scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Analysis of continuous h −1 least-squares methods for the steady navier-stokes system

02 Mar 2021-Applied Mathematics and Optimization (Springer US)-Vol. 83, Iss: 1, pp 461-488
TL;DR: Two methods for the steady Navier–Stokes system of incompressible viscous fluids are analyzed and the convergence of minimizing sequences for the least-squares functional toward solutions is shown.
Abstract: We analyse two H −1 least-squares methods for the steady Navier-Stokes system of incompressible viscous fluids. Precisely, we show the convergence of minimizing sequences for the least-squares functional toward solutions. Numerical experiments support our analysis.

Summary (4 min read)

1. Introduction

  • From a purely analytical perspective, the following is a well-known existence theorem.
  • The minimization of this functional leads to a so-called continuous H−1-least-squares type method, following the terminology of [5] and later use in [3].
  • Least-squares methods to solve non linear boundary value problems have been the subject of intensive developments in the last decades, as they present several advantages, notably on computational and stability viewpoints.
  • The main reason of this work is to show that, under the assumption of Theorem 1.1, minimizing sequences for this so-called Date: 23-04-2018.
  • Laboratoire de Mathématiques, Université Clermont Auvergne, UMR CNRS 6620, Campus des Cézeaux, 63177 Aubière, France.

2 JÉROME LEMOINE, ARNAUD MÜNCH, AND PABLO PEDREGAL

  • Error functional do actually converge strongly to the solution of (1.1).
  • The authors first consider in Section 2 the case where minimizing sequences live in V × L20(Ω) with V defined below by (2.1), a set of divergence free fields.
  • Then, in Section 3, the authors discuss the general case where the field y is not a priori assumed to be divergence free.
  • In the two cases, the authors provide a sufficient condition of the convergence of the values of the error functional E in terms of the convergence of the values of its derivative.
  • Section 5 describes the conjugate gradient algorithm associated to the error functional E while section 6 discusses numerically the celebrated exemple of the 2D channel with a backward facing step.

2. Steady case under the div-free constraint

  • As indicated in the Introduction, in order to solve the boundary value problem (1.1), the authors use a least-squares type approach.
  • This computation is performed below in the proof of Proposition 2.4.
  • In that direction, their main theorem is the following.
  • H−1-LEAST-SQUARES METHODS FOR NAVIER-STOKES 3 Theorem 2.1.
  • This proposition very clearly establishes that as the authors take down the error E to zero, they get closer, in the strong norm, to the solution of the problem, and so, it justifies why a promising strategy to find good approximations of the solution of problem (1.1) is to look for global minimizers of (2.4).

4 JÉROME LEMOINE, ARNAUD MÜNCH, AND PABLO PEDREGAL

  • By the second part of Lemma 2.3, the last term above vanishes, while for the third term, the first part of the same lemma leads to∫.
  • A practical way of taking a functional to its minimum is through some use of descent directions, i.e. the use of its derivative.
  • The authors computations here follow closely those in [15].
  • Before proceeding to their second step for a full proof of Theorem 2.1, the authors stress that the error functional E(y) is coercive in the sense E(y)→∞ if ‖y‖H10(Ω) →∞. Indeed, from (2.3), and using y itself as a test function, it is elementary to arrive, using∫.
  • The relevant issue is to check coercivity.

6 JÉROME LEMOINE, ARNAUD MÜNCH, AND PABLO PEDREGAL

  • According to their brief discussion before the statement of the proposition the authors are proving, vector fields y remain in a bounded set of V.
  • The authors need to check that then the corresponding solutions given by Lemma 2.5 remain in a bounded set as well.

3. Steady case without the div-free constraint

  • In practice, implementing the div-free constraint, as done in [3], is possibly expansive as it requires at each iteration several resolutions of the steady Stokes equation.
  • The authors would like to explore to what extent a similar approach can be implemented that allows for fields without the div-free constraint.
  • This new framework forces us to take into account the pressure field.
  • Before getting into the proof of similar results as in the preceding section, it is instructive to spend some time with the following interesting discussion.
  • Focus next on the operator taking F in (3.4) into the corresponding minimizer π0.

8 JÉROME LEMOINE, ARNAUD MÜNCH, AND PABLO PEDREGAL

  • The authors are now ready to show a main result as in the previous section in this more general context.
  • The authors strategy proceeds again in two steps.
  • The theorem is, however, exactly the same.
  • The first step of the proof involves an upper bound of the difference y − y0 in terms of the quantity E(y, π).
  • METHODS FOR NAVIER-STOKES 9 Recall, however, that y0 is divergence-free.

10 JÉROME LEMOINE, ARNAUD MÜNCH, AND PABLO PEDREGAL

  • It is elementary then to have the statement in the proposition from this inequality.
  • Concerning the second step, there are just minor changes in the proof.
  • As before, the authors plan to use several appropriate choices for the direction (Y,Π).
  • To this end, by definition, the corrector v solves the variational formulation∫ METHODS FOR NAVIER-STOKES 11 for some Π ∈ L2(Ω) and for every w ∈ H10(Ω), provided fields y are taken from the same ball in the statement of that lemma.
  • This together with (3.13) finishes the proof.

4. Minimizing sequence

  • In practice, however, one would typically use a gradient method to calculate iteratively such sequences.
  • Given that the exact solution y0 of the problem corresponds to an absolute minimum of the smooth functional E, for a certain small positive constant c2, one can ensure that ‖y0 − y0‖H10(Ω) ≤ c2 implies that the sequences computed through a gradient procedure starting from y0 will converge to y0.
  • It would be interesting to have more explicit information about the size of the constant c2 that, eventually, could be of some help to decide in practice how to select the initial guess.
  • The authors strategy is to show that the quantity E′(y) · (y0 − y) becomes non-positive, if y is sufficiently close, in a precise quantitative way, to the exact solution y0.

12 JÉROME LEMOINE, ARNAUD MÜNCH, AND PABLO PEDREGAL

  • For some constant C provided ν−2‖f‖H−1(Ω) is small enough.
  • It remains, hence, to quantify the continuity of E at the solution y0.
  • It is a matter of keeping track of the constants in all those inequalities used above in the proof to have an expression of the constant C4 guaranteeing the claimed convergence.
  • There are four quantities involved: viscosity ν, size of the source term ‖f‖ = ‖f‖H−1(Ω), Poincaré’s constant C for Ω, and the constant C(n) of the Sobolev compact embedding of H1(Ω) into L4(Ω).
  • The non-divergence free situations is a bit more involved though a parallel proof would proceed along the same lines as in the case the authors have just explored.

5. Conjugate gradient algorithm

  • The introduction of the ε term allows to fix the constant of the pressure π.
  • The appropriate tool to produce minimizing sequence for the functional Eε is gradient method.
  • Among them, the Polak-Ribière version of the conjugate gradient (CG for short in the sequel) algorithm (see [8]) have shown its efficiency in the similar context analyzed in [13, 14, 12].

14 JÉROME LEMOINE, ARNAUD MÜNCH, AND PABLO PEDREGAL

  • The CG algorithm associated to the minimization over V of the functional E, defined in section 2, is very similar: the Poisson problems are simply replaced by Stokes problems (the authors refer to [3]).
  • In both cases, the matrix (to be invert) associated to those four problems is the same and does not change from an iteration to the next one.

6. Numerical illustration: Two dimensional channel with a backward facing step

  • The authors consider the celebrated test problem of a two-dimensional channel with a backward facing step, described for instance in Section 45 of [6] (see also [10]).
  • The authors use exactly the geometry and boundary conditions from this reference.
  • The authors now comment some computations performed with the FreeFem++ package developed at the University Paris 6 (see [9]).
  • Similarly, Table 3 reports the norms obtained when minimizing the functional E over A (section 3).
  • In both cases, the Polak-Ribiere version of the conjugate gradient algorithm, initialized with the solution of corresponding Stokes problem.

16 JÉROME LEMOINE, ARNAUD MÜNCH, AND PABLO PEDREGAL

  • Finite element approximation P1/P1 (which do not satisfies the Ladyzenskaia-Babushka-Brezzi condition) provides similar results in term of accuracy and convergence.
  • Figure 4 depicts the evolution of the norm of the gradient with respect to the iterates.
  • The values of the cost √ E(yk, πk) in both cases are however similar, which suggests that the functional E is flat near local minima.
  • The last line of Table 2 also displays the results of the BB algorithm for the minimization of E over V leading to similar results than CG method in term of speed of convergence.
  • The corresponding mesh is composed of 19714 triangles and 10208 vertices.

18 JÉROME LEMOINE, ARNAUD MÜNCH, AND PABLO PEDREGAL

  • Method leads to similar values but does not allow a reduction of the computational cost.
  • The BB algorithm (from Stokes to ν = 1/700) converges after 510 iterates and leads to similar values: again its allows a reduction of the computation costs (2× 510 resolution of Stokes problems for BB whereas CG requires 4×357 resolution of Stokes problems).
  • Eventually, as expected from their observations for ν = 1/50, the minimization of the functional E of Section 3 defined over A using CG and BB algorithms requires more iterates (1020 and 439 respectively) leading to a larger computational cost but similar numerical values.
  • For ν = 1/700, the minimization of E using P1/P1 finite element approximation remains stable, contrary to the previous cases.

7. Conclusions and perspectives

  • The authors have analyzed two H−1-least-squares methods and shown that they allow the construction of strong convergent sequences toward the solution (assumed unique) of the steady Navier-Stokes system.
  • This study justifies in particular the least-squares approach introduced without proof in [3], which assume that the sequences are divergence free.
  • Numerical experiments on a 2D channel with Poiseuille flow confirms their analysis and highlights the robustness of such methods with respect to the initial guess and also with respect to the approximation.
  • The second least-squares functional coupled with a conjugate gradient method requires however much more iterates to achieve a satisfactory approximation.
  • A natural extension of this study is the unsteady case : using ideas from [15], the authors may, at least in the divergence free situation of Section 2, obtain a result similar to Theorem 2.1 in the dynamic situation, and then, adapting [13, 12], examine the corresponding controllability issue.

Did you find this useful? Give us your feedback

Content maybe subject to copyright    Report

HAL Id: hal-01774607
https://hal.archives-ouvertes.fr/hal-01774607v2
Submitted on 9 May 2018
HAL is a multi-disciplinary open access
archive for the deposit and dissemination of sci-
entic research documents, whether they are pub-
lished or not. The documents may come from
teaching and research institutions in France or
abroad, or from public or private research centers.
L’archive ouverte pluridisciplinaire HAL, est
destinée au dépôt et à la diusion de documents
scientiques de niveau recherche, publiés ou non,
émanant des établissements d’enseignement et de
recherche français ou étrangers, des laboratoires
publics ou privés.
ANALYSIS OF CONTINUOUS H –1
LEAST-SQUARES METHODS FOR THE STEADY
NAVIER-STOKES SYSTEM
Jerôme Lemoine, Arnaud Munch, Pablo Pedregal
To cite this version:
Jerôme Lemoine, Arnaud Munch, Pablo Pedregal. ANALYSIS OF CONTINUOUS H –1 LEAST-
SQUARES METHODS FOR THE STEADY NAVIER-STOKES SYSTEM. Applied Mathematics
and Optimization, Springer Verlag (Germany), In press, 83 (1), pp.461-488. �hal-01774607v2�

ANALYSIS OF CONTINUOUS H
1
-LEAST-SQUARES METHODS FOR THE
STEADY NAVIER-STOKES SYSTEM
J
´
EROME LEMOINE, ARNAUD M
¨
UNCH, AND PABLO PEDREGAL
Abstract. We analyze two H
1
-least-squares methods for the steady Navier-Stokes system
of incompressible viscous fluids. Precisely, we show the convergence of minimizing sequences
for the least-squares functional toward solutions. Numerical experiments support our analysis.
Key Words. Steady Navier-Stokes system, Least-squares approach, Gradient method.
1. Introduction
Let R
N
, N = 2 or N = 3 be a bounded connected open set whose boundary is
Lipschitz. We denote by n = n(x) the outward unit normal to at any point x Ω. Bold
letters and symbols denote vector-valued functions and spaces; for instance L
2
(Ω) is the Hilbert
space of the functions v = (v
1
, . . . , v
N
) with v
i
L
2
(Ω) for all i.
This work is concerned with the (numerical) approximation for the steady Navier-Stokes
system
(1.1)
(
νy + (y · )y + π = f , · y = 0 in ,
y = 0 on ,
which describes a viscous incompressible fluid flow in the bounded domain Ω, submitted to the
external force f. Our strategy is to use a least-squares approach, much in the spirit of [2], [3],
[7], but in a systematic way as in [15], having in mind some applications to control problem as
described in [12, 13] for the Stokes system. From a purely analytical perspective, the following
is a well-known existence theorem.
Theorem 1.1 ([17]). For any f H
1
(Ω), there exists at least (y, π) H
1
0
(Ω) ×L
2
0
(Ω) solution
of (1.1). Moreover, if ν
2
kfk
H
1
(Ω)
is small enough, then the couple (y, π) is unique.
We put L
2
0
(Ω) for the space of functions in L
2
(Ω) with zero mean. Assuming f H
1
(Ω),
the solution of (1.1) may be investigated by considering the following functional
(y, π) k νy + (y ·)y + π fk
2
H
1
(Ω)
+ k∇ · yk
2
L
2
(Ω)
over the space H
1
0
(Ω)×L
2
0
(Ω). The minimization of this functional leads to a so-called continuous
H
1
-least-squares type method, following the terminology of [5] and later use in [3].
Least-squares methods to solve non linear boundary value problems have been the subject
of intensive developments in the last decades, as they present several advantages, notably on
computational and stability viewpoints. We refer to the book [2]. The main reason of this work
is to show that, under the assumption of Theorem 1.1, minimizing sequences for this so-called
Date: 23-04-2018.
Laboratoire de Math´ematiques, Universit´e Clermont Auvergne, UMR CNRS 6620, Campus des C´ezeaux,
63177 Aubi`ere, France. e-mail: jerome.lemoine@uca.fr.
Laboratoire de Math´ematiques, Universit´e Clermont Auvergne, UMR CNRS 6620, Campus des ezeaux,
63177 Aubi`ere, France. e-mail: arnaud.munch@uca.fr.
INEI. Universidad de Castilla La Mancha. Campus de Ciudad Real (Spain). e-mail: pablo.pedregal@uclm.es.
Research supported by MTM2017-83740-P, by PEII-2014-010-P of the Conserjer´ıa de Cultura (JCCM), and by
grant GI20152919 of UCLM.
1

2 J
´
EROME LEMOINE, ARNAUD M
¨
UNCH, AND PABLO PEDREGAL
error functional do actually converge strongly to the solution of (1.1). We first consider in Section
2 the case where minimizing sequences live in V × L
2
0
(Ω) with V defined below by (2.1), a set
of divergence free fields. Then, in Section 3, we discuss the general case where the field y is
not a priori assumed to be divergence free. In the two cases, we provide a sufficient condition
of the convergence of the values of the error functional E in terms of the convergence of the
values of its derivative. Then, in Section 4, we show that gradient methods, under hypothesis of
Theorem 1.1, produce converging sequences to the unique solution of (1.1). Section 5 describes
the conjugate gradient algorithm associated to the error functional E while section 6 discusses
numerically the celebrated exemple of the 2D channel with a backward facing step.
2. Steady case under the div-free constraint
As indicated in the Introduction, in order to solve the boundary value problem (1.1), we
use a least-squares type approach. If we insist in keeping explicitly the div-free constraint for
fields, then the pressure field does not play a specific role, so that we can eliminate it from the
formulation.
We consider the Hilbert space
(2.1) V := H
1
0,div
(Ω) = {y H
1
0
(Ω) : · y = 0 in },
endowed with the norm of the gradient k∇yk
L
2
(Ω)
and define the functional E : V R
+
by
putting
(2.2) E(y) =
1
2
Z
|∇v|
2
dx
where the corrector v is the unique minimizer in V for the variational problem
Minimize in v V :
Z
1
2
|∇v|
2
+ (νy y y) · v
dx hf, vi
H
1
(Ω),H
1
0
(Ω)
.
Notice that the differential constrained imposed on V, leads to the existence of a multiplier π,
the pressure, such that
(2.3)
(
v + π + (νy + div(y y) f ) = 0, · v = 0 in ,
v = 0 on ,
with div(y y) = (y.)y + y · y = (y.)y since y is here divergence free.
The existence of a unique solution for this quadratic, constrained problem in v is standard.
Optimality conditions lead directly to the weak form of (2.3).
The functional E is a so-called error functional which measures, through the corrector variable
v, the deviation of the pair (y, π) from being a solution of the underlying system (1.1). We then
consider the following extremal problem
(2.4) inf
yV
E(y).
Note that the error functional E is differentiable as a functional defined on the Hilbert space V,
because the operator y 7→ v taking each y V into its associated corrector v, as stated above,
is a differentiable operation. Indeed, E
0
(y) can always be identified with an element of V itself.
This computation is performed below in the proof of Proposition 2.4.
From Theorem 1.1, the infimum is equal to zero, and is reached by a solution of (1.1). Beyond
this statement, we would like to argue why we believe it is a good idea to use a (minimization)
least-squares approach to approximate the solution of (1.1) by solving (2.4).
The least-squares problem (2.2)-(2.3)-(2.4) has been actually introduced in [3], section 4.2 (and
numerically discussed in [4]), in order to solve one step in time of an implicit Euler scheme, time
approximation for the unsteady Navier-Stokes system. However, the analysis of the convergence
of the method was not given there. In that direction, our main theorem is the following.

H
1
-LEAST-SQUARES METHODS FOR NAVIER-STOKES 3
Theorem 2.1. There is a positive constant C
2
, such that if {y
j
}
j>0
is a sequence in
B := {y V : ky k
H
1
0
(Ω)
C
2
}
with E
0
(y
j
) 0 as j , then the whole sequence y
j
converges strongly as j in H
1
0
(Ω)
to the unique solution y
0
of (1.1) guaranteed by Theorem 1.1, if ν
2
kfk
H
1
(Ω)
is small enough.
We divide the proof in two main steps.
(1) First, we use a typical a priori bound to show that leading the error functional E down
to zero implies strong convergence to the unique solution of (1.1).
(2) Next, we will show that taking the derivative E
0
to zero actually suffices to take E to
zero.
Proposition 2.2. Assume that ν
2
kfk
H
1
(Ω)
is small enough and let y
0
V be the unique
solution of (1.1), mentioned in Theorem 1.1. If dimension n 4, there is a positive constant
C C(ν, n, kf k
H
1
(Ω)
) such that for every y V, we have
ky y
0
k
2
H
1
0
(Ω)
CE(y).
This proposition very clearly establishes that as we take down the error E to zero, we get closer,
in the strong norm, to the solution of the problem, and so, it justifies why a promising strategy
to find good approximations of the solution of problem (1.1) is to look for global minimizers of
(2.4).
The proof of this proposition basically amounts to a typical a priori estimate which is essen-
tially the same that the proof of uniqueness in page 112 in [17]. Recall the following basic fact
which will be utilized several times in what follows.
Lemma 2.3. For all u, v V, we have
Z
(v v : u + u v : v) dx =
Z
u v : u dx = 0.
Proof. (of Proposition 2.2) Bear in mind (2.3)
v + (νy + div(y y) f ) + π = 0, in ,
in addition to the boundary condition v = 0 on Ω. Since (y
0
, π
0
) is the unique solution of
problem (1.1), we can write
v + (νy + div(y y)) (νy
0
+ div(y
0
y
0
)) + (π π
0
) = 0 in .
If we put Y y
0
y, reorganizing terms it is also true that
(2.5) v νY + (π
0
π) div(y y y
0
y
0
) = 0 in .
In a linear situation, we would immediately achieve the bound in the statement. But the pres-
ence of the non-linear (quadratic) term that is so essential to Navier-Stokes requires a bit more
analysis.
Let us focus on the difference
Div div(y y y
0
y
0
).
We first put
Div = div(y
0
Y) + div(Y y).
If we take this identity back to (2.5), multiply by Y, and integrate by parts, taking into account
boundary conditions and bearing in mind that divY = 0, we are led to
Z
v · Y dx + ν
Z
|∇Y|
2
dx
Z
y
0
Y : Y dx
Z
Y y : Y dx = 0.

4 J
´
EROME LEMOINE, ARNAUD M
¨
UNCH, AND PABLO PEDREGAL
By the second part of Lemma 2.3, the last term above vanishes, while for the third term, the
first part of the same lemma leads to
Z
y
0
Y : Y dx =
Z
Y Y : y
0
dx C
2
(n)kYk
2
H
1
0
(Ω)
ky
0
k
H
1
0
(Ω)
,
where C(n) is the constant of the Sobolev embedding of H
1
(Ω) into L
4
(Ω) provided n 4 (see
Lemmas 1.1 and 1.2 in pages 108 and 109, respectively, of [17]). By putting together all of this
information, we have
ν
Z
|∇Y|
2
dx =
Z
y
0
Y : Y dx +
Z
v · Y dx,
and then
(2.6) νkYk
2
H
1
0
(Ω)
C
2
(n)kYk
2
H
1
0
(Ω)
ky
0
k
H
1
0
(Ω)
+ kvk
H
1
0
(Ω)
kYk
H
1
0
(Ω)
.
But y
0
is precisely the solution of problem (1.1), and using it as a test function in its own system,
it is again immediate to check that
(2.7) ky
0
k
H
1
0
(Ω)
C
ν
kfk
H
1
(Ω)
for some C = C(Ω) > 0. Altogether, we find that
ν
1 C C
2
(n) ν
2
kfk
H
1
(Ω)
kYk
H
1
0
(Ω)
kvk
H
1
0
(Ω)
,
and under our hypotheses and the fact that 2E(y) = kvk
H
1
0
(Ω)
, this is the statement in the
proposition.
A practical way of taking a functional to its minimum is through some (clever) use of descent
directions, i.e. the use of its derivative. In doing so, the presence of local minima is always
something that may dramatically spoil the whole scheme. The unique structural property that
discards this possibility is the strict convexity of the functional. However, for non-linear equations
like (1.1), one cannot expect this property to hold for the functional E in (2.2). Nevertheless, we
insist in that for a descent strategy applied to our extremal problem (2.4), numerical procedures
cannot converge except to a global minimizer leading E down to zero. In doing so, thanks to
Proposition 3.4, we are establishing the strong convergence of approximations to the unique
solution of (1.1).
Indeed, we would like to show that the only critical points for E correspond to solutions of
(1.1). In such a case, the search for an element y solution of (1.1) is reduced to the minimization
of E, as indicated in the preceding paragraph. Precisely, we would like to prove the following
proposition, in the spirit of [16]. Our computations here follow closely those in [15].
Before proceeding to our second step for a full proof of Theorem 2.1, we stress that the error
functional E(y) is coercive in the sense
E(y) if kyk
H
1
0
(Ω)
.
Indeed, from (2.3), and using y itself as a test function, it is elementary to arrive, using
Z
y y : y dx = 0
according to Lemma 2.3, that
(2.8) νkyk
H
1
0
(Ω)
C(n, Ω)
kfk
H
1
(Ω)
+ kvk
H
1
(Ω)
for some constant C(n, Ω) > 0, leading to the coercivity of E.

Citations
More filters
Journal ArticleDOI
TL;DR: This work analyzes a least-squares method in order to solve implicit time schemes associated to the 2D and 3D Navier-Stokes system, introduced in 1979 by Bristeau, Glowinksi, Periaux, Perrier and Pironneau.
Abstract: This work analyzes a least-squares method in order to solve implicit time schemes associated to the 2D and 3D Navier–Stokes system, introduced in 1979 by Bristeau, Glowinksi, Periaux, Perrier and Pironneau. Implicit time schemes reduce the numerical resolution of the Navier–Stokes system to multiple resolutions of steady Navier–Stokes equations. We first construct a minimizing sequence (by a gradient type method) for the least-squares functional which converges strongly and quadratically toward a solution of a steady Navier–Stokes equation from any initial guess. The method turns out to be related to the globally convergent damped Newton approach applied to the Navier–Stokes operator. Then, we apply iteratively the analysis on the fully implicit Euler scheme and show the convergence of the method uniformly with respect to the time discretization. Numerical experiments for 2D examples support our analysis.

11 citations

Journal ArticleDOI
TL;DR: In this article, a space-time least-squares method associated with the unsteady Navier-Stokes system is introduced and analyzed, and the convergence is quadratic.
Abstract: We introduce and analyze a space-time least-squares method associated with the unsteady Navier–Stokes system. Weak solution in the two dimensional case and regular solution in the three dimensional case are considered. From any initial guess, we construct a minimizing sequence for the least-squares functional which converges strongly to a solution of the Navier–Stokes system. After a finite number of iterations related to the value of the viscosity coefficient, the convergence is quadratic. Numerical experiments within the two dimensional case support our analysis. This globally convergent least-squares approach is related to the damped Newton method.

10 citations

Posted Content
TL;DR: In this paper, a constructive proof and algorithm for the controllability of semilinear 1D wave equations with Dirichlet boundary conditions is presented. But the proof is based on the Leray-Schauder fixed point theorem, which is not constructive.
Abstract: It has been proved by Zuazua in the nineties that the internally controlled semilinear 1D wave equation ∂tty − ∂xxy + g(y) = f 1ω, with Dirichlet boundary conditions, is exactly controllable in H 1 0 (0, 1) ∩ L 2 (0, 1) with controls f ∈ L 2 ((0, 1) × (0, T)), for any T > 0 and any nonempty open subset ω of (0, 1), assuming that g ∈ C 1 (R) does not grow faster than β|x| ln 2 |x| at infinity for some β > 0 small enough. The proof, based on the Leray-Schauder fixed point theorem, is however not constructive. In this article, we design a constructive proof and algorithm for the exact controllability of semilinear 1D wave equations. Assuming that g does not grow faster than β ln 2 |x| at infinity for some β > 0 small enough and that g is uniformly Holder continuous on R with exponent s ∈ [0, 1], we design a least-squares algorithm yielding an explicit sequence converging to a controlled solution for the semilinear equation, at least with order 1 + s after a finite number of iterations.

8 citations

Journal ArticleDOI
TL;DR: A constructive proof yielding an explicit sequence converging strongly to a controlled solution for the semilinear equation, at least with order 1 + p after a finite number of iterations is designed.
Abstract: The null distributed controllability of the semilinear heat equation $\partial_t y-\Delta y + g(y)=f \,1_{\omega}$ assuming that $g\in C^1(\mathbb{R})$ satisfies the growth condition $\limsup_{\vert r\vert\to \infty} g(r)/(\vert r\vert \ln^{3/2}\vert r\vert)=0$ has been obtained by Fern\'andez-Cara and Zuazua in 2000. The proof based on a non constructive fixed point theorem makes use of precise estimates of the observability constant for a linearized heat equation. Assuming that $g^\prime$ is bounded and uniformly H\"older continuous on $\mathbb{R}$ with exponent $p\in (0,1]$, we design a constructive proof yielding an explicit sequence converging strongly to a controlled solution for the semilinear equation, at least with order $1+p$ after a finite number of iterations. The method is based on a least-squares approach and coincides with a globally convergent damped Newton methods: it guarantees the convergence whatever be the initial element of the sequence. Numerical experiments in the one dimensional setting illustrate our analysis.

7 citations

Posted Content
TL;DR: A minimizing sequence for the least-squares functional is constructed which converges strongly to a solution of the Navier-Stokes system and the convergence is quadratic.
Abstract: We introduce and analyze a space-time least-squares method associated to the unsteady Navier-Stokes system. Weak solution in the two dimensional case and regular solution in the three dimensional case are considered. From any initial guess, we construct a minimizing sequence for the least-squares functional which converges strongly to a solution of the Navier-Stokes system. After a finite number of iterates related to the value of the viscosity constant, the convergence is quadratic. Numerical experiments within the two dimensional case support our analysis. This globally convergent least-squares approach is related to the damped Newton method when used to solve the Navier-Stokes system through a variational formulation.

5 citations

References
More filters
Journal ArticleDOI
TL;DR: First the freefem++ software deals with mesh adaptation for problems in two and three dimension, second, it solves numerically a problem with phase change and natural convection, and finally to show the possibilities for HPC the software solves a Laplace equation by a Schwarz domain decomposition problem on parallel computer.
Abstract: This is a short presentation of the freefem++ software. In Section 1, we recall most of the characteristics of the software, In Section 2, we recall how to to build the weak form of a partial differential equation (PDE) from the strong form. In the 3 last sections, we present different examples and tools to illustrated the power of the software. First we deal with mesh adaptation for problems in two and three dimension, second, we solve numerically a problem with phase change and natural convection, and the finally to show the possibilities for HPC we solve a Laplace equation by a Schwarz domain decomposition problem on parallel computer.

2,867 citations


"Analysis of continuous h −1 least-s..." refers methods in this paper

  • ...We now comment some computations performed with the FreeFem++ package developed at the University Paris 6 (see [9])....

    [...]

Journal ArticleDOI
TL;DR: Etude de nouvelles methodes de descente suivant le gradient for the solution approchee du probleme de minimisation sans contrainte. as mentioned in this paper.
Abstract: Etude de nouvelles methodes de descente suivant le gradient pour la solution approchee du probleme de minimisation sans contrainte. Analyse de la convergence

2,585 citations

01 Jan 2005
TL;DR: In this article, the development of dierent versions of nonlinear conjugate gradient methods, with special attention given to global convergence properties, is reviewed, with a focus on the convergence properties of the dierent methods.
Abstract: This paper reviews the development of dierent versions of nonlinear conjugate gradient methods, with special attention given to global convergence properties.

775 citations


"Analysis of continuous h −1 least-s..." refers methods in this paper

  • ...Among the various CG versions reported in [8], we observe that the Polak-Ribière leads to the best results in term of speed of convergence....

    [...]

  • ...Among them, the Polak-Ribière version of the conjugate gradient (CG for short in the sequel) algorithm (see [8]) have shown its efficiency in the similar context analyzed in [13, 14, 12]....

    [...]

Journal ArticleDOI
TL;DR: Results indicate that the global Barzilai and Borwein method may allow some significant reduction in the number of line searches and also in theNumber of gradient evaluations.
Abstract: The Barzilai and Borwein gradient method for the solution of large scale unconstrained minimization problems is considered. This method requires few storage locations and very inexpensive computations. Furthermore, it does not guarantee descent in the objective function and no line search is required. Recently, the global convergence for the convex quadratic case has been established. However, for the nonquadratic case, the method needs to be incorporated in a globalization scheme. In this work, a nonmonotone line search strategy that guarantees global convergence is combined with the Barzilai and Borwein method. This strategy is based on the nonmonotone line search technique proposed by Grippo, Lampariello, and Lucidi [SIAM J. Numer. Anal., 23 (1986), pp. 707--716]. Numerical results to compare the behavior of this method with recent implementations of the conjugate gradient method are presented. These results indicate that the global Barzilai and Borwein method may allow some significant reduction in the number of line searches and also in the number of gradient evaluations.

731 citations

Book ChapterDOI

455 citations


Additional excerpts

  • ...Numerical illustration: Two dimensional channel with a backward facing step We consider the celebrated test problem of a two-dimensional channel with a backward facing step, described for instance in Section 45 of [6] (see also [10])....

    [...]

Frequently Asked Questions (13)
Q1. What are the contributions in "Analysis of continuous h –1 least-squares methods for the steady navier-stokes system" ?

The authors analyze two H−1-least-squares methods for the steady Navier-Stokes system of incompressible viscous fluids. Precisely, the authors show the convergence of minimizing sequences for the least-squares functional toward solutions. 

Bold letters and symbols denote vector-valued functions and spaces; for instance L2(Ω) is the Hilbert space of the functions v = (v1, . . . , vN ) with vi ∈ L2(Ω) for all i. 

Their strategy is to show that the quantityE′(y) · (y0 − y)becomes non-positive, if y is sufficiently close, in a precise quantitative way, to the exact solution y0. 

The main interest of the BB algorithm is on the computational viewpoint as it requires only two resolutions of Poisson problem, namely (5.4), (5.5) per iterate. 

The more recent Barzilai-Borwein algorithm allows to reduce significantly the number of iterations together with the computational cost. 

Least-squares methods to solve non linear boundary value problems have been the subject of intensive developments in the last decades, as they present several advantages, notably on computational and stability viewpoints. 

A practical way of taking a functional to its minimum is through some (clever) use of descent directions, i.e. the use of its derivative. 

(Ω) ≤ 10−3 (gk denotes the residual at iterates k) is achieved in 39 iterates and leads to results very close to those from the resolution of (6.1), see table 2. 

Their strategy is to use a least-squares approach, much in the spirit of [2], [3], [7], but in a systematic way as in [15], having in mind some applications to control problem as described in [12, 13] for the Stokes system. 

19The authors have analyzed two H−1-least-squares methods and shown that they allow the construction of strong convergent sequences toward the solution (assumed unique) of the steady Navier-Stokes system. 

In such a case, the search for an element y solution of (1.1) is reduced to the minimization of E, as indicated in the preceding paragraph. 

For the functional Eε, the CG algorithm reads as follows :• Step 0: Initialization - Given any η > 0 and any z0 = (y0, π0) ∈ A, compute the residual g0 = (y0, π0) ∈ A solution of(5.1) (g0, (Y,Π))A = E ′ ε(y 0, π0) · (Y,Π), ∀(Y,Π) ∈ A. 

Figure 3 depicts the evolution (in log scale)of the norm ‖gk‖H1(Ω) of the gradient and √ E(yk) = |vk|H10(Ω) with respect to the iterates: theconvergence to zero is sur-linear as the authors observe √ E(yk) = O(e−0.15k).