• Aucun résultat trouvé

Non-analytic perturbations

Dans le document On Landau damping (Page 152-157)

Tdf dx.

Takingk=0 in (12.8) shows that, for anyη∈Rd,

|˜g(t, η)−f˜0(η)|Cδe−2πλ|η|. (12.11) Also, from the non-linear Vlasov equation, for anyη∈Rd we have

˜

g(t, η) = ˜fi(0, η)−

t 0

Td

RdF(τ, x)·∇vf(τ, x, v)e−2iπη·vdv dx dτ

= ˜fi(0, η)−2iπ

l∈Zd

t 0

F(τ, l)·η f(τ,˜ −l, η)dτ.

Using the bounds (12.7) and (12.10), it is easily shown that the above time-integral converges exponentially fast ast!∞, with rateO(e−λt) for anyλ, to its limit

˜

g(η) = ˜fi(0, η)−2iπ

l∈Zd

0

F(τ, l)·ηf˜(τ,−l, η)dτ. (12.12)

By passing to the limit in (12.11) we see that

|˜g(η)−f˜0(η)|Cδe−2πλ|η|, and this concludes the proof of Theorem 2.6.

13. Non-analytic perturbations

Although the vast majority of studies of Landau damping assume that the perturbation is analytic, it is natural to ask whether this condition can be relaxed. As we noticed in Remark 3.5, this is the case for the linear problem. As for non-linear Landau damping, once the analogy with KAM theory has been identified, it is anybody’s guess that the

answer might come from a Moser-type argument. However, this is not so simple, be-cause the “loss of convergence” in our argument is much more severe than the “loss of regularity” which Moser’s scheme allows one to overcome.

For instance, the second-order correctionh2 satisfies

th2+v·∇xh2+F[f1]·∇vh2+F[h2]·∇vf1=−F[h1]·∇vh1.

The action of F[f1] is to curve trajectories, which does not help in our estimates. Dis-carding this effect and solving by Duhamel’s formula and Fourier transform, we obtain, withS=−F[h1]·∇vh1 and2=

Rdh2dv, ˆ

2(t, k) t

0 K0(t−τ, k)ˆ2(τ, k) +2iπ

t 0

l∈Zd

(k−l)W(k−l)ˆ2(τ, k−l)vh1(τ, l, k(t−τ))

+ t

0

S(τ, k, k(t−τ))! dτ.

(13.1)

(The term withK0 includes the contribution ofvf0.)

Our regularity/decay estimates onh1will never be better than those on the solution of the free transport equation, i.e., hi(x−vt, v), where hi=fi−f0. Let us forget about the effect ofK0 in (13.1), replace the contribution of S by a decaying termA(kt). Let us choose d=1 and assume that ˆhi(l,·)=0 if l=±1. For k>0, let us use the long-time approximation

˜hi(−1, k(t−τ)−τ) 1[0,t](τ) c

k+1δkt/(k+1), c=

R

h˜i(−1, s)ds= ˆhi(−1,0).

Note that c=0 in general. Plugging all these simplifications into estimate (13.1) and choosingW(k)=1/|k|1+γ suggests the a-priori simpler equation

ϕ(t, k) =A(kt)+ ckt (k+1)γ+2ϕ

kt k+1, k+1

. (13.2)

Replacingϕ(t, k) byϕ(t, k)/A(kt) reduces toA=1, and then we can solve this equation by power series as in§7.1.3, obtaining

ϕ(t, k)A(kt)e(ckt)1/(γ+2). (13.3) With a polynomial deterioration of the rate, we could use a regularization argument, but the fractional exponential is much worse.

However, our bounds are still good enough to establish decay for Gevrey perturba-tions. Let us agree that a functionf=f(x, v) lies in the Gevrey classGν, ν1, if

|f˜(k, η)|=O(exp(−c|(k, η)|1/ν)) for some c >0;

in particularG1means analytic. (An alternative convention would be to require thenth derivative to beO((n!)ν).) As we shall explain, we can still get non-linear Landau damp-ing if the initial datumfi lies in Gν forν close enough to 1. Although this is still quite demanding, it already shows that non-linear Landau damping is not tied to analyticity or quasi-analyticity, and holds for a large class of compactly supported perturbations.

Theorem 13.1. Let λ>0. Let f0=f0(v)0 be an analytic homogeneous profile such that

n∈Nd0

λn

n!∇nvf0L1(Rd)<∞,

and let W=W(x)satisfy |W(k)|=O(1/|k|2). Assume that condition(L)from§2.2holds.

Let ν∈(1,1+θ)with θ=1/ξ(d, γ),where ξwas defined in (11.6). Let β >0 and α<1/ν.

Then there is ε>0 such that if δ:= sup

k∈Zd η∈Rd

|( ˜fi−f˜0)(k, η)|eλ|η|1/νeλ|k|1/ν+

Td

Rd|(fi−f0)(x, v)|eβ|v|dv dxε,

then as t!+∞ the solution f=f(t, x, v) of the non-linear Vlasov equation on Td×Rd with interaction potential W and initial datum fi satisfies

|f˜(t, k, η)−f˜(η)|=O(δectα) for all (k, η)Zd×Rd and

F(t,·)Cr(Td)=O(δectα) for all r∈N

for some c>0 and some homogeneous Gevrey profile f, where F stands for the self-consistent force.

Remark 13.2. In view of (13.3), one may hope that the result remains true forθ=2.

Proving this would require much more precise estimates, including among other things a qualitative improvement of the constants in Theorem 4.20 (recall Remark 4.23).

Remark 13.3. One could also relax the analyticity off0, but there is little incentive to do so.

Sketch of proof. We first decomposehi=fi−f0, using truncation by a smooth par-tition of unity in Fourier space,

hi= n=0

F−1hiχn) n=0

hni,

where F is the Fourier transform. Each bump function χn should be localized around the domain (in Fourier space)

Dn={(k, η)∈Zd×Rd:nK|(k, η)|(n+1)K},

for some exponentK >1; but at the same timeF−1n) should be exponentially decreas-ing inv. To achieve this, we let

χn= 1Dn∗γ and γ(η) =eπ|η|2.

ThenF−1n)=F−1(1Dn)γ has Gaussian decay, independently ofn; so there is a

uni-form bound on

Td

Rd|hni(x, v)|eβ|v|dv dx for someβ >0.

On the other hand, if (k, η)∈Dn and (k, η)∈D/ n−1∪Dn∪Dn+1, then

|k−k|+|η−η|cnK−1

for somec>0. From this one obtains, after simple computations,

n(k, η)|1(n−1)K|(k,η)|(n+2)K+Cecn2(K−1)ec(|k|2+|η|2). So (with constantsCandc changing from line to line)

|˜hni(k, η)|Ceλ|k|1/νeλ|η|1/ν1(n−1)K|(k,η)|(n+2)K+Cecn2(K−1)ec(|k|+|η|) Cmax"

eλ(n−1)K/ν/2, ecn2(K−1)$

e¯λn(|k|+|η|), where

λ¯n12λ(n+2)−(1−1/ν)K. IfK2 then 2(K−1)>K/ν; so

hniY¯λn,¯λnCe−λnK/ν/2.

Then we may apply Theorem 4.20 to get a bound on hni in the space Zˆλn,ˆλn;1 with λˆn=12¯λn, at the price of a constant exp(C(n+2)(1−1/ν)K). Assuming Kν >(1−1/ν)K, i.e.,ν <2, we end up with

hniZˆλn,ˆλn;1=O

e−cnK/ν

, ˆλn=12¯λn. (13.4) Then we run the iteration scheme of§8 with the following modifications: (1) instead ofhn(0,·)=0, choose hn(0,·)=hni, and (2) choose regularity indicesλn∼λˆn which tend to zero asntends to infinity. This generates an additional error term of sizeO

ecnK/ν , and imposes that λn−λn+1 be of order n−[(1−1/ν)K+1]. When we apply the bilinear estimates from §6, we can take ¯λ−λ to be of the same order; soα=αn and ε=εn can be chosen proportional to n−[(1−1/ν)K+1]. Then the large constants coming from the time-response will be, as in§11, of ordernqecnr, withq∈Nandr=[(1−1/ν)K+1]ξ, and the scheme will still converge likeO(ecns) for anys<K/ν, provided thatK/ν >r, i.e.,

ν−1+ν K<1

ξ.

The rest of the argument is similar to what we did in§§10–12. In the end the decay rate of any non-zero mode of the spatial densityis controlled by

n=0

e−cnse−λnt

n=0

e−cns

supn0e−cnse−cn−(1−1/ν)tCe−cts/K, and the result follows sinces/K is arbitrarily close to 1/ν.

Remark 13.4. An alternative approach to Gevrey regularity consists in rewriting the whole proof with the help of Gevrey norms such as

fCνλ= n=0

λnf(n)

n!ν and fFνλ=

k∈Z

e2πλ|k|1/ν|fˆ(k)|,

which satisfy the algebra property for any ν1. Then one can hybridize these norms, rewrite the time-response in this setting, estimate fractional exponential moments of the kernel, etc. The only part which does not seem to adapt to this strategy is§9 where the analyticity is crucially used for the local result.

Remark 13.5. In a more general Cr context, we do not know whether decay holds for the non-linear Vlasov–Poisson equation. Speculations about this issue can be found in [55] where it is shown that (unlike in the linearized case) one needs more than one derivative on the perturbation. As a first step in this direction, we mention that our methods imply a bound like O(δ/(1+t)r−¯r) for times t=O(1/δ), where ¯r is a constant

andr>¯r, as soon as the initial perturbation has normδin a functional spaceWrinvolving rderivatives in a certain sense. The reason why this is non-trivial is that the natural time scale for non-linear effects in the Vlasov–Poisson equation is notO(1/δ), butO(1/√

δ), as predicted by O’Neil [75] and very well checked in numerical simulations [61].(17) Let us sketch the argument in a few lines. Assume that (for some positive constantscandC)

hi= n=0

hni, hniZλn,λn;1Cn

2rn and λn=cn

2n. (13.5)

Then we may run the Newton scheme again choosingαn∼cn/2n,cn=O(δ2−(rr1)n) and εn=cδ. Over a time-interval of lengthO(1/δ), Theorem 7.7 (ii) only yields a multiplica-tive constantO(ecδt9n)=O(210n). In the end, after Sobolev injection again, we recover a time-decay on the forceF like

δ n=0

2nr22nreλntsup

n0(2n(rr3)eλnt) (1+t)rr4,

as desired. Equation (13.5) means thathi is of sizeO(δ) in a functional spaceWrwhose definition is close to the Littlewood–Paley characterization of a Sobolev space with r derivatives. In fact, if the conjecture formulated in Remark 4.23 holds true, then it can be shown thatWrcontains all functions in the Sobolev spaceWr+r0,2 satisfying an adequate moment condition, for some constantr0.

Dans le document On Landau damping (Page 152-157)