
Citation: Andrea Manzoni, Alfio Quarteroni, Sandro Salsa. A saddle point approach to an optimal boundary control problem for steady Navier-Stokes equations[J]. Mathematics in Engineering, 2019, 1(2): 252-280. doi: 10.3934/mine.2019.2.252
[1] | Ivan Fumagalli . Discontinuous Galerkin method for a three-dimensional coupled fluid-poroelastic model with applications to brain fluid mechanics. Mathematics in Engineering, 2025, 7(2): 130-161. doi: 10.3934/mine.2025006 |
[2] | Lars Eric Hientzsch . On the low Mach number limit for 2D Navier–Stokes–Korteweg systems. Mathematics in Engineering, 2023, 5(2): 1-26. doi: 10.3934/mine.2023023 |
[3] | Giulio Ortali, Nicola Demo, Gianluigi Rozza . A Gaussian Process Regression approach within a data-driven POD framework for engineering problems in fluid dynamics. Mathematics in Engineering, 2022, 4(3): 1-16. doi: 10.3934/mine.2022021 |
[4] | Gennaro Ciampa, Gianluca Crippa, Stefano Spirito . Propagation of logarithmic regularity and inviscid limit for the 2D Euler equations. Mathematics in Engineering, 2024, 6(4): 494-509. doi: 10.3934/mine.2024020 |
[5] | Kyungkeun Kang, Dongkwang Kim . Existence of generalized solutions for Keller-Segel-Navier-Stokes equations with degradation in dimension three. Mathematics in Engineering, 2022, 4(5): 1-25. doi: 10.3934/mine.2022041 |
[6] | Daniele Cerroni, Florin Adrian Radu, Paolo Zunino . Numerical solvers for a poromechanic problem with a moving boundary. Mathematics in Engineering, 2019, 1(4): 824-848. doi: 10.3934/mine.2019.4.824 |
[7] | Giuseppe Procopio, Massimiliano Giona . Bitensorial formulation of the singularity method for Stokes flows. Mathematics in Engineering, 2023, 5(2): 1-34. doi: 10.3934/mine.2023046 |
[8] | Camilla Nobili . The role of boundary conditions in scaling laws for turbulent heat transport. Mathematics in Engineering, 2023, 5(1): 1-41. doi: 10.3934/mine.2023013 |
[9] | Yangyang Qiao, Qing Li, Steinar Evje . On the numerical discretization of a tumor progression model driven by competing migration mechanisms. Mathematics in Engineering, 2022, 4(6): 1-24. doi: 10.3934/mine.2022046 |
[10] | Tommaso Tassi, Alberto Zingaro, Luca Dede' . A machine learning approach to enhance the SUPG stabilization method for advection-dominated differential problems. Mathematics in Engineering, 2023, 5(2): 1-26. doi: 10.3934/mine.2023032 |
Optimal control problems (OCPs) for partial differential equations is a research area that features challenging theoretical questions for mathematical analysis, the need of devising efficient numerical algorithms, with a broad range of meaningful applications in engineering, like, e.g., aerodynamics [29,30] and biofluid dynamics [33,40]. In this paper we address one of the most classical problems in this field, namely the boundary control for the Navier-Stokes equations. Much has been done for this problem, both on the mathematical analysis and on its numerical approximation; there are, however, some issues that have not yet been addressed – the case of the state problem with mixed Dirichlet and Neumann conditions is one of those. At the best of our knowledge, a comprehensive framework to address both (i) the well posedness analysis for the OCP related to this system and (ii) the derivation of a system of optimality conditions, is still missing. Setting this framework, and exploiting it in view of the numerical approximation of the OCP, is the scope of our paper.
More precisely, we study the mixed boundary value problem for the incompressible steady Navier-Stokes equations, both in two and three dimensions, where we prescribe homogeneous Neumann conditions (on the normal stress) on a portion of the boundary, homogeneous Dirichlet conditions on the velocity field on another boundary portion, and a velocity control on the remaining portion of the domain boundary*. For the sake of application, we consider the minimization of quadratic cost functionals involving either the square of the L2 norm of the vorticity, or a velocity tracking-type term, on the whole computational domain. For the problem at hand, we introduce an ad-hoc saddle point formulation of the state problem which, besides velocity and pressure fields, requires a third variable representing the normal Cauchy stress and a Lagrange multiplier penalizing the Dirichlet data (this latter involving the control function, too). This saddle point formulation is more apt for numerical approximation than the one with a single field u that is obtained after projecting the Navier-Stokes equations onto a divergence free space. For the control problem, the saddle point formulation enables us to derive a system of first-order optimality conditions, and to carry out a rigorous mathematical analysis of the OCP.
Splitting the boundary ∂Ω into portions on which different boundary conditions are set is rather common in several fluid dynamics applications. Dirichlet conditions are used for the problem at hand to impose a velocity profile on the inflow boundary and on the control boundary, and a no-slip condition at the lumen boundary, since it is assumed that there fluid particles perfectly adhere to the vessel wall. Neumann conditions prescribe a force per unit area as the normal component of the stress tensor; if the force is vanishing, the Neumann boundary portion is a free outflow – this is indeed the condition we assign on the outflow boundary. Note also that when Dirichlet boundary conditions are specified on the entire boundary ∂Ω, the pressure is not uniquely defined.
Similar problems have been studied by several authors. What follows is an account of the most relevant works on the subject. In a classical paper by Gunzburger [21], the same control problem is addressed with however full Dirichlet boundary conditions (and a different cost functional), by a saddle-point formulation; a similar approach has been applied to viscous drag reduction in [19]. In a series of other papers the problem is faced working on the subspace of divergence free velocity fields. In [12], the unsteady problem is studied, with Dirichlet conditions on the whole boundary; in [13], an exterior problem in two dimensions for the steady Navier-Stokes equations with boundary Neumann control is considered for the minimization of a drag functional; an inequality state constraint is also enforced on the velocity field. Distributed and Neumann controls are also considered, e.g., in [20]. Boundary control problems for vorticity minimization have been analyzed in [32], in the case of steady Stokes equations, and in [31], considering instead unsteady Navier-Stokes equations; in both cases, however, a pure Dirichlet problem is considered. In [17] the analysis of existence of optimal boundary control for the Navier-Stokes equations with mixed boundary conditions is carried out in three dimensions, without deriving a system of optimality conditions. Regarding the case of unsteady Navier-Stokes equations, a boundary OCP similar to the one studied in our paper has been analyzed in [24], however dealing with a pure Dirichlet problem. The analysis and the approximation of OCPs involving distributed controls and in the case of unsteady Navier-Stokes flows has been considered in [22,23], again dealing with full Dirichlet boundary conditions. Mixed Dirichlet-Neumann boundary conditions have been instead considered in [4], however without carrying out the well-posedness analysis. We believe that our presentation is more comprehensive, and allows the reader to have a better picture of the whole mathematical analysis for both the state and the control problem and the derivation of a system of optimality conditions.
The structure of the paper is as follows. In Section 2 we introduce a ad-hoc saddle point formulation of the state problem with a Lagrange multiplier that represents the normal Cauchy stress on the whole boundary and carry out its well-posedness analysis in Section 3; some technical results related to the Stokes solver are reported in Section 4; In Section 5 we set up the ground for the analysis of the control problem; we then investigate the differentiability of the control-to-solution map in Section 6, and derive first order optimality conditions for the OCP in Section 7; In Section 8 we illustrate a possible way to approximate our control problem, by relying on a descent iterative method for the minimization of the cost functional, and using the Galerkin finite element method to discretize both the state and the adjoint problem; Numerical results shown in Section 9 concern the solution of an interior flow problem, involving a simplified two-dimensional tract of blood vessel where a bypass is inserted. The goal is to regularize the flow by minimizing a tracking-type quadratic functional, expressing the distributed observation of the (square of) the L2 norm of the difference v−zd between the fluid velocity v and the target velocity zd, this latter being provided by the solution of the corresponding Stokes problem over the whole domain. Conclusions will then follow in Section 10.
In this paper we consider a state system given by the steady Navier-Stokes equations, in velocity and pressure formulation, for steady, incompressible and viscous fluid flows. For this state system, we analyze a minimization problem for quadratic functionals of the velocity field through a velocity control operated on a portion of the inflow boundary. For a concrete example we denote by Ω⊂Rd, d=2,3 the domain in Figure 1 below, occupied by the fluid and by Γ=∂Ω its boundary. The inflow boundary Γ1=Γc∪Γin is composed by Γc, the control boundary, and by Γin, where a Dirichlet data is assigned; Γout and Γw are the outflow and the wall boundary portions, where zero stress and zero velocity are assigned, respectively.
Let
X=H1Γw(Ω)d={v∈H1(Ω)d:v=0 on Γw} |
denote the state space, endowed with the norm ‖v‖X=‖∇v‖0 where
‖∇v‖0=(d∑i,j=1∫Ω(∂vi∂xj)2dx)1/2. |
Our goal is to minimize either the vorticity functional
˜J1(v,u)=12∫Ω|∇×v|2dx+τ2∫Γc|∇u|2dσ | (2.1) |
or a velocity tracking-type cost functional
˜J2(v,u)=12∫Ω|v−zd|2dx+τ2∫Γc|∇u|2dσ | (2.2) |
where τ≥0, u∈U=H10(Γc)d is a control variable and (v,π)=(v(u),π(u))∈X×L2(Ω) is a weak solution of the following system:
{−νΔv+(v⋅∇)v+∇π=0in Ωdivv=0in Ωv=vinon Γinv=uon Γcv=0on Γwπn−ν∂v∂n=0on Γout, | (2.3) |
where v and π denote the fluid velocity and pressure, respectively, ν is a given kynematic viscosity, and n is the unit outward normal direction. In the case of ˜J2, a flow matching problem is solved, aiming at driving the velocity field toward a target velocity profile zd∈X; this latter can be given, e.g., by the Stokes flow in the same domain, whenever the goal is to minimize recirculation and vortex generation for the sake of regularization of flows.
If τ>0, the second term in (2.1) or (2.2), called penalization term, besides helping in the well-posedness analysis, prevents using "too large" controls in the minimization of the cost functional.
We handle the Dirichlet boundary conditions on Γ1 (involving the control function) by a Lagrange multiplier approach. Let e and b be the following bilinear forms
e(v,w)=∫Ων∇v:∇wdx,b(v,q)=−∫Ωqdivvdx | (2.4) |
and c be the trilinear form
c(v,w,z)=∫Ων(v⋅∇)w⋅zdx=d∑i,j=1∫Ωννi∂wj∂xizjdx. |
Using Hölder inequality, and the Sobolev inequality
‖v‖L4(Ω)≤CS‖∇v‖0=CS‖v‖X, |
we have that
|c(v,w,z)|≤‖v‖L4(Ω)‖∇w‖0‖z‖L4(Ω)≤C2S‖v‖X‖w‖X‖z‖X∀v,w,z∈X. | (2.5) |
Let us now define
g=g(u)={uon Γc0on Γwvinon Γin; |
assuming that vin∈H1/200(Γin)d, we have that g∈H1/200(Γ1)d, since u∈H10(Γc)d⊂Tc with continuous injection, that is, for some positive constant c0,
‖u‖H1/200(Γc)d≤c0‖u‖H10(Γc)d. | (2.6) |
For the definition of these functional spaces, see, e.g., [7,Vol. Ⅲ,Chap. Ⅶ]. Observe that if v∈X, then v|Γ1∈H1/200(Γ1)d.
For ease of notation, hereon we denote by T1=H1/200(Γ1)d and by ⟨⋅,⋅⟩Γ1 the duality pairing between T1 and its dual T∗1. In the same way, we will denote by Tc=H1/200(Γc)d, and by ⟨⋅,⋅⟩Γc the duality pairing between Tc and its dual T∗c.
Let V=X×L2(Ω)×T∗1. The weak formulation of (2.3) that we consider reads as follows: find (v,π,t)∈V such that
{e(v,φ)+c(v,v,φ)+b(φ,π)+⟨t,φ⟩Γ1=0∀φ∈Xb(v,ϕ)=0∀ϕ∈L2(Ω)⟨λ,v⟩Γ1=⟨λ,g⟩Γ1∀λ∈T∗1. | (2.7) |
Note that a formal integration by parts in the first equation gives for the boundary stress πn−ν∂v∂n,
t=πn−ν∂v∂non Γ1, | (2.8) |
and
0=πn−ν∂v∂non Γout. |
Thus, the first equation in (2.7) enforces in a natural way the homogeneous Neumann condition on Γout.
By grouping together the last two equations in (2.7), we obtain the following equivalent formulation: find v∈X, (π,t)∈Q=L2(Ω)×T∗1 such that
{e(v,φ)+c(v,v,φ)+b(φ,π)+⟨t,φ⟩Γ1=0∀φ∈Xb(v,ϕ)+⟨λ,v⟩Γ1=⟨λ,g⟩Γ1∀(ϕ,λ)∈Q. | (2.9) |
Upon defining the bilinear form ˜b:X×Q→R as
˜b(v,(ϕ,λ))=b(v,ϕ)+⟨λ,v⟩Γ1 |
and the norm
‖(ϕ,λ)‖Q=√‖ϕ‖2L2(Ω)+‖λ‖2T∗1, |
we reformulate problem (2.9) as: find v∈X,(π,t)∈Q such that
{e(v,φ)+c(v,v,φ)+˜b(φ,(π,t))=0∀φ∈X˜b(v,(ϕ,λ))=⟨G,(ϕ,λ)⟩Q∗,Q∀(ϕ,λ)∈Q, | (2.10) |
with G∈Q∗ given by
⟨G,(ϕ,λ)⟩Q∗,Q=⟨λ,g⟩Γ1. | (2.11) |
Despite their ubiquitous use when modeling viscous incompressible flows, Navier-Stokes equations endowed with mixed boundary conditions have been seldom analyzed from a theoretical standpoint. If homogeneous Neumann conditions are set on a boundary portion, energy inequalities or a priori estimates of a weak solution to the Navier-Stokes system cannot be obtained because there is no control over the energy flux on this boundary portion. Some results can be found, e.g., in [41] regarding the existence and uniqueness of solution in weighted Lp Sobolev spaces and Hölder spaces for problems set over polyhedral domains; see also, e.g., [3,35,36] for the time-dependent case and [2] for a well-posedness analysis in divergence-free spaces. Additional conditions that estimate backward flows over the Neumann boundary have been considered, e.g., in [34].
In this paper we derive a well-posedness result for the steady Navier-Stokes system in presence of mixed Dirichlet-Neumann boundary conditions including the pressure field among the state variables, in non-weighted Sobolev spaces. We show that if ‖g‖T1 is small enough, then (2.10) is well posed. We first need to introduce the following continuous operators, related to the bilinear form ˜b: B:X→Q∗ defined by
⟨Bv,(ϕ,λ)⟩Q∗,Q=˜b(v,(ϕ,λ))∀(ϕ,λ)∈Q, |
and B∗:Q↦X∗ defined by
⟨B∗(ϕ,λ),φ⟩X∗,X=˜b(φ,(ϕ,λ))∀φ∈X. |
We denote by NB the kernel of B, by N⊥B its orthogonal (in X) and by N0B the polar of NB, that is
N0B={L∈X∗:⟨L,φ⟩X∗,X=0for all φ∈NB}. |
The so called inf-sup condition holds for the bilinear form ˜b:
Lemma 3.1. There exists a constant ˜β>0 such that
inf0≠(ϕ,λ)∈Qsup0≠v∈X˜b(v,(ϕ,λ))‖v‖X‖(ϕ,λ)‖Q≥˜β. | (3.1) |
Proof. Given (˜ϕ,˜λ)∈Q, according to Riesz Theorem there exists a unique λ∗∈T1 such that
⟨˜λ,θ⟩Γ1=(λ∗,θ)T1∀θ∈T1, |
with ‖λ∗‖T1=‖˜λ‖T∗1. Then we look for ˜v∈X that satisfies
{b(˜v,ϕ)=(˜ϕ,ϕ)L2(Ω)∀ϕ∈L2(Ω)˜v=−λ∗on Γ1 |
with
‖˜v‖X≤˜C(‖˜ϕ‖L2(Ω)+‖λ∗‖T1). |
Such ˜v can be obtained as follows. Let r∈X be the lifting of
h={−λ∗on Γ10on Γ0h0on Γout |
where
h0=(∫Ω˜ϕ+∫Γ1λ∗⋅n)g0 |
with ∫Γoutg0⋅n=1. Let us now set ˜v=w+r, with w∈H10(Ω)d to be determined. We have
div˜v=divw+divr=˜ϕ, |
that is, divw=˜ϕ−divr. Moreover,
∫Ω˜ϕ−∫Ωdivr=∫Ω˜ϕ+∫Γ1λ∗⋅n−∫Γouth0⋅n=0. |
It follows (see, e.g., [15]) that there exists w∈H10(Ω)d with divw=˜ϕ−divr so that ˜v=w+r meets the required assumptions. Thus we can write
˜b(˜v,(˜ϕ,˜λ)))=b(˜v,˜ϕ)−⟨˜λ,˜v⟩Γ1=‖˜ϕ‖2L2(Ω)+‖λ∗‖2T1≥12(‖˜ϕ‖L2(Ω)+‖λ∗‖T1)2≥12˜C(‖˜ϕ‖L2(Ω)+‖λ∗‖T1)‖˜v‖X≥12˜C(‖˜ϕ‖2L2(Ω)+‖˜λ‖2T∗1)1/2‖˜v‖X=12˜C‖(˜ϕ,˜λ)‖Q‖˜v‖X |
from which (3.1) follows with ˜β=12˜C, by taking the supremum over ˜v and the infimum over (˜ϕ,˜λ).
The following lemma is well known (see [15,Chap. 1,Thm. 4.1]).
Lemma 3.2. Condition (3.1) is equivalent to each one of the following properties:
(ⅰ) the operator B is a continuous isomorphism from N⊥B onto Q∗ and
‖w‖X≤˜β−1‖Bw‖Q∗ for all w∈N⊥B; | (3.2) |
(ⅱ) the operator B∗ is a continuous isomorphism from Q onto N0B and
‖(ϕ,λ)‖Q≤˜β−1‖B∗(ϕ,λ)‖X∗ for all (ϕ,λ)∈Q. | (3.3) |
We are ready to prove the following result.
Theorem 3.3. There exists η>0, depending only on ν, ˜β,CS, such that if ‖g(u)‖T1≤η, then problem (2.10) has a unique solution (v,π,t)=(v(u),π(u),t(u))∈X×L2(Ω)×T∗1. Moreover,
‖v‖X≤C1‖g(u)‖T1 |
and
‖π‖L2(Ω)+‖t‖T∗1≤C2(‖g(u)‖T1+‖g(u)‖2T1) | (3.4) |
where C1=5˜β−1 and C2 depend only on ν, ˜β,CS.
Remark 3.4. A possible choice of η is
η=˜β60C2Smin{ν,ν−1}. | (3.5) |
Proof. Write v=z+v0 with v0∈N⊥B, z∈NB. Define
˜e(z,ψ,φ)=e(ψ,φ)+c(z,ψ,φ)+c(v0,ψ,φ)+c(ψ,v0,φ). |
To solve problem (2.10), we first determine v0 by solving
˜b(v0,(ϕ,λ))=⟨G,(ϕ,λ)⟩Q∗,Q∀(ϕ,λ)∈Q, | (3.6) |
where G∈Q∗ is defined in (2.11). Then we determine z∈NB from the equation
˜e(z,z,φ)=⟨F0,φ⟩N∗B,NB∀φ∈NB, | (3.7) |
where F0∈N∗B is the restriction to NB of F∈X∗, given by†
†Rigorously, F0=F∘P where P∈L(X∗,N∗B) is defined as ⟨Pf,φ⟩N∗B,NB=⟨f,φ⟩X∗,X∀f∈X∗,φ∈NB. Clearly, ‖F0‖NB∗≤‖F‖X∗.
⟨F0,φ⟩N∗B,NB=⟨F,φ⟩X∗,X=−e(v0,φ)−c(v0,v0,φ)∀φ∈NB. |
Finally, with v=z+v0, we find (π,t)∈Q from
˜b(ψ,(π,t))=−e(v,ψ)−c(v,v,ψ) ∀ψ∈X. | (3.8) |
Note that
‖F0‖N∗B≤‖F‖X∗≤ν‖v0‖X+C2S‖v0‖2X | (3.9) |
and that ⟨F,ψ⟩−˜e(z,z,ψ)=−e(v,ψ)−c(v,v,ψ), ψ∈X. We divide the proof in four steps.
Step 1. Equation (3.6) is equivalent to
˜Bv0=G in Q∗ |
which, by Lemma 3.2, (i), has a unique solution v0∈N⊥B. Moreover, from (3.2) and (2.11),
‖v0‖X≤˜β−1‖G‖Q∗≤˜β−1‖g(u)‖T1. | (3.10) |
Step 2. Now we reduce (3.7) to a fixed point equation. Let z∈NB be given and consider the linear equation
˜e(z,w,φ)=⟨F0,φ⟩N∗B,NB∀φ∈NB. | (3.11) |
For each z, the bilinear form ˜e(z;⋅,⋅) is continuous, by (2.5). Moreover,
˜e(z,φ,φ)≥(ν−C2S‖z‖X−2C2S‖v0‖X)‖φ‖2X. |
Thus, if
‖z‖X≤4˜β−1‖g(u)‖T1 |
then
˜e(z,φ,φ)≥(ν−6˜β−1C2S‖g(u)‖T1)‖φ‖2X. |
Therefore, assuming
‖g(u)‖T1≤η | (3.12) |
with η as in (3.5), we have
˜e(z,φ,φ)≥(ν−6˜β−1C2S‖g(u)‖T1)‖φ‖2X≥ν2‖φ‖2X |
and ˜e(z;⋅,⋅) is NB−coercive. Let us define
δu=˜β−1‖g(u)‖T1 | (3.13) |
and, for m integer, define
Smδu={z∈NB:‖z‖X≤mδu}. | (3.14) |
By the Lax-Milgram Theorem, for each z∈S4δu, we can define a solution map operator T(z)∈L(N∗B,NB) that to each L∈N∗B associates the unique solution w=T(z)L of the equation
˜e(z,w,φ)=⟨L,φ⟩N∗B,NB∀φ∈NB. | (3.15) |
Moreover,
‖T(z)L‖NB≤2ν‖L‖N∗B, |
from which
‖T(z)‖L(N∗B,NB)≤2ν. |
If z∗ is a fixed point for the operator z↦T(z)F0, then
˜e(z∗,z∗,φ)=⟨F0,φ⟩N∗B,NB∀φ∈NB |
which is nothing but equation (3.7).
Step 3. We show that, if ‖L‖N∗B is small enough, z↦T(z)L is a strict contraction from S4δu into itself. From (3.15), if
‖L‖N∗B≤2νδu, |
then z↦T(z)L maps S4δu into itself. We now estimate ‖(T(z1)−T(z2))L‖NB. Thanks to the continuity of ˜e, we can rewrite equation (3.15) as
⟨A(z)w,φ⟩N∗B,NB=⟨L,φ⟩N∗B,NB∀φ∈NB |
or
A(z)w=L, |
where A(z)∈L(NB,N∗B). Clearly T(z)=A(z)−1 and therefore we have the identity
T(z1)−T(z2)=T(z1)[A(z2)−A(z1)]T(z2) |
from which
‖T(z1)−T(z2)‖L(N∗B,NB)≤4ν2‖A(z2)−A(z1)‖L(NB,N∗B). |
Now
|⟨A(z2)ψ,φ⟩N∗B,NB−⟨A(z1)ψ,φ⟩N∗B,NB|=|c(z2−z1;ψ,φ)|≤C2S‖ψ‖X‖φ‖X‖z2−z1‖X |
whence
‖A(z2)−A(z1)‖L(NB,N∗B)≤C2S‖z2−z1‖X |
and
‖(T(z1)−T(z2))L‖X≤C2S‖L‖N∗B‖z2−z1‖X. |
We deduce that if
‖L‖N∗B<min{C−2S,2νδu} |
then z↦T(z)L is a strict contraction and therefore has a unique fixed point in Sδ.
If L=F0, we have (g=g(u))
2ν‖F0‖N∗B≤2‖v0‖X+2C2Sν‖v0‖2X≤2˜β−1‖g‖T1+2C2Sν˜β‖g‖2T1=2c0˜β−1‖g‖T1(1+C2Sν˜β‖g‖T1)≤4δu=4˜β−1‖g‖T1 |
since
1+C2S˜βν‖g‖T1≤1+112≤2. |
Moreover,
C2S‖F0‖N∗B≤νC2S‖v0‖X+C4S‖v0‖2X≤ν˜β−1C2S‖g‖T1+C4S˜β−2‖g‖2T1=˜β−1C2S‖g‖T1(ν+C2S˜β−1‖g‖T1)≤1312ν˜β−1C2S‖g‖T1<1 |
since
‖g‖T1≤η<1213βC2S1ν. |
Thus, T(z)F0: S4δu→S4δu is a strict contraction and has a unique fixed point z∗∈NB with
‖z∗‖≤4˜β−1‖g‖T1. | (3.16) |
Step 4. It remains to solve equation (3.8). Observe that equation (3.7) means that the operator E(v)∈X∗ given by
⟨E(v),ψ⟩X∗,X=e(v,ψ)+c(v,v,ψ)=˜e(z∗,z∗,ψ)+e(v0,ψ)+c(v0,v0,ψ) |
where v=z∗+v0, belongs to N0B. Therefore, by Lemma 3.2, (ii), there exists a unique (π,t)∈Q such that
⟨B∗((π,t)),ψ⟩X∗,X=˜b(ψ,(π,t))=−e(v,ψ)−c(v,v,ψ) ∀ψ∈X |
with
‖(π,t)‖Q=√‖π‖2L2(Ω)+‖t‖2T∗1≤˜β−1‖E(v)‖X∗≤˜β−1(ν‖v‖X+C2S‖v‖2X)≤C(‖g(u)‖T1+‖g(u)‖2T1) |
and the proof is complete with C1=5˜β−1 and C2 depending only on the constants ν, ˜β,CS.
Remark 3.5. Theorem 3.3 holds with a right hand side f∈X∗ with small enough ‖f‖X∗ norm. Then the unique solution (v,π,t) satisfies the estimates
‖v‖X≤C1{‖f‖X∗+‖g(u)‖T1} | (3.17) |
and
‖π‖L2(Ω)+‖t‖T∗1≤C2(‖f‖X∗+‖g(u)‖T1+‖g(u)‖2T1) | (3.18) |
where C1 and C2 only depend on the constants ν, ˜β,CS.
Under the hypotheses of Theorem 3.3, stressing the dependence on u of the solution to problem (2.10), we define the solution map S:u↦(v(u),π(u),t(u)). We will need the following result, concerning the map u↦v(u), since only this component is involved in the cost functional.
Proposition 3.6. Under the hypotheses of Theorem 3.3, possibly with a smaller η depending only on ν, ˜β,CS, there holds
‖v(u2)−v(u1)‖X≤CL‖u2−u1‖Tc | (3.19) |
with CL depending only on ν, ˜β,CS.
Proof. Write v(uj)=z(uj)+v0(uj)=zj+v0j, with zj∈NB, v0j∈N⊥B, j=1,2. Recall that
‖zj‖X,‖v0j‖X≤5˜β−1η | (3.20) |
From (3.10), since the equation for v0j is linear and v2−v1=0 on Γin, we get
‖v02−v01‖X≤˜β−1‖u2−u1‖Tc. | (3.21) |
The equation for z=z2−z1 reads, after routine calculations,
e(z,φ)+c(z2,z2,φ)−c(z1,z1,φ)+c(v01z,φ)+c(z,v02,φ)=−e(v02−v01,φ)−c(z1,v02−v01,φ)−c(v02−v01,z1,φ)−c(v02,v02,φ)+c(v01,v01,φ) |
for all φ∈X. Letting φ=z, we have, using ‖zj‖X≤5˜β−1η
|c(z2,z2,z)−c(z1,z1,z)|=|c(z2,z2,z)−c(z2,z1,z)+c(z2,z1,z)−c(z1,z1,z)|≤|c(z2,z,z)|+|c(z,z1,z)|≤10˜β−1ηC2S‖z‖2X. |
Similarly, using (3.21) and ‖v0j‖X≤5˜β−1η,
|c(v02,v02,z)−c(v01,v01,z)|≤|c(v02,v02−v01,z)|+|c(v02−v01,v01,z)|≤10˜β−1ηC2S‖z‖X‖v02−v01‖X≤10˜β−2ηC2S‖z‖X‖u2−u1‖Tc |
and
|c(v01,z,z)+c(z,v02,z)|≤10˜β−1ηC2S‖z‖2X. |
Similar calculations give
|c(v02−v01,z1,z)|≤5˜β−2ηC2S‖z‖X‖u2−u1‖Tc. |
Collecting all the above inequalities, and noting that
|e(v02−v01,z)|≤ν˜β−1‖z‖X‖u2−u1‖Tc, |
we deduce
ν‖z‖2X≤30˜β−1ηC2S‖z‖2X+(10˜β−1C2S+ν)˜β−1‖z‖X‖u2−u1‖Tc. |
Thus, if
η=ν60C2Smin{ν,ν−1} | (3.22) |
then (3.19) follows with CL depending only on ν,˜β,CS.
The same technique used in the proof of Theorem 3.3 provides existence and uniqueness of the weak solution (v,π,t)∈V of the Stokes problem
{e(vs,φ)+b(φ,πs)+⟨ts,φ⟩Γ1=⟨F,φ⟩X∗,X∀φ∈Xb(vs,ϕ)=(h,ϕ)L2(Ω)∀ϕ∈L2(Ω)⟨λ,vs⟩Γ1=⟨λ,G⟩Γ1∀λ∈T∗1, | (4.1) |
for any (F,h,G)∈V∗=X∗×L2(Ω)×T1. Moreover (v,π,t) satisfies the estimate
‖v‖X+‖π‖L2(Ω)+‖t‖T∗1≤C{‖F‖X∗+‖h‖L2(Ω)+‖G‖T1} | (4.2) |
where C only depends on the constants ν,CS. Let us introduce the Stokes solver
R:V∗→V, |
defined as follows: given (F,h,G)∈V∗,
R(F,h,G)=(vs,πs,ts)∈V |
is the unique solution of (4.1). By (4.2), R is a (linear) continuous isomorphism between V∗ and V. Moreover, the following result holds.
Lemma 4.1. The Stokes solver R is self-adjoint.
Proof. Let (˜F,˜h,˜G)∈V∗ and R(˜F,˜h,˜G)=(˜vs,˜πs,˜ts). We have to show that
⟨(vs,πs,ts),(˜F,˜h,˜G)⟩V,V∗=⟨(F,h,G),(˜vs,˜πs,˜ts)⟩V,V∗. | (4.3) |
Indeed, we have
{e(˜vs,φ)+b(φ,˜πs)+⟨˜ts,φ⟩Γ1=⟨˜F,φ⟩X∗,X∀φ∈Xb(˜vs,ϕ)=(˜h,ϕ)L2(Ω)∀ϕ∈L2(Ω)⟨λ,˜vs⟩Γ1=⟨λ,˜G⟩Γ1∀λ∈T∗1. | (4.4) |
Substituting (φ,ϕ,λ)=(˜vs,˜πs,˜ts) into (4.1) and (φ,ϕ,λ)=(vs,πs,ts) into (4.4) we get
{e(vs,˜vs)+b(˜vs,πs)+⟨ts,˜vs⟩Γ1=⟨F,˜vs⟩X∗,Xb(vs,˜πs)=(h,˜πs)L2(Ω)⟨˜ts,vs⟩Γ1=⟨˜ts,G⟩Γ1 | (4.5) |
and
{e(vs,˜vs)+b(vs,˜πs)+⟨˜ts,vs⟩Γ1=⟨˜F,vs⟩X∗,Xb(˜vs,πs)=(˜h,πs)L2(Ω)⟨ts,˜vs⟩Γ1=⟨ts,˜G⟩Γ1 | (4.6) |
Thus,
⟨F,˜vs⟩X∗,X−b(˜vs,πs)−⟨ts,˜vs⟩Γ1=⟨˜F,vs⟩X∗,X−b(vs,˜πs)−⟨˜ts,vs⟩Γ1 |
or, using the other equations,
⟨F,˜vs⟩X∗,X−(˜h,πs)L2(Ω)−⟨ts,˜G⟩Γ1=⟨˜F,vs⟩X∗,X−(h,˜πs)L2(Ω)+⟨˜ts,G⟩Γ1 |
which easily gives (4.3).
We are now ready for the well-posedness analysis of the optimal control problem
J(v,u)→minsubject to(2.7), v∈X, (π,t)∈Q, u∈Uad. | (5.1) |
Here J is given by either (2.1) or (2.2) and
Uad={u: c0‖u‖H10(Γc)d≤η/2}, |
is a closed, convex set in H10(Γc)d, where η denotes the constant defined in (3.5). Note that the definition of Uad depends on the positive constant c0. If we assume that ‖vin‖T1≤η/2 then
‖g(u)‖T1≤‖vin‖T1+c0‖u‖H10(Γc)d≤η |
and there exists a unique solution v(u)∈X, (π(u),t(u))∈Q, with ‖v(u)‖X≤C1‖g(u)‖T1≤Cη. The following result holds.
Theorem 5.1. If ‖vin‖T1≤η/2, there exists an optimal control ˆu∈Uad and a corresponding optimal state ˆv∈X, (ˆπ,ˆt)∈Q.
Proof. Define the set of feasible points by
F={(v,π,t,u)∈Sδ×Q×Uad such that (2.7) is satisfied}, |
where Sδ={v∈X:‖v‖X≤δ}. The proof follows from classical arguments (see, e.g., [27,Sect. 1.5]), after proving the following four properties:
(1) J(v,u)≥0 and F is nonempty by Theorem 3.3.
(2) Let {vk,πk,tk,uk}∈F be a minimizing sequence. Then {vk,πk,tk,uk} is bounded. Indeed {vk,uk}⊂Sδ×Uad and therefore is bounded in X×H10(Γ1)d. From (3.4),
‖πk‖L2(Ω)+‖tk‖T∗1≤C(η+η2) | (5.2) |
and hence {πk,tk} is bounded in Q.
(3) F is weakly sequentially closed. We only need to check that if {vk,πk,tk,uk} is a solution to
{e(vk,φ)+c(vk,vk,φ)+b(φ,πk)+⟨tk,φ⟩Γ1=0φ∈Xb(vk,ϕ)=0∀ϕ∈L2(Ω)⟨λ,vk⟩Γ1=⟨λ,g(uk)⟩Γ1∀λ∈T∗1 | (5.3) |
and (vk,πk,tk)⇀(v,π,t) in V, uk⇀u in H10(Γc)d, then (v,π,t,u) is a solution of (2.7). Since vk⇀v in X and the trace operator v↦v|Γ1 is continuous from X into T1, it follows that, for every φ∈X, ϕ∈L2(Ω) and λ∈T∗1,
e(vk,φ)→e(v,φ),b(vk,ϕ)→b(v,ϕ) and ⟨λ,vk⟩Γ1→⟨λ,v⟩Γ1. |
Also, πk⇀π in L2(Ω) implies that b(φ,πk)→b(φ,π). Moreover, the embedding of H10(Γ1)d into Tc is continuous and hence uk⇀u in H10(Γc)d implies that ⟨λ,g(uk)⟩Γ1→⟨λ,g(u)⟩Γ1 and ⟨tk,φ⟩Γ1→⟨t,φ⟩Γ1.
Let us write
c(vk,vk,φ)−c(v,v,φ)=c(vk−v,vk,φ)−c(v,vk−v,φ). |
and recall that X is compactly embedded into L4(Ω)d, so that vk⇀v in X implies that vk→v in L4(Ω)d. Thus
|c(vk,vk,φ)−c(v,v,φ)|≤‖vk‖X‖φ‖L4(Ω)d‖vk−v‖L4(Ω)d≤δ‖φ‖L4(Ω)d‖vk−v‖L4(Ω)d→0. |
Finally, thanks to the weak convergence of {vk} in X, we have
c(v,vk−v,φ)→0. |
Passing to the limit for k→∞ into (5.3), we deduce that (v,π,t,u) is a solution of (2.7).
(4) J is continuous and convex and therefore also sequentially weakly lower semicontinuous.
In order to derive first order optimality conditions, we need to show that the solution map S:u↦(v(u),π(u),t(u)) is Fréchet differentiable in Uad. Precisely, we show the following result.
Theorem 6.1. Let ˉu∈Uad,k∈H10(Γc)d such that ˉu+k∈Uad. The solution map S is Fréchet-differentiable at ˉu and its derivative S′(ˉu) is given by
S′(ˉu)k=(v′(ˉu)k,π′(ˉu)k,t′(ˉu)k)=(w,q,s) |
where, setting ˉv=v(ˉu), (w,q,s)∈V is the solution to
{e(w,φ)+c(w,ˉv,φ)+c(ˉv,w,φ)+b(φ,q)+⟨s,φ⟩Γ1=0∀φ∈Xb(w,ϕ)=0∀ϕ∈L2(Ω)⟨λ,w⟩Γ1=⟨λ,χΓck⟩Γ1∀λ∈T∗1. | (6.1) |
Remark 6.2. Since ‖ˉu‖X<η, the bilinear form
a(w,φ)=e(w,φ)+c(w,ˉv,φ)+c(ˉv,w,φ) | (6.2) |
is X−coercive hence the linear problem (6.1) is well posed.
Proof. Let ˉu∈Uad, k∈H10(Γc)d such that ˉu+k∈Uad. Denote by S(ˉu+k)=(vk,πk,tk) and S(ˉu)=(ˉv,ˉπ,ˉt) the corresponding solutions in V to (2.7). Write
c(vk,vk,φ)−c(ˉv,ˉv,φ)=c(vk,vk−ˉv,φ)−c(ˉv−vk,ˉv,φ)=−c(ˉv−vk,ˉv,φ)−c(ˉv,vk−ˉv,φ)+c(ˉv−vk,ˉv−vk,φ). |
Then the difference (d,p,r)=(vk−ˉv,πk−ˉπ,tk−ˉt)=S(ˉu+k)−S(ˉu) satisfies the system
{e(d,φ)−c(d,ˉv,φ)−c(ˉv,d,φ)+b(φ,p)+⟨r,φ⟩Γ1=−c(d,d,φ)b(d,ϕ)=0⟨λ,d⟩Γ1=⟨λ,χΓck⟩Γ1 | (6.3) |
for all (φ,ϕ,λ)∈V. Therefore,
(d,p,r)−(w,q,s)=S(ˉu+k)−S(ˉu)−S′(ˉu)k |
is a solution to the problem
{e(z,φ)+c(z,ˉv,φ)+c(ˉv,z,φ)+b(φ,p−q)+⟨r−s,φ⟩Γ1=−c(d,d,φ)b(z,ϕ)=0⟨λ,z⟩Γ1=0. | (6.4) |
for all (φ,ϕ,λ)∈V, where z=d−w. From estimates (3.17), (3.18) we obtain
‖z‖X≤C‖f‖X∗ | (6.5) |
and
‖p−q‖L2(Ω)+‖r−s‖T∗1≤C‖f‖X∗ | (6.6) |
where C depends only on ν, ˜β,c0,CS and f is the functional φ⟼−c(d,d,φ).
Using (3.19) we obtain
|c(d,d,φ)|≤C2S‖d‖2X‖φ‖X≤C2SCL‖k‖2Tc‖φ‖X |
from which
‖f‖X∗≤C2SC2‖k‖2Tc. | (6.7) |
From (6.6) and (6.7) we have
‖S(ˉu+k)−S(ˉu)−S′(ˉu)k‖V≤C3‖k‖2Tc≤c0C3‖k‖2H10(Γc)d |
where C3 depends only on the constants ν, ˜β,CS.
Remark 6.3. From the previous proof, it is clear that S is differentiable also with respect to the Tc-norm.
We are now ready to write the first order optimality conditions for our OCP (5.1). Precisely, we can show the following result.
Theorem 7.1. Let ˆu be an optimal control and (ˆv,ˆπ,ˆt) the corresponding optimal state. There exists a unique multiplier (ˆz,ˆq,ˆs)∈V such that ˆu and (ˆz,ˆq,ˆs) satisfy the following optimality system:
● state problem (2.7);
● adjoint problem: for every (ψ,ϕ,λ)∈V,
{e(ψ,ˆz)+c(ˆv,ψ,ˆz)+c(ψ,ˆv,ˆz)+b(ψ,ˆq)+⟨ˆs,ψ⟩Γ1=˜J′v(ˆv,ˆu)ψb(ˆz,ϕ)=0⟨λ,ˆz⟩Γ1=0; | (7.1) |
● variational inequality: for every k∈Uad,
J′(ˆu)(k−ˆu)=(τˆu,k−ˆu)H10(Γc)d+⟨ˆs,k−ˆu⟩Γc≥0 | (7.2) |
where J(u)=˜J(v(u),u) is the reduced cost functional.
Proof. Define the nonlinear operator N:V×Tc→V∗ by
N(v,π,t,u)=(v⋅∇v,−h,−g(u)). |
The nonlinear component of N is a continuous bilinear operator and therefore N is Fréchet differentiable‡, with
‡Since H10(Γc) is continuously embedded into H1/200(Γc) it follows that N is differentiable also as an operator from H10(Γc)d into V∗ and the two differentials coincide.
N′(v,π,t)(v,π,t,u)(ψ,q,r)=((v⋅∇)ψ+(ψ⋅∇)v,0,0) | (7.3) |
and
N′u(v,π,t,u)k=(0,0,−g0(k)) | (7.4) |
where
g0(k)={kon Γc0on Γw∪Γin. |
Consider now the Stokes solver R:V∗→V defined in Section 4. In terms of R and N, the state problem can be written as
G(v,π,t,u)=R−1(v,π,t)+N(v,π,t,u)=0 | (7.5) |
with G:V×T1→V∗.
In Lemma 7.3 below we prove that:
(a) the adjoint operator N′(v,π,t)(v,π,t,u)∗::V→V∗ is given by
N′(v,π,t)(v,π,t,u)∗:(z,q,s)=(−(v⋅∇)z+(∇v)⊤z+[(v⋅n)z]|Γ1,0,0) | (7.6) |
where n is the exterior unit normal to Γc;
(b) the adjoint operator N′u(v,π,t,u)∗:V→T∗c is given by
N′u(v,π,t,u)∗:(z,q,s)=−s. | (7.7) |
By a well known theorem (see, e.g., [27,Thm. 1.45]) the existence of a unique multiplier (ˆz,ˆq,ˆs)∈V follows from the properties (i),(ii),(iii) listed below.
(i) ˜J and G are Fréchet differentiable, with
˜J′v(v,u)w={(∇×v,∇×w)L2(Ω)d in the case of J1(v,u)(v−zd,w)L2(Ω)d in the case of J2(v,u)∀w∈X˜J′u(v,u)k=τ(u,k)H10(Γc)d∀k∈H10(Γc)d |
and ˜J′π(v,u)=0, ˜J′t(v,u)=0. Moreover, since R is self-adjoint,
G′(v,π,t,u)(w,q,r,k)=R−1(w,q,r)+N′(v,π,t,u)(w,q,r,k). |
(ii) From Theorem 6.1 we know that, for every u∈Uad, the state equation (7.5) defines a unique solution map S(u) which is Fréchet-differentiable.
(iii) For every ˉu∈Uad, the partial derivative G′(v,q,s)(ˉv,ˉπ,ˉt,ˉu), with (ˉv,ˉπ,ˉt)=(v(ˉu),π(ˉu),t(ˉu)) is a continuous isomorphism between V and V∗. Indeed, given (F,h,G)∈V∗, the equation
G′(v,π,t)(ˉv,ˉπ,ˉt,ˉu)(ψ,η,r)=R−1(ψ,η,r)+N′(v,π,t)(ˉv,ˉπ,ˉt,ˉu)(ψ,η,r)=(F,h,G), | (7.8) |
in weak form reads:
{e(ψ,φ)+c(ψ,ˉv,φ)+c(ˉv,ψ,φ)+b(φ,η)+⟨r,φ⟩Γ1=⟨F,φ⟩X∗,X∀φ∈Xb(ψ,ϕ)=(h,ϕ)L2(Ω)∀ϕ∈L2(Ω)⟨λ,ψ⟩Γ1=⟨λ,G⟩Γ1∀λ∈T∗1. | (7.9) |
Since ˉu∈Uad, the bilinear form
a(ψ,φ)=e(ψ,φ)+c(ψ,ˉv,φ)+c(ˉv,ψ,φ) |
is X−coercive, therefore problem (7.9) is well posed and G′(v,q,s)(ˉv,ˉπ,ˉt,ˉu) is a continuous isomorphism between V and V∗.
Introducing the Lagrangian functional
L((v,π,t),u,(z,q,s))=˜J(v,u)−⟨G(v,π,t,u),(z,q,s)⟩V∗,V, |
where (z,q,s) is a multiplier, the first order optimality conditions are obtained by differentiating L. Precisely, the adjoint problem reads (ˆL′(⋅)=L′(⋅)((ˆv,ˆπ,ˆt),ˆu,(ˆz,ˆq,ˆs))):
ˆL′(v,π,t)(ψ,η,r)=⟨˜J′(v,π,t)(ˆv,ˆu),(ψ,η,r)⟩V∗,V−⟨G′(v,π,t)(ˆv,ˆπ,ˆt,ˆu)(ψ,η,r),(ˆz,ˆq,ˆs)⟩V∗,V=0 |
or, equivalently, upon introducing the adjoint operator G′(v,π,t)(ˆv,ˆπ,ˆt,ˆu)∗,
⟨˜J′(v,π,t)(ˆv,ˆu)−G′(v,π,t)(ˆv,ˆπ,ˆt,ˆu)∗(ˆz,ˆq,ˆs),(ψ,η,r)⟩V∗,V=0 | (7.10) |
for every (ψ,η,r)∈V.
Since R is self-adjoint, we have (see the proof of Lemma 7.3)
⟨G′(v,π,t)(ˆv,ˆπ,ˆt,ˆu)∗(ˆz,ˆq,ˆs),(ψ,η,r)⟩V∗,V=⟨R−1(ˆz,ˆq,ˆs)+N′(v,π,t)(ˆv,ˆπ,ˆt,ˆu)∗(ˆz,ˆq,ˆs),(ψ,η,r)⟩V∗,V=⟨R−1(ˆz,ˆq,ˆs),(ψ,η,r)⟩V∗,V+c(ˆv,ψ,ˆz)+c(ψ,ˆv,ˆz)+(η,0)L2(Ω)+⟨r,0⟩Γ1=0 |
and therefore (7.10) corresponds to the following system:
{e(ψ,ˆz)+c(ˆv,ψ,ˆz)+c(ψ,ˆv,ˆz)+b(ψ,ˆq)+⟨ˆs,ψ⟩Γ1=˜J′v(v,u)ψb(ˆz,η)=0⟨r,ˆz⟩Γ1=0 | (7.11) |
for every (ψ,η,r)∈V, which is (7.1). The variational inequality reads, recalling (7.7),
ˆL′u(k−ˆu)=⟨˜J′u(ˆv,ˆu),k−ˆu⟩U∗,U+⟨N′u(ˆv,ˆπ,ˆt,ˆu)∗(ˆz,ˆq,ˆs),k−ˆu⟩V∗,V=(τˆu,k−ˆu)H10(Γc)2+⟨ˆs,k−ˆu⟩Γc≥0∀k∈Uad |
which is (7.2).
As in Remark 6.2, since ˆu∈Uad, the bilinear form (adjoint of the bilinear form (6.2))
a∗(z,ψ)=e(ψ,z)+c(ψ,ˆv,z)+c(ˆv,ψ,z) |
is X−coercive and the linear problem (7.1) is well posed. To write the adjoint system in strong form, note that
c(ψ,ˆv,ˆz)=∫Ω(ψ⋅∇)ˆv⋅ ˆz dx=d∑i,j=1∫Ωψi∂ˆvj∂xiˆzj dx=∫Ω(∇ˆv)⊤ˆz⋅ψ dx | (7.12) |
and
c(ˆv,ψ,ˆz)=∫Ω(ˆv⋅∇)ψ⋅ˆz dx=∫∂Ω(ˆv⋅n)(ˆz⋅ψ) dσ−∫Ω(ˆv⋅∇)ˆz⋅ψ dx−∫Ωdivˆv (ˆz⋅ψ) dx=∫Γout(ˆv⋅n)(ˆz⋅ψ) dσ−c(ˆv,ˆz,ψ) | (7.13) |
since div ˆv=0, ˆz=0 on ∂Ω∖Γout.
Moreover, in the case of the vorticity functional ˜J1,
∫Ω(∇׈v)⋅(∇×ψ) dx=∫Ω(∇×∇׈v)⋅ψ dx+∫∂Ω(∇ˆv−(∇ˆv)⊥)n⋅ψ dσ. |
Therefore, the adjoint system in strong form reads, setting the rotation tensor ∇ˆv−(∇ˆv)⊥=ˆW,
{−νΔˆz−(ˆv⋅∇)ˆz+(∇ˆv)⊤ˆz+∇ˆq=∇×∇׈vin Ωdivˆz=0in Ωˆz=0on Γ0∪Γ1ˆqn−ν∂ˆz∂n−(ˆv⋅n)ˆz+ˆWn=0on Γout. | (7.14) |
We can also deduce that, regarding the multiplier,
ˆs=ˆqn−ν∂ˆz∂n+ˆWnon Γ1. | (7.15) |
In the case of the velocity tracking-type cost functional ˜J2, the strong form of the adjoint problem is straightforward to obtain, and reads as follows:
{−νΔˆz−(ˆv⋅∇)ˆz+(∇ˆv)⊤ˆz+∇ˆq=ˆv−zdin Ωdivˆz=0in Ωˆz=0on Γ0∪Γ1ˆqn−ν∂ˆz∂n−(ˆv⋅n)ˆz=0on Γout. | (7.16) |
In this case, we find instead that
ˆs=ˆqn−ν∂ˆz∂non Γ1. | (7.17) |
Remark 7.2. From (7.2) we deduce that the derivative of the reduced functional J(u)=˜J(v(u),u) at u=ˆu∈Uad is given by
J′(ˆu)k=(τˆu,k)H10(Γc)d+⟨ˆs,k⟩Γc∀k∈Uad. |
Then, introducing the embedding operator B:T∗c→H−1(Γc)d defined by
⟨Bˆs,k⟩H−1(Γc)d,H10(Γc)d=⟨ˆs,k⟩Γc∀k∈H10(Γc)d, |
we find that
∇J(ˆu)=τˆu+ˆw |
where ˆw∈H10(Γc)d is the unique weak solution to
−Δˆw=Bˆs in Γc. | (7.18) |
The variational inequality (7.2) can be written in the form
(τˆu+ˆw,k−ˆu)H10(Γc)d≥0∀k∈Uad. | (7.19) |
If Bˆs≠0 and τ=0, (7.19) gives
ˆu=−η2c0ˆw‖ˆw‖H10(Γc)d. | (7.20) |
If Bˆs≠0 and τ>0, we have ˆu=−ˆw/τ if ˆw/τ∈Uad, otherwise ˆu is given by (7.20).
Lemma 7.3. The following expressions hold:
(a) N′(v,π,t)(v,π,t,u)∗:V→V∗ is given by
N′(v,π,t)(v,π,t,u)∗(z,q,s)=(−(v⋅∇)z+(∇v)⊤z+[(v⋅n)z]|Γ1,0,0). | (7.21) |
where n is the exterior unit normal to Γc.
(b) N′u(v,π,t,u)∗:V→T∗c is given by
N′u(v,π,t,u)∗:(z,q,s)=−s. | (7.22) |
Proof. (a) We check (7.21). Using (7.12), (7.13) and recalling (7.3), we can write
⟨N′(v,π,t)(v,π,t,u)(ψ,η,r),(z,q,s)⟩V∗,V=c(v,ψ,z)+c(ψ,v,z)=∫Ω(v⋅∇)ψ⋅zdx+∫Ω(ψ⋅∇)v ⋅ zdx=∫Ω{−(v⋅∇)z⋅ψ+(∇v)⊤z ⋅ ψ}dx+∫Γ1(v⋅n)(z ⋅ ψ)dσ+(η,0)L2(Ω)+⟨r,0⟩Γ1=⟨N′(v,π,t)(v,π,t,u)∗(z,q,s),(ψ,η,r)⟩V∗,V | (7.23) |
for every (z,q,s)∈V and every (ψ,η,r)∈V∗.
(b) We check (7.22). We have, using (7.4),
⟨N′u(v,π,t,u)k,(z,q,s)⟩V∗,V=⟨s,−k⟩Γc=⟨−s,k⟩Γc=⟨N′u(v,π,t,u)∗(z,q,s),k⟩T∗c,Tc |
for every (z,q,s)∈V and k∈Tc.
For the numerical solution of the OCP (5.1) we use an iterative method for the minimization of the cost functional, and we exploit the Galerkin finite element method for the numerical discretization of the state equation, the adjoint equation and the optimality equation. In this way we can take advantage of the system of first-order (necessary) optimality conditions we have derived in the previous sections. Several alternatives have been proposed in literature; see, e.g., [14,25,28,43] for the case of sequential quadratic programming methods applied to boundary control of steady Navier-Stokes flows, [9] for an augmented Lagrangian method, and [5] for inexact Newton methods. A review of numerical strategies for OCPs can be found, e.g., in [1,18,26].
The state system consists of the steady Navier-Stokes equations, while the adjoint system is given by a steady Oseen problem. Following the saddle-point approach described so far, for both these problems Dirichlet boundary conditions are enforced in weak form, by means of the Lagrange multipliers t and s, respectively. These two latter variables are indeed related with the boundary stress over the Dirichlet boundary, see, e.g., equations (2.8) and (7.15)–(7.17). As a matter of fact, for the numerical approximation, we can rewrite the state and the adjoint equations in a conventional manner, by introducing appropriate functional spaces and suitable lifting terms to take into account non-homogeneous Dirichlet conditions. This option allows us to avoid the explicit computation of the Lagrange multipliers t and s.
For the numerical approximation of both these problems, let us introduce two subspaces Xh⊂X and Qh⊂L2(Ω), of dimension NV,NQ<+∞, respectively, and let us denote by vh∈Xh and ph∈Qh the FE approximations for the velocity and the pressure fields [44]. Similarly, zh∈Xh and qh∈Qh will denote the FE approximations for the adjoint velocity and pressure fields, respectively. We also denote by uh∈Uh∩Uad the FE approximation of the control variable, where Uh⊂U is a subspace of the control space U=H10(Γc)d of dimension NU<+∞.
The Galerkin-FE approximation of the state system (2.7) thus reads as follows: we seek for (vh,ph)∈Xh×Qh such that
{˜e(vh,φh)+c(vh,vh,φh)+b(φh,πh)=F1(φh)∀φh∈Xhb(vh,ϕh)=F2(ϕh)∀ϕh∈Qh, | (8.1) |
where vDh=vDh(uh)∈Xh is a discrete (lifting) function interpolating the Dirichlet data,
˜e(vh,φh)=e(vh,φh)+c(vDh,vh,φh)+c(vh,vDh,φh),F1(φh)=−e(vDh,φh)−c(vDh,vDh,φh),F2(ϕh)=−b(vDh,ϕh). |
All the forms are continuous over the discrete spaces Xh and Qh. Conversely, the approximation stability is ensured by imposing that the coercivity and inf-sup conditions are still valid at the discrete level. In particular, the coercivity of e over Xh is inherited from that over X; on the other hand, we require that b is inf-sup stable over Xh×Qh, so that the following discrete Brezzi inf-sup condition, holds [6]:
∃ ˜βh>0: βh=infq∈Qhsupw∈Xhb(q,w)‖w‖V‖q‖Q≥˜βh. | (8.2) |
This last property is ensured e.g., by choosing Xh×Qh as the space of Taylor-Hood P2−P1 finite elements for the velocity and the pressure, respectively (in both the state and the adjoint problem); however, this choice is not restrictive --- the whole construction keeps holding for other spaces combinations as well [16].
We can now derive the matrix formulation corresponding to the Galerkin-FE approximation (8.1). Let us denote by {ϕvi}Nvi=1 and {ϕpi}Npi=1 the Lagrangian basis of the FE spaces Xh and Qh, respectively, so that we can express the FE velocity and pressure as
vh=Nv∑i=1uiϕvi,ph=Np∑i=1piϕpi. | (8.3) |
We remark that the solution to (8.1) is vanishing on the whole Dirichlet boundary, so that the corresponding velocity approximation fulfilling the boundary conditions is given by vh+vDh. Thus, by denoting as v_h∈RNv and p_h∈RNp the vectors of the degrees of freedom appearing in (8.3), problem (8.1) can be rewritten as:
[Eh+Ch(v_h)BThBh0][v_hp_h]=[f1,hf2,h] | (8.4) |
where, for 1≤i,j≤Nv and 1≤k≤Np,
(Eh)ij=˜e(ϕvj,ϕvi),(Bh)ki=b(ϕpk,ϕvi),(Ch(w_)h)ij=Nv∑m=1wmc(ϕvm,ϕvj,ϕvi),(f1,h)i=F1(ϕvi),(f2,h)k=F2(ϕpk), | (8.5) |
We solve the nonlinear saddle-point problem (8.4) by means of a fixed-point method (also referred to as Picard iteration, see, e.g., [11,Sect. 8.2] it has a larger ball of convergence than Newton's method (see e.g., [11], Chapter 7.2 and references therein). Moreover, if a small data condition (like, e.g., (3.3) for the case at hand) is satisfied, the fixed-point method is globally convergent. Thus, starting from an initial guess (v_(0)h,p_(0)h), for k≥1 we solve
[Eh+Ch(v_(k−1)h)BThBh0][v_(k)hp_(k)h]=[f1,hf2,h], | (8.6) |
to obtain (v_(k)h,p_(k)h), until ‖v_(k)h−v_(k−1)h‖V≤εNStol, given a small tolerance εNStol>0. As initial guess, we take the Stokes solution of (8.4). Each Oseen system (8.6) is solved by means of a sparse LU factorization.
Regarding instead the numerical approximation of the adjoint variables, the adjoint system (7.11) is linear with respect to both the adjoint velocity and pressure. Moreover, it depends on the state variables only through the term c(ˆv,ψ,ˆz)+c(ψ,ˆv,ˆz) and the right-hand side of (7.11); hence, the FE arrays corresponding to these two terms must be reassembled at each step during the optimization procedure. Given the approximation vh=vh(uh) of the (state) velocity and uh of the control function, the Galerkin-FE approximation of the adjoint system (7.11) thus reads as follows: we seek for (zh,qh)∈Xh×Qh such that
{e(φh,zh)+c(vh,zh,φh)+c(zh,vh,φh)+b(φh,qh)=G1(vh,φh)∀φh∈Xhb(zh,ϕh)=0∀ϕh∈Qh, | (8.7) |
where
G1(vh,φh)=˜J′v(vh,uh)φh. |
Problem (8.7) can be rewritten algebraically as:
[ETh+Dh(v_h)BThBh0][z_hq_h]=[g1,h(vh)0] | (8.8) |
where, for 1≤i,j≤Nv and 1≤k≤Np,
(Dh(w_)h)ij=Nv∑m=1wm(c(ϕvm,ϕvj,ϕvi)+c(ϕvj,ϕvm,ϕvi)),(g1,h(vh))i=G1(vh,ϕvi). | (8.9) |
We now describe the numerical optimization method exploited for the solution of our OCP. We rely on a gradient-based algorithm exploiting the gradient ∇J(uh) to iteratively update the control until a suitable convergence criterion is fulfilled; a similar approach can be found, e.g., in [8]. Notable instances are descent methods, such as the gradient, (non-linear) conjugate gradient, quasi-Newton or Newton methods. In the simplest case of a gradient (or steepest descent) method, starting from an initial guess u(0), we iteratively generate a sequence
u(k+1)h=PUad(u(k)h−λ(k)∇J(u(k)h)),k≥0, |
where λ(k)>0 is a step size, until, e.g., a suitable stopping criterion is fulfilled, and PUad denotes the projection onto Uh∩Uad, to be evaluated similarly to what we have done in (7.20).
As stopping criterion, we might require that either
‖∇J(u(k)h)‖L2(Ω)d<ε |
or
|J(u(k)h)−J(u(k−1)h)|<ε | (8.10) |
for a given tolerance ε>0.
At each step of the gradient method, we solve the state Navier-Stokes system (8.1), from which we obtain v(k), and the value of the cost functional ˜J(v(k),u(k)). Then, we solve the adjoint system (8.7) (v(k) and u(k) being given) and obtain (z(k),q(k)). From these variable, we determine the boundary stress
s(k)h=q(k)hn−ν∂z(k)h∂non Γc, |
and approximate (by the Galerkin finite element method) the solution of the following problem,
∫Γc∇w(k)h:∇φhdσ=∫Γcs(k)h⋅φhdσ∀φh∈H10(Γc)d |
in order to evaluate the gradient
∇J(u(k)h)=τu(k)h+w(k)h. |
We highlight that at each step of the steepest descent method, we take as initial guess of the Newton method used to solve the state Navier-Stokes system the state velocity computed at the previous step of the optimization procedure, instead of computing it by solving a Stokes system, for the sake of computational efficiency.
In this section we present the numerical results related with an OCP for the optimal control of blood flows – described through a steady Navier-Stokes model – through a bypass graft. This latter usually provides blood flow through an alternative bridging path in order to overcome critically occluded arteries; one of the most dangerous cases is related to coronary arteries, which supply the oxygen-rich blood perfusion to the heart muscle. The design of the connection between the graft vessel and the host arterial vessel is a critical factor in avoiding post-operative recurrence of the occlusion, since fluid dynamic phenomena such as recirculation, oscillating or untypical high/low shear rates, and stagnation areas, can cause the growth of another stenosis downstream the arterial-graft connection. Hence, a typical design of a bypass graft aims at minimizing some cost functionals related to haemodynamic quantities [10,42]. However, a rigorous model for blood circulation should take into account (ⅰ) the flow unsteadiness, (ⅱ) the arterial wall deformability, and possibly (ⅲ) complex rheological model to characterize the agammaegated nature of blood. Optimal control (and optimal design) problem require the repeated simulation of these flow equations (and the evaluation of the cost functional to be minimized); to reduce the computational burden, we adopt steady incompressible Navier-Stokes equations for laminar Newtonian flows.
We focus on the minimization of the tracking-type functional (2.2) in order to drive the blood velocity towards a specified velocity target state zd, featuring a regular pattern; this latter is given by the solution of the Stokes system in the same domain. Other possible cost functionals for the problem at hand have been employed, e.g., in [37,38]; the effect of uncertainty (e.g., affecting either the residual flow across Γin, or the geometrical configuration of the bypass) has been explored in [38,39,45].
Here a two-dimensional geometrical configuration is considered, although the whole theoretical analysis and the computational pipeline are perfectly valid for three-dimensional problems as well. For the case at hand, we consider a fluid flow in a tract of blood vessel where a bypass is inserted. As a matter of fact, the whole bypass graft is not simulated; its action is represented via a velocity control u acting on the boundary Γc⊂∂Ω, the interface where the final portion of the graft and the host vessel meet (see Figure 1). We consider two different scenarios: (i) a total occlusion of the host artery (setting vin=0 on Γin, and (ii) the presence of a residual blood flow (vin≠0) on that boundary. We choose ρ=1, ˉv=25cms−1, ˉd=|Γc|≈0.75cm from which, defining the Reynolds number Re=ρˉvˉd/μ, which is set equal to Re=200, the dynamic viscosity coefficient μ can be deduced. We first compute the solution to the Stokes system, to determine the target velocity profile zd; we set the relaxation parameter λ(k)=τ=0.1 and the regularization parameter τ=0.05. We consider the stopping criterion based on the difference of two successive iterates (8.10), for which we select ε=0.01; the tolerance for the stopping criterion of the Newton method is instead given by εNStol=10−5. For our computations we use a mesh with 27,233 triangular elements and 14,048 vertices.
We first consider the case of total occlusion of the host arterial vessel, that is, vin=0. In Figure 2 we report the velocity (both its magnitude and the streamlines) and pressure fields corresponding to the initial control function u(0)h and the optimal control ˆuh obtained when the steepest descent algorithm stops, after Nit=157 iterations. In this case, the cost functional is reduced of about 37, decreasing from J(u(0)h)=288.241 to J(ˆuh)=181.445. A reduction in the magnitude of the adjoint velocity field, which can be remarked by comparing the plots displayed in Figure 3, also shows that the derivative of the cost functional evaluated at uh=ˆuh is smaller than the corresponding quantity calculated when uh=u(0)h.
The magnitude of the vorticity is reported in Figure 4: a global decrease of this quantity is indeed obtained also by minimizing the tracking-type cost functional, as it results by comparing the vorticity magnitude for the initial and the optimal control.
Similar results are obtained when a residual flow across Γin is present; in this case, we set
vin=(aexp(−(y−ˉy)22b2),0)T |
with a=50 and b=10−2, to model a (limited) blood flow across the occluded vessel. Its presence actually reduces the recirculation around the graft-artery junction, (compare Figure 5 and Figure 2). Also in this case a remarkable reduction of the value of the cost functional is obtained: the steepest descent algorithm stops after Nit=151 iterations, and the cost functional is reduced of about 37, decreasing from J(u(0)h)=289.796 to J(ˆuh)=183.565. Similar considerations can be obtained by comparing the adjoint velocity and the vorticity fields calculated for the initial and the optimal controls (Figures 6 and 7).
In this paper we have proposed a saddle point approach to deal with boundary control problems for the steady Navier-Stokes equations in presence of mixed Dirichlet-Neumann boundary conditions. Both the well posedness analysis for the optimal control problem and the derivation of a system of first-order optimality conditions have been carried out, taking advantage of a suitable treatment of boundary Dirichlet controls realized by means of Lagrange multipliers. The minimization of quadratic cost functionals of the velocity has been considered for the sake of flow regularization. A steepest descent method has then been applied for numerical optimization, exploiting the Galerkin finite element method for the discretization of the derived system of optimality conditions. Numerical results have been shown for simplified two-dimensional fluid flows in a tract of blood vessel where a bypass is inserted, proving the efficacy of our approach.
The authors declare no conflict of interest.
[1] | Akcelik V, Biros G, Ghattas O, et al. (2006) Parallel algorithms for PDE-constrained optimization, In: Heroux, M.A., Raghavan, P., Simon, H.D., Editors, Parallel Processing for Scientific Computing, Philadelphia: Society for Industrial and Applied Mathematics, 291–322. |
[2] |
Beneš M, Kučera P, (2012) On the Navier–Stokes flows for heat-conducting fluids with mixed boundary conditions. J Math Anal Appl 389: 769–780. doi: 10.1016/j.jmaa.2011.12.017
![]() |
[3] |
Beneš M, Kučera P, (2016) Solutions to the Navier–Stokes equations with mixed boundary conditions in two-dimensional bounded domains. Math Nachr 289: 194–212. doi: 10.1002/mana.201400046
![]() |
[4] |
Berggren M, (1998) Numerical solution of a flow-control problem: vorticity reduction by dynamic boundary action. SIAM J Sci Comput 19: 829–860. doi: 10.1137/S1064827595294678
![]() |
[5] | Biros G, Ghattas O, (2005) Parallel Lagrange–Newton–Krylov–Schur methods for PDEconstrained optimization. Part II: The Lagrange–Newton solver and its application to optimal control of steady viscous flows. SIAM J Sci Comput 27: 714–739. |
[6] | Brezzi F, (1974) On the existence, uniqueness, and approximation of saddle point problems arising from Lagrangian multipliers. RAIRO 2: 129–151. |
[7] | Dautray R, Lions JL, (2000) Mathematical Analysis and Numerical Methods for Science and Technology. Berlin Heidelberg: Springer-Verlag. |
[8] |
Dedè L, (2007) Optimal flow control for Navier-Stokes equations: Drag minimization. Int J Numer Meth Fluids 55: 347–366. doi: 10.1002/fld.1464
![]() |
[9] |
Desai M, Ito K, (1994) Optimal controls of Navier–Stokes equations. SIAM J Control Optim 32: 1428–1446. doi: 10.1137/S0363012992224972
![]() |
[10] |
Do H, Owida AA, Morsi YS, (2012) Numerical analysis of coronary artery bypass grafts: An over view. Comput Meth Prog Bio 108: 689–705. doi: 10.1016/j.cmpb.2011.12.005
![]() |
[11] | Elman H, Silvester D, Wathen A, (2005) Finite Elements and Fast Iterative Solvers with Applications in Incompressible Fluid Dynamics. Series in Numerical Mathematics and Scientific Computation, Oxford University Press. |
[12] |
Fursikov A, Gunzburger MD, Hou L, (1998) Boundary value problems and optimal boundary control for the Navier–Stokes system: The two-dimensional case. SIAM J Control Optim 36: 852–894. doi: 10.1137/S0363012994273374
![]() |
[13] | Fursikov A, Rannacher R, (2010) Optimal neumann control for the two-dimensional steadystate Navier-Stokes equations, In: Fursikov, A.V., Galdi, G.P., Pukhnachev, V.V., Editors, New Directions in Mathematical Fluid Mechanics: The Alexander V. Kazhikhov Memorial Volume, Basel: Birkhäuser Basel. |
[14] |
Ghattas O, Bark J, (1997) Optimal control of two- and three-dimensional incompressible Navier- Stokes flows. J Comput Phys 136: 231–244. doi: 10.1006/jcph.1997.5744
![]() |
[15] | Girault V, Raviart PA, (1986) Finite Element Methods for Navier-Stokes Equations: Theory and Algorithms. Berlin and New York: Springer-Verlag. |
[16] | Gresho P, Sani R, (1998) Incompressible Flow and the Finite Element Method: Advection- Diffusion and Isothermal Laminar Flow. John Wiley & Sons. |
[17] |
Guerra T, Sequeira A, Tiago J, (2015) Existence of optimal boundary control for the Navier- Stokes equations with mixed boundary conditions. Port Math 72: 267–283. doi: 10.4171/PM/1968
![]() |
[18] | Gunzburger MD, (2003) Perspectives in Flow Control and Optimization. Series in Advances in Design and Control. Philadephia: Society for Industrial and Applied Mathematics. |
[19] |
Gunzburger MD, Hou L, Svobodny T, (1992) Boundary velocity control of incompressible flow with an application to viscous drag reduction. SIAM J Control Optim 30: 167–181. doi: 10.1137/0330011
![]() |
[20] |
Gunzburger MD, Hou L, Svobodny TP, (1991) Analysis and finite element approximation of optimal control problems for the stationary Navier-Stokes equations with distributed and Neumann controls. Math Comput 57: 123–151. doi: 10.1090/S0025-5718-1991-1079020-5
![]() |
[21] |
Gunzburger MD, Hou LS, Svobodny TP, (1991) Analysis and finite element approximation of optimal control problems for the stationary Navier-Stokes equations with Dirichlet controls. ESAIM: Math Model Numer Anal 25: 711–748. doi: 10.1051/m2an/1991250607111
![]() |
[22] |
Gunzburger MD, Manservisi S, (1999) The velocity tracking problem for Navier-Stokes flows with bounded distributed controls. SIAM J Control Optim 37: 1913–1945. doi: 10.1137/S0363012998337400
![]() |
[23] |
Gunzburger MD, Manservisi S, (2000) Analysis and approximation of the velocity tracking problem for Navier-Stokes flows with distributed control. SIAM J Numer Anal 37: 1481–1512. doi: 10.1137/S0036142997329414
![]() |
[24] |
Gunzburger MD, Manservisi S, (2000) The velocity tracking problem for Navier-Stokes flows with boundary control. SIAM J Control Optim 39: 594–634. doi: 10.1137/S0363012999353771
![]() |
[25] | Heinkenschloss M, (1998) Formulation and analysis of a sequential quadratic programming method for the optimal Dirichlet boundary control of Navier-Stokes flow, In: Hager, W.W., |
[26] | Pardalos, P.M., Authors, Optimal Control: Theory, Algorithms, and Applications, Springer US, 178–203. |
[27] |
Herzog R, Kunisch K, (2010) Algorithms for PDE-constrained optimization. GAMM Mitteilungen 33: 163–176. doi: 10.1002/gamm.201010013
![]() |
[28] | Hinze M, Pinnau R, Ulbrich M, et al. (2009) Optimization with PDE Constraints, Series in Mathematical Modelling: Theory and Applications, Springer Netherlands. |
[29] | Hou L, Ravindran SS, (1999) Numerical approximation of optimal flow control problems by a penalty method: Error estimates and numerical results. SIAM J Sci Comput 20: 1753–1777. |
[30] |
Jameson A, (1988) Aerodynamic design via control theory. J Sci Comput 3: 233–260. doi: 10.1007/BF01061285
![]() |
[31] | Jameson A, (1995) Optimum aerodynamic design using CFD and control theory. In: Proceedings of the 12th AIAA Computational Fluid Dynamics Conference 1995, 926–949. |
[32] | Kim H, (2006) A boundary control problem for vorticity minimization in time-dependent 2D Navier-Stokes equations. Korean J Math 23: 293–312. |
[33] |
Kim H, Kwon O, (2006) On a vorticity minimization problem for the stationary 2D Stokes equations. J Korean Math Soc 43: 45–63. doi: 10.4134/JKMS.2006.43.1.045
![]() |
[34] |
Koltukluoğlu T, Blanco P, (2018) Boundary control in computational haemodynamics. J Fluid Mech 847: 329–364. doi: 10.1017/jfm.2018.329
![]() |
[35] |
Kračmar S, Neustupa J, (2001) A weak solvability of a steady variational inequality of the Navier- Stokes type with mixed boundary conditions. Nonlin Anal 47: 4169–4180. doi: 10.1016/S0362-546X(01)00534-X
![]() |
[36] |
Kračmar S, Neustupa J, (2018) Modeling of the unsteady flow through a channel with an artificial outflow condition by the Navier–Stokes variational inequality. Math Nachr 291: 1801–1814. doi: 10.1002/mana.201700228
![]() |
[37] | Kučera P, Skalák Z, (1998) Local solutions to the Navier–Stokes equations with mixed boundary conditions. Acta Appl Math 54: 275–288. |
[38] |
Kunisch K, Vexler B, (2007) Optimal vortex reduction for instationary flows based on translation invariant cost functionals. SIAM J Control Optim 46: 1368–1397. doi: 10.1137/050632774
![]() |
[39] |
Lassila T, Manzoni A, Quarteroni A, et al. (2013) Boundary control and shape optimization for the robust design of bypass anastomoses under uncertainty. ESAIM-Model Num 47: 1107–1131. doi: 10.1051/m2an/2012059
![]() |
[40] |
Lassila T, Manzoni A, Quarteroni A, et al. (2013) A reduced computational and geometrical framework for inverse problems in haemodynamics. Int J Numer Meth Bio 29: 741–776. doi: 10.1002/cnm.2559
![]() |
[41] |
Lei M, Archie J, Kleinstreuer C, et al. (1997) Computational design of a bypass graft that minimizes wall shear stress gradients in the region of the distal anastomosis. J Vasc Surg 25: 637–646. doi: 10.1016/S0741-5214(97)70289-1
![]() |
[42] |
Maz'ya V, Rossmann J, (2009) Mixed boundary value problems for the stationary Navier–Stokes system in polyhedral domains. Arch Ration Mech Anal 194: 669–712. doi: 10.1007/s00205-008-0171-z
![]() |
[43] |
Migliavacca F, Dubini G, (2005) Computational modeling of vascular anastomoses. Biomech Model Mechanobiol 3: 235–250. doi: 10.1007/s10237-005-0070-2
![]() |
[44] |
Prudencio EE, Byrd R, Cai XC, (2006) Parallel full space SQP Lagrange–Newton–Krylov- Sschwarz algorithms for PDE-constrained optimization problems. SIAM J Sci Comput 27: 1305– 1328. doi: 10.1137/040602997
![]() |
[45] | Quarteroni A, Valli A, (1994) Numerical Approximation of Partial Differential Equations. Berlin- Heidelberg: Springer-Verlag. |
[46] |
Sankaran S, Marsden A, (2010) The impact of uncertainty on shape optimization of idealized bypass graft models in unsteady flow. Phys Fluids 22: 121902. doi: 10.1063/1.3529444
![]() |
1. | Pierre-Henri Cocquet, Michaël Rakotobe, Delphine Ramalingom, Alain Bastide, Error analysis for the finite element approximation of the Darcy–Brinkman–Forchheimer model for porous media with mixed boundary conditions, 2021, 381, 03770427, 113008, 10.1016/j.cam.2020.113008 | |
2. | Andrea Manzoni, Alfio Quarteroni, Sandro Salsa, 2021, Chapter 10, 978-3-030-77225-3, 325, 10.1007/978-3-030-77226-0_10 | |
3. | Carolina Introini, Stefano Riva, Stefano Lorenzi, Simone Cavalleri, Antonio Cammi, Non-intrusive system state reconstruction from indirect measurements: A novel approach based on Hybrid Data Assimilation methods, 2023, 182, 03064549, 109538, 10.1016/j.anucene.2022.109538 | |
4. | Pedro Nogueira, Ana L. Silvestre, Jorge Tiago, Steady Navier–Stokes Equations with Regularized Directional Do-Nothing Boundary Condition: Optimal Boundary Control for a Velocity Tracking Problem, 2025, 91, 0095-4616, 10.1007/s00245-024-10216-4 | |
5. | Pedro Nogueira, Ana L. Silvestre, Regularized directional do-nothing boundary conditions for the Navier-Stokes equations: Analytical and numerical study, 2025, 499, 00963003, 129398, 10.1016/j.amc.2025.129398 |