• Aucun résultat trouvé

The Stokes conjecture for waves with vorticity

N/A
N/A
Protected

Academic year: 2022

Partager "The Stokes conjecture for waves with vorticity"

Copied!
25
0
0

Texte intégral

(1)

Ann. I. H. Poincaré – AN 29 (2012) 861–885

www.elsevier.com/locate/anihpc

The Stokes conjecture for waves with vorticity

Eugen Varvaruca

a

, Georg S. Weiss

b,

aDepartment of Mathematics and Statistics, University of Reading, Whiteknights, PO Box 220, Reading RG6 6AX, UK bDepartment of Mathematics, Heinrich Heine University, 40225 Düsseldorf, Germany

Received 10 October 2011; received in revised form 6 March 2012; accepted 7 May 2012 Available online 23 May 2012

Abstract

We study stagnation points of two-dimensional steady gravity free-surface water waves with vorticity.

We obtain for example that, in the case where the free surface is an injective curve, the asymptotics at any stagnation point is given either by the “Stokes corner flow” where the free surface has acorner of120, or the free surface ends in ahorizontal cusp, or the free surface ishorizontally flatat the stagnation point. The cusp case is a new feature in the case with vorticity, and it is not possible in the absence of vorticity.

In a second main result we exclude horizontally flat singularities in the case that the vorticity is 0 on the free surface. Here the vorticity may have infinitely many sign changes accumulating at the free surface, which makes this case particularly difficult and explains why it has been almost untouched by research so far.

Our results are based on calculations in the original variables and do not rely on structural assumptions needed in previous results such as isolated singularities, symmetry and monotonicity.

©2012 Elsevier Masson SAS. All rights reserved.

1. Introduction

The classical hydrodynamical problem of traveling two-dimensional gravity water waves with vorticity can be described mathematically as a free-boundary problem for a semilinear elliptic equation: given an open connected setΩ in the(x, y)plane and a functionγ of one variable, find a non-negative functionψinΩ such that

ψ= −γ (ψ ) inΩ∩ {ψ >0}, (1.1a)

ψ (x, y)2= −y onΩ{ψ >0}. (1.1b)

The present paper is an investigation by geometric methods of the singularities of the free boundary{ψ >0}.

Let us briefly describe, following[4], the connection between problem(1.1)and the nonlinear governing equations of fluid motion. Consider a wave of permanent form moving with constant speed on the free surface of an incompress- ible inviscid fluid, acted on by gravity. With respect to a frame of reference moving with the speed of the wave, the flow is steady and occupies a fixed regionDin the plane. The boundary∂Dof the fluid region contains a partaD

* Corresponding author.

E-mail addresses:e.varvaruca@reading.ac.uk(E. Varvaruca),weiss@math.uni-duesseldorf.de(G.S. Weiss).

0294-1449/$ – see front matter ©2012 Elsevier Masson SAS. All rights reserved.

http://dx.doi.org/10.1016/j.anihpc.2012.05.001

(2)

which is free and in contact with an air region. Under the assumption that the fluid regionDis simply connected, the incompressibility condition shows that the flow can be described by astream functionψ:DR, so that the relative fluid velocity isy,ψx). The Euler equations imply that thevorticityω:= −ψsatisfies

ωxψy=ωyψx inD. (1.2)

It is easy to see that (1.2) is satisfied whenever

ω=γ (ψ ) inD (1.3)

for some (smooth) function γ of variableψ, which will be referred to as avorticity function. (Conversely, under additional assumptions, see[4], (1.2) implies the existence of such a functionγ.) The kinematic boundary condition that the same particles always form the free surfaceaDis equivalent to

ψis locally constant onaD.

Also, in the presence of (1.3), Bernoulli’s Theorem and the fact that on the fluid–air interfaceaDthe pressure in the fluid equals the constant atmospheric pressure imply that

1

2|∇ψ|2+gyis locally constant onaD,

whereg >0 is the gravitational constant. We therefore obtain, after some normalization, and at least in the case when

aDis connected, that the following equations and boundary conditions are satisfied:

ψ=γ (ψ ) inD, (1.4a)

ψ=0 onaD, (1.4b)

|∇ψ|2+2gy=0 onaD. (1.4c)

Eqs. (1.4) are usually supplemented by suitable boundary conditions on the rest of the boundary of D, or some conditions on the flow at infinity if the fluid domain is unbounded. Classical types of waves which have received most attention in the literature are periodic and solitary waves of finite depth (in which the fluid domainDhas a fixed flat bottomy= −d, at whichψis constant), and periodic waves of infinite depth (in which the fluid domain extends to y= −∞and the condition limy→−∞ψ (x, y)=(0,c)holds, wherecis the speed of the wave). Conversely, for any vorticity functionγ, any solution of(1.4)gives rise to a traveling free-surface gravity water wave, irrespective of whether D is simply connected or aD is connected. Problem(1.1)is a local version of problem (1.4), under the additional assumption that ψ >0 in the fluid region, and whereψ has been extended by the value 0 to the air region. In(1.1), the domainΩ is a neighborhood of a point of interest on the fluid–air interface, the fluid regionD is identified with the set {(x, y): ψ (x, y) >0}(in short{ψ >0}) and the fluid–air interface aD with{ψ >0}, while the gravitational constantg has been normalized by scaling. Note that problem(1.1)is also relevant for the description of more general steady flow configurations (for example, the fluid domain could have a non-flat bottom, and there could be some further external forcing acting at the boundary of the fluid region which is not in contact with the air region).

The theory of traveling water waves with vorticity has a long history, whose highlights include the pioneering paper of Gerstner[10], the first rigorous proof of existence of periodic waves of small amplitude by Dubreil-Jacotin [6], and the foundation[4]of Constantin and Strauss, which proved existence of smooth waves of large amplitude for the periodic finite-depth problem. The paper[4]has generated substantial interest and follow-up work on steady water waves with vorticity, see[20]for a survey of recent results.

In this paper we investigate the shape of the free boundary{ψ >0}atstagnation points, which are points where the relative fluid velocityy,ψx)is the zero vector. The Bernoulli condition (1.1b) shows that such points are on the real axis, while the rest of the free boundary is in the lower half-plane. Stokes[19]conjectured that, in the irrotational caseγ≡0, at any stagnation point the free surface has a (symmetric) corner of 120, and formal asymptotics suggest that the same result might be true also in the general case of waves with vorticityγ ≡0. (See Fig.1.) In the irrotational case, the Stokes conjecture was first proved, under isolatedness, symmetry, and monotonicity assumptions, by Amick, Fraenkel and Toland[3]and Plotnikov[15](see also[21]for a simplification of the proof in[3]), while a geometric proof has recently been given in[23]without any such structural assumptions.

(3)

Fig. 1. Stokes corner. Fig. 2. Cusp. Fig. 3. Horizontally flat stagnation point.

In the caseγ ≡0, the only rigorous results available on waves with stagnation points are very recent and require in an essential way symmetry and monotonicity of the free surface: in[22]it was proved that, at stagnation points, a symmetric monotone free boundary has either a corner of 120or a horizontal tangent. Moreover, it was also shown there that, ifγ0 close to the free surface, then the free surface necessarily has a corner of 120. (On the other hand, ifγ (0) <0, there exist very simple examples where the free surface is the real axis, a line of stagnation points.) The existence of waves, with non-zero vorticity, having stagnation points has been obtained in the setting of periodic waves of finite depth over a flat horizontal bottom, in the following cases in the paper[17]submitted simultaneously with the present paper: for any non-positive vorticity functionγ and any period of the wave, and under certain restrictions on the size ofγ and the wave period (roughly speaking, the vorticity has to be sufficiently small and the period sufficiently large) ifγ is positive somewhere. The extreme waves constructed in[17]are obtained as weak limits of large-amplitude smooth waves whose existence was proved by Constantin and Strauss[4], and they are symmetric and monotone. It was shown in[17]that the free surface of any symmetric monotone wave with stagnation points which is a limit of smooth waves cannot have a horizontal tangent at the stagnation points (in particular, the free surface cannot be horizontally flat), irrespective of the vorticity functionγ, and therefore, as a consequence of[22], the free surface of such a wave necessarily has corners of 120at stagnation points.

The present paper is the first study of stagnation points of steady two-dimensional gravity water waves with vor- ticity in the absence of structural assumptions of isolatedness of stagnation points, symmetry and monotonicity of the free boundary, which have been essential assumptions in all previous works. We obtain for example that, in the case when the free surface is an injective curve, the asymptotics at any stagnation point is given either by the “Stokes corner flow” where the free surface has acorner of120, or the free surface ends in ahorizontal cusp(see Fig.2), or the free surface ishorizontally flatat the stagnation point (see Fig.3).

The cusp case is a new feature in the case with vorticity, and it is not possible without the presence of vorticity [23]. It is interesting to point out that Gerstner[10]constructed an explicit example of a steady wave with vorticity whose free surface has avertical cuspat a stagnation point. However, that vertical cusp is due to the fact that in his example the vorticity is infinite at the free surface, while in the present paper we only consider the case of vorticities which are smooth up to the free surface. We conjecture the cusps in our paper — the existence of which is still open

— to be due to the break-down of the Rayleigh–Taylor condition in the presence of vorticity.

The second half of our paper is devoted toexcluding horizontally flat singularitiesin the case that the vorticity is non-negative at the free surface. (Horizontally flat singularities are possible if the vorticity is negative at the free surface.) Of particular difficulty is the case when the vorticity is 0 at the free surface, and may have infinitely many sign changes accumulating there.

Let us briefly state our main result and give a plan of the paper:

Main Result.Letψbe a suitable weak solution of (1.1)(compare to Definition3.2)satisfyingψ (x, y)2Cmax(−y,0) locally inΩ,

let the free boundary∂{ψ >0}be a continuous injective curveσ =1, σ2)such thatσ (0)=(x0,0), and assume that the vorticity function satisfies either|γ (z)|Cz, orγ (z)0, for allzin a right neighborhood of0.

(4)

(i) If the Lebesgue density of the set{ψ >0}at(x0,0)is positive, then the free boundary is in a neighborhood of (x0,0)the union of twoC1-graphs of functionsη1:(x0δ, x0] →Randη2: [x0, x0+δ)Rwhich are both continuously differentiable up tox0and satisfyη1(x0)=1/√

3andη2(x0)= −1/√ 3.

(ii) Elseσ1(t) =x0in(t1, t1)\ {0},σ1x0does not change its sign att=0, and

tlim0

σ2(t) σ1(t)x0=0.

If we assume in addition that either{ψ >0}is a subgraph of a function in they-direction or that{ψ >0}is a Lipschitz set, then the set of stagnation points is locally in Ω a finite set, and at each stagnation point (x0,0)the statement in(i)holds.

1.1. Plan of the paper

The flow of the paper follows[23]with new aspects and difficulties which we are going to point out:

After gathering some notation in Section2, in Section3we introduce suitable weak solutions and prove a mono- tonicity formula. Consequences of the monotonicity formula (Section4) make a blow-up analysis of singularities possible. The general case (without the injective curve assumption) is stated in Theorem4.5. Different from the zero vorticity case handled in[23], there appears a new case in which the Lebesgue density of the set{ψ >0}is 0. As- suming the free surface to be an injective curve in a neighborhood of the singularity we obtain in Theorem4.6a more precise description: in the new case the free surface formscuspspointing in thex- or−x-direction. As in[23]we are able to show that Stokes corner singularities are isolated points (Section5).

Starting with Section 6, the focus of our analysis is on points at which the set{ψ >0}has full Lebesgue density. In the caseγ (0)=0, an extension of thefrequency formula(Theorem6.7) introduced by the authors in[23]leads here to aBessel differential inequality(see the proof of Theorem6.12) which shows that the right-hand side of the frequency formula is integrable. This part is substantially different from[23]. It is then possible (Sections7–9) to do a blow-up analysis in order to exclude horizontally flat singularities (Theorem10.1). All our results are based on calculations in the original variables.

2. Notation

We denote byχAthe characteristic function of a setA. For any real numbera, the notationa+stands for max(a,0).

We denote byx·y the Euclidean inner product in Rn×Rn, by|x| the Euclidean norm inRn and byBr(x0):=

{xRn: |xx0|< r}the ball of centerx0and radiusr. We will use the notationBr forBr(0), and denote byωnthe n-dimensional volume ofB1. Also,Lnshall denote then-dimensional Lebesgue measure andHs thes-dimensional Hausdorff measure. Byνwe will always refer to the outer normal on a given surface. We will use functions of bounded variationBV(U ), i.e. functionsfL1(U )for which the distributional derivative is a vector-valued Radon measure.

Here|∇f|denotes the total variation measure (cf.[12]). Note that for a smooth open setERn,|∇χE|coincides with the surface measure on∂E.

3. Notion of solution and monotonicity formula

In some sections of the paper we work with an n-dimensional generalization of the problem described in the Introduction. LetΩ be a bounded domain inRnwhich has a non-empty intersection with the hyperplane{xn=0}, in which to consider the combined problem for fluid and air. We study solutionsu, in a sense to be specified, of the problem

u= −f (u) inΩ∩ {u >0},

|∇u|2=xn onΩ{u >0}. (3.1)

Note that, compared to the Introduction, we have switched notation fromψ tou and fromγ tof, and we have

“reflected” the problem at the hyperplane{xn=0}. The nonlinearityf is assumed to be a continuous function with

(5)

primitiveF (z)=z

0f (t ) dt. Since our results are completely local, we do not specify boundary conditions on∂Ω. In view of the second equation in (3.1), it is natural to assume throughout the rest of the paper thatu≡0 inΩ∩ {xn0}.

We begin by introducing our notion of avariational solutionof problem (3.1).

Definition 3.1(Variational solution).We defineuWloc1,2(Ω)to be avariational solutionof (3.1) ifuC0(Ω)C2∩ {u >0}),u0 inΩ andu≡0 inΩ∩ {xn0}, and the first variation with respect to domain variations of the functional

J (v):=

Ω

|∇v|2−2F (v)+xnχ{v>0} dx

vanishes atv=u, i.e.

0= − d dJ

u

x+φ(x)

=0

=

Ω

|∇u|2−2F (u)

divφ−2∇uDφu+xnχ{u>0}divφ+χ{u>0}φn

dx

for anyφC01;Rn).

Note for future reference that for each open set there isCD<+∞such thatu+CD is a non-negative Radon measure inD, the support of the singular part of which (with respect to the Lebesgue measure) is contained in the set{u >0}: by Sard’s theorem{u=δ} ∩Dis for almost everyδ a smooth surface. It follows that for every non-negativeζC0(D)

D

∇max(u−δ,0)· ∇ζCDζ dx=

D

ζ (χ{u>δ}u+CD) dx

D{u>δ}

ζu·ν dHn10, provided that|f (u)|CDinD. Lettingδ→0 and using thatuis continuous and non-negative inΩ, we obtain

D

(u· ∇ζCDζ ) dx0.

Thusu+CDis a non-negative distribution inD, and the stated property follows.

Since we want to focus in the present paper on the analysis of stagnation points, we will assume that everything is smooth away fromxn=0, however this assumption may be weakened considerably by using in{xn>0}regularity theory for the Bernoulli free boundary problem (see [2] for regularity theory in the case f =0 — which could effortlessly be perturbed to include the case of bounded f — and see [5] for another regularity approach which already includes the perturbation).

Definition 3.2(Weak solution).We defineuWloc1,2(Ω)to be aweak solutionof (3.1) if the following are satisfied:

uis avariational solutionof (3.1) and the topological free boundary{u >0}∩Ω∩{xn>0}is locally aC2,α-surface.

Remark 3.3.(i) It follows that in{xn>0}the solution is a classical solution of (3.1).

(ii) For any weak solutionuof (3.1) such that

|∇u|2Cxn+ locally inΩ,

uis a variational solution of (3.1),χ{u>0}is locally in{xn>0}a function of bounded variation, and the total variation measure|∇χ{u>0}|satisfies

r1/2n

Br(y)

xnd|∇χ{u>0}|C0

for allBr(y)Ωsuch thatyn=0 (see[23, Lemma 3.4]).

(6)

The first tool in our analysis is an extension of the monotonicity formula in[25],[24, Theorem 3.1]to the boundary case. The roots of those monotonicity formulas are harmonic mappings[18,16]and blow-up[14].

Theorem 3.4(Monotonicity formula). Letube a variational solution of (3.1), letx0Ω such thatxn0=0, and let δ:=dist(x0, ∂Ω)/2. Let, for anyr(0, δ),

Ix0,u(r)=I (r)=rn1

Br(x0)

|∇u|2uf (u)+xnχ{u>0}

dx, (3.2)

Jx0,u(r)=J (r)=rn2

∂Br(x0)

u2dHn1, (3.3)

Mx0,u(r)=M(r)=I (r)−3

2J (r) (3.4)

and

Kx0,u(r)=K(r)=r

∂Br(x0)

2F (u)−uf (u)

dHn1+

Br(x0)

(n−2)uf (u)−2nF (u)

dx. (3.5)

Then, for a.e.r(0, δ), I(r)=rn2

2r

∂Br(x0)

(u·ν)2dHn1−3

∂Br(x0)

uu·ν dHn1

+rn2K(r), (3.6)

J(r)=rn3

2r

∂Br(x0)

uu·ν dHn1−3

∂Br(x0)

u2dHn1

(3.7)

and

M(r)=2rn1

∂Br(x0)

u·ν−3 2 u r

2

dHn1+rn2K(r). (3.8)

Proof. The identity (3.7) can be easily checked directly, being valid for any functionuWloc1,2(Ω)(not necessarily a variational solution of (3.1)).

For small positiveκandηκ(t):=max(0,min(1,rκt)), we take after approximationφκ(x):=ηκ(|xx0|)(xx0) as a test function in the definition of a variational solution. We obtain

0=

Ω

|∇u|2−2F (u)+xnχ{u>0}

κx−x0+ηκx−x0xx0dx

−2

Ω

|∇u|2ηκx−x0+ ∇u· xx0

|xx0|∇u· xx0

|xx0|ηx−x0x−x0 dx

+

Ω

ηκx−x0xnχ{u>0}dx.

Passing to the limit asκ→0, we obtain, for a.e.r(0, δ), 0=n

Br(x0)

|∇u|2−2F (u)+xnχ{u>0} dxr

∂Br(x0)

|∇u|2−2F (u)+xnχ{u>0} dHn1 +2r

∂Br(x0)

(u·ν)2dHn1−2

Br(x0)

|∇u|2dx+

Br(x0)

xnχ{u>0}dx. (3.9)

(7)

Also observe that letting→0 in

Br(x0)

u· ∇max(u−,0)1+dx=

Br(x0)

f (u)max(u−,0)1+dx+

∂Br(x0)

max(u−,0)1+u·ν dHn1

for a.e.r(0, δ), we obtain the integration by parts formula

Br(x0)

|∇u|2uf (u) dx=

∂Br(x0)

uu·ν dHn1 (3.10)

for a.e.r(0, δ).

Note also that

I(r)= −(n+1)rn2

Br(x0)

|∇u|2uf (u)+xnχ{u>0} dx

+rn1

∂Br(x0)

|∇u|2uf (u)+xnχ{u>0}

dHn1. (3.11)

Using (3.9) and (3.10) in (3.11), we obtain (3.6). Finally, (3.8) follows immediately by combining (3.6) and (3.7). 2 4. Densities

From now on we assume Assumption 4.1.Letusatisfy

|∇u|2Cxn+ locally inΩ.

Remark 4.2.Note that Assumption4.1implies that u(x)C1

xn+3/2

and that in the casexn0=0, rn2K(r)C2

√1 r,

whereC2depends onx0but is locally uniformly bounded.

Remark 4.3.Unfortunately the combination of vorticity and gravity makes it hard to obtain the estimate

|∇u|2+2F (u)−xn+0 (4.1)

related to the Rayleigh–Taylor condition in the time-dependent problem, but the weaker estimate Assumption4.1has been verified under certain assumptions in[22].

We first show that the functionMx0,uhas a right limitMx0,u(0+), of which we derive structural properties.

Lemma 4.4.Letube a variational solution of (3.1)satisfying Assumption4.1. Then:

(i) Letx0Ω be such thatxn0=0. Then the limitMx0,u(0+)exists and is finite.(Note thatu=0 in{xn=0}by assumption.)

(8)

(ii) Letx0Ωbe such thatxn0=0, and let0< rm→0+asm→ ∞be a sequence such that the blow-up sequence um(x):=u(x0+rmx)

rm3/2

(4.2) converges weakly inWloc1,2(Rn)to a blow-up limitu0. Thenu0is a homogeneous function of degree3/2, i.e.

u0(λx)=λ3/2u0(x) for anyxRnandλ >0.

(iii) Letumbe a converging sequence of (ii). Thenumconverges strongly inWloc1,2(Rn).

(iv) Letx0Ωbe such thatx0n=0. Then Mx0,u(0+)= lim

r0+rn1

Br(x0)

xn+χ{u>0}dx,

and in particularMx0,u(0+)∈ [0,+∞). Moreover,Mx0,u(0+)=0implies thatu0=0inRnfor each blow-up limitu0of (ii).

(v) The functionxMx,u(0+)is upper semicontinuous in{xn=0}.

(vi) Letumbe a sequence of variational solutions of (3.1)with nonlinearityfmin a domainΩm, where Ω1Ω2⊂ · · · ⊂ΩmΩm+1⊂ · · · and

m=1

Ωm=Rn,

such thatum converges strongly tou0inWloc1,2(Rn),χ{um>0} converges weakly inL2loc(Rn)toχ0, andfm(um) converges to0locally uniformly inRn. Thenu0is a variational solution of (3.1)with nonlinearityf =0inRn and satisfies the monotonicity formula(withf =0), but withχ{u0>0}replaced byχ0. Moreover, for eachx0Rn such thatxn0=0, and all instances ofχ{u0>0}replaced byχ0,

Mx0,u0(0+)lim sup

m→∞ Mx0,um(0+).

Proof. (i) By Remark4.2, u(x)C1

xn+3/2

locally inΩ

and rn2K(r)C2r1/2 for eachx0Ωsatisfyingxn0=0. (4.3) Thusrrn2K(r)is integrable at such pointsx0, and from Theorem3.4we infer that the functionMx0,u has a finite right limitMx0,u(0+).

(ii) For each 0< σ <∞the sequenceumis by assumption bounded inC0,1(Bσ). For any 0< < σ <∞, we write the identity (3.8) in integral form as

2 σ

rn1

∂Br(x0)

u·ν−3 2 u r

2

dHn1dr=M(σ )M()σ

rn2K(r) dr. (4.4)

It follows by rescaling in (4.4) that 2

Bσ(0)\B(0)

|x|n3

um(x)·x−3 2um(x)

2

dx

M(rmσ )M(rm)+

rmσ

rm

rn2K(r)dr→0 asm→ ∞, which yields the desired homogeneity ofu0.

(9)

(iii) In order to show strong convergence ofuminWloc1,2(Rn), it is sufficient, in view of the weakL2-convergence of∇um, to show that

lim sup

m→∞

Rn

|∇um|2η dx

Rn

|∇u0|2η dx

for eachηC01(Rn). Letδ:=dist(x0, ∂Ω)/2. Then, for eachm,umis a variational solution of um= −rm1/2f

rm3/2um

inBδ/rm∩ {um>0},

|∇um|2=xn onBδ/rm{um>0}. (4.5)

Sinceumconverges tou0locally uniformly, it follows from (4.5) thatu0 is harmonic in{u0>0}. Also, using the uniform convergence, the continuity ofu0and its harmonicity in{u0>0}we obtain as in the proof of (3.10) that

Rn

|∇um|2η dx= −

Rn

um

um· ∇ηrm1/2f rm3/2um

η dx

→ −

Rn

u0u0· ∇η dx=

Rn

|∇u0|2η dx

asm→ ∞. It therefore follows thatumconverges tou0strongly inWloc1,2(Rn)asm→ ∞.

(iv) Let us take a sequencerm→0+such thatumdefined in (4.2) converges weakly inWloc1,2(Rn)to a functionu0. Note that by the definition of a variational solution,um andu0 are identically zero inxn0. Using (iii) and the homogeneity ofu0, we obtain that

mlim→∞Mx0,u(rm)=

B1

|∇u0|2dx−3 2

∂B1

u20dHn1+ lim

r0+rn1

Br(x0)

xn+χ{u>0}dx

= lim

r0+rn1

Br(x0)

xn+χ{u>0}dx.

ThusMx0,u(0+)0, and equality implies that for eachτ >0,umconverges to 0 in measure in the set{xn> τ}as m→ ∞, and consequentlyu0=0 inRn.

(v) For eachδ >0 we obtain from the monotonicity formula (Theorem3.4), Remark4.2as well as the fact that limxx0Mx,u(r)=Mx0,u(r)forr >0, that

Mx,u(0+)Mx,u(r)+C2

rMx0,u(r)+δ

2Mx0,u(0+)+δ, if we choose for fixedx0firstr >0 and then|xx0|small enough.

(vi) The fact thatu0is a variational solution of (3.1) and satisfies the monotonicity formula in the sense indicated follows directly from the convergence assumption. The proof for the rest of the claim follows by the same argument as in (v). 2

In the two-dimensional case, we identify the possible values ofMx0,u(0+), and classify the blow-up limits atx0in terms of the value ofMx0,u(0+), which leads to the proof of asymptotic homogeneity of the solution.

Theorem 4.5(Two-dimensional case). Letn=2, letube a variational solution of (3.1)satisfying Assumption4.1, letx0Ω be such thatx20=0, and suppose that

r3/2

Br(x0)

x2d|∇χ{u>0}|C0

for allr >0such thatBr(x0)Ω. Then the following hold:

(10)

(i) M(0+)

0,

B1

x2+χ{x:π/6<θ <5π/6}dx,

B1

x2+dx

.

(ii) IfM(0+)=

B1x2+χ{x:π/6<θ <5π/6}dx, then u(x0+rx)

r3/2

√2

3 ρ3/2cos 3

2

min

max

θ,π 6

,

6

π 2

asr→0+, strongly inWloc1,2(R2)and locally uniformly onR2, wherex=cosθ, ρsinθ ).

(iii) IfM(0+)∈ {0,

B1x2+dx}, then u(x0+rx)

r3/2 →0 asr→0+,

strongly inWloc1,2(R2)and locally uniformly onR2.

Proof. Consider a blow-up sequenceumas in Lemma4.4(ii), whererm→0+, with blow-up limitu0. Because of the strong convergence ofumtou0inWloc1,2(R2)and the compact embedding fromBVintoL1,u0is a homogeneous solution of

0=

R2

|∇u0|2divφ−2∇u0u0 dx+

R2

(x2χ0divφ+χ0φ2) dx (4.6)

for anyφC01(R2;R2), whereχ0is the strongL1loc-limit ofχ{um>0}along a subsequence. The values of the function χ0are almost everywhere in{0,1}, and the locally uniform convergence ofumtou0implies thatχ0=1 in{u0>0}.

The homogeneity ofu0and its harmonicity in{u0>0}show that each connected component of{u0>0}is a cone with vertex at the origin and of opening angle 120. Sinceu=0 in{x20}, this shows that{u0>0}has at most one connected component. Note also that (4.6) implies thatχ0is constant in each open connected setG⊂ {u0=0}that does not intersect{x2=0}.

Consider first the case when{u0>0}is non-empty, and is therefore a cone as described above. Letzbe an arbitrary point in{u0>0} \ {0}. Note that the normal to{u0>0}has the constant valueν(z)inBδ(z){u0>0}for some δ >0. Plugging inφ(x):=η(x)ν(z)into (4.6), whereηC01(Bδ(z))is arbitrary, and integrating by parts, it follows that

0=

{u0>0}

−|∇u0|2+x2(1− ¯χ0)

η dH1. (4.7)

Hereχ¯0denotes the constant value ofχ0in the respective connected component of{u0=0}∩ {x2 =0}. Note that by Hopf’s principle,∇u0·ν =0 onBδ(z){u0>0}. It follows therefore thatχ¯0 =1, and hence necessarilyχ¯0=0.

We deduce from (4.7) that |∇u0|2=x2 on {u0>0}. Computing the solutionu0 of the corresponding ordinary differential equation on∂B1yields that

u0(x)=

√2

3 ρ3/2cos 3

2

min

max

θ,π 6

,

6

π 2

, wherex=cosθ, ρsinθ ), and thatM(0+)=

B1x2+χ{x:π/6<θ <5π/6}dxin the case under consideration.

Consider now the caseu0=0. It follows from (4.6) thatχ0is constant in{x2>0}. Its value may be either 0 in which caseM(0+)=0, or 1 in which caseM(0+)=

B1x2+dx.

Since the limit M(0+) exists, the above proof shows that it can only take one of the three distinct values {0,

B1x+2χ{x:π/6<θ <5π/6}dx,

B1x2+dx}. The above proof also yields, for each possible value ofM(0+), the ex- istence of auniqueblow-up limit, as claimed in the statement of the theorem. 2

Under the assumption that the free boundary is locally an injective curve, we now derive its asymptotic behavior as it approaches a stagnation point.

(11)

Fig. 4. Stokes corner. Fig. 5. Full density singularity.

Fig. 6. Left cusp. Fig. 7. Right cusp.

Theorem 4.6(Curve case). Letn=2, letube a weak solution of (3.1)satisfying Assumption4.1, and letx0Ω be such that x20=0. Suppose in addition that {u >0} is in a neighborhood of x0 a continuous injective curve σ:(t0, t0)R2such thatσ=1, σ2)andσ (0)=x0. Then the following hold:

(i) If M(0+)=

B1x2+χ{x:π/6<θ <5π/6}dx, then (cf. Fig.4) σ1(t) =x10 in(t1, t1)\ {0} and, depending on the parametrization, either

tlim0+

σ2(t)

σ1(t)x10= 1

√3 and lim

t0

σ2(t)

σ1(t)x10= − 1

√3, or

tlim0+

σ2(t)

σ1(t)x10= − 1

√3 and lim

t0

σ2(t)

σ1(t)x10= 1

√3. (ii) IfM(0+)=

B1x2+dx, then(cf. Fig.5)σ1(t) =x10in(t1, t1)\ {0},σ1x10changes sign att=0and

tlim0

σ2(t) σ1(t)x10=0.

(iii) IfM(0+)=0, then(cf. Fig.6and Fig.7)σ1(t) =x10in(t1, t1)\ {0}1x10does not change its sign att=0, and

tlim0

σ2(t) σ1(t)x10=0.

(12)

Proof. We may assume thatx10=0. Moreover, for eachyR2we define argy as the complex argument ofy, and we define the sets

L±:=

θ0∈ [0, π]:there istm→0±such that argσ (tm)θ0asm→ ∞ . Step 1:BothL+andLare subsets of{0, π/6,5π/6, π}.

Indeed, suppose towards a contradiction that a sequence 0 =tm→0, m→ ∞exists such that argσ (tm)θ0(L+L)\ {0, π/6,5π/6, π}, letrm:= |σ (tm)|and let

um(x):=u(rmx) rm3/2

.

For eachρ >0 such thatB˜:=Bρ(cosθ0,sinθ0)satisfies

∅ = ˜B

(x,0):xR

(x,|x|/

3):xR ,

we infer from the formula for the unique blow-up limitu0(see Theorem4.5) that the signed measure um(B)˜ →u0(B)˜ =0 asm→ ∞.

On the other hand, um= −rm1/2f

rm3/2um +√

x2H1{um>0},

which implies, sinceB˜∩{um>0}contains a curve of length at least 2ρ−o(1), that 0←um(B)˜ c(θ0, ρ)C1rm1/2 asm→ ∞,

wherec(θ0, ρ) >0, a contradiction. Thus the property claimed in Step 1 holds.

Step 2: It follows thatσ1(t) =0 for all sufficiently smallt =0. Now a continuity argument yields that bothL+ andLare connected sets. Consequently

+:= lim

t0+argσ (t )

exists and is contained in the set{0, π/6,5π/6, π}, and := lim

t0argσ (t )

exists and is contained in the set{0, π/6,5π/6, π}.

Step 3: In the case M(0+)=

B1x2+χ{x:π/6<θ <5π/6}dx, we know now from the formula for u0 that u0(B1/10(

3/2,1/2)) >0 and thatu0(B1/10(−√

3/2,1/2)) >0. It follows that the set {+, }contains both π/6 and 5π/6. But then the sets{+, }and{π/6,5π/6}must be equal, and the fact thatu=0 onx2=0 implies case (i) of the theorem.

Step 4: In the case M(0+)∈ {0,

B1x2+dx}, we have that u0(B1/10(±√

3/2,1/2))=0, which implies that +, ∈ {/ π/6,5π/6}. Thus+, ∈ {0, π}. Using the fact that u=0 onx2=0, we obtain in the case+ = that M(0+)=

B1x2+dx and in the case+=thatM(0+)=0. Together, the last two properties prove case (ii) and case (iii) of the theorem. 2

Remark 4.7. In[23]we used a strong version of the Rayleigh–Taylor condition (which is always valid in the case of zero vorticity) in order to prove that the cusps of case (iii) are not possible. Unfortunately we do not have the Rayleigh–Taylor condition (4.1) in the case with non-zero vorticity, and the method of[23]breaks down here. Still weconjecturethat the cusps in case (iii) arenotpossible when assuming the Rayleigh–Taylor condition.

5. Partial regularity at non-degenerate points

Definition 5.1(Stagnation points).Letu be a variational solution of (3.1). We callSu:= {xΩ:xn=0 andx

{u >0}}the set ofstagnation points.

Throughout the rest of this section we assume thatn=2.

Références

Documents relatifs

This article is organized as follows: in the first section, we present the work of Bouton on the GLDE and his first theorem on the properties of invariants and isobaric polynomials..

The first existence proof for waves whose slope need not be small is due to Krasovskii [6], who applied positive-operator methods to a non-linear integral equation, but

Under the assumption that the fundamental solution of the two- dimensional vorticity equation with initial data αδ is unique, she constructs a family of initial data λu 0 (λx)

This suggests that one may actually be able to obtain Mordell type results for bigger moduli, in the sense that the perfect squares residues appear essentially in bigger numbers,

In Sections 2 and 3, we recall some properties of weighted Sobolev spaces, and some results concerning the Laplace operator and vector potentials in the half-space.. The fourth

a problem the unknown angular velocity co for the appropriate coordinate system in which the solution is stationary can be determined by the equilibrium

We are interested in the spectral element discretization of the Stokes problem in a two- or three-dimensional bounded domain, when provided with boundary conditions on the

We analyze the spectral-Tau method for the Helmholtz équations in section 2 where an optimal error estimate is given. In section 3, we introducé and analyze a new