• Aucun résultat trouvé

Global well-posedness for an advection–diffusion equation arising in magneto-geostrophic dynamics

N/A
N/A
Protected

Academic year: 2022

Partager "Global well-posedness for an advection–diffusion equation arising in magneto-geostrophic dynamics"

Copied!
19
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpc

Global well-posedness for an advection–diffusion equation arising in magneto-geostrophic dynamics

Susan Friedlander

a,

, Vlad Vicol

b

aDepartment of Mathematics, University of Southern California, 3620 S. Vermont Ave., Los Angeles, CA 90089, United States bDepartment of Mathematics, University of Chicago, 5734 University Ave., Chicago, IL 60637, United States

Received 10 July 2010; received in revised form 3 January 2011; accepted 11 January 2011 Available online 31 January 2011

Abstract

We use De Giorgi techniques to prove Hölder continuity of weak solutions to a class of drift-diffusion equations, with L2 initial data and divergence free drift velocity that lies inLt BMOx1. We apply this result to prove global regularity for a family of active scalar equations which includes the advection–diffusion equation that has been proposed by Moffatt in the context of magnetostrophic turbulence in the Earth’s fluid core.

©2011 Elsevier Masson SAS. All rights reserved.

Résumé

Nous utilisons des techniques de De Giorgi pour démontrer la continuité Hölder de solutions faibles pour une classe d’équations de dérive-diffusion, avec données initialesL2et champ de vitesse incompressible appartenant àLt BMOx1. Nous appliquons ce résultat pour démontrer la régularité globale pour une famille d’équations du scalaire actif qui comprend l’équation d’advection–

diffusion qui a été proposée par Moffatt dans le contexte de la turbulence magnétostrophique dans le noyau fluide de la Terre.

©2011 Elsevier Masson SAS. All rights reserved.

MSC:76D03; 35Q35; 76W05

Keywords:Global regularity; Weak solutions; De Giorgi; Parabolic equations; Magneto-geostrophic equations

1. Introduction

Active scalar evolution equations have been a topic of considerable study in recent years, in part because they arise in many physical models. In particular, such equations are prevalent in fluid dynamics. In this paper we first examine a class of drift-diffusion equations for an unknown scalar fieldθ (t, x), of the form

tθ+(v· ∇=θ, (1.1)

wherev(t, x)is a given divergence free vector field that lies in the function space L2tL2xLt BMOx1,t >0, and x∈Rd. In Theorem 2.1 we prove that weak solutions to (1.1) are Hölder continuous. Note that this result is new for

* Corresponding author.

E-mail addresses:susanfri@usc.edu (S. Friedlander), vicol@math.uchicago.edu (V. Vicol).

0294-1449/$ – see front matter ©2011 Elsevier Masson SAS. All rights reserved.

doi:10.1016/j.anihpc.2011.01.002

(2)

such linear parabolic equations with very singular coefficients [1,18,19,25,27,34,35]. We then use this result to prove in Theorem 3.1 that Leray–Hopf weak solutions of the active scalar equation

tθ+(u· ∇=θ, (1.2)

divu=0, (1.3)

uj=iTijθ (1.4)

are classical solutions. In (1.4), the velocity vectoruis obtained fromθvia{Tij}, ad×dmatrix of Calderón–Zygmund singular integral operators (that is, they are boundedL2L2andLBMO) such that∂ijTij≡0. Note that in (1.4) we have used the summation convention on repeated indices, andi, j∈ {1, . . . , d}.

Our motivation for addressing the system (1.2)–(1.4) comes from a model proposed by Moffatt [22] for magne- tostrophic turbulence in the Earth’s fluid core. This model is derived from the full magnetohydrodynamic equations (MHD) in the context of a rapidly rotating, density stratified, electrically conducting fluid. After a series of approxima- tions relevant to the geodynamo model, a linear relationship is established between the velocity and magnetic vector fields, and the scalar “buoyancy”θ. The sole remaining nonlinearity in the system occurs in the evolution equation forθ, which has the form

tθ+(u· ∇=S+κθ, (1.5)

whereSis a source term, andκis the coefficient of thermal diffusivity. Here the three-dimensional velocityuis such that divu=0, and it is obtained from the buoyancy via

u=M[θ], (1.6)

whereMis a nonlocal differential operator of order 1. We describe the precise form of the operatorMin Section 4.

An important feature of this operator is the spatial inhomogeneity that occurs due to the underlying mean magnetic field. We call (1.5)–(1.6) themagnetogeostrophic equation(MG). We show that the MG system satisfies the conditions under which we prove Theorem 3.1, and hence obtain (cf. Theorem 4.1) global well-posedness for (1.5)–(1.6).

An active scalar equation that has received much attention in the mathematical literature following its presentation by Constantin, Majda, and Tabak [9], as a two-dimensional toy model for the three-dimensional fluid equations, is the so called surface quasi-geostrophic equation (SQG) (see, for example, [2,6,7,10,11,16,29,33] and references therein).

The dissipative form of this equation for which there is a physical derivation is

tθ+(u· ∇= −()1/2θ, (1.7)

where

u= ∇()1/2θ(R2θ,R1θ ) (1.8)

andRi represents theith Riesz transform. It was recently proved by Caffarelli and Vasseur [2] that solutions of (1.7)–

(1.8) withL2initial data are smooth (see also the review article [3]). Well-posedness for (1.7)–(1.8) in the case of smooth periodic initial data was also obtained by Kiselev, Nazarov, and Volberg [16]. See also Constantin and Wu [10, 11] for the super-critically dissipative SQG.

We note that the magnetogeostrophic equation MG and the critically dissipative SQG equation (1.7)–(1.8) are both derived from the Navier–Stokes equations in the context of a rapidly rotating fluid in a thin shell. For both systems the Coriolis force is dominant in the momentum equation. In the case of the SQG equation the relation (1.8) is derived via a projection of the three-dimensional problem onto the two-dimensional horizontal bounding surface. In the case of the MG equation the coupling with the magnetic induction equation closes the three-dimensional linear system that produces the operators{Tij}, withuj=iTijθ.

Systems (1.2)–(1.4) and (1.7)–(1.8) have strong similarities. In particular, they have the same relative order of the spatial derivatives between the advection term and the diffusive term. Moreover, ifθ (t, x)is a solution of (1.2)–

(1.4), thenθλ(t, x)=θ (λ2t, λx)is also a solution, and henceL(Rd)is the critical Lebesgue space with respect to the natural scaling of the equation. We note thatLis also the critical Lebesgue space for the critically dissipative surface quasi-geostrophic equation (1.7)–(1.8), and for the modified surface quasi-geostrophic equation (cf. Constantin, Iyer, and Wu [8]). The advantage of system (1.2)–(1.4) over the critical SQG equation is that the diffusive term is given

(3)

via a local operator. The tradeoff is that the drift velocity in (1.2)–(1.4) is more singular, i.e., the derivative of aBMO function (see Koch and Tataru [17] for the Navier–Stokes equations inBMO1).

Our proof of Theorem 2.1 and Theorem 3.1 is along the lines of the proof of Caffarelli and Vasseur [2, Theorem 3]

for the critical SQG equation. The primary technique employed in [2,11,32], and in the present paper, is the De Giorgi iteration [12]. This consists of first showing that a weak solution is bounded by proving that the function max{θh,0}

has zero energy ifhis chosen large enough. Then a diminishing oscillation result implies smoothness of the solution in a subcritical space, namely Cα, for someα(0,1). The proof of Hölder continuity for solutions of (1.1) with vLt BMOx1does not follow directly either from [2], wherevLt BMOx, or from [8], wherevLt Cx1α and α(0,1). The crucial step in the proof of Theorem 2.1 is the local energy and uniform estimates. The main obstruction to applying the classical parabolic De Giorgi estimates via anLp-based Caccioppoli inequality (1< p <∞), is that v(t,·)BMO1. In Section 2 we give details as to how we overcome this difficulty.

Eq. (1.1) is in the class of parabolic equations in divergence form that have been studied extensively, including in the classical papers of Nash [24], Moser [23], Aronson and Serrin [1]. Osada [25] allowed for singular coefficients and proved Hölder continuity of solutions to (1.1) whenvLt Wx1,is divergence free. Hence Theorem 2.1 may be also viewed as an improvement of the results of Osada, since iffBMO(Rd)L2(Rd), then it does not follow that fL(Rd)(cf. [30]). In the same spirit, Zhang [34,35] and Semenov [27] give strong regularity results for parabolic equations of the type (1.1), where the singular divergence free velocity satisfies a certain form boundedness condition.

We note that this form boundedness condition does not cover the casevLt BMOx1, and hence Theorem 2.1 does not follow from the results in [27,34,35], and vice-versa. The overall conclusion of the body of work on parabolic equations with a singular drift velocity is that the divergence free structure ofvproduces a dramatic gain in regularity of the solution, compared to the classical theory (cf. [18]).

Organization of the paper. In Section 2 we prove Hölder regularity for the linear drift-diffusion equation (1.1), with vbeing a given divergence free vector field in the function spaceL2t,xLt BMOx1. In Section 3 we apply this result to prove that a Leray–Hopf weak solutionθ of the nonlinear active scalar system (1.2)–(1.4) is Hölder smooth for positive time. Since Hölder regularity is subcritical for the natural scaling of (1.2)–(1.4) we can bootstrap to prove higher regularity and hence conclude that the solution is a classical solution. In Section 4 we describe an active scalar equation that arises as a model for magneto-geostrophic dynamics in the Earth’s fluid core. We show that this three- dimensionalMG equationis an example of the general system (1.2)–(1.4). In Appendix A we prove the existence of weak solutions to (1.2)–(1.4) evolving fromL2(Rd)initial data.

2. Regularity for a parabolic equation with singular drift Consider the evolution of an unknown scalarθ (t, x)given by

tθ+(v· ∇=θ (2.1)

where the velocity vector v(t, x)=(v1(t, x), . . . , vd(t, x))L2((0,)×Rd)is given, and(t, x)∈ [0,∞)×Rd. Additionally letvsatisfy

jvj(t, x)=0 (2.2)

in the sense of distributions. We expressvjas

vj(t, x)=iVij(t, x) (2.3)

in [0,∞)×Rd, where we have used the summation convention on repeated indices, and we denoted Vij =

()1ivj. The matrix{Vij}di,j=1is given, and satisfies VijL

(0,);L2 Rd

L2

(0,); ˙H1 Rd

(2.4) for alli, j∈ {1, . . . , d}.

Theorem 2.1(The linear problem). Given θ0L2(Rd)and {Vij} satisfying(2.4), let θL([0,∞);L2(Rd))L2((0,); ˙H1(Rd))be a global weak solution of the initial value problem associated to(2.1)–(2.3). If additionally

(4)

we have VijL([t0,);BMO(Rd))for all i, j ∈ {1, . . . , d}and somet0>0, then there existsα >0such that θCα([t0,)×Rd).

In analogy with the constructions in [2,11], the proof of Theorem 2.1 consists of two steps. Fort0>0 fixed, we first prove thatθL([t0,);L(Rd)). The main challenge is to prove the Hölder regularity of the solution, which is achieved by using the method of De Giorgi iteration (cf. [12,14,19]). Note that for divergence-free vL2t,x, the existence of a weak solutionθto (2.1)–(2.3), evolving fromθ0L2, is known (for instance, see [27] where the more generalvL1locis treated, also [2], and references therein). Moreover, this weak solution satisfies the classical energy inequality and the level set energy inequalities (2.6) below.

Remark 2.2.The conclusion of Theorem 2.1 holds if the Laplacian on the right side of (2.1) is replaced by a generic second-order strongly elliptic operatori(aijj), with bounded measurable coefficients{aij}.

Remark 2.3.We note that the De Giorgi techniques used here to prove Hölder regularity for solutions to (2.1)–(2.3) can also be used to prove Hölder regularity for the problem with a forcing term Son the right side of (2.1). In this case we considerSLrt,xto be an externally given force, withr >1+d/2 (cf. [19]).

Remark 2.4.In a very recent preprint, Seregin, Silvestre, Šverák, and Zlatoš [28] also use De Giorgi techniques to prove Hölder regularity of solutions to a parabolic equation with drift velocities inLt BMOx1.

Notation.In the following we shall use the classical function spaces:Lp – Lebesgue spaces,BMO– functions with bounded mean oscillation, BMO1 – derivatives of BMO functions, H˙s – homogeneous Sobolev spaces, andCα – Hölder spaces. To emphasize the different integrability in space and time we shall denoteLp([0,∞);Lq(Rd))by LptLqxfor 1p, q∞, and similarly forLptH˙x1andLpt BMOx. AlsoLpt,x(I×B)=Lp(I;Lp(B))for anyI⊂Rand B⊂Rd. The ball inRdand the parabolic cylinder inRd+2are classically denoted byBρ(x0)= {x∈Rd:|xx0|< ρ} andQρ(t0, x0)= [t0ρ2, t0] ×Bρ(x0)forρ >0. Lastly, we shall write(fk)+=max{fk,0}.

2.1. Boundedness of the solution

The first step is to show that a weak solution is bounded for positive time.

Lemma 2.5 (FromL2 toL). Let θL([0,∞);L2(Rd))L2((0,); ˙H1(Rd)) be a global weak solution of (2.1)–(2.3)evolving fromθ0L2(Rd), wherevL2((0,);L2(Rd)). Then for allt >0we have

θ (t,·)

L(Rd)0L2(Rd)

td/4 , (2.5)

for some sufficiently large positive dimensional constantC.

Proof. The proof of this lemma is mutatis-mutandis as in [2,11], and requires only the fact thatvis divergence free.

The main idea is that sincey(yh)+is convex, for allh >0 we have

th)+h)++(v· ∇)(θh)+0,

and hence, multiplying byh)+integrating by parts, and using that divv=0, we obtain the energy inequality

Rd

θ (t2,·)h

+2dx+2

t2

t1

Rd

h)+2dx dt

Rd

θ (t1,·)h

+2dx, (2.6)

for allh >0 and 0< t1< t2<∞. Fort0>0, andH >0 to be chosen sufficiently large, we definetn=t0t0/2n, hn=HH /2n, and

cn=sup

ttn

Rd

θ (t,·)hn

+2dx+2 tn

Rd

hn)2dx dt,

(5)

wheren0. The inequality (2.6), the Gagliardo–Nirenberg–Sobolev inequality, and Riesz interpolation then imply that

cn+1 C

t0H4/d2n(1+4/d)cn1+2/d.

LettingH=Cc1/20 /t0d/40L2(Rd)/t0d/4, for some sufficiently large dimensional constantC, implies thatcn→0 exponentially asn→ ∞, and thereforeθ (t0,·)H. Applying the same procedure to−θconcludes the proof of the lemma. We refer the reader to [2,11] for further details. 2

2.2. Local energy and uniform inequalities

In proving the boundedness of the solution we only required thatvL2t,x, and divv=0. For the rest of the section we use the additional assumptionvLt BMOx1.

Lemma 2.6(First energy inequality). LetθLt L2xL2tH˙x1be a global weak solution of the initial value problem associated to(2.1)–(2.3). Furthermore, assume thatVijL((0,);BMO(Rd))for alli, j∈ {1, . . . , d}, and(2.4) holds. Then for any0< r < Randh∈R, we have

(θ−h)+2

Lt L2x(Qr)+∇h)+2

L2t,x(Qr) CR

(Rr)2(θ−h)+2d+22

L2t,x(QR)(θ−h)+d+22

Lt,x(QR), (2.7) whereC=C(d,VijLt BMOx)is a fixed positive constant, and we have denotedQρ= [t0ρ2, t0] ×Bρ(x0)for ρ >0and an arbitrary(t0, x0)(0,)×Rd. Moreover, estimate(2.7)also holds withθreplaced byθ.

Remark 2.7.Note that from Lemma 2.5 we have thatθLt,x, and hence the right side of (2.7) is finite.

Remark 2.8. The classical local energy inequality (cf. [14,19,25], see also [2,11]) does not contain the term h)+Lt,x(QR) on the right, since the velocity fieldvis not as singular as in our case. In this section we prove that since in (2.7) the exponent 2/(d+2)ofh)+Lt,x(QR)is “small enough”, the De Giorgi program may still be carried out to obtain the Hölder regularity of weak solutions.

Proof of Lemma 2.6. Fixh∈Rand let 0< r < Rbe such thatt0/2R2>0. Letη(t, x)C0((0,)×Rd)be a smooth cutoff function such that

0η1 in(0,)×Rd,

η≡1 inQr(x0, t0), and η≡0 in cl

QcR(x0, t0)

(t, x): tt0 ,

|∇η| C

Rr, |∇∇η| C

(Rr)2, |tη| C

(Rr)2 inQR(x0, t0)\Qr(x0, t0),

for some positive dimensional constantC. Definet1=t0R2>0 and lett2∈ [t0r2, t0]be arbitrary. Multiply (2.1) byh)+η2and then integrate on[t1, t2] ×Rdto obtain

t2

t1

Rd

t

h)2+

η2dx dt−2

t2

t1

Rd

jjh)+h)+η2dx dt+

t2

t1

Rd

iVijj

h)2+ η2dx dt

=0. (2.8)

The main obstruction to applying classical the de Giorgi estimates (via theLp-based Caccioppoli inequality, cf. [14, 19]) is thatiVijLt BMOx1, as opposed to the case Lt Wx1, considered by Osada [25] (see also [27]). We overcome this difficulty by subtracting fromVij(t,·)its spatial mean over{t} ×BR, namelyVij,BR(t )(this does not introduce any lower order terms becausexiVij,BR(t )=0), and by appealing to the John–Nirenberg inequality. More precisely, we define

(6)

Vij,R(t, x)=Vij(t, x)Vij,BR(t )=Vij(t, x)− 1

|BR|

BR

Vij(t, y) dy, (2.9)

and note thatiVij =iVij,R. Therefore, the third term on the left of (2.8) may be replaced by

t2

t1

Rd

iVij,Rj

h)2+

η2dx dt.

We integrate by parts intthe first term on the left of (2.8), and useη(t1,·)≡0. The second term we integrate twice by parts inxj, and the third term on the left of (2.8) we integrate by parts first inxj (and usej(∂iVij,R)=i(∂jVij)=

jvj=0) and then integrate by parts inxi, to obtain 1

2

Rd

θ (t2,·)h2

+η(t2,·)2dx+

t2

t1

Rd

h)+2η2dx dt

=

t2

t1

Rd

h)2+η∂tη dx dt+

t2

t1

Rd

h)2+j(η∂jη) dx dt

t2

t1

Rd

Vij,Rh)2+i(η∂jη) dx dt

−2

t2

t1

Rd

Vij,Rih)+h)+η∂jη dx dt. (2.10)

Using the bounds on the time and space derivatives ofη, the fact thatη≡1 onQr,t2t0, the Hölder andε-Young inequalities, we obtain from (2.10)

Br

θ (t2,·)h2 +dx+2

t2

t1

Rd

h)+2η2dx dt

C

(Rr)2

QR

h)2+dx dt+ C (Rr)2

QR

|Vij,R|h)2+dx dt

+

t2

t1

Rd

h)+2η2dx dt+ C (Rr)2

QR

|Vij,R|2h)2+dx dt. (2.11)

After absorbing the third term on the right of (2.11) into the left side, we take the supremum overt2∈ [t0r2, t0], to obtain

h)+2

Lt L2x(Qr)+∇h)+2

L2t,x(Qr)

C

(Rr)2h)+2

L2t,x(QR)+ C (Rr)2

QR

|Vij,R|h)2+dx dt

+ C

(Rr)2

QR

|Vij,R|2h)2+dx dt. (2.12)

As a corollary of the celebrated John–Nirenberg inequality (cf. [14,30]) we have that for any fixed R >0, t ∈ [t0R2, t0], and 1< p <∞,

Vij,R(t,·)

Lp(BR)=Vij(t,·)Vij,BR(t )

Lp(BR)CVij(t,·)

BMO(Rd)|BR|1/p,

(7)

whereC=C(d, p) >0 is a fixed constant (recall thatC(d, p)→ ∞asp→ ∞). The fact thatVijL([t0/2,); BMO(Rd))implies that for allt∈ [t0R2, t0]we have

Vij,R(t,·)

Lp(BR)C0|BR|1/p (2.13)

for a positive constantC0=C0(VijL([t0/2,);BMO(Rd)), d, p). We fix 0< ε <2 to be chosen later, and using (2.13) and the Hölder inequality we bound

QR

|Vij,R|h)2+dx dt=

t0

t0R2 BR

Vij,R(t, x)(θh)2+(t, x) dx

dt

C0|BR|ε/2

t0

t0R2

θ (t,·)h

+2

L4/(2ε)(BR)dt.

Using the interpolation inequalityfLpCf2/pL2 f1L2/p, withp=4/(2−ε), we obtain from the above esti- mate that

QR

|Vij,R|h)2+dx dtC0|BR|ε/2

t0

t0R2

θ (t,·)h

+2ε

L2(BR)θ (t,·)h

+ε

L(BR)dt

C0Rε(d+2)/2h)+2ε

L2t,x(QR)h)+ε

Lt,x(QR). (2.14)

Similarly, from (2.13), the Hölder inequality andLpinterpolation, we obtain

QR

|Vij,R|2h)2+dx dtC0Rε(d+2)/2h)+2ε

L2t,x(QR)h)+ε

Lt,x(QR). (2.15)

Combining estimates (2.12) with (2.14), (2.15), and the Hölder inequality, we conclude that h)+2

Lt L2x(Qr)+∇h)+2

L2t,x(Qr)

C0Rε(d+2)/2

(Rr)2 (θ−h)+2ε

L2t,x(QR)(θ−h)+ε

Lt,x(QR). (2.16)

The proof of the lemma is concluded by lettingε=2/(d+2)in (2.16). 2 By applying the Hölder inequality to the right side of (2.7) we then obtain:

Corollary 2.9.Letθbe as in Lemma2.6. Then we have (θ−h)+2

Lt L2x(Qr)+∇h)+2

L2t,x(Qr) CRd+2

(Rr)2(θ−h)+2

Lt,x(QR), (2.17)

for some positive constantC=C(d,VijLt BMOx).

We now fix a point(t0, x0)(0,)×Rdand we prove the Hölder continuity ofθat this point. Throughout the following we denote byQρthe cylinderQρ(t0, x0), for anyρ >0.

The following lemma gives an estimate on the supremum ofθon a half cylinder, in terms of the supremum on the full cylinder. A similar statement may be proven for−θ.

Lemma 2.10.Letθbe as in Lemma2.6. Assume thath0supQr

0θ, wherer0>0is arbitrary. We have sup

Qr

0/2

θh0+C

|{θ > h0} ∩Qr0|1/(d+2) r0

1/2 sup

Qr

0

θh0

(2.18) for some positive constantC=C(d,VijLt BMOx).

(8)

The above estimate differs from the classical one cf. [19, Theorem 6.50] in that the power of|{θ > h0}∩Qr0|/|Qr0| is 1/(2d+4)instead of 1/(d+2). However, the key feature of (2.18) is that the coefficient of (supQr

0θh0) does not scale withr0. It is convenient to introduce the following notation:

A(h, r)= {θ > h} ∩Qr,

a(h, r)= |A(h, r)|,

b(h, r)= h)+2L2 t,x(Qr),

M(r)=supQrθ,

m(r)=infQrθ,

• osc(Q)=supQθ−infQθ.

Proof of Lemma 2.10. Let 0< r < Rand 0< h < H. We have b(h, r)= θh2L2

t,x(A(h,r))θh2L2

t,x(A(H,r))(Hh)2a(H, r). (2.19)

Let η(t, x)C0(R×Rd)be a smooth cutoff such thatη≡1 onQr,η≡0 onQc(r+R)/2∩ {t t0}, and |∇η| C/(Rr)for some universal constantC >0. Then, by Hölder’s inequality and the choice ofηwe obtain

b(h, r)=h)+2

L2t,x(Qr)

a(h, r)2/(d+2)(θ−h)+2

L2(dt,x+2)/d(Qr)

a(h, r)2/(d+2)η(θ−h)+2

L2(dt,x+2)/d((−∞,t0)×Rd). (2.20)

Using the Gagliardo–Nirenberg–Sobolev inequality and Riesz interpolation f2L2(d+2)/d((−∞,t0)×Rd)Cf2L

t L2x((−∞,t0)×Rd)+Cf2L2

t,x((−∞,t0)×Rd), estimate (2.20) implies that

b(h, r)Ca(h, r)2/(d+2)η(θ−h)+2

Lt L2x((−∞,t0)×Rd)+∇

η(θh)+2

L2t,x((−∞,t0)×Rd)

Ca(h, r)2/(d+2)h)+2

Lt L2x(Q(r+R)/2)+∇h)+2

L2t,x(Q(r+R)/2)

+ 1

(Rr)2(θ−h)+2

L2t,x(Q(r+R)/2)

for some positive dimensional constantC. Using the first energy inequality, i.e., Lemma 2.6, and the Hölder inequality, we bound the far right side of the above and obtain

b(h, r)Ca(h, r)2/(d+2) R

(Rr)2h)+22/(d+2)

L2t,x(QR) h)+2/(d+2)

Lt,x(QR)

Ca(h, r)2/(d+2) R

(Rr)2b(h, R)11/(d+2)h)+2/(d+2)

Lt,x(QR) (2.21)

for some sufficiently large positive constantC=C(d,VijLt BMOx). By combining estimates (2.19) and (2.21) we obtain the main consequence of Lemma 2.6, that is

b(H, r) CR

(Hh)4/(d+2)(Rr)2b(h, R)1+1/(d+2)H )+2/(d+2)

Lt,x(QR). (2.22)

The above estimates give the proof of the lemma as follows. Let rn =r0/2+r0/2n+1r0/2, hn =h(hh0)/2nh, and bn=b(hn, rn+1), for all n0, where r0 andh0 are as in the statement of the lemma, whileh>0 is to be chosen later. By lettingH=hn+1,h=hn,r=rn+2, andR=rn+1in (2.22), we obtain

(9)

bn+1 Crn+1

(hh0)4/(d+2)r022n(2+4/(d+2))b1n+1/(d+2)hn+1)+2/(d+2)

Lt,x(Qrn+1)

C(M(r0)h0)2/(d+2)

(hh0)4/(d+2)r0 2n(2+4/(d+2))b1n+1/(d+2), (2.23) by using

(θ−hn+1)+

Lt,x(Qrn+1)= sup

A(hn+1,rn+1)

θhn+1 sup

Qrn+1

θhn+1sup

Qr0

θh0=M(r0)h0 which holds sinceM(rn)M(r0)andhnh0. LetB=24+2(d+2). We choosehlarge enough so that

C(M(r0)h0)2/(d+2)

(hh0)4/(d+2)r0 b1/(d0 +2) 1

B, (2.24)

then by induction we obtain from (2.23) that bn b0/Bn, and therefore bn→0 as n→ ∞. This implies that supQr

0/2θh. A simple calculation shows that if we let h=h0+CB(d+2)/4(M(r0)h0)1/2b1/40

r0(d+2)/4 (2.25)

then (2.24) holds. Lastly,b0=b(h0,3r0/4)may be bounded via (2.21) and the Hölder inequality as b0Ca(h0, r0)1/(d+2)r0d+2

M(r0)h02

. (2.26)

The proof of the lemma is concluded by combining supQr

0/2θhwith (2.25) and (2.26). From the above proof it follows that inequality (2.18) also holds withθis replaced by−θ. 2

As opposed to the elliptic case, in the parabolic theory we need an additional energy inequality to control the possible growth of level sets of the solution.

Lemma 2.11(Second energy inequality). LetθLt L2xL2tH˙x1be a global weak solution of the initial value problem associated to(2.1)–(2.3). Furthermore, assume thatVijLt BMOx for alli, j∈ {1, . . . , d}, and(2.4)holds. Fix an arbitraryx0∈Rd, leth∈R,0< r < R, and0< t1< t2. Then we have

θ (t2,·)h

+2

L2(Br)θ (t1,·)h

+2

L2(BR)+C Rd(t2t1)

(Rr)2 h)+2

Lt,x((t1,t2)×BR) (2.27) for some sufficiently large positive constantC=C(d,VijLt BMOx), where we have denotedBρ=Bρ(x0)forρ >0.

Proof. Note that by Lemma 2.5 we have that θLt,x and hence the right side of (2.27) is finite. LetηC0(Rd) be a smooth cutoff such thatη≡1 onBr,η≡0 onBRc, and|∇η(x)|C/(Rr), for allx∈Rd, for some constant C >0. Multiply (2.1) byη2h)+and integrate fromt1tot2to obtain

t2

t1

Rd

t

h)+2

η2dx dt−2

t2

t1

Rd

jjh)+h)+η2dx dt

= −

t2

t1

Rd

iVijj

h)+2

η2dx dt

= −

t2

t1

Rd

iVij,Rj

h)+2

η2dx dt,

where, as in (2.9), we have denotedVij,R(t, x)=Vij(t, x)|B1R|

BRVij(t, y) dy. After integrating by parts we get

Références

Documents relatifs

(2) is discretized with an implicit Eulerian centered scheme and direct MUMPS [1] or preconditioned MILU-BICGSTAB II iterative solvers are used to obtain solutions. In addition to

[11] A. Grünrock, Bi- and trilinear Schrödinger estimates in one space dimension with applications to cubic NLS and DNLS, International Mathematics Research Notices, 41 ,

We prove the global well-posedness and we study the linear response for a system of two coupled equations composed of a Dirac equation for an infinite rank operator and a nonlinear

To give a qualitative picture, we report in figure 6 the final numerical solutions after one full turn obtained using the SD-ROM with online a-posteriori stabilization for r = 30,

[1] Gregory R. Baker, Xiao Li, and Anne C. Analytic structure of two 1D-transport equations with nonlocal fluxes. Bass, Zhen-Qing Chen, and Moritz Kassmann. Non-local Dirichlet

We give some sufficient conditions in order that the uniqueness and comparison properties hold for the associated solutions; these conditions are expressed in terms of the moduli

The method of [22] turned out to be useful in the context of the KP-I equation (and some other models such as the Schr¨ odinger maps [15, 19]) and the first global well-posedness

In [18] this approach combined with estimates in Bourgain type spaces leads to a global well-posedness result in the energy space H 1/2 ( T ).. For a Banach space X, we denote by k ·