Uniqueness for the two dimensional Navier Stokes equation with a measure as initial vorticity
32 pages
English

Uniqueness for the two dimensional Navier Stokes equation with a measure as initial vorticity

-

Le téléchargement nécessite un accès à la bibliothèque YouScribe
Tout savoir sur nos offres
32 pages
English
Le téléchargement nécessite un accès à la bibliothèque YouScribe
Tout savoir sur nos offres

Description

Niveau: Supérieur, Licence, Bac+1
Uniqueness for the two-dimensional Navier-Stokes equation with a measure as initial vorticity Isabelle Gallagher Institut de Mathematiques de Jussieu Universite de Paris 7 Case 7012, 2 place Jussieu 75251 Paris Cedex 05, France Thierry Gallay Institut Fourier Universite de Grenoble I BP 74 38402 Saint-Martin d'Heres, France November 25, 2004 Abstract We show that any solution of the two-dimensional Navier-Stokes equation whose vorticity distribution is uniformly bounded in L1(R2) for positive times is entirely determined by the trace of the vorticity at t = 0, which is a finite measure. When combined with previous existence results by Cottet, by Giga, Miyakawa & Osada, and by Kato, this uniqueness property implies that the Cauchy problem for the vorticity equation in R2 is globally well- posed in the space of finite measures. In particular, this provides an example of a situation where the Navier-Stokes equation is well-posed for arbitrary data in a function space that is large enough to contain the initial data of some self-similar solutions. 1 Introduction We consider the two-dimensional incompressible Navier-Stokes equation ∂u ∂t + (u · ?)u = ∆u??p , div u = 0 , x ? R 2 , t > 0 , (1.1) where u(x, t) ? R2 denotes the velocity field of the fluid and p(x, t) ? R the pressure field.

  • similar solution

  • navier- stokes equation

  • function spaces

  • sufficiently small

  • all real-valued

  • vorticity

  • initial measure


Sujets

Informations

Publié par
Nombre de lectures 22
Langue English

Extrait

Uniqueness for the two-dimensional Navier-Stokes equation a measure as initial vorticity
with
Isabelle Gallagher Thierry Gallay Institut de Mathematiques de Jussieu Institut Fourier UniversitedeParis7UniversitedeGrenobleI Case 7012, 2 place Jussieu BP 74 75251ParisCedex05,France38402Saint-MartindHeres,France gallagher@math.jussieu.fr Thierry.Gallay@ujf-grenoble.fr
November 25, 2004
Abstract
We show that any solution of the two-dimensional Navier-Stokes equation whose vorticity distribution is uniformly bounded inL1(R2) for positive times is entirely determined by the trace of the vorticity att nsahiicsueaemitcnehW.erwdenibmoithprevious,0hw= existence results by Cottet, by Giga, Miyakawa & Osada, and by Kato, this uniqueness property implies that the Cauchy problem for the vorticity equation inR2is globally well-posed in the space of nite measures. In particular, this provides an example of a situation where the Navier-Stokes equation is well-posed for arbitrary data in a function space that is large enough to contain the initial data of some self-similar solutions.
1 Introduction
We consider the two-dimensional incompressible Navier-Stokes equation tu+ (u r)u= u rp divu= 0 xR2 > t0(1.1) whereu(x t)R2 eldcityvelostheaddneiufohtedetonp(x t)Rthe pressure eld. Since this system is very well-known, we do not comment here on its derivation and rather refer to the monographs [7], [21], [27] for a general introduction. The  rst mathematical result on the Cauchy problem is due to Leray [22] who proved that, for any initial datau0L2(R2), system (1.1) has a unique global solutionuC0([0+) L2(R2)) such thatu(0) =u0 andruL2((0+) L2(R2 space)). TheL2(R2) is naturally associated with the Navier-Stokesequationfortwodi erentreasons.Firstitistheenergy space, because the square of theL2norm ofuis the  is nonincreasing with time.total (kinetic) energy of the uid, which Next, the s2ap(Rc2e)L2(R2) isscale invariantesehtni,thaetnsthkeutr0a(nsfo)krLm2(aRt2i)o=nku0kL2(R2)for anyu0Land any > is important because0. This u(x t)7→u(x 2t) p(x t)7→2p(x 2t)  >0(1.2)
is a symmetry of (1.1). This invariance was used by Kato [18] to prove that the Navier-Stokes equation in thed-dimensional spaceRdis locally well-posed for arbitrary data inLd(Rd) and evengloballywell-posedforsucientlysmalldatainthatspace,seealso[28],[14].Katosresult was subsequently extended to larger scale invariant function spaces, such as the homogeneous
1
Besov spaceB˙qsp,(Rd) withs= andp q <anchon [3], [4] and Meyer[24].Asimilaranalysiswas1c+arprohtveroirdeuoftsecaybrroMpsyeetaiounaPilnnodnceiet,isnenqCtay Giga and Miyakawa [16]. One interest of dealing with larger function spaces is that they may contain initial data which are homogeneous of degree 1 and therefore give rise to self-similar solutions of (1.1). This is the case of the Besov space above ifq=, or of the larger space BMO 1(Rd such spaces, however, it is not known) introduced by Koch and Tataru [20]. In whether the Cauchy problem is well-posed for large data, even locally in time. We now return to the two-dimensional cased First,= 2 which is simpler for several reasons. the a priori estimates allow in that case to prove that all solutions are global. For instance, in [10], F.Planchonandthe rstauthorprovedthat,forarbitrarydatainB˙,pqs(R2) withs= 1 +p andp q <, there exists a unique global solution to the Navier-Stokes equation (1.1). This result was recently extended by Germain [13] to the larger spaceVMO 1(R2), which is the closure ofS(R2) inBMO 1(R2argestspisisthelvelecoticaferoht.T)gdelht,eruoowonkwnidle yhcih one can solve the Navier-Stokes equation for arbitrary data. Note however thatVMO 1(R2) does not contain any non-trivial homogeneous function of degree 1. Another speci cit y of the two-dimensional case is that the vorticityωd=ef1u2 2u1is a scalarquantitywhichsatis esaremarkablysimpleequation,namely ωt+u rω=  xω R2 t >0.(1.3)
Thevelocity eldu(x t) can be reconstructed from the vorticity distributionω(x t) by the Biot-Savart law u(x t) = 21ZR2(|xx  yy)|2ω(y t) dy  wherex= (x1 x2)d=ef( x2 x1). In terms of the vorticity, the invariance (1.2) reads ω(x t)7→2ω(x 2t).(1.4) A natural scale invariant space for the vorticity is thusL1(R2 Cauchy problem for (1.3)). The inL1(R2) was studied for instance in [1], where results analogous to Leray’s and Kato’s theorems forthevelocity eldareobtained.However,itisimportanttorealizethatavorticityinL1(R2) does not imply a velocity eld inL2(R2). Indeed, ifuL2(R2) and ifω=1u2 2u1L1(R2), then it is easy to verify that necessarilyRR2ωdx= 0. Since the integral ofω(which is the circulationofthevelocity eldatin nity)isconservedundertheevolutionof(1.3),itfollows thatiftheinitialvorticityhasnonzerointegralthentheassociatedvelocity eldwillneverbe of nite energy. This “discrepancy” between function spaces for the vorticity and the velocity is speci c to the two-dimensional case. Indeed, if for instanceωsolves the vorticity equation in L2(R3), then the associated velocity eld does solve the Navier-Stokes equation inL3(R3). In this paper, we study the Cauchy problem for the vorticity equation (1.3) inM(R2), the space of all nite real measures onR2. If∈ M(R2), the total variation ofis de ned by kkM= supZR2ϕdϕC0(R2)kϕkL1whereC0(R2of all real-valued continuous functions on) is the set R2it ninytW.evanishinga recall thatM(R2) equipped with the total variation norm is a Banach space, whose norm is invariant under the scaling transformation (1.4). Another useful topology onM(R2) is the weak-topology which can be characterized as follows: a sequence{n}inM(R2) converges weakly toifRR2ϕdnRR2ϕdasn→ ∞for allϕC0(R2 that case, we write). Inn* .
2
Existence of solutions of (1.3) with initial data inM(R2tet[8],vedbyCot ssrptor)aw and independently by Giga, Miyakawa, and Osada [15]. Uniqueness is a more dicult problem. Using a Gronwall-type argument, it is shown in [15] that uniqueness holds if the atomic part oftheinitialvorticityissucientlysmall,seealso[19].Thefactthatthesizeconditiononly involves the atomic part of the measure is a consequence of the key estimate (see [15]) lim supt1 qketkLqCqkkpp1< q+t0 wherekkppdenotes the total variation of the atomic part of∈ M(R2). On the other hand, the case of a large Dirac mass was solved recently by C.E. Wayne and the second author [12] usingacompletelydi erentapproach,whichwenowbrieydescribe.We rstobservethat, given any R, equation (1.3) has an exact self-similar solution given by ω(x t) =Gt tx u(x t) =vt Gtx xR2 t >0(1.5)
where = 1 .6) G(4)e||2/4 vG( 2) =1||21 e||2/4 R2.(1 This solution is often called theLamb-Oseen vortexwith total circulation . In factω(x t) is also a solution of the linear heat equationtω= ω, because the nonlinearity in (1.3) vanishes identically due to radial symmetry (this is again speci c to the two-dimensional case). The strategy of [12] consists in rewriting (1.3) into self-similar variables as in (2.9) below. Using a pair of Lyapunov functions, the authors show that the Oseen vortices G( R) are the only equilibria of the rescaled equation. By compactness arguments, they deduce that all solutions converge inL1(R2) to Oseen vortices ast+, and as a byproduct that (1.5) is the unique solution of (1.3) such thatkω( t)kL1Kfor allt >0 andω( t)* 0ast0, where0is the Dirac mass at the origin. Another proof of the same result is proposed in [9]. The goal of the present paper is to solve the uniqueness problem in the general case by combining the result of [15], which works when the initial measure has small atomic part, with the method of [12], which allows to handle large Dirac masses. Our main result is the following: Theorem 1.1Let∈ M(R2), and xT >0,K >0. Then the vorticity equation (1.3) has at most one solution ωC0((0 T) L1(R2)L(R2))
such thatkω( t)kL1Kfor allt(0 T)andω( t)* ast0. Here and in the sequel, we say thatω(t)ω( t) is a (mild) solution of (1.3) on (0 T) if the associated integral equation ω(t) = e(t t0)ω(t0) Zt0tr e(t s)u(s)ω(s)ds(1.7) is satis ed for all 0< t0< t < T. If we combine Theorem 1.1 with the existence results in [8], [15], [19], we conclude that there is a unique global solution to (1.3) for any initial measure inM(R2 fact the method we use). In to prove uniqueness also implies that this solution depends continuously on the data, so that the Cauchy problem for the vorticity equation (1.3) is globally well-posed in the spaceM(R2). If in addition we use the results in [12] on the long-time behavior of the solutions, we obtain the following nalstatement:
3
Theorem 1.2
For any
∈ M(R2), the vorticity equation (1.3) has a unique global solution ωC0((0) L1(R2)L(R2))
such thatkω( t)kL1 kkMfor allt >0andω( t)* ast0 solution depends. This continuously on the initial measurein the norm topology ofM(R2), uniformly in time on compact intervals. Moreover, ZR2x= d=ef(R2)for allt >0ω(x t) d and tlimt1 pω(x t) Gt txLxp= 0for allp[1].(1.8) Remark that the spaceM(R2) does contain nontrivial homogeneous distributions (the Dirac masses at the origin), hence Theorem 1.2 gives an example of a situation where the Navier-Stokes equation is well-posed for arbitrary data in a function space that is large enough to contain the initial data of some self-similar solutions (the Oseen vortices). Remark also that Oseen’s vortex plays a double role in Theorem 1.2: it is the unique solution of (1.3) when the initial vorticity is a Dirac mass at the origin, and on the other hand it describes the long-time behavior of all solutions, see (1.8). In fact, it is possible to show that (1.8) is a consequence of the uniqueness of the solution when= 0 with measure-valued initial data is Uniqueness, see [6] and [17]. also a key tool for proving convergence of stochastic approximations of the vorticity equation, see [23]. The rest of this paper is devoted to the proof of Theorem 1.1 and of the continuity statement in Theorem 1.2. Before entering the details, let us give a short idea of the argument. Previous works on the subject assumed that the initial vorticityeither has a small atomic part [15], [19], or consists of a single Dirac mass [12]. So it is natural to decomposemosufot ainetni isolated Dirac masses, and a remainder whose atomic part is arbitrarily small (depending on the number of terms in the previous sum). The idea is then to use the methods of [12] to deal with the large Dirac masses, and the argument of [15], [19] to treat the remainder. The dicult y is of course that equation (1.3) is nonlinear so that the interactions between the various terms have to be controlled. To implement these ideas, we start in Section 2.1 by recalling some general properties of convection-di usionequations,oftheheatsemi-groupinself-similarvariables,andoftheBiot-Savart law. The proof of Theorem 1.1 begins in Section 3, where we decompose the initial measure as explained above and show that the solutionω(x t) also admits a natural decom-position into a sum of Oseen vortices and a remainder. In Section 4, we derive the integral equations satis ed by the remainder terms, and we state a few crucial estimates that will be proved in an appendix (Section 6). These results are used in Section 5, where Theorem 1.1 is proved by a Gronwall-type argument. The same techniques also establish the continuity claim in Theorem 1.2.
Notations.We denote byK0 K1 . . .ts,thevaluesofwhuomriacnnotsnah-ahcix erhtdeguor out the paper. In contrast, we denote byC0 C1 . . .olactnere idekatnahcicwhtsanstonlc valuesindi erentparagraphs.Otherpositiveconstants(whicharenotusedanywhereelsein the text) will be generically denoted byC. As a general rule, we do not distinguish between scalars and vectors in function spaces: althoughu(x tsa)ictvee row,dlirweetuL2(R2) and notuL2(R2)2. To simplify the notation, we denote the mapx7→ω(x t) byω( t) or just byω(t).
Acknowledgements.The authors thank D. Iftimie for fruitful discussions.
4
2
Preliminaries
This section is a collection of known results that will be used in the proof of Theorem 1.1.
2.1Fundamentalsolutionofaconvection-di usionequation
Weconsiderthefollowinglinearconvection-di usionequation
tω(x t) +U(x t) rω(x t) = ω(x t)(2.1) wherexR2,t(0 T), andU:R2(0 T)R2is a (given) time-dependent divergence-free vector eld.TheresultscollectedhereareduetoOsada[25],andtoCarlenandLoss[5]. Following [5], we suppose thatUC0((0 T) L(R2)) and that kU( t)kL(R2)tK00< t < T (2.2) for someK0>to [25], we also assume that 0. According d=ef1U2 2U1C0((0 T) L1(R2)) with k ( t)kL1(R2)K00 .< t < T(2.3) Then any solutionω(x t) of (2.1) can be represented as ω(x t) =Z2 U(x t;y s)ω(y s) dy  xR20< s < t < T (2.4) R whereUauitno2(1..)hTfeollowingsiundathefalsomentfonoitulevnocehtdin-iocteqonsi u properties ofUwill be useful: For any (01) there existsK1>0 (depending only onK0and ) such that 0< U(x t;y s)tK1exp  |xt y|s2)(2.5)  s4( forx yR2and 0< s < t < T also have a similar Gaussian lower bound, see [25]., see [5]. We There exists(01) (depending only onK0) and, for any >0, there existsK2>0 (depending only onK0and) such that | U(x t;y s)  U(x0 t0;y0 s0)| K2|x x0|+|t t0| /2+|y y0|+|s s0| /2(2.6) 0 whenevert sandt0 s, see [25]. For 0< s < t < Tandx yR2, ZR2 U(x t;y s) dx= 1ZR2 U(x t;y s) dy= 1.(2.7) For 0< s < r < t < Tandx yR2, ZR2  U(x t;y s) =U(x t;z r)U(z r;y s) dz .(2.8) Remark 2.1Ifx yR2andt >0, it follows from (2.6) that the functions7→ U(x t;y s) can be continuously extended up tos= 0and that this extension (still denoted by,  U) satis es properties (2.5) to (2.8) withs= 0.
5
  • Univers Univers
  • Ebooks Ebooks
  • Livres audio Livres audio
  • Presse Presse
  • Podcasts Podcasts
  • BD BD
  • Documents Documents