• Aucun résultat trouvé

Finite element approximation of finitely extensible nonlinear elastic dumbbell models for dilute polymers

N/A
N/A
Protected

Academic year: 2022

Partager "Finite element approximation of finitely extensible nonlinear elastic dumbbell models for dilute polymers"

Copied!
30
0
0

Texte intégral

(1)

DOI:10.1051/m2an/2011062 www.esaim-m2an.org

FINITE ELEMENT APPROXIMATION OF FINITELY EXTENSIBLE NONLINEAR ELASTIC DUMBBELL MODELS FOR DILUTE POLYMERS

John W. Barrett

1

and Endre S¨ uli

2

Abstract. We construct a Galerkin finite element method for the numerical approximation of weak solutions to a general class of coupled FENE-type finitely extensible nonlinear elastic dumbbell models that arise from the kinetic theory of dilute solutions of polymeric liquids with noninteracting polymer chains. The class of models involves the unsteady incompressible Navier–Stokes equations in a bounded domain Ω⊂Rd,d= 2 or 3, for the velocity and the pressure of the fluid, with an elastic extra-stress tensor appearing on the right-hand side in the momentum equation. The extra-stress tensor stems from the random movement of the polymer chains and is defined through the associated probability density function that satisfies a Fokker–Planck type parabolic equation, a crucial feature of which is the presence of a centre-of-mass diffusion term. We require no structural assumptions on the drag term in the Fokker–Planck equation; in particular, the drag term need not be corotational. We perform a rigorous passage to the limit as first the spatial discretization parameter, and then the temporal discretization parameter tend to zero, and show that a (sub)sequence of these finite element approximations converges to a weak solution of this coupled Navier–Stokes–Fokker–Planck system. The passage to the limit is performed under minimal regularity assumptions on the data: a square-integrable and divergence-free initial velocity datumu0for the Navier–Stokes equation and a nonnegative initial probability density function ψ0 for the Fokker–Planck equation, which has finite relative entropy with respect to the Maxwellian M.

Mathematics Subject Classification. 35Q30, 35K65, 65M12, 65M60, 76A05, 82D60.

Received March 19, 2011. Revised August 3, 2011.

Published online February 13, 2012.

1. Introduction

This paper is concerned with the construction and convergence analysis of a finite element method for the numerical approximation of weak solutions to a system of nonlinear partial differential equations that arises from the kinetic theory of dilute polymer solutions. The mathematical model under consideration couples the incompressible Navier–Stokes equations, with the divergence of the elastic extra-stress tensor appearing on its right-hand side, to a high-dimensional Fokker–Planck equation with an unbounded drift term, whose transport coefficients depend on the fluid velocity and whose solution is a probability density function featuring in the

Keywords and phrases.Finite element method, convergence analysis, existence of weak solutions, kinetic polymer models, FENE dumbbell, Navier–Stokes equations, Fokker–Planck equations.

1 Department of Mathematics, Imperial College London, London SW7 2AZ, UK.j.barrett@ imperial.ac.uk

2 Mathematical Institute, University of Oxford, 24–29 St. Giles’, OX1 3LB Oxford, UK.endre.suli@ maths.ox.ac.uk

Article published by EDP Sciences c EDP Sciences, SMAI 2012

(2)

J.W. BARRETT AND E. S ¨ULI

definition of the elastic extra stress tensor. The existence and exponential equilibration of global weak solutions to the system was proved in two substantial recent papers, [7,9], for both finitely extensible nonlinear bead- spring chain type polymer models and Hookean type bead-spring chain models, using a delicate combination of entropy estimates and weak compactness arguments applied to sequences of approximations that were generated by a temporal semidiscretization of the system. Weak solutions to the Navier–Stokes–Fokker–Planck system, whose existence was thus shown, were also shown to satisfy the natural energy inequality originating from the formal balance law for the system. In this paper, we construct a fully-discrete numerical approximation of the model and show that, as the spatial mesh size tends to zero, a subsequence of the sequence of numerical solutions converges to a weak solution of the temporally semidiscrete scheme, a subsequence of whose solutions is, in turn, known from [7] to converge to a weak solution of the coupled Navier–Stokes–Fokker–Planck system as the time step tends to zero. Hence we deduce that a subsequence of the sequence of numerical approximations converges to a weak solution of the problem. To the best of our knowledge, this is the first example of a general family of fully-discrete, convergent numerical methods for the coupled Navier–Stokes–Fokker–Planck system. The key contributions of the paper are the following: (1) the finite element method constructed here reproduces at the discrete level the energy inequality satisfied by weak solutions of the Navier–Stokes–Fokker–Planck system; and (2) convergence to weak solutions is established under minimal regularity assumptions on the data: a square- integrable and divergence-free initial velocity datum for the Navier–Stokes equation and a nonnegative initial probability density function for the Fokker–Planck equation, which has finite relative entropy with respect to the Maxwellian of the model.

We shall now describe the Navier–Stokes–Fokker–Planck system under consideration, arising from the kinetic theory of dilute polymer solutions. The solvent is an incompressible, viscous, isothermal Newtonian fluid confined to a bounded open setΩ⊂Rd,d= 2 or 3, with a Lipschitz-continuous boundary∂Ω. For the sake of simplicity of presentation we shall suppose thatΩhas ‘solid boundary’∂Ω; the velocity fielduwill then satisfy the no-slip boundary conditionu= 0on∂Ω. The polymer chains, which are suspended in the solvent, are assumed not to interact with each other. The conservation of momentum and mass equations for the solvent then have the form of the incompressible Navier–Stokes equations in which the elasticextra-stresstensorτ(i.e., the polymeric part of the Cauchy stress tensor), appears as a source term:

GivenT R>0, findu : (x, t)∈Ω×[0, T]→u(x, t)∈Rdandp : (x, t)∈Ω×(0, T]→p(x, t)∈Rsuch that

∂u

∂t + (u

· ∇

x)u

−ν Δxu

+

xp=f

+

x ·τ

inΩ×(0, T], (1.1a)

x ·u

= 0 inΩ×(0, T], (1.1b)

u= 0

on∂Ω×(0, T], (1.1c)

u(x

,0) =u

0(x

) ∀x

∈Ω. (1.1d)

It is assumed that each of the equations above has been written in its nondimensional form; u denotes a nondimensional velocity, defined as the velocity field scaled by the characteristic flow speedU0;ν∈R>0is the reciprocal of the Reynolds number,i.e.the ratio of the kinematic viscosity coefficient of the solvent andL0U0, whereL0is a characteristic length-scale of the flow;pis the nondimensional pressure andfis the nondimensional density of body forces.

In abead-spring chain model, consisting ofK+ 1 beads coupled withKelastic springs to represent a polymer chain, the extra-stress tensorτis defined by theKramers expression as a weighted average ofψ, the probability density function of the (random) conformation vector q

:= (q

T1, . . . , q

TK)TRKd of the chain (cf.(1.7) below, in the case ofK= 1), withq

i representing thed-component conformation/orientation vector of theith spring.

The Kolmogorov equation satisfied by ψ is a second-order parabolic equation, the Fokker–Planck equation, whose transport coefficients depend on the velocity fieldu. The domain D of admissible conformation vectors D RKd is aK-fold Cartesian product D1× · · · ×DK of balanced convex open setsDi Rd,i = 1, . . . , K;

the term balanced means that q

i Di if, and only if, −q

i Di. Hence, in particular, 0 ∈Di, i = 1, . . . , K.

(3)

Typically Di is the whole of Rd or a bounded open d-dimensional ball centred at the origin 0 Rd for each i= 1, . . . , K. WhenK= 1, the model is referred to as thedumbbell model.

Henceforth we shall confine ourselves to the case of K = 1 – the dumbbell model; then, D =D1, and we shall simply write D instead of D1. Concerning extensions to the case of K≥1, the reader is referred to the concluding remarks in Section5.

LetO ⊂[0,∞) denote the image ofDunder the mappingq

∈D→ 12|q

|2, and consider thespring potentialU∈ C2(O;R≥0). Clearly, 0∈ O. We shall suppose thatU(0) = 0 and thatU is monotonic increasing and unbounded onO. The elastic spring-forceF : D⊆RdRdof the spring is defined by

F(q

) =U(1 2|q

|2)q

. (1.2)

Remark 1.1. In the Hookean dumbbell model the spring force is defined by F(q

) = q

, with q

D = Rd, corresponding toU(s) =s,s∈ O= [0,∞). This model is physically unrealistic as it admits an arbitrarily large extension.

We shall therefore assume in what follows thatD is aboundedopen ball, centred at the origin 0Rd, with boundary∂D. We shall further suppose that there exist constantscj>0,j= 1,2,3,4, andγ >1 such that the (normalized) MaxwellianM, defined by

M(q

) = 1

Ze−U(12|q

|2)

, Z:=

D

e−U(12|q

|2)

dq, (1.3)

and the associated spring potentialU satisfy c1[dist(q

, ∂D)]γ≤M(q

) ≤c2[dist(q

, ∂D)]γ ∀q

∈D, (1.4a)

c3[dist(q

, ∂D)]U(1 2|q

|2)≤c4 ∀q

∈D. (1.4b)

Observe that

M(q

)q[M(q

)]−1=−[M(q

)]−1qM(q

) =qU(1

2|q|2) =U(1 2|q|2)q

. (1.5)

Since [U(12|q

|2)]2 = (logM(q

) + Const.)2, it follows from (1.4a), (1.4b) that (ifγ >1, as has been assumed

here,)

D

1 + [U(1 2|q

|2)]2+ [U(1 2|q

|2)]2

M(q

) dq

<∞. (1.6)

Remark 1.2. In the FENE (finitely extensible nonlinear elastic) dumbbell model the spring force is given by F(q

) = (1− |q

|2/b)−1q

, q

D = B(0, b12), corresponding to U(s) = 2blog 12sb

, s ∈ O = [0,2b).

Here B(0, b12) is a bounded open ball inRd centred at the origin 0 Rd and of fixed radius b12, with b >0.

Direct calculations show that the Maxwellian M and the elastic potential U of the FENE model satisfy the conditions (1.4a), (1.4b) withγ:=2b provided thatb >2. Thus, (1.6) also holds forb >2.

The governing equations of the general FENE-type dumbbell model with centre-of-mass diffusion are (1.1a)–

(1.1d), where the extra-stress tensor τ, depending on the probability density function ψ, is defined by the Kramers expression:

τ(ψ) =k C(ψ)

−k ρ(ψ)I. (1.7)

Here the dimensionless constantk >0 is a constant multiple of the product of the Boltzmann constantkB and the absolute temperatureT,Iis the unitd×dtensor, and

C(ψ)(x, t) =

D

ψ(x, q

, t)q

q

TU 1

2|q

|2

dq, and ρ(ψ)(x, t) =

D

ψ(x, q

, t) dq

, (1.8)

(4)

J.W. BARRETT AND E. S ¨ULI

whereρ(ψ)(x, t) is the density of polymer chains located atx at timet. The probability density functionψis a solution of the Fokker–Planck equation

∂ψ

∂t + (u· ∇x)ψ+q · σ(u)q

ψ

=ε Δxψ+ 1 2λ∇q ·

M∇q

ψ M

inΩ×D×(0, T], (1.9) with σ(v) ≡ ∇xv, where (∇xv)(x, t) Rd×d and {∇xv}ij = ∂x∂vi

j. In (1.9), ε > 0 is the centre-of-mass diffusion coefficient defined asε:= (0/L0)2/(8λ) with0:=

kBT/Hsignifying the characteristic microscopic length-scale and λ:= (ζ/4H)(U0/L0), whereζ >0 is a friction coefficient and H>0 is a spring-constant. The dimensionless parameter λ∈R>0, called the Weissenberg number (and usually denoted byWi), characterizes the elastic relaxation property of the fluid.

We impose the following boundary and initial conditions onψ:

M 1

2λ∇

q

ψ M

−σ

(u

)q

ψ M

· q

|q

| = 0 onΩ×∂D×(0, T], (1.10a) ε∇

xψ · n

= 0 on∂Ω×D×(0, T], (1.10b)

ψ(·,·,0) =ψ0(·,·)0 onΩ×D, (1.10c)

whereq

is normal to∂D, asD is a bounded ball centred at the origin inRd, andnis normal to∂Ω. The initial conditionψ0 is nonnegative, defined onΩ×D, with

Dψ0(x, q

) dq

= 1 for a.e. x ∈Ω, and assumed to have finite relative entropy with respect to the MaxwellianM;i.e.

Ω×Dψ0(x, q

) log(ψ0(x, q

)/M(q

)) dq

dx<∞. The boundary and initial conditions forψhave been chosen so as to ensure that

D

ψ(x, q

, t) dq

=

D

ψ(x, q

,0) dq

= 1 ∀(x, t)∈Ω×(0, T]. (1.11) Remark 1.3. The collection of equations and structural hypotheses ((1.1a)–(1.1d))–((1.10a)–(1.10c)) will be referred to throughout the paper as model (P), or as thegeneral FENE-type dumbbell model with centre-of-mass diffusion.

A noteworthy feature of equation (1.9) in the model (P) compared to classical Fokker–Planck equations for dumbbell models in the literature is the presence of thex-dissipative centre-of-mass diffusion termε Δxψon the right-hand side of the Fokker–Planck equation (1.9). We refer to Barrett and S¨uli [3] for the derivation of (1.9);

see also the article by Schieber [28] concerning generalized dumbbell models with centre-of-mass diffusion, and the recent paper of Degond and Liu [14] for a careful justification of the presence of the centre-of-mass diffusion term through asymptotic analysis. In standard derivations of bead-spring models the centre-of-mass diffusion term is routinely omitted on the grounds that it is several orders of magnitude smaller than the other terms in the equation. Indeed, when the characteristic macroscopic length-scaleL01, (for example,L0= diam(Ω)), Bhave et al.[10] estimate the ratio20/L20 to be in the range of about 10−9to 10−7. However, the omission of the term ε Δxψfrom (1.9) in the case of a heterogeneous solvent velocityu(x, t) is a mathematically counterproductive model reduction. Whenε Δxψis absent, (1.9) becomes a degenerate parabolic equation exhibiting hyperbolic behaviour with respect to (x, t). Since the study of weak solutions to the coupled problem requires one to work with velocity fieldsuthat have very limited Sobolev regularity (typicallyu∈L(0, T;L2(Ω))∩L2(0, T;H10(Ω))), one is then forced into the technically unpleasant framework of hyperbolically degenerate parabolic equations with rough transport coefficients (cf. Ambrosio [1] and DiPerna and Lions [15]). The resulting difficulties are further exacerbated by the fact that a typical spring forceF(q

) for a finitely extensible model (such as FENE) explodes asq

approaches∂D; see Remark1.2above. For these reasons, here we shall retain the centre-of-mass diffusion term in (1.9).

Lions and Masmoudi [24] proved the global existence of weak solutions for the simplifiedcorotational FENE dumbbell model, i.e. with σ(u) = xu replaced by its skew-symmetric part 12(∇xu(∇xu)T), and with

(5)

ε = 0; see also the work of Masmoudi [25]. Under very general assumptions on the finite-dimensional spaces used for the purpose of spatial discretization, including, in particular, classical conforming finite element spaces and spectral Galerkin subspaces, Barrett and S¨uli [5] showed the convergence of a (sub)sequence of numerical approximations to a weak solution of the coupled Navier–Stokes–Fokker–Planck system (P), for a large class of unbounded spring potentials, including the FENE potential, in the case of the simplified corotational model.

Recently, Masmoudi [26] has extended the analysis of Lions and Masmoudi [24] to the noncorotational case. For a fuller literature survey on the mathematical analysis of FENE-type dumbbell models we refer the reader to Barrett and S¨uli [7]. In the rest of this section we concentrate on those references that are relevant to the finite element approximation developed and analyzed in this paper.

In Barrett and S¨uli [4] we showed the existence of global-in-time weak solutions to the general class of noncorotational FENE type dumbbell models (including the standard FENE dumbbell model) with centre-of- mass diffusion and microscopic cut-off (cf.(1.13) and (1.14) below) in the drag term

q ·(u)q

ψ) =∇q ·

σ(u)q

M ψ

M

· (1.12)

We observe that ifψ/M is bounded above then, forL∈R>0sufficiently large, the drag term (1.12) is equal to

q ·

σ(u)q

M βL ψ

M

, (1.13)

whereβL∈C(R) is a cut-off function defined as

βL(s) := min(s, L). (1.14)

It then follows that, forL1, any solutionψof (1.9), such thatψ/M is bounded above, also satisfies

∂ψ

∂t + (u

· ∇

x)ψ+

q ·

σ

(u

)q

M βL ψ

M

=ε Δxψ+ 1 2λ∇

q ·

M∇

q

ψ M

inΩ×D×(0, T], (1.15) and the following boundary and initial conditions:

M 1

2λ∇

q

ψ M

−σ

(u

)q

βL ψ

M

· q

|q

| = 0 onΩ×∂D×(0, T], (1.16a) ε∇

xψ · n

= 0 on∂Ω×D×(0, T], (1.16b)

ψ(·,·,0) =M(·)βL0(·,·)/M(·))0 onΩ×D. (1.16c) Clearly, if there existsL >0 such that 0≤ψ0≤L M, thenM βL0/M) =ψ0. HenceforthL >1 is assumed.

Remark 1.4. The coupled problem (1.1a)–(1.1d), (1.7), (1.8), (1.15), (1.16a)–(1.16c) will be referred to as model (PL), or as thegeneral FENE-type dumbbell model with centre-of-mass diffusion and microscopic cut-off, with cut-off parameterL >1.

In order to highlight the dependence on L, in subsequent sections the solution to (1.15), (1.10a)–(1.10c) will be labelled ψL, and we work with the variable ψL := ψL/M. Due to the coupling of (1.15) to (1.1a)–

(1.1d) through (1.7), the velocity and the pressure will also depend on Land we shall therefore denote them in subsequent sections by uL and pL. As has been already emphasized earlier, the centre-of-mass diffusion coefficientε >0 is a physical parameter and is regarded as being fixed, so we do not highlight its presence in the model through our notation.

(6)

J.W. BARRETT AND E. S ¨ULI

Barrett and S¨uli [8] constructed a Galerkin finite element approximation, and proved (sub)sequence conver- gence, to a weak solution of a system similar to (PL), where ψL in the convective term, in addition to the drag term, is replaced byβL(ψL). Finally, Barrett and S¨uli proved in [7,9] the existence and exponential equili- bration of global-in-time weak solutions to Navier–Stokes–Fokker–Planck systems for general classes of finitely extensible nonlinear elastic and Hookean-type bead-spring chains models.

After this brief literature survey, we now turn our attention to the description of the formal energy bounds on which the constructions of the temporally semidiscrete and fully-discrete approximations are based.

2. Formal energy bounds for (P) and (P

L,δ

)

In this section we identify formally the energy structure for (P), and related regularized models. Before doing so, we note that the notation| · |will be used to signify one of the following. When applied to a real numberx,

|x|will denote the absolute value of the numberx; when applied to a vectorv,|v|will stand for the Euclidean norm of the vectorv; and, when applied to a square matrixA,|A|will signify the Frobenius norm, [tr(ATA)]12, of the matrixA, where, for a square matrixB,tr(B) denotes the trace ofB.

Multiplying (1.1a) byu, integrating overΩ, and noting (1.1b), (1.1c) yields that 1

2 d dt

Ω

|u|2 dx

+ν

Ω

|∇xu

|2dx

Ω

f

·u

dx

=

Ω

τ(Mψ) :

xu

dx

=−k

Ω

C(Mψ) :

xu

dx

, (2.1)

where ψ := ψ/M. Let F(s) := (lns−1)s+ 1 for s > 0, with F(0) := 1. Multiplying the Fokker–Planck equation (1.9) by F(ψ) lnψ, on assuming that ψ > 0, integrating overΩ×D and noting (1.10a), (1.10b) and (1.1b), (1.1c) yields that

d dt

Ω×DMF(ψ) dq

dx

+ 1

2λ

Ω×DM∇

qψ· ∇

q[F(ψ)] dq

dx

+ε

Ω×DM∇

xψ· ∇

x[F(ψ)] dq

dx=

Ω×D[(∇

xu

)q

]· ∇

q[F(ψ)] dq

dx. (2.2) It follows, on noting that F(s) =s−1 >0 for s > 0 and hence that ψ∇q[F(ψ)] = qψ, (1.5), (1.1b) and M = 0 on∂D that

Ω×D

[(

xu

)q

]· ∇

q[F(ψ)] dq

dx=

Ω×D

M[(

xu

)q

]· ∇

qψdq

dx

=

Ω×DM U(1 2|q

|2)q

·[(∇

xu

)q

]ψdq

dx

=

Ω

C(Mψ) :

xu

dx

, (2.3)

on recalling (1.8). Combining (2.1)–(2.3), we obtain the following energy law for (P):

d dt

1 2

Ω

|u|2 dx

+k

Ω×DMF(ψ) dq

dx

+ν

Ω

|∇xu

|2 dx

+k ε

Ω×DMψ|∇

x[F(ψ)] |2dq

dx

+ k 2λ

Ω×DMψ|∇

q[F(ψ)]| 2dq

dx=

Ω

f

·u

dx

. (2.4)

(7)

To make the above rigorous, and for computational purposes, we replace the convex function F ∈C(R≥0) C(R>0) by its convex regularization FδL∈C2,1(R) defined, for anyδ∈(0,1) andL >1, as follows:

FδL(s) :=

⎧⎪

⎪⎩

s2−δ2

2δ + (lnδ−1)s+ 1 s≤δ, F(s)≡(lns−1)s+ 1 δ≤s≤L,

s2−L2

2L + (lnL−1)s+ 1 L≤s.

(2.5)

Hence, we have that [FδL](s) =

⎧⎨

s

δ + lnδ−1 s≤δ,

lns δ≤s≤L,

s

L+ lnL−1 L≤s,

and [FδL](s) =

⎧⎨

δ−1 s≤δ, s−1 δ≤s≤L, L−1 L≤s.

(2.6)

In addition, we introduce

βδL(s) := [[FδL](s)]−1=

⎧⎨

δ s≤δ, s δ≤s≤L, L L≤s.

(2.7)

It follows from (2.7), for any sufficiently smoothϕ, that βδL(ϕ)∇

x([FδL](ϕ) ) =∇

xϕ and βδL(ϕ)∇

q([FδL](ϕ) ) =

qϕ. (2.8)

Let {uL,δL,δ} solve problem (PL,δ), which is a regularization of the problem (P), similar to (PL), where βL(·) in (1.15) and (1.16a) is replaced byβLδ(·). Multiplying the Fokker–Planck equation in (PL,δ) by [FδL](ψL,δ), integrating overΩ×D, noting the boundary conditions and (2.8) yields, similarly to (2.2) and (2.3), that

d dt

Ω×DMFδL(ψL,δ) dq

dx

+ 1

2λ

Ω×DM∇

qψL,δ· ∇

q

[FδL](ψL,δ)

dq

dx

+ε

Ω×DM∇

xψL,δ· ∇

x

[FδL](ψL,δ)

dq

dx=

Ω

C

(MψL,δ) :

xu

L,δdx

. (2.9) Combining (2.9) and the (PL,δ) version of (2.1), we obtain the following energy law for (PL,δ), a regularized analogue of (2.4):

d dt

1 2

Ω

|uL,δ|2dx

+k

Ω×DMFδL(ψL,δ) dq

dx

+ν

Ω

|∇xu

L,δ|2 dx

+k ε

Ω×DM βδL(ψL,δ)|∇

x

[FδL](ψL,δ) |2dq

dx

+ k 2λ

Ω×DM βδL(ψL,δ)|∇

q

[FδL](ψL,δ) |2dq

dx=

Ω

f

·u

L,δ dx

. (2.10)

On noting that [FδL]≥L−1, and

min{FδL(s), s[FδL](s)} ≥ s2

2δ ifs≤0,

s2

4L−C(L) ifs≥0, (2.11)

one deduces from (2.10) that sup

t∈(0,T)

Ω

|u

L,δ|2dx

+ν

ΩT

|∇xu

L,δ|2dx

dt+δ−1 sup

t∈(0,T)

Ω×D

M|[ψL,δ]|2dq

dx

≤C(L). (2.12)

(8)

J.W. BARRETT AND E. S ¨ULI

In addition, one can show that sup

t∈(0,T)

Ω×DM|ψL,δ|2dq

dx

+1

λ T

0

Ω×DM

qψL,δ2dq

dxdt +ε

T

0

Ω×D

M

xψL,δ2 dq

dxdt+ sup

t∈(0,T)

Ω

|C

(MψL,δ)|2dx

≤C(L, T). (2.13) The above formal bounds can be made rigorous and the existence of a global-in-time weak solution{uL,δL,δ} to (PL,δ) can be established, see [4]. Moreover, one can take the limit δ→ 0+ in problem (PL,δ) to establish the existence of a global-in-time weak solution{uLL} to problem (PL) with ψL 0 a.e. in Ω×D×(0, T).

Once again, see [4].

The aim of this paper is to construct a finite element approximation, (PΔt,hL,δ ), of problem (PL,δ), which mimics the energy law (2.10) at a discrete level, and to show that a (sub)sequence of this approximation converges to a weak solution of (PΔtL ), as the spatial discretization parameterh, as well as the regularization parameterδ, tend to zero. Here (PΔtL ) is a time discretization of (PL). Barrett and S¨uli [7] showed that for a specific time discretization (PΔtL ) a (sub)sequence of this approximation converges to a weak solution of (P), as the cut-off parameterLtends to infinity with the time discretization parameterΔt= o(L−1).

The outline of this paper is as follows. In the next section, we introduce the necessary function spaces. In addition, we introduce the particular time discretization, (PΔtL ), of (PL) and state the relevant convergence results from Barrett and S¨uli [7]. In Section 4, we introduce our finite element approximation, (PΔt,hL,δ ), of problem (PL,δ) and show that a (sub)sequence of this approximation converges to a weak solution of (PΔtL ), as the spatial discretization parameterh, as well as the regularization parameterδ, tend to zero. Hence combining this with the convergence result in Section 3, we obtain the desired result that a (sub)sequence of our finite element approximation (PΔt,hL,δ ) converges to a weak solution of (P) as firsth, δ 0+ and thenL→ ∞, with Δt= o(L−1). In Section 5 we discuss possible extensions of our results. The paper closes with an Appendix, containing the proofs of some technical bounds required in the convergence analysis of the initialization of the scheme.

3. The discrete-in-time approximation (P

ΔtL

)

Let

H :={w ∈L2(Ω) :x ·w = 0} and V :={w ∈H 10(Ω) :x ·w = 0}, (3.1) where the divergence operatorx· is to be understood in the sense of distributions onΩ. LetV be the dual ofV. More generally, letVσdenote the closure of the set of all divergence-freeC0 (Ω) functions in the norm of H 1

0(Ω)∩Hσ(Ω),σ≥1, equipped with the Hilbert space norm, denoted by · Vσ, inherited fromH σ(Ω), and letVσsignify the dual space ofVσ, with duality pairing·,·Vσ. AsΩis a bounded Lipschitz domain, we have thatV1=V (cf.Temam [30], Chap. 1, Thm. 1.6). Similarly,·,·H10(Ω)will denote the duality pairing between (H10(Ω)) andH 10(Ω). The norm on (H10(Ω)) will be that induced from takingx · L2(Ω) to be the norm on H 1

0(Ω).

For later purposes, we recall the following well-known Gagliardo–Nirenberg inequality. Letr∈[2,∞) ifd= 2, andr∈[2,6] ifd= 3 andθ=d1

21r

. Then, there is a constantC=C(Ω, r, d), such that, for allη∈H1(Ω):

ηLr(Ω)≤Cη1−θL2(Ω)ηθH1(Ω). (3.2) Let F ∈C(R>0) be defined by F(s) :=s(logs−1) + 1, s > 0. As lims→0+F(s) = 1, the function F can be considered to be defined and continuous on [0,∞), where it is a nonnegative, strictly convex function with F(1) = 0.

(9)

We recall our assumptions on the data:

(A1) ∂Ω ∈C0,1; D=B(0

, rD) withrD>0 andU satisfying (1.4a), (1.4b) withγ >1;

u0∈H

; ψ0:= ψ0

M 0 a.e.onΩ×D with F(ψ0)∈L1M×D) and

D

M(q

)ψ0(x

, q

) dq

= 1 for a.e.x

∈Ω; and f

∈L2(0, T; (H

10(Ω))). (3.3)

Here,LpM×D), forp∈[1,∞), denotes the Maxwellian-weightedLpspace overΩ×D with norm ϕ LpM(Ω×D):=

Ω×DM|ϕ|pdq

dx 1p

.

Similarly, we introduceLpM(D), the Maxwellian-weightedLpspace overD. Letting ϕ HM1(Ω×D):=

Ω×DM

|ϕ|2+|∇xϕ|2+|∇qϕ|2 dqdx

12

, (3.4)

we then set

X≡HM1×D) :=

ϕ∈L1loc×D) :ϕ HM1 (Ω×D)<∞

. (3.5)

It is shown in Appendix C of [6] (the extended version of Barrett and S¨uli [7]) that

C×D) is dense in X. (3.6)

In addition, we note that the embeddings

HM1 (D)→L2M(D), (3.7a)

HM1×D)≡L2(Ω;HM1 (D))∩H1(Ω;L2M(D))→L2M×D)≡L2(Ω;L2M(D)) (3.7b) are compact if γ 1 in (1.4a), (1.4b); see Appendix D in [6]. Throughout we will assume that (3.3) hold, so that (1.6) and (3.7a), (3.7b) hold. We note for future reference that (1.8) and (1.6) yield that, forϕ∈L2M(Ω×D),

Ω

|C(Mϕ)| 2dx

=

Ω

D

Mϕ U q

q

Tdq

2 dx

≤C

Ω×DM|ϕ|2dq

dx

, (3.8)

whereC is a positive constant.

We now formulate our discrete-in-time approximation of problem (PL) for a fixed parameterL >1. For any T >0 andN 1, letN Δt=T andtn=n Δt,n= 0, . . . , N. To prove existence of a solution under minimal smoothness requirements on the initial datum u0 (recall (3.3)), we introduce u0 = u0(Δt) V, the unique solution of

Ω

u

0·v

+Δt∇

xu

0:

xv

dx=

Ω

u0·v

dx

∀v

∈V

; (3.9)

and so

Ω

[|u0|2+Δt|∇xu0|2] dx

Ω

|u0|2dx≤C. (3.10)

In addition, we have thatu0 converges tou0 weakly inH in the limit ofΔt→0+.

(10)

J.W. BARRETT AND E. S ¨ULI

Analogously to defining u0 for a given initial velocity field u0, we shall assign a certain ‘smoothed’ initial datum, ψ0=ψ0(L, Δt)∈HM1×D), to the initial datumψ0 such that

Ω×DM

ψ0ϕ+Δt

xψ0· ∇

xϕ+

qψ0· ∇

qϕ

dq

dx=

Ω×DM βL(ψ0)ϕdq

dx ∀ϕ∈HM1×D).

(3.11) Forp∈[1,∞), let

Zp:=

ϕ∈LpM×D) :ϕ≥0 a.e. onΩ×D and

D

M(q

)ϕ(x

, q

) dq

1 for a.e.x

∈Ω

. (3.12) It is proved in the appendix that there exists a uniqueψ0∈HM1×D) satisfying (3.11); furthermore,

ψ0∈Z1;

Ω×DMF(ψ0) dq

dx+ 4Δt

Ω×DM∇x

ψ02+q

ψ02

dqdx

Ω×DMF(ψ0) dq

dx; (3.13a) and

ψ0≡βL(ψ0)→ψ0 weakly inL1M×D) as L→ ∞ and Δt→0+. (3.13b) It follows from (3.13a), (3.13b) and (1.14) thatψ0∈Z2; in fact,ψ0∈L×D)∩HM1×D).

Our discrete-in-time approximation of (PL) is then defined as follows.

(PΔtL )Let u0L :=u0 ∈V and ψ0L := ψ0 Z2. Then, forn = 1, . . . , N, given (un−1L Ln−1) V ×Z2, find (unLnL)∈V ×(X∩Z2) such that

Ω

u

nL−u

n−1L

Δt + (u

n−1L · ∇

x)u

nL

· w

dx

+ν

Ω

xu

nL:

xw

dx

=f

n, w

H01(Ω)−k

Ω

C(MψnL) :

xw

dx

∀w

∈V

, (3.14a)

Ω×DM ψnL−ψn−1L Δt ϕdq

dx+

Ω×DM 1

2λ∇

qψLn[σ

(u

nL)q

]βL(ψnL)

· ∇qϕdq

dx

+

Ω×DM

ε∇

xψLn−u

n−1 L ψnL

· ∇

xϕdq

dx= 0 ∀ϕ∈X; (3.14b) where, fort∈[tn−1, tn), andn= 1, . . . , N,

f

Δt,+(·, t) =f

n(·) := 1 Δt

tn

tn−1

f

(·, t) dt(H

10(Ω))⊂V

. (3.15)

It follows from (3.3) and (3.15) that N

n=1

Δtf

nrH−1(Ω) T

0 f

rH−1(Ω)dt≤C for anyr∈[1,2], (3.16a) f

Δt,+→f

strongly inL2(0, T; (H

10(Ω))) as Δt→0+. (3.16b)

Références

Documents relatifs

Barrett and S¨ uli have introduced, and proved the existence of global-in-time weak solutions for d = 2 and 3 to, microscopic-macroscopic dumbbell models of dilute polymers

Concerning a priori error estimates for the contact problem of linearly elastic bodies approximated by a standard affine finite-element method, a rate of convergence of order h 3/4 can

However, our current proof of convergence of a subsequence of finite element approximations to a weak solution of the coupled Navier–Stokes–Fokker–Planck system with

Numerical tests are presented, validating the model from both a numerical (convergence in time and space) and a physical (comparison with analytical solutions for the pressure) point

Finite element method, degenerate parabolic equation, nonlinear semigroup.. 1 Department of Mathematics, Faculty of Science, Gakushuin University, 1-5-1 Mejiro, Toshima-ku,

Also, this section strives to critically review techniques designed to boost the L 2 -convergence rate without excessive additional computational effort An a priori and an a

[17] M. VANMAELE, A numerical quadrature finite element method for 2nd-order eigenvalue problems with Dirichlet-Robin boundary conditions. VAN KEER, Error estimâtes for a finite

The numerical results obtained with the pro- posed FEM approximation are compared with those obtained considering the Laplace-Beltrami operator (i.e., its limit when &#34; goes