• Aucun résultat trouvé

NEW EXAMPLES OF PROBABILISTIC WELL-POSEDNSS FOR NONLINEAR WAVE EQUATIONS

N/A
N/A
Protected

Academic year: 2021

Partager "NEW EXAMPLES OF PROBABILISTIC WELL-POSEDNSS FOR NONLINEAR WAVE EQUATIONS"

Copied!
38
0
0

Texte intégral

(1)

HAL Id: hal-02064322

https://hal.archives-ouvertes.fr/hal-02064322

Preprint submitted on 11 Mar 2019

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

NEW EXAMPLES OF PROBABILISTIC WELL-POSEDNSS FOR NONLINEAR WAVE

EQUATIONS

Chenmin Sun, Nikolay Tzvetkov

To cite this version:

Chenmin Sun, Nikolay Tzvetkov. NEW EXAMPLES OF PROBABILISTIC WELL-POSEDNSS FOR NONLINEAR WAVE EQUATIONS. 2019. �hal-02064322�

(2)

NONLINEAR WAVE EQUATIONS

CHENMIN SUN, NIKOLAY TZVETKOV

Abstract. We consider fractional wave equations with exponential or arbitrary poly- nomial nonlinearities. We prove the global well-posedness on the support of the cor- responding Gibbs measures. We provide ill-posedness constructions showing that the results are truly super-critical in the considered functional setting. We also present a result in the case of a general randomisation in the spirit of the work by N. Burq and the second author.

1. Introduction

Our goal in this work is to give new examples of probabilistic well-posedness for non- linear wave equations with data of super-critical regularity. More precisely, we consider fractional wave equations with exponential or arbitrary polynomial nonlinearities. We will prove the global well-posedness on the support of the corresponding Gibbs measures (and also a result for more general random initial data). We will also provide ill-posedness constructions showing that the considered problem is super-critical in the sense that the obtained solutions crucially depend on the particular regularisations of the initial data.

Let us recall that in the case of a deterministic low regularity well-posedness for dispersive PDE’s, the obtained solutions can be seen as limits of approximated smooth solutions, independently of the choice of the approximation of the low regularity initial data (see e.g.

[13, 16, 17]).

1.1. The case of exponential nonlinearity. Let (M, g) be a compact smooth riemann- ian manifold of dimensiondwithout boundary. Let ∆gbe the associated Laplace-Beltrami operator. For σ ∈ R, we set Dσ = (1−∆g)σ/2.Consider the following (fractional) wave equation

(1.1) ∂t2u+Du+eu = 0,

whereu :R×M → R. The case α = 1 corresponds to the usual wave (or Klein-Gordon) equation posed on M. Let u be a smooth solution of (1.1). If we multiply (1.1) by ∂tu and integrate over M, we get

d dt

hZ

M

1

2(∂tu)2+1

2(Dαu)2+eui

= 0.

Let (ϕn)n≥0 be an orthonormal basis ofL2(M) of eigenfunctions of −∆g associated with increasing eigenvalues (λ2n)n≥0. By the Weyl asymptotics λn ≈nd−12 . With v =∂tu, we rewrite (1.1) as the following first order system:

(1.2) ∂tu=v, ∂tv=−Du−eu.

1

(3)

The system (1.2) is a Hamiltonian system of PDEs with the Hamiltonian:

(1.3) H(u, v) = 1

2 Z

M

(Dαu)2+v2 +

Z

M

eu.

The Hamiltonian H(u, v) controls the Hα norm of (u, v), where we denote Hs≡Hs(M)×Hs−α(M)

for any s∈R, and Hs(M) is the classical Sobolev space of orders. Forα≥ d2,

Z

M

eu

≤CeCkuk2Hα(M)

and therefore for these values ofα the potential energy can be seen as a perturbation. In particular we can show the global well-posedness of (1.2) for data inHα.

Theorem 1.1. Letα≥ d2. The Cauchy problem associated with (1.2)is (deterministically) globally well-posed for data in Hs, s≥α.

Let (gn, hn)n≥0 be a family of independent standard gaussians on the probability space (Ω,F,P). The gaussian measure µ is the image measure under the map ω 7→ (uω, vω) defined by

(1.4) uω(x) =

X

n=0

gn(ω)

niαϕn(x), vω(x) =

X

n=0

hn(ω)ϕn(x).

We can see µ as a probability measure on Hs(M), σ < α−d/2. One has the following key property ofµ.

Proposition 1.2. Let α > d/2. For θ ∈ [0, α−d/2), kDθukL(M) is finite µ-almost surely. More precisely

µ{u: kDθukL(M)> λ} ≤C e−cλ2 for some C, c >0 independent of λ≥1.

Applying Proposition 1.2 with θ= 0 we obtain that R

Meu is finiteµalmost surely and we can define the Gibbs measure ρ associated with (1.2) as

dρ(u, v) =eRMeudµ(u, v). Indeed, using that ifu=P

ncnϕn then 1

2 Z

M

(Dαu)2=

X

n=0

nic2n,

we deduce that one may interpretµ as a renormalisation of the formal measure Z−1e12

R

M((Dαu)2+v2) du dv

and therefore ρ becomes Z−1e−H(u,v)dudv which reminds the Gibbs measure for finite dimensional systems. We have the following result.

(4)

Theorem 1.3. Let α > d2 and 0 < σ < α− d2. The problem (1.2) is µ almost surely globally well-posed. Moreover ρ is invariant under the resulting flow Φ(t) in the following sense. There exists a measurable set Σ⊂ Hσ with full µ measure, such that Φ(t)Σ = Σ and for any measurable set A⊂Σ, we have ρ(A) =ρ(Φ(t)A) for allt∈R.

The main difficulty in Theorem 1.3 comes from the fact that the random data (1.4) is not in the scope of applicability of the deterministic well-posedness result of Theorem 1.1.

We however have the following connection between Theorem 1.1 and Theorem 1.3.

Theorem 1.4. Let (uωN(t, x), vωN(t, x))be the solution of (1.2)given by Theorem 1.1 with smooth data

uωN(x) = X

λn≤N

gn(ω)

niαϕn(x), vNω(x) = X

λn≤N

hn(ω)ϕn(x).

Then almost surely in ω, (uωN(t, x), vNω(t, x)) converges to the solution (u, v) of (1.2) con- structed in Theorem 1.3, inC(R;Hσ)∩Lloc(R×M).

The restrictionα > d/2 in Theorem 1.3 and Theorem 1.4 is optimal in the sense that for α=d/2 the construction of the measure ρ fails because R

eu is ill defined on the support ofµ. However, ford= 2 one may suitably renormaliseR

eu. Such a renormalisation would unfortunately lead to a change of the equation. In the case of the renormalisation used in [14] one would obtain the wave equation, without the mass term, but with a source term, related to the curvature ofM. One can also use a renormalisation as in [11] which would avoid the source term in the equation but the mass term should be kept. In the case of both renormalisations we have just mentioned, one can apply compactness techniques as employed in [11, 20]. The obtained solutions would however be non unique and an approximation result as the one of Theorem 1.4 is completely out of reach of the scope of the applicability of these weak solution techniques. We believe that obtaining a result as Theorem 1.4 in the case α= 1 (d= 2) for the above mentioned renormalised equations is an interesting and challenging problem.

A novelty in the proof of Theorem 1.3 and Theorem 1.4 compared to [2, 3, 9, 10, 25]

is that because of the exponential nonlinearity, we need to prove probabilistic Strichartz estimates involvingL norms with respect to the time variables.

1.2. The case of arbitrary polynomial nonlinearity. The strategy to prove Theo- rem 1.3 and Theorem 1.4 works equally well for power type nonlinearity of an arbitrary degree, as follows

(1.5) ∂t2u+Du+u2k+1= 0.

We have the following statement in the context of (1.5).

Theorem 1.5. Let α > d2 and 0 < σ < α−d2. Then (1.5) is µ almost surely globally well-posed.

Note that from the scaling consideration: u(t, x) 7→ λαku(λαt, λx), the critical index is sc = d2αk. Therefore, if s < d2αk, (1.5) is super-critical with respect to Hs. As a

(5)

consequence, we have an ill-posedness result, Proposition 6.1, proved in Section 6.

We underline that Theorem 1.5 is really a super-critical result, in the sense that the way to approximate the solution inC(R;Hσ) by smooth solutions is very sensitive. More precisely, as a consequence of Theorem 1.5 and Proposition 6.1, we have the following remarkable statement.

Corollary 1.6. Assume that (2k−1)d

2k+ 1 < α < kd

2 , k∈N, and d

2− 2α

2k−1 ≤σ < α−d 2.

• For almost every (uω0, uω1)∈ Hσ, there exists a sequence uωN(t, x)∈C(R;C(M)), N = 1,2,· · · of global solutions to (1.5) such that

N→∞lim k(uωN(0), ∂tuωN(0))−(uω0, uω1)kHσ →0, while for every T >0,

N→∞lim kuωn(t)kL([0,T];Hσ(M))→ ∞.

• Let (uωN(t, x), vNω(t, x)) be the solution of (1.5) with smooth data uωN(x) = X

λn≤N

gn(ω)

niαϕn(x), vNω(x) = X

λn≤N

hn(ω)ϕn(x).

Then almost surely in ω, (uωN(t, x), vNω(t, x)) converges to the solution (u, v) of (1.5) constructed in Theorem 1.5, in C(R;Hσ)∩Lloc(R×M).

Remark 1.7. For fixed k∈N, if

(2k−1)d

2k+ 1 < α < kd 2 , then

d

2 − 2α 2k−1,d

2 −α k

0, α−d 2

6=∅.

Remark 1.8. The first assertion of this corollary will follow from the strong ill-posedness result of Proposition 6.1. Unlike usual ill-posedness construction near the zero initial data, we prove norm-inflation near any smooth data of arbitrary size. The restriction α > (2k−1)d2k+1 here is only a technical assumption (see case 2 in the proof of Lemma 6.5 for detailed discussion). It would be interesting to decide whther the same conclusion holds for the full range d2 < α < kd2 .

(6)

1.3. General randomisations. We remark that for the polynomial nonlinearity, if the underlying manifold M = Td, we could also treat the general randomization introduced in [9]. More precisely, for any (u0, v0)∈Hs(Td)×Hs−1(Td),

u0(x) =a0+ X

n∈Zd\{0}

(an,1cos(n·x) +an,2sin(n·x)),

v0(x) =b0+ X

n∈Zd\{0}

(bn,1cos(n·x)x+bn,2sin(n·x)), we consider the randomization around (u0, v0):

uω0(x) =a0g0(ω) + X

n∈Zd\{0}

(an,1gn,1(ω) cos(n·x) +an,2gn,2(ω) sin(n·x)), v0ω(x) =b0h0(ω) + X

n∈Zd\{0}

(bn,1hn,1(ω) cos(n·x)x+bn,2hn,2(ω) sin(n·x)), (1.6)

where {g0(ω), gn,j(ω), h0(ω), hn,j(ω) : n ∈ Zd\ {0}, j = 1,2} are independent standard Gaussian variables. Denote by

uω0,N(x) =a0g0(ω) + X

|n|≤N,n6=0

(an,1gn,1(ω) cos(n·x) +an,2gn,2(ω) sin(n·x)), v0,Nω (x) =b0h0(ω) + X

|n|≤N,n6=0

(bn,1hn,1(ω) cos(n·x)x+bn,2hn,2(ω) sin(n·x)), We have the following almost surely global existence as well as uniqueness theorem:

Theorem 1.9. Assume that M = Td, α > d2. Let (u0, v0) ∈ Hs with (k−1)αk < s <

α. Then almost surely in ω ∈ Ω, (1.5) with initial data (uω0, v0ω) is globally well-posed.

Moreover, the sequence of smooth solutions uN(t) to (1.5) with initial datum (uω0,N, v0,Nω ) converges in C(R;Hs(Td)×Hs−1(Td)) to the solutionuω(t) with initial data (uω0, v0ω).

We will sketch the proof in Section 7. The only additional ingredient in the proof of Theorem 1.9 is an energy a priori estimate, following the method of Oh-Pocovnicu [21]

(see also [23]). The crucial fact we use to prove the energy estimate is the almost sure Lbound for the linear evolution of the Gaussian random initial data. This is the reason to restrict our consideration to M =Td in Theorem 1.9. For general randomizations on arbitrary manifold, the almost sureLbound does not always hold true (see for example [1]). However, as in [22], using the idea of Burq-Lebeau [6], such an L bound can be achieved by imposing some assumptions on the variations of the Fourier coefficients of (u0, v0).

It is worth mentioning that there are many situations when the energy method of Oh- Pocovnicu does not cover the results obtained by exploiting the Gibbs measure. Indeed, for general randomizations, we needs→α ask→ ∞while for data on the support of the Gibbs measure we need 0< s < α−d2,independentlyof k.

(7)

Acknowledgement. The authors are supported by the ANR grant ODA (ANR-18-CE40- 0020-01). The problem considered in this paper is inspired by a talk of Vincent Vargas at the University of Cergy-Pontoise and by a discussion of the second author with R´emi Rhodes and Vincent Vargas at ENS Paris.

2. Construction of the Gibbs measure

2.1. Notations. Denote by ΠN the sharp spectral projector onEN = span{ϕnn≤N}.

LetπN be a smooth projector where πN

X

n=0

cnϕn

:=

X

n=0

ψ(λn/N)cnϕn,

where ψ ∈ C0(R) ψ(r) ≥ 0 andψ(r) ≡1 if r ≤1/2, ψ(r) ≡ 0 if r > 1. By convention π= Id. Clearly, we have

πNΠNN, πNΠN/2= ΠN/2. We will also use the notations:

πN := Id−πN, ΠN := Id−ΠN.

2.2. Definition of Gibbs measure. Denote by µeN the distribution of the EN ×EN valued random variable

ω7→

 X

λn≤N

gn(ω)

niαϕn(x), X

λn≤N

hn(ω)ϕn(x)

.

Consider the Gaussian measureµeN induced by this map, which is the probability measure on R2 dim(EN) defined by

deµN = Y

λn≤N

niα 2π e

2 n+1)αa2

n 2 b2n

2 dandbn= 1 ZN

e−H0(an,bn) Y

λn≤N

dandbn, whereH0 is the free Hamiltonian

H0(u, v) = 1 2

Z

M

(|Dαu|2+|v|2).

Now we define a Gaussian measure on Hσ(M)(σ < α−d/2) be the induced probability measure by the map

ω 7→

X

n=0

gn(ω) hλniαϕn(x),

X

n=0

hn(ω)ϕn(x)

! .

The measureµcan be decomposed intoµ=µN⊗µeN for allN, whereµN is the distribution of the random variable onEN×EN. Now we define the Gibbs measure ρ by

dρ(u) = exp

− Z

M

eu

dµ(u).

We denote by

FN(u) = Z

M

eπNu, F(u) = Z

M

eu

(8)

and

G(u) = exp

− Z

M

eu

, GN(u) = exp

− Z

M

eπNu

. To be precise, we firstly define its finite dimensional approximations

(2.1) dρeN(u) =GN(u)dµeN(u), dρN(u) =GN(u)dµ(u) =dµN⊗dρeN. The following proposition justifies our definition ofdρ(u).

Proposition 2.1. We have the following statements :

(1) The sequences (FN(u))N≥1 and (GN(u))N≥1 converge to the limits F(u), G(u) in Lp(dµ(u)), 2≤p <∞, respectively. In particular, G(u) exists almost surely with respect to µ.

(2) G(u)−1 = exp R

Meu

is almost surely finite with repsect to µ.

(3) limN→∞ρN(A) =ρ(A), for every Borel set A⊂ Hσ. We need the following Moser-Trudinger type inequality.

Lemma 2.2. There exists C >0 such that for all f ∈Hd2(M), we have Z

M

ef ≤exp

Ckfk2

Hd2(M)

. Proof. From

ef(x)

X

k=0

|f(x)|k k! , we have that

Z

M

ef

X

k=0

kfkkLk(M) k! . Note that the inequality (see [5])

kfkLp(Rd)≤C√ pkfk

Hd2(Rd)

also holds true by replacingRd toM, because its proof is only based on local arguments.

Therefore, we have

Z

M

ef

X

k=0

kk2kfkk

Hd2(M)

k! .

From the simple observation

kk/2

k! ≤ 2k/2 b k/2c!,

we obtain that Z

M

ef ≤exp

Ckuk2

Hd2

.

This completes the proof of Lemma 2.2.

Proof of Proposition 2.1. (1) As mentioned in Remark 3.8 of [24], in order to prove that FN(u), GN(u) converge toF(u), G(u) in Lp(dµ), it will be sufficient to show that

(9)

• kFN(u)kLp(dµ),kGN(u)kLp(dµ) are uniformly bounded.

• FN(u), GN(u) converge to F(u), G(u) in measure.

Now we verify the boundeness inLp(dµ). Note thatGN(u)≤1, we only need to check for FN(u). We write

|FN(u)| ≤vol(M)eNukL(M). Therefore using Proposition 1.2, we write forλ≥1,

µ(u:|FN(u)|> λ)≤µ u:kπNukL(M)> log(λ) C

≤C0e−C(log(λ))2 ≤Clλ−l, for every l ≥1. This proves the uniform in N boundedeness of kFN(u)kLp(dµ) for every p <∞. Next, we claim thatFN(u) converges in measure toF(u). Once this is justified, the convergence in measure for GN(u) would follow automatically sinceGN =e−FN. For N1 ≥N2, we observe that

|FN1(u)−FN2(u)| ≤ Z

M

N1u−πN2u| ·eN1u|+|πN2u|

≤CkπN1u−πN2ukL2(M)exp kπN1ukL(M)+kπN2ukL(M)

. Therefore by Cauchy-Schwarz,kFN1(u)−FN2(u)kL1(dµ) is bounded by

C

N1u−πN2ukL2(M)

L2(dµ)kexp kπN1ukL(M)+kπN2ukL(M)

kL2(dµ). The first factor is clearly going to zero as N1, N2 go to infinity. One can show that the second factor is uniformly bounded, exactly as in the proof of the uniform boundedeness of kFN(u)kLp(dµ) . Therefore (FN(u)) is a Cauchy sequence in L1(dµ) which implies its convergence in mesure. This in turn implies the convergence in measure of the sequence (GN(u)) (GN is a continuous function of FN).

(2) To show thatR

Meu is almost surely finite, it will be sufficient to verify that Eµ

Z

M

eu

<∞.

From the proof of (1), Eµ[eNukL(M)] is uniformly bounded inN. Thus we conclude by the dominated convergence.

(3) It will be sufficient to check that, for all Borel set A⊂Hσ(M), we have

N→∞lim Z

Hσ(M)

1u∈A|GN(u)−G(u)|dµ(u) = 0.

This is a simple consequence ofL2(dµ(u)) convergence, since Z

Hσ(M)

1u∈A|GN(u)−G(u)|dµ(u)≤ kGN(u)−G(u)kL2(dµ)µ(A)1/2 →0,

asN → ∞. This completes the proof of Proposition 2.1.

(10)

3. Probabilistic local well posedness

3.1. Deterministic local well-posedness result. Consider the following truncated ver- sion of (1.2)

(3.1) ∂tu=v, ∂tv=−Du−πN(eπNu), with initial data

(3.2) u|t=t0 =u0, v|t=t0 =v0. Let us next define the free evolution. The solution of

tu=v, ∂tv =−Du, subject to initial data

u(0, x) =u0(x), v(0, x) =v0(x) is given by

S(t)(u¯ 0, v0) = (S(t)(u0, v0), ∂tS(t)(u0, v0)), where

S(t)(u0, v0) = cos(tDα)u0+D−αsin(tDα)v0

and

tS(t)(u0, v0) =−Dαsin(tDα)u0+ cos(tDα)v0. Assume that01, 1r0 <∞ such that

d r0

< 0 < α−d 2,

which is possible sinceα > d2. In order to establish the probabilistic local well-posednesss as well as globalizing the dynamics, we need some auxillary functional spacesXs,β,Yβ,Zβ, definining via the norms

k(u, v)kYβ :=X

l∈Z

(1 +|l|)−β sup

l≤t<l+1

kS(t)(u, v)kW0,r0(M),

k(u, v)kXs,β :=k(u, v)kHs +k(u, v)kYβ

and

k(u, v)kZβ :=X

l∈Z

(1 +|l|)−β sup

l≤t≤l+1

kS(t)(u, v)kL(M). Note that

α−s0

d = 1

2 − 1 r0

withs0=α−d 2 + d

r0

> 0,

thus Hα(M) ⊂W0,r0(M)⊂ L(M) (with continuous inclusions). Moreover, Yβ ⊂ Zβ. The definition of this weighted in time spaces Yβ,Zβ is inspired by the work of N. Burq and the second author (a similar definition is appeared in [9]). The weight β >1 will be fixed in the sequel to ensure the l1 summation, without other importance. These norms are only designed to treat the linear evolution part of the solution, since unlike [10], the linear evolution is not periodic in time.

Denote by ΦN(t) the flow of the truncated equation (3.1),(3.2). In components, we write

ΦN(t)(u0, v0) = (Φ1N(t)(u0, v0),Φ2N(t)(u0, v0))

(11)

with Φ1N(t)(u0, v0) =uN(t),Φ2N(t)(u0, v0) =∂tuN(t). Similarly, we denote by Φ(t)(u0, v0) = (Φ1(t)(u0, v0),Φ2(t)(u0, v0)),

where Φ1(t)(u0, v0) =u(t),Φ2(t)(u0, v0) =∂tu(t) for solutions of the untruncated equation

t2u+Du+eu= 0, (u, ∂tu)|t=0 = (u0, v0).

We also denote the nonlinear evolution part by

W(t)(u0, v0) := Φ1(t)(u0, v0)−S(t)(u0, v0),

WN(t)(u0,N, v0,N) := Φ1N(t)(u0,N, v0,N)−S(t)(u0,N, v0,N).

The next proposition contains the local theory for (3.1) and the original system (1.2).

Proposition 3.1. LetN ∈N∪ {∞}andβ >1. There existc >0andκ >0such that the following holds true. The Cauchy problems (3.1)-(3.2)and (1.2) are locally well-posed for data (u0, v0) such that S(t)(u0, v0) ∈ L(M). More precisely for every R ≥1 if (u0, v0) satisfies

(3.3) k(u0, v0)kZβ < R,

then there is a unique solution of (3.1)-(3.2)on [t0−τ, t0+τ], where

(3.4) τ =ce−κR

which can be written as

(u, v) = ¯S(t−t0)(u0, v0) + (˜u,˜v), with

k(˜u,v)k˜ Hα ≤C .

In particular, the Cauchy problems (3.1)-(3.2) and (1.2) are locally well-posed for data (u0, v0)∈ Xσ,β, for allt∈[t0−τ, t0+τ], provided that σ < α−d2.

Proof. We can write (3.1)-(3.2) as

(u(t), v(t)) = ¯S(t−t0)(u0, v0) + F1(u), F2(u) , where

F1(u) =− Z t

t0

D−αsin((t−τ)DαN eπNu(τ) dτ,

F2(u) =− Z t

t0

cos((t−τ)DαN eπNu(τ) dτ.

If we write (u, v) = ¯S(t−t0)(u0, v0) + (˜u,v), we obtain that (˜˜ u,v) solves˜

tu˜= ˜v, ∂tv˜=−Du˜−πN(eπNu+S(t)(u0,v0))) with zero initial data. Therefore ˜u solves

(3.5) u(t) =˜ F1(˜u+S(t−t0)(u0, v0)).

Once we solve (3.5), we recover ˜v by ˜v=∂tu. Define the map Φ˜ u0,v0 by Φu0,v0(u) :=F1(u+S(t−t0)(u0, v0)).

(12)

It follows from the definition that

u0,v0(u)kL([t0,t0+T];Hα(M))≤CTkeπN(u+S(t−t0)(u0,v0))kL(I;L2(M))

≤CT(vol(M))12keπN(u+S(t−t0)(u0,v0))kL(I;L(M))

(3.6)

where I = [t0, t0 +T]. Since πN is bounded on L (see e.g. [4]), a use of the Sobolev embedding Hα(M)⊂L(M) yields the estimate

u0,v0(u)kL([I;Hα(M))≤CT eCkS(t−t0)(u0,v0))kL(I×M)ekukL(I;Hα(M)). Note that

kS(t−t0)(u0, v0)kL(I×M)≤Ck(u0, v0)kZβ, hence under (3.3) we deduce that

(3.7) kΦu0,v0(u)kL(I;Hα(M))≤CT eCRekukL(I;Hα(M)). Similarly we obtains that

u0,v0(u)−Φu0,v0(v)kL(I;Hα(M))

≤CT eCRexp kukL(I;Hα(M))+kvkL(I;Hα(M))

ku−vkL(I;Hα(M)). (3.8)

Define the space XT as

XT ={u∈C(I;Hα(M)) : kukL(I;Hα(M))≤1}.

Using (3.7), we obtain that forT as in (3.4) the map Φu0,v0 enjoys the property Φu0,v0(XT)⊂XT.

Under the same restriction on T, thanks to (3.8), we obtain that the map Φu0,v0 is a contraction on XT. The fixed point of this contraction is the solution of (3.5). This

completes the proof of Proposition 3.1.

In the same spirit of the proof, we establish a local convergence result, which will be needed to construct global dynamics in Section 4.

Lemma 3.2. Assume that 0 < σ < α− d2 and β >1. There exist R0 >0, c > 0, κ > 0 such that the following holds true. Consider a sequence (u0,Np, v0,Np) ∈ ENp ×ENp and (u0, v0)∈ Xσ, withNp→ ∞. Assume that there existsR > R0 such that

k(u0,Np, v0,Np)kXσ,β ≤R, k(u0, v0)kXσ,β ≤R, and

p→∞lim kπNp(u0,Np, v0,Np)−(u0, v0)kHσ = 0.

Then if we set τ = ce−κR, the flow ΦNp(t)(u0,N0, v0,Np),Φ(t)(u0, v0) exist for t ∈ [−τ, τ]

and satisfy

Np(t)(u0,Np, v0,Np)kL([−τ,τ];Xσ,β) ≤R+ 1, kΦ(t)(u0, v0,)kL([−τ,τ];Xσ,β)≤R+ 1.

Furthermore,

p→∞lim kπNpΦNp(t)(u0,Np, v0,Np)−Φ(t)(u0, v0)kL([−τ,τ];Hσ)= 0, and

p→∞lim kπNpWNp(t)(u0,Np, v0,Np)−W(t)(u0, v0)kL([−τ,τ];Hα(M))= 0.

(13)

Proof. The existence of ΦNp(t)(u0,Np, v0,Np) and Φ(t)(u0, v0) as well as the bound on [−τ, τ]

are guranteed by the local well-posedness result, Propsotion 3.1. We only need to prove the convergence.

Denote by

(up(t), ∂tup(t)) = ΦNp(t)(u0,Np, v0,Np), (u(t), ∂tu(t)) = Φ(t)(u0, v0).

From local theory, we can write

up(t) =S(t)(u0,Np, v0,Np) +wp(t), u(t) =S(t)(u0, v0) +w(t), such that

k(wp, ∂twp)kL([0,τ];Hα) ≤R+ 1, k(w, ∂tw)kL([0,τ];Hα) ≤R+ 1.

The convergence of the linear part

NpS(t)(u0,Np, v0,Np)−S(t)(u0, v0)kL([0,τ];Hα)= 0 follows from the assumption and the boundeness of πNp on W0,r0(M).

Next we estimate the nonlinear part

Npwp(t)−w(t)kHα(M).

Writting wp, w by the Duhamel formula and using triangle inequality, we can bound the quantity above by the three contributions:

I(t) :=

Z t 0

eπNp(wp(t0)+S(t0)(u0,Np,v0,Np))−eπNp(w(t0)+S(t0)(u0,v0))

L2(M)dt0, II(t) :=

Z t

0

ew(t0)+S(t0)(u0,v0)−eπNp(w(t0)+S(t0)(u0,v0))

L2(M)dt0, III(t) :=

Z t 0

πNew(t0)+S(t0)(u0,v0)

L2(M)dt0. Since

ew(t)+S(t)(u0,v0)∈L([−τ, τ]×M), we have that

kIIIkL([−τ,τ])=op→∞(1).

For I(t) and II(t), we can bound them by C(R, τ)

op→∞(1) + Z t

0

Npwp(t0)−w(t0)kL2(M)dt0

.

By applying Grwonwall inequality, the proof of Lemma 3.2 is complete.

3.2. Large deviation estimate for linear evolutions. Proposition 3.1 is deterministic.

The probabilistic part of the analysis comes from the following statement.

Proposition 3.3. Assume that β >1. There are positive constantsC andcsuch that for every R≥1,

µ (u0, v0) : k(u0, v0)kYβ ≥R

≤C e−cR2. As a consequence, we also have a similar bound for the Gibbs measure

ρ (u0, v0) : k(u0, v0)kYβ ≥R

≤C e−cR2.

(14)

Observe that in the case of an exponential nonlinearity, we need a large deviation estimate for L norms in time (in [9] onlyLp in time for a finite p were established).

Proof of Proposition 3.3. Letη∈C0be a bump function localising in the interval [−2,2].

Denote by ηl(t) :=η(t−l) for l∈Z. We need to show

X

l∈Z

hli−βl(t)S(t)(u0, v0)kL(R;W0,r0(M)

Lp(dµ)

≤C√ p .

Coming back to the definition of S(t), we observe that it suffices to prove the bound

X

l∈Z

hli−β ηl(t)

X

n=0

cngn(ω)

niα eitλαnϕn(x)

L(R;W0,r0(M)

Lp(Ω)

≤C√ p ,

where |cn|is bounded. Thanks to β >1, we have P

lhli−β <∞, thus it suffices to prove the bound

ηl(t)

X

n=0

cngn(ω)

niα eitλαnϕn(x)

L(R;W0,r0(M))

Lp(Ω)

≤C√ p,

uniformly in l∈Z.

To simplify the notation, we only write down the case where l = 0. For other l, the arguments are exactly the same. Using the Sobolev embedding with a largeq and a small δ (depending on q) we obtain that its is sufficient to obtain

DδtD0 η(t)

X

n=0

cngn(ω)

niα eitλαnϕn(x) Lq(

R;Lr0(M)

Lp(Ω)≤C√ p ,

whereDδt = (1−∂t2)δ/2. Since Dδϕn=hλniδϕn, we are reduced to prove the bound

X

n=0

cngn(ω)

niα−δαn(t)ϕn(x)

Lq(R;Lr0(M)

Lp(Ω)

≤C√ p ,

where αn(t) =Dδt(η(t) exp(itλαn)).Applying the Minkowski inequality, we deduce that it suffices to prove that forp≥q ≥r0,

X

n=0

cngn(ω)

niα−δαn(t)ϕn(x) Lp(Ω)

Lr0(M;Lq(R))

≤C√ p .

Observe that in this discussionqis large but fixed andpgoes to∞. Now for a fixed (t, x), we can apply the Khinchin inequality and write

X

n=0

cngn(ω)

niα−δαn(t)ϕn(x)

Lp(Ω) ≤C√ p

X

n=0

n(t)|2n(x)|2ni2(α−δ)

12 . Therefore, we reduce the matters to the deterministic bound

X

n=0

n(t)|2n(x)|2ni2(α−δ)

Lr0/2(M;Lq/2(R)

≤C .

(15)

For a fixed x, we apply the triangle inequality to obtain that (3.9)

X

n=0

n(t)|2n(x)|2ni2(α−δ)

Lq/2(R)

X

n=0

n(x)|2

ni2(α−δ)nk2Lq(R). Lemma 3.4. There is C such that for every n≥1, kαnkLq(R) ≤Cλαδn . Proof. Letβ(t) =η(t)eitλαn, we have that

nkLq(R) ≤CkβkLq(R)+CkβkB˙δ q,2(R). From the characterization of Besov spaces (see [5]), we have that

kβk2˙

Bq,2δ ∼ Z

R

Z

R

|β(t+τ)−β(t)|qdt 2q

|τ|1+2δ. We write

η(t) exp(itλαn)−η(τ) exp(iτ λαn) = exp(itλαn)(η(t)−η(τ)) +η(τ)(exp(itλαn)−exp(iτ λαn)) which yields two contributions. The first one is again uniformly bounded. Therefore the issue is to check that

Z

R

Z

R

|η(t)(ei(t+τ)λαn −eitλαn)|qdt 2q

|τ|1+2δ ≤(Cλαδn )2. For|τ| ≤cλ−αn , we use

|eiτ λαn −1| ≤ |τ|λαn, and thus

Z

|τ|<cλ−αn

1

|τ|1+2δ Z

R

|η(t)(ei(t+τ)λαn−eitλαn)|qdt 2q

≤ Z

|τ|<cλ−αn

1

|τ|1+2δ Z

R

|η(t)|q|τ|qλαqn 2q

dτ ≤Cλ2αδn . The other contribution for |τ|> cλ−αn can be bounded by

Z

|τ|>cλ−αn

Z

R

|2η(t)|qdt 2q

dτ ≤Cλ2αδn . This completes the proof of Lemma 3.4.

Coming back to (3.9) and using Lemma 3.4, we deduce that it suffices to majorize

X

n=0

n(x)|2ni2(α−δ−αδ)

Lr0/2(M).

Using the compactness of M, it suffices to get the following estimate sup

x∈M

X

n=0

n(x)|2

ni2(α−δ−αδ) <∞.

(16)

Fix δ sufficiently small such that β := 2(α−δ−αδ) > d (here we fix the value of q as well). Therefore we need to show that

(3.10) sup

x∈M

X

n=0

n(x)|2

niβ <∞, β > d .

Estimate (3.10) is direct if|ϕn(x)|are uniformly bounded (this is the case of the torus). In the case of a general manifold, it is not true that|ϕn(x)|are uniformly bounded. However (3.10) is true thanks to [15]. More precisely for a dyadic N, we can write

X

N≤hλni≤2N

n(x)|2

niβ ≤CN−β X

N≤hλni≤2N

n(x)|2.

Thanks to [15] there is C such that for every dyadic N ≥1 and every x∈M,

(3.11) X

N≤hλni≤2N

n(x)|2 ≤CNd.

This readily implies (3.10). The proof of Proposition 3.3 is completed.

We complete this section by proving following probabilistic bound for the tails of sharp spectral truncation, which will be used in the proof of Theorem 1.4.

Proposition 3.5. Assume that 2≤r0 <∞, 0 < α−d2, 0< s < α−d20 and β >1.

Then there exist C >0, c >0,such that for any R >0, we have µ

(u0, v0) :NsN(u0, v0)kYβ > R

≤Ce−cR2.

Proof. Following the same notations as in the proof of the Proposition 3.3. It suffices to prove that

(3.12)

η(t) X

λn≥N

cngn(ω)

niα eitλαnϕn(x)

L(R;W0,r0(M))

Lp(Ω)

≤CN−s√ p,

forp large enough. From Sobolev embedding, we are reduced to prove the bound (3.13)

X

λn≥N

cngn(ω)

niα−0αn(t)ϕn(x)

Lq(R;Lr0(M))

Lp(Ω)

≤CN−s√ p,

whereαn(t) =Dtδ(η(t)eitλαn). From Minkowski inequality, it will be sufficient to prove that forp≥q, p≥r0,

(3.14)

X

λn≥N

cngn(ω)

niα−0αn(t)ϕn(x) Lp(Ω)

Lq(

R;Lr0(M))

≤CN−s√ p

Références

Documents relatifs

In view of the result of Kappeler and Topalov for KdV it thus appears that even if the dissipation part of the KdV-Burgers equation allows to lower the C ∞ critical index with

The aim of this paper is to establish positive and negative optimal results on the local Cauchy problem in Sobolev spaces for the Korteweg-de Vries- Burgers (KdV-B) equation posed

Sharp well-posedness results for the generalized Benjamin- Ono equation with high nonlinearity, 2007.

The Cauchy problem for dissipative Ko- rteweg de Vries equations in Sobolev spaces of negative order. On the low regularity of the Korteweg-de

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

The key point is that we cut the initial data into two parts : the first part being regular enough to apply theorem 3, and the second one being H ˙ 1 2 with small initial data, in

To overcome the difficulties caused by the weakly dispersive effect, our specific strategy is to benefit from the “almost” transport effect of these singular parts and to exploit

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