμ Δ ξ ¯ i0tg(tσ )Δ ξ ¯ i(σ )dσ =α iφ ¯ χ Oi in Q,div(φ ¯ )=0,div(ξ ¯ i)=0 in Q,φ ¯ =0,ξ ¯ i=0 on Σ ,ξ ¯ i(x,0)=0,φ ¯ (x,T)=f^ (x)f(x)+i=1Nρ i2[ξ ¯ i(x,T)] in Ω ,(51)
we deduce that:
(u(.,T),φ̂(T)φ(T))H+0T(p(t),φ̂(t)φ(t))dt+Qi=1N(φ̂(t)φ(t))αiψiχ𝒪idxdt=Qφ(φ̂φ)χ𝒪dxdt(52)
and
(ψi(T),ξ̂i(T)ξi(T))HQψiαi(φ̂(t)φ(t))χ𝒪idxdt+0Tp̂(t),ξ̂iξidt=0,(53)
where i=1,2,,N. From Remark 0.3, we have that the expressions in (52) and (53) becomes:
(u(.,T),φ̂(T)φ(T))H+Qi=1N(φ̂(t)φ(t))αiψiχ𝒪idxdt=Qφ(φ̂φ)χ𝒪dxdt(54)
and
(ψi(T),ξ̂i(T)ξi(T))HQψiαi(φ̂(t)φ(t))χ𝒪idxdt=0,(55)
where i=1,2,,N. Summing (55) from 1 to N and substituting into (54), we get:
(u(.,T),φ̂(T)φ(T))H+i=1N(ψi(T),ξ̂i(T)ξi(T))H=Qφ(φ̂φ)χ𝒪dxdt.(56)
From (48)5 and (51)5, it follows that:
φ̂(T)φ(T)=f̂f+i=1Nρi2(ξ̂i(T)ξi(T)),ψi(T)=ρi2u(.,T).(57)
Combining (56) and (57), we find that:
(u(.,T),f̂f)H=𝒪×(0,T)φ(φ̂φ)dxdt.(58)
Substituting (58) in (47), we get:
(u(.,T)uT,f̂f)H+εf̂εf0,for allf̂H.(59)
More precisely, we summarize these results in the following theorem:

Theorem 2. The best control function v of the leader, that is, the function that minimizes

12𝒪×(0,T)v2dxdt
subject to u(.,T,v,𝐰(v))uT+εB, is given by:
v=φχ𝒪,
where φ is given from unique solution {u,ψi,φ,ξi} of the optimality system:
|utμ Δ u0tg(tσ )Δ u(σ )dσ +p+i=1Nα iψ iχ Oi=φ χ O in Q,ψ itμ Δ ψ itTg(η t)Δ ψ i(η )dη +p^ =0 in Q,φ tμ Δ φ tTg(η t)Δ φ (η )dη +p^ =0 in Q,ξ itμ Δ ξ i0tg(tσ )Δ ξ i(σ )dσ +p=α iφ χ Oi in Q,div(φ )=0,div(ξ i)=0 in Q,div(u)=0,div(ψ i)=0 in Q,φ =0,ξ i=0,u=0,ψ i=0 on Σ ,ξ i(x,0)=0,φ (x,T)=f(x)+i=1Nρ i2[ξ i(x,T)] in Ω ,u(x,0)=0,ψ i(x,T)=ρ i2[u(x,T,v)] in Ω ,
with f being the unique solution of the variational inequality:
(u(.,T,f)uT,f̂f)H+εf̂εf0,for allf̂H.(60)

Remark 7. In (60) we rewrite u(.,T,f)to emphasize the fact that the solutions{u,ψi,φ,ξi}of the optimality system also depends onf.

Conclusions

In this this section we make some comments and briefly discuss some possible extensions of our results and also indicate open issues on the subject.

Extensions to various analogous scenarios

We have seen in this paper that it is possible to obtain approximate controllability result for the Oldroyd equation following a Stackelberg-Nash strategy. The point of adding secondary controls w1,...,wN consists in the task of being a low cost control that ensures that the solution u of (8) is not far from ideal state uT. Importantly, the concepts and techniques employed in this paper can be applied to various analogous scenarios, such as hierarchical controllability for linear and semilinear parabolic and hyperbolic equations, different types of non-cylindrical control domains, similar boundary control problems, among others. Indeed, concerning these themes, several prior studies are noteworthy. For example, in Límaco et al. (2009) the authors present the hierarchical approximate controllability for the linear heat equation in a non-cylindrical domain. Jesus (2015) demonstrated hierarchical approximate controllability for linearized micropolar fluids in moving domains, employing a Stackelberg-Nash strategy. In the case of an Oldroyd fluid system, the exact controllability of Galerkin’s approximations is proposed and analyzed in Marinho et al. (2014). In Araruna et al. (2015), the authors established a Stackelberg-Nash strategy with exact controllability for the leader control in semilinear parabolic equations, utilizing Carleman inequalities. For semilinear hyperbolic equations, a Stackelberg-Nash strategy with exact controllability for the leader control is proved in Araruna et al. (2018).

Extensions to Navier-Stokes equations

The controllability of Navier-Stokes equations have received significant attention in recent years, as evidenced by the extensive study documented in Coron et al. (2020) along with the latest advancements. However, to our knowledge, there exists a gap in the literature regarding null controllability for these equations. For further insights, see Araruna et al. (2015), p. 20, or more recent discussions in Araruna et al. (2024), p. 15. This unresolved issue presents an intriguing and pertinent avenue for future research.

Extension to nonlinear Oldroyd system

The theoretical analysis of nonlinear Oldroyd systems has been a focal point of research efforts in recent years, as evidenced by the works of Lions & Masmoudi (2000), Fernández-Cara et al. (2002), Galdi (2008), and Renardy (2009). Notably, Fernández-Cara et al. (2020), p. 4, emphasize the significance of investigating Oldroyd fluids governed by nonlinear partial differential equations. Consequently, it is interesting to explore how the findings of this paper might be applicable in the context of the nonlinear Oldroyd fluids system, for instance, in the context of the equations (5)– (7). Note that this system in question is more difficult to solve than the system (8). The main reason is the presence of the nonlinear term (u.)u in (5). However, whether or not the results in this paper can be extended to this framework is at present an open question.

The case with a different definition for cost functionals (10)

In Guillén-González et al. (2013), the authors solve an approximate control problem for the Stokes equation using the Stackelberg-Nash strategy. In this article, the cost functionals for the followers controls vi,(i=1,2), are defined by:

Ji(f,v1,v2):=αi2ωi,d×(0,T)|uui,d|2dxdt+μi2ωi×(0,T)|vi|2dxdt,(61)
where αi,μi are positive constants, and ui,d are functions given in L2(ωi,d×(0,T)).

Thus, a priori, the leader and the followers have different tasks, with the followers’s task being to prevent the state function from deviating from a given function.

Note that the situation considered in this article is different: here, all the controls have the same objective. However, we assume a situation where depending on the region, the control may need to have a different configuration, that is, if we have two regions, we will need two controls. Therefore, similar to the article by Guillén-González et al. (2013), we assume a hierarchy among the controls and employ with the Stackelberg optimization (cooperative) strategy, with a leader and followers. In the case of three or more regions, we will have one leader and N followers (N>1), and they will act without collaboration among themselves, hence we use the Nash strategy.

Moreover, if we follow the same procedures as in this article but with the functionals defined as in (61), we will have the followers controls characterized by the following optimality system:

|utμΔu0tg(tσ)Δu(σ)dσ+p=vχ𝒪i=121μiqiχωiinQ,qitμΔqitTg(σt)Δqi(σ)dσ+p̂=αi(uui,d)χωi,dinQ,div(u)=0,div(qi)=0inQ,u=0,qi=0onΣ,u(x,0)=u0,qi(x,T)=0inΩ.(62)
Finally, we can solve the approximate control problem for the system (62) following the same ideas as in this article.

Non null controllability for the system (8)

It is expectable that the system (8) (with a nonzero initial data for u) is not null controllable, in view of other previous results on the controllability of parabolic systems with memory; see Guerrero & Imanuvilov (2013) for more details. Thus, it would also be quite interesting to obtain a result asserting that null controllability for the system 8 does not hold in general, which we plan to present in a forthcoming paper.

On the other hand, exact controllability problems also can be considered in this context. In general, exact controllability does not hold. Another interesting issue is what happens as the viscosity coefficient goes to zero. It is known that, for μ=0, exact controllability holds at least under some geometric control conditions. For interested readers on this subject, we cite for instance Boldrini et al. (2012).

ACKNOWLEDGMENTS

The authors want to express their gratitude to the anonymous reviewers for their questions and commentaries; they were very helpful in improving this article. Isaías Pereira de Jesus was supported by CNPq/Brazil grants 307488/2019-5, 305394/2022-3 and PRPG (UFPI).

References

  • ABERGEL F & TEMAM R. 1990. On some control problem in fluid mechanics. Theor Comput Fluid Dyn 1: 303-325.
  • ARARUNA FD, DE MENEZES SD & ROJAS-MEDAR MA. 2014. On the approximate controllability of Stackelberg-Nash strategies for linearized micropolar fluids. Applied Math & Optimization 70: 373-393.
  • ARARUNA FD, FERNÁNDEZ-CARA E & DA SILVA LC. 2018. Hierarchic control for the wave equation. Journal of Optimization Theory and Applications 178: 264-288.
  • ARARUNA FD, FERNÁNDEZ-CARA E & DA SILVA LC. 2024. Bi-objective and hierarchical control for the Burgers equation. J Evol Equations 24: 30.
  • ARARUNA FD, FERNÁNDEZ-CARA E & SANTOS MC. 2015. Stackelberg-Nash exact controllability for linear and semilinear parabolic equations. ESAIM Control Optim Calc Var 21: 835-856.
  • ASTARITA G & MARRUCI G. 1976. Principles of non-Newtonian fluid mechanics. McGraw-Hill.
  • AUBIN JP. 1984. L’analyse non linéaire et ses économiques. Paris: Masson.
  • BETTS J. 2001. Practical methods for optimal control using nonlinear programming. Philadelphia: SIAM.
  • BLIEGER L. 2007. Real-time PDE-constraints optimization. Philadelphia: SIAM.
  • BOLDRINI JL, DOUBOVA A, FERNÁNDEZ-CARA E & GONZÁLEZ-BURGOS M. 2012. Some Controllability Results for Linear Viscoelastic Fluids. SIAM Journal on Control and Optimization 50: 900-924.
  • BREZIS H. 2010. Functional analysis, Sobolev spaces and partial differential equations. Springer-Verlag.
  • CORON JM. 1996. On the controllability of 2-D incompressible perfect fluids. J Math Pures Appl 9(75): 155-188.
  • CORON JM, MARBACH F & SUEUR F. 2020. Small-time global exact controllability of the Navier-Stokes equation with Navier slip-with-friction boundary conditions. J Eur Math Soc 22 (5): 1625-1673.
  • DÍAZ JI & LIONS J-L. 2005. On the approximate controllability of Stackelberg-Nash strategies. In: Díaz JI (Ed), Ocean Circulation and Pollution Control Mathematical and Numerical Investigations. p. 17-27. Berlin: Springer.
  • DOUBOVA A & FERNÁNDEZ-CARA E. 2012. On the control of viscoelastic Jeffreys fluids. Systems & Control Letters 61: 573-579.
  • EKELAND I & TEMAM R. 1974. Analyse convexe et problèmes variationnels. Paris: Dunod, Gauthier-Villars.
  • FABRE C. 1996. Uniqueness results for Stokes equations and their consequences in linear and nonlinear control problems. ESAIM: COCV 1: 267-302.
  • FERNÁNDEZ-CARA E & GUERRERO S. 2006. Null controlabillity of the Burgers system with distributed controls. Systems Controls Letters 56: 366-372.
  • FERNÁNDEZ-CARA E, GUILLÉN-GONZÁLEZ F & ORTEGA RR. 2002. Mathematical modeling and analysis of viscoelastic fluids of the Oldroyd kind. Amsterdam: North-Holland, p. 543-566.
  • FERNÁNDEZ-CARA E, GUERRERO S, YMANUVILOV OY & PUEL JP. 2004. Local exact controllability of the Navier Stokes system. J Math Pures Appl 83: 1501-1542.
  • FERNÁNDEZ-CARA E, GUERRERO S, YMANUVILOV OY & PUEL JP. 2006. Some controllability results for the N-dimensional Navier-Stokes and Boussinesq system with N - 1 scalar controls. SIAM J Control Optim 45: 146-173.
  • FERNÁNDEZ-CARA E, MACHADO JL & SOUSA DA. 2020. Non null controllability of Stokes equations with memory. ESAIM: COCV doi:10.1051/cocv/2019067.
  • FURSIKOV AV. 2000. Optimal control of distributed systems, Theory and Applications. Transl. Math. Monogr. 187. Providence, RI: American Mathematical Society.
  • FURSIKOV AV & IMANUVILOV OY. 1999. Exact controllability of the Navier-Stokes and Boussinesq equations. Russian Math Surveys 54 (3): 565-618.
  • GALDI GP. 2008. Mathematical Problems in Classical and Non-Newtonian Fluid Mechanics. Oberwolfach Semin, Vol. 37. Basel: Birkhäuser. 121-273 p.
  • GLASS O. 2000. Exact Boundary Controllability of 3-D Euler Equation. ESAIM: COCV 5: 1-44.
  • GUERRERO S & IMANUVILOV OY. 2013. Remarks on Non Controllability of the Heat Equation with Memory. ESAIM: COCV 19: 288-300.
  • GUILLÉN-GONZÁLEZ F, MARQUES-LOPES FP & ROJAS-MEDAR MA. 2013. On the Approximate Controllability of Stackelberg-Nash Strategies for Stokes Equations. Proc Amer Math Soc 141: 1759-1773.
  • GUNSBURGER M. 2003. Perspectives in Flow Control and Optimization. Philadelphia: SIAM.
  • IMANUVILOV OY. 2001. Remarks on Exact Controllability for the Navier-Stokes Equations. ESAIM: COCV 6: 39-72.
  • JESUS IP. 2015. Remarks on Hierarchic Control for a Linearized Micropolar Fluids System in Moving Domains. Applied Math & Optimization 72: 493-521.
  • JESUS IP & DE MENEZES SDB. 2015. On the Approximate Controllability of Stackelberg-Nash Strategies for Linearized Micropolar Fluids in Moving Domains. Computational and Applied Mathematics 34: 587-606.
  • KELVIN W. 1875. On the Theory Viscoelastic Fluids. Math a Phys Pap 3: 27-84.
  • LIMACO J, CLARK HR & MEDEIROS LA. 2009. Remarks on Hierarchic Control. J Math Anal Appl 359: 368-383.
  • LIONS JL. 1994. Hierarchic Control. Mathematical Science, Proc Indian Academic Science 104: 295-304.
  • LIONS JL & ZUAZUA E. 1998. Exact Boundary Controllability of Galerkin’s Approximations of Navier-Stokes Equations. Ann Scuola Norm Sup Pisa Cl Sci 4: 605-621.
  • LIONS PL & MASMOUDI N. 2000. Global Solutions for Some Oldroyd Models of Non-Newtonian Flows. Chin Ann Math 21: 131-146.
  • MARINHO AO, LOUREDO AT & MIRANDA MM. 2014. Exact Controllability of Galerkin’s Approximations for the Oldroyd Fluid System. International Journal of Modeling and Optimization 4: 126-132.
  • MAXWELL JC. 1859. On the Dynamical Theory of Gases. Philos Magazine p. 129-145.
  • MAXWELL JC. 1868. On the Dynamical Theory of Gases. Philos Trans Roy Soc London 157: 49-88.
  • NASH J. 1951. Noncooperative Games. Ann Math 54: 286-295.
  • OLDROYD JG. 1950. On the Formulation of Rheological Equations of State. Proc Roy Soc London Ser A 200: 523-541.
  • OLDROYD JG. 1953. The Elastic and Viscous Properties of Emulsions and Suspensions. Proc Roy Soc London Ser A 218: 122-132.
  • OLDROYD JG. 1959. Non-Newtonian Flow of Liquids and Solids. Rheology: Theory and Applications 1: 653-682.
  • OLDROYD JG. 1964. Non-Linear Stress, Rate of Strain Relations at Finite Rates of Shear in So-Called Linear Elastico-Viscous Liquids. In: Second-Order Effects in Elasticity, Plasticity and Fluid Dynamics. p. 520-529. Jerusalem: Jerusalem Academic Press.
  • OSKOLKOV AP. 1989. Initial Boundary Value Problems for the Equations of Motion of Kelvin-Voigt Fluids and Oldroyd Fluids. Proc of the Steklov Institute of Mathematics 179: 137-182.
  • OSKOLKOV AP & AKHMATOV MM. 1974. Convergent Difference Schemes for Equations of Motion of Oldroyd Fluids. J Sov Math 47: 2926-2933.
  • OSKOLKOV AP, KARAZEEVA NA & KOTSIOLIS AA. 1987. Dynamical System Generated by the Equations of Motion of an Oldroyd Fluid of Order L. Inst V A Steklova AN SSSR 164: 47-53.
  • OSKOLKOV AP & KOTSIOLIS AA. 1986. Solvability of the Basic Initial Boundary Problem for the Equations of Motion of an Oldroyd Fluid on (0,∞) and the Behavior of its Solutions as t→∞. Inst V A Steklova AN SSSR 150: 48-52.
  • PARETO V. 1896. Cours d’Économie Politique. Lausanne, Switzerland: Rouge.
  • RENARDY M. 2009. Global Existence of Solutions for Shear Flow of Certain Viscoelastic Fluids. J Math Fluid Mech 11: 91-99.
  • ROCKAFELLAR RT. 1967. Duality and Stability in Extremum Problems Involving Convex Functions. Pacif J Math 21: 167-187.
  • ROCKAFELLAR RT. 1969. Convex Analysis. Princeton, N. J.: Princeton University Press.
  • SERRIN J. 1959. Mathematical Principles of Classical Fluid Mechanics. In: Handbuch der Physik. Vol. 8. p. 125-263. Springer-Verlag.
  • STAVRE R. 2002. The Control of the Pressure for a Micropolar Fluid. Z Angew Math Phys 53: 912-922. Dedicated to Eugen Soós.
  • TEMAM R. 1979. Navier-Stokes Equations. 2. North-Holland.
  • VOIGT W. 1889. Uber die Innere Reinbung der Feste Korper. Abh Konigl Ges Wiss Gottingen .
  • VOIGT W. 1892. Uber Innere Feste Korper, Insbessondere der Metalle. Ann Phys 283: 671-693.
  • VON STACKELBERG H. 1934. Marktform und Gleichgewicht. Berlin, Germany: Springer.
  • WILKINSON WL. 1960. Non-Newtonian Fluids. Oxford: Pergamon Press.

APPENDIX A

On existence and uniqueness of solutions

0 This appendix aims to prove the existence and uniqueness of solutions to the coupled system (22). To simplify the notation, we will make the change of variable ϕ(x,t)=φ(x,Tt) into (22) which yields the following equivalent system:

|ϕ tμ Δ ϕ  0tg(tη )Δ ϕ (η )dη +p^ =0    in   Q,ξ itμ Δ ξ i0tg(tσ )Δ ξ i(σ )dσ +p=α iφ χ Oi  in   Q,div(ϕ )=0,div(ξ i)=0     in     Q,ϕ =0,ξ i=0     on     Σ ,ξ i(x,0)=0,ϕ (x,0)=f(x)+i=1Nρ i2[ξ i(x,T)]    in    Ω .(A.1)
Then, the following result holds:

Theorem 3. For each 1,2,,N,assume thatρiL(Ω), fH, and that i=1Nαi is sufficiently small. Let g:[0,)[0,) be given by g(t)=γeδt, where γ=δ(νμ),μ=kλ1, and δ=λ1 are positive constants. Then, system\((A.1)\)admits a unique solution

{ϕ,ξi,p̂,p}C0([0,T];H)×C0([0,T];V)×L2(0,T;L2(Ω))×L2(0,T;H1(Ω))
such that
ϕ tμ Δ ϕ tTg(η t)Δ ϕ (η )dη +p^ =0inL2(0,T;L2(Ω )),ξ itμ Δ ξ i0tg(tσ )Δ ξ i(σ )dσ +p=α iϕ χ OiinL2(0,T;L2(Ω )),ξ i(0)=0andϕ (0)=f+i=1Nρ i2[ξ i(T)]inΩ .

Proof: We proceed by using the Faedo-Galerkin method for a special basis {wj}j of V. We refer to the book of Temam (1979) for a complete description of this method. Let Vm=[w1,....,wm] be the subspace generated by the first m vectors {wj}j. The approximate problem consists in finding functions ϕm(x,t)=j=1mqjm(t)wj(x), and ξim(x,t)=j=1mhjmi(t)wj(x), where qjm(t) and hjmi(t) are real functions defined in [0,T], such that:

|(ϕ m(t)t,wj)+μ ((ϕ m(t),wj))+tTg(η t)((ϕ m(η ),wj))dη =0,(ξ im(t)t,wj)+μ ((ξ im(t),wj))+0tg(tσ )((ξ im(σ ),wj))dσ =α i(ϕ m(t)χ Oi,wj),ξ im(0)=0,ϕ m(0)i=1Nρ i2ξ im(T)=λ =1m(f,wj)wj,for allwjVm.(A.2)
Multiplying \((A.2)\) 1 by qjm(t) and summing for j=1,.....m, we find that:
ddt|ϕm(t)|2+μ||ϕm(t)||2+0tg(tσ)((ϕm(σ),ϕm(t)))dσ|ϕm(t)|2.(A.3)
Integrating \((A.6)\) from 0 to t, and applying Fubini’s formula and Gronwall’s inequality, we get that:
|ϕm(t)|2C(T)|ϕm(0)|2,for allm,t0.(A.4)
Analogously, multiplying \((A.2)\) 2 by λjhjmi(t) and summing from j=1,......,m, we obtain:
ddt||ξim(t)||2+μ2|Δξim(t)|2+0tg(tσ)(Δξm(σ),Δξm(t))dσ|ϕm(t)|2.(A.5)
Thus, combining (𝐴.4), and (𝐴.5) it follows that:
ddt||ξim(t)||2+μ2|Δξim(t)|2+0tg(tσ)(Δξm(σ),Δξm(t))dσC(T)|ϕm(0)|2.(A.6)
Integrating (A.3) from 0 to t, and combining again Fubini’s formula and Gronwall’s inequality, we have that:
ξim(t)2C̃(T)|ϕm(0)|2for allm,t[0,T].(A.7)
In particular,
ξim(T)2C̃(T)|ϕm(0)|2for allm.(A.8)

From the definition of ϕm(0), (see \((A.2)\)), we see that:

|ϕm(0)|2c0C1(N)(i=1N|ρi|L(Ω)4)(i=1Nξim(T)2)+|f|2,(A.9)
where c0 is the constant of immersion of the V into H.

From (𝐴.8) and (𝐴.9), we conclude that:

[1c0C1(N)(i=1N|ρi|L(Ω)4)i=1Nαi2μ]i=1Nξim(T)2i=1Nαi2μ|f|2.(A.10)

Notice that we can assume i=1Nαi2 sufficiently small so that:

β=1c0C1(N)(i=1N|ρi|L(Ω)4)i=1Nαi2μ>0.

Hence, from ((A.10)), we obtain that:

i=1Nξim(T)21βμi=1Nαi2|f|2.(A.11)

Now, combining (𝐴.9) and (𝐴.11), we deduce that:

|ϕm(0)|2[c0C1(N)(i=1N|ρi|L(Ω)4)1βμi=1Nαi2+1]|f|2,for allm.(A.12)

From (𝐴.3) − (𝐴.7), it follows that:

(ϕm)mis bounded inL2(0,T;V),(ϕm)mis bounded inL(0,T;H),(ξim)mis bounded inL(0,T;V),(Δξim)mis bounded inL2(0,T;H),(A.13)
for i=1,2,3,.....,N. In particular, from (A.13)4 we conclude that:
(ξim)mis bounded inL2(0,T;𝐇2(Ω)).(A.14)

Moreover, from (𝐴.13)1 − (𝐴.13)3, and (𝐴.14), we can extract subsequences of (ϕm)m and (ξim)m such that:

ϕmϕinL2(0,T;V),ϕm*ϕinL(0,T;H),ξim*ξiinL(0,T;V),ξimξiinL2(0,T;𝐇2(Ω)),(A.15)
for i=1,2,....,N.

From (𝐴.15)1 and (𝐴.15)2, we find that:

ddt(ϕ(t),v)+μ((ϕ(t),v))+0tg(tσ)((ϕ(σ),v))dσ=0for allvVin𝒟(0,T).

This means that:

ϕ tL2(0,T;V),ϕ t μ Δ ϕ tTg(η t)Δ ϕ (η )dη =0in the sense of L2(0,T;V).(A.16)

Moreover,

ϕtμΔϕ(t)tTg(ηt)Δϕ(η)dη=0,(A.17)
in the sense of V in [0,T]. In particular, we have that:
ϕtμΔϕ(t)tTg(ηt)Δϕ(η)dη,v=0,(A.18)
for all v𝒱 in [0,T]. Hence, as in Temam (1979), there exists p̂(t)L2(Ω) such that:
|p̂(t)|L2(Ω)C||p̂(t)||𝐇1(Ω),(A.19)
and
p̂(t)=ϕtμΔϕ(t)tTg(ηt)Δϕ(η)dη.(A.20)
Therefore, from (𝐴.19) and (𝐴.20), we conclude that:
p̂L2(0,T;𝐇1(Ω)),(A.21)
and
p̂L2(0,T;L2(Ω)).(A.22)
Since ϕtL2(0,T;V), ϕL2(0,T;V), and VHV, then as in Temam (1979), we see that:
ϕC0([0,T];H).(A.23)

Finally, we deduce (A.15) that:

ξiL2(0,T;𝐋2(Ω)),(A.24)
pL2(0,T;H1(Ω)),(A.25)
ξiC0([0,T];V),(A.26)
ϕ tμ Δ ϕ tTg(η t)Δ ϕ (η )dη +p^ =0inL2(0,T;L2(Ω )),(A.27)
ξ itμ Δ ξ i0tg(tσ )Δ ξ i(σ )dσ +p=α iχ OiinL2(0,T;L2(Ω )),(A.28)
for all i=1,2,....,N. The uniqueness is obtained of standard form.

Remark 0.8. The proof of the existence, uniqueness, and regularity of solutions for the system (8) follows similarly to the demonstration of the previous theorem, with some suitable adaptations. Due to that, we will omit this proof.

Publication Dates

  • Publication in this collection
    20 Sept 2024
  • Date of issue
    2024

History

  • Received
    21 Apr 2024
  • Accepted
    07 June 2024
location_on
Academia Brasileira de Ciências Rua Anfilófio de Carvalho, 29, 3º andar, 20030-060 Rio de Janeiro RJ Brasil, Tel: +55 21 3907-8100 - Rio de Janeiro - RJ - Brazil
E-mail: aabc@abc.org.br
rss_feed Acompanhe os números deste periódico no seu leitor de RSS
Acessibilidade / Reportar erro