• Keine Ergebnisse gefunden

On Linear Damping Around Inhomogeneous Stationary States of the Vlasov-HMF Model

N/A
N/A
Protected

Academic year: 2022

Aktie "On Linear Damping Around Inhomogeneous Stationary States of the Vlasov-HMF Model"

Copied!
47
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

https://doi.org/10.1007/s10884-021-10044-y

On Linear Damping Around Inhomogeneous Stationary States of the Vlasov-HMF Model

Erwan Faou1 ·Romain Horsin1·Frédéric Rousset2 Dedicated to the memory of Walter Craig

Received: 31 October 2020 / Revised: 18 May 2021 / Accepted: 30 June 2021 / Published online: 6 August 2021

© The Author(s), under exclusive licence to Springer Science+Business Media, LLC, part of Springer Nature 2021

Abstract

We study the dynamics of perturbations around an inhomogeneous stationary state of the Vlasov-HMF (Hamiltonian Mean-Field) model, satisfying a linearized stability criterion (Penrose criterion). We consider solutions of the linearized equation around the steady state, and prove the algebraic decay in time of the Fourier modes of their density. We prove moreover that these solutions exhibit a scattering behavior to a modified state, implying a linear damping effect with an algebraic rate of damping.

Keywords Vlasov equations·Damping effects·HMF model·Hamiltonian systems· Angle-action variables

Mathematics Subject Classification 35Q83·35P25

1 Introduction

In this paper we consider the Vlasov-HMF (Hamiltonian Mean-Field) model. It is an ideal toy model that keeps several features of more complex kinetic equations, such as the Vlasov–Poisson system. It is moreover rather easy to do numerical simulations and ana- lytic calculations on it, and it has thus received much interests in the physics literature (see

This work was partially supported by the ERC starting Grant GEOPARDI No. 279389.

B

Erwan Faou

Erwan.Faou@inria.fr Romain Horsin Romain.Horsin@inria.fr Frédéric Rousset

frederic.rousset@universite-paris-saclay.fr

1 INRIA-Rennes Bretagne Atlantique and IRMAR (UMR 6625), Université de Rennes I, Rennes, France

2 Laboratoire de Mathématiques d’Orsay (UMR 8628), Université Paris-Saclay, Gif-sur-Yvette, France

(2)

[2–6,13–19]). This model exhibits also analogies with the Kuramoto model of coupled oscil- lators in its continuous limit [11,21,23]. A long time analysis of the Vlasov-HMF model around spatially homogeneous stationary states has been performed in [22], where a nonlin- ear Landau damping result is proved in Sobolev regularity. In this paper, we consider the case of inhomogeneous steady states and study the long time behavior of the linearized equation.

The Vlasov-HMF equation, with an attractive potential, reads

tf(t,x, v)+ {f,H[f]}(t,x, v)=0, H[f](t,x, v)= v2

2 −φ[f](t,x), φ[f](t,x)=

T×Rcos(x−y)f(t,y, v)dydv,

(1.1)

with(t,x, v)∈R×T×R,whereT=R/Z,and where

{f,g} =xf∂vgvf∂xg (1.2) is the Poisson bracket. It is rather easy to prove that this equation is globally well-posed, in Sobolev regularity for instance, using standard tools for transport equations associated with a divergence-free vector field.

The potential can be also expressed as the following trigonometric polynomial

φ[f](x)=C[f]cos(x)+S[f]sin(x), (1.3) with

C[f] =

T×Rcos(y)f(y, v)dydv and S[f] =

T×Rsin(y)f(y, v)dydv, where we use the normalized Lebesgue measure on the torus.1

The previous equation possesses stationary solutions of the form η(x, v)=G(h0(x, v)) , h0(x, v)= v2

2 −M0cos(x), M0>0. (1.4) for some functionG:R→R. The constantM0(called the magnetization) has to fulfill the condition

M0=C

G v2

2 −M0cos(x)

. (1.5)

Up to translationxx+x0, these are essentially the only stationary solutions, see Sect.6 where examples of couple(M0,G)satisfying the previous condition and the necessary sta- bility condition ensuring damping effects are studied.

For such a stationary states, if we seek solution of (1.1) under the form f(t,x, v) = η(x, v)+r(t,x, v)with initial conditionr(0,x, v)=r0(x, v), we obtain the equation

tr(t,x, v)− {η, φ[r]}(t,x, v)+ {r,H[η]}(t,x, v)− {r, φ[r]}(t,x, v)=0.

In this paper, we will retain the linear part of this equation, namely the linearized equation aroundη, given by

tr(t,x, v)− {η, φ[r]}(t,x, v)+ {r,h0}(t,x, v)=0. (1.6)

1The Lebesgue measure on[−π, π]divided by the length of the torus 2π.

(3)

The goal of this paper is the analysis of the long time behavior of this equation.

In the homogeneous caseM0 =0 where the steady steady statesηdepends only on the velocity variablev, the situation both for the linear and nonlinear equation has been widely studied for general Vlasov equations. In this case,h0(x, v)= v22 the flow of the Hamilto- nianh0is trivially calculated: without the potential termφ[r]in (1.6), the solution is given explicitly byr(t,x, v)=r0(xtv, v)and gives rise to damping effect (see Landau [30]) implying a weak convergence ofr(t,x, v)towards the average ofr0with respect tox. It then turns out that under a stability condition onηcalled the Penrose condition, then the flow of the full linear Eq. (1.6) behaves like the transport part for large times. This is well expressed as a scattering result whereg(t,x, v):=r(t,x+tv, v)is shown to converge for large times towards a smooth functiong(x, v)depending onr0(x, v). With this result in hand, the weak limit ofr(t,x, v)can be identified whent→ +∞. The scattering convergence rate depends on the regularity of the solution and is expected to be typically exponential for Gevrey or analytic functions and polynomial in time for finite Sobolev regularity. This linear scattering under a Penrose condition is the starting point of the nonlinear results of [8,24,33] showing that this scattering behavior persists in nonlinear equations. These result were proven for Gevrey initial data and general Vlasov equation including in particular the Vlasov–Poisson system. For the Vlasov-HMF a similar result can be proved under Sobolev regularity in the homogeneous case, see [22]. The question of Landau damping in Sobolev regularity for the Vlasov–Poisson system has been recently addressed, for instance in [7,35], where Landau damping results are proved in a weakly collisional regime, or in [9,10,27,28] in the case of unconfined systems.

In the non-homogeneous caseM0 >0 studied in this paper, the situation has been recently investigated (in particular in the physics literature see [3,4] and in [20]). In this paper, we propose to follow the same strategy as in the homogeneous case where the free flowxx+tvis replaced by the flowψt(x, v)of the Hamiltonianh0,associated with the ordinary differential equation

x˙ =vh0(x, v) =v

˙

v= −∂xh0(x, v)= −M0sin(x), (1.7) which is the classical dynamical system for the motion of a Pendulum. The flowψtis globally well defined and symplectic. In particular it preserves the Poisson bracket

t∈R, {f,g} ◦ψt(x, v)= {fψt,gψt}(x, v).

Note that for a given function f(x, v), the function(t,x, v)f(t,x, v) = f(ψt(x, v)) solves the equationtf = {f,h0}andrψtis the solution of the free flow in (1.6).

In this paper, we shall prove that under appropriate assumptions onη(of Penrose type) andr0(a natural orthogonality condition), the solution of the linear Eq. (1.6) also exhibit a scattering behavior:g=r◦ψtconverges towards a functiong, implying that the coefficients S[r(t)]andC[r(t)]decay in time, with algebraic rates of damping which depends on the regularity of the initial data and on the behavior of the function in the vicinity of the origin (x, v)=(0,0)(the center of the “eye" of the pendulum).

The main ingredient of the proof is the use of action-angle variables(θ,a)for the integrable flow (1.7) and for which the flowψt can be calculatedψt(θ,a) = θ +ω(a)t for some frequency functionω. We will use this “explicit" formula (up to the knowledge of elliptic functions) to prove thatψt gives rise to damping effect and that for smooth functionsϕ, the function of the formϕψt has a weak limit that can be nicely expressed in terms of action-angle variables, and with a convergence rate in timetdepending on the “flatness" of

(4)

ϕand of the observable near the origin. This last particularity reflects the singularity of the action-angle change of variable. The full statement of this result is given in Theorem2.2 which is proven in Sect. 7and requires the use of precise asymptotics of Jacobi elliptic functions. This makes the proof seemingly technical, but the arguments are in fact simple for a reader familiar with this literature (we make a crucial use of many formulas in the book [12]).

With this result in hand, our main results give the decay of the functionsS(t)andC(t) and the convergence ofg = rϕt, (and weak convergence ofr, see Theorem2.4and Corollary 2.5) under an orghogonality assumption forr0 well expressed in action-angle variable, and a Penrose condition onη, (2.13).

We then conclude by showing the existence of stationary statesηi.e. of couple(M0,G) satisfying (1.5)) fulfilling the stability condition and relate it with more classical stability condition from the physics literature [3] that was also used in [31] conditioning the nonlinear orbital stability of the inhomogeneous steady states of Vlasov-HMF. Strikingly enough, the key argument relies on explicit formulae in the action-angle change of variable and the direct verification that some terms do not vanish from known Fourier expansions of elliptic functions that can be found in [12]. Note finally that the extension of our scattering result to the nonlinear case remains for the moment an open question.

To finish this introduction, we would like to stress that the extension of this work to the Vlasov–Poisson equation is probably achievable but would demand certainly some efforts.

Stationary states similar to the pendulum can be found, with a Hamiltonian of the form

v2

2 +V(x)whereVis the solution of a nonlinear partial differential equation (degenerating to (1.5) in the HMF case). Obtaining damping results in this case would require that this Hamiltonian is integrable. If this will always be the case in the 1D case where(x, v)∈T×R, the phase portait of the Hamiltonian is in general more complicated than the pendulum, and the singularities of the action-angle change of variable less documented. As explained above, our analysis relies heavily on explicit formulae given by elliptic functions. In higher dimensions, the situation is even much more delicate. Note however that local results could probably be obtained near elliptic points ofV where the dynamics can be conjugated “generically" to some integrable flow, at least to some order of approximation by using Birkhoff reduction technics.

2 Statements of the Main Results

We now fix the notations and give the main results. The first part shows the dispersive effect of the flow of the Pendulum, and the second part gives the main application for the long time behavior of the linear Eq. (1.6).

2.1 Damping in Action-Angle Variables

As a one-dimensional Hamiltonian system, the system associated with the Hamiltonian h0(x, v)is integrable. We will need relatively precise informations about the correspond- ing action-angle change of variable. Let us split the space into three chartsU+,UandU as follows:

(5)

Fig. 1 Phase portrait of the pendulum

U+= {(x, v)∈T×R|v >0 and h0(x, v) >M0}, U= {(x, v)∈T×R|v <0 and h0(x, v) >M0}, and U= {(x, v)∈T×R|h0(x, v) <M0}.

(2.1)

We have thath0(x, v) ≥ −M0 and the center of the “eye" U corresponds to the point (x, v)=(0,0)which minimizesh0. The set

{(x, v)∈T×R | h0(x, v)=M0} will usually be called the “separatix" (Fig.1).

Let us first recall the following Theorem:

Theorem 2.1 Setting h(x, v)= v22M0cos(x), then for∗ ∈ {±,◦}, there exists a symplectic change of variable(x, v)(ψ,h)from Uto the set

V:= {(ψ,h)∈R2|hI, ψ(−r(h),r(h))},

where r(h)is a positive function, I±=(M0,+∞)and I=(−M0,M0)such that the flow of the pendulum in the variable(ψ,h)is h(t)=h(0)andψ(t)=t+ψ(0).

Moreover, there exists a symplectic change of variables(ψ,h)(θ,a)from Vto W= {(θ,a)∈R2|θ(−π, π),aJ} =T×J,

with J±=(π4

M0,+∞)and J =(0,π8

M0)such that θ(ψ,h)=ω(h)ψ, and ha(h)= 1

ω(h) = π r(h),

so that the flow of the pendulum in the variables(θ,a)in Wis a(t) =a(0)andθ(t)= (a(0))+θ(0).

This Theorem is explicit in the sense that the changes of variables express in terms of Jacobi elliptic functions. Asθ is a variable in a fixed torus, Fourier series in variableθ are

(6)

well defined on each setWcorresponding toU. For a given function f(x, v)we can define the restriction fof f to the setU,and the Fourier coefficients

f(a)= 1 2π

π

−π f(x(θ,a), v(θ,a))eiθdθ, ∈Z, aJ (2.2) wherex(θ,a)andv(θ,a)are given by the change of variable onU. Note that for given functions f andϕ, we have the decomposition

U

f(x, v)ϕ(x, v)dxdv=

∈Z

J

f(a)ϕ(a)da, (2.3) for all∗ ∈ {◦,±}.Finally, let us notice that the Jacobian of the change of variableha(h) isha(h)= ω1(h), and we have in particular

J

f(a)ϕ (a)da=

I

f(a(h))ϕ (a(h)) 1

ω(h)dh. (2.4)

We will usually write f(h)for the quantity f(a(h)), and several times consider functions f as depending on(x, v),(θ,h)and(θ,a)by keeping the same notation. For example a stationnary stateη depends only onh and hence on a and will be writtenη = G(h) or η=G(a).

In fact the singularities of the relevant functions in action-angle variables are better expressed in variables(θ,h), which are not symplectic, but on which integrals and flow of the system are easy to calculate. Moreover, in this case,

f0(a)= 1 2π

π

−π f(x(θ,h), v(θ,h))dθ

can be seen as an average off on the isocurve{(x, v)|h0(x, v)=h}, whileψis the arclength on this curve, the jacobianω1(h)appearing in the standard co-area formula, which is another way to see (2.3)–(2.4).

The notationsC(a)andS(a)will be used for the Fourier coefficients of the functions θ →cos(x(θ,a)) and θ→sin(x(θ,a)),

respectively, and both restricted toU.These coefficients can be calculated explicitly using elliptic functions (see Propositions7.5and7.12), and we shall write

cos(x(θ,a))=

∈Z

C(a)e and sin(x(θ,a))=

∈Z

S(a)e, (2.5) for(θ,a)J×(−π, π).

Before stating our first result, let us fix some notations. We use the classical notation v =(1+|v|2)1/2, for anyv∈Rd, and for a two-dimensional integerα=1, α2)∈N2,we set|α| =α12. We shall also writex,vα for the operator acting on functionsf :T×R→C by the formula

x,vα f(x, v)=xα1vα2f(x, v).

In Sect.7we prove the following result:

Theorem 2.2 (Damping effect of the pendulum flow) Consider f(x, v)andϕ(x, v)two functions such that

|α|≤maxmvμxα,vf(x, v)LCm,μ and max

|α|≤Mxα,vϕ(x, v)LCM,

(7)

for some m, M andμ >2. Let p and q be defined by

p=max{n≥1, ∂xα,vf(0,0)=0,∀α,1≤ |α| ≤n}, q=max{n≥1, ∂x,vα ϕ(0,0)=0,∀α,1≤ |α| ≤n},

with the convention that these number are0is the corresponding sets are empty. Then, if m≥5+p+ p+q

2 and M ≥max

7+q+ p+q 2 ,m+2

, there exists C>0such that for all t≥0,we have

T×R f(x, v)ϕ(ψt(x, v))dxdv−

∗∈{±,◦}

J

f0(a)ϕ0(a)daC

tp+q2 +2.

Let us explain this Theorem as follows: the starting point of the proof is the Fourier expansion (2.3), which yields

T×R f(x, v)ϕ(ψt(x, v))dxdv=

∗∈{±,◦}∈Z

J

f(a)ϕ (a)ei tω(a)da

=

∗∈{±,◦}∈Z

I

f(h)ϕ(h)ei tω(h) 1 ω(h)dh. Now we can use a stationary phase argument by integrating with respect tohto gain a decay with respect tot. Typically, this kind of analysis depends on the possible cancellation of

hω(h). In our case, the situation seems to be very favourable, ashω(h)never approaches zero, as shown in Sect.7. The stationary phase argument also relies on cancellations of f∗ andϕat the boundary points, and there the problems come from the singularities of the action-angle variables.

We can distinguish two zones, starting with the separatix hM0. In this case, the action-angle variables induce logarithmic singularities. Essentially it means that the Fourier coefficients f, ϕinvolve logarithmic singularities nearh=M0. However, near this point, ω(h)also exhibits a logarithmic singularity, and it can be shown thathω(h)goes to infinity fast enough to ensure a decay in time which is essentially driven by the regularity of f and ϕ. So the problems are not at the separatix.

Near the pointh = −M0, the situation is more delicate: in this zone, the pendulum Hamiltonian is essentially a perturbation of the Harmonic oscillator, for which no damping is expected (ωbeing constant). However, we can prove thathω(h)does not vanish near this point. But this is not enough: indeed the action-angle variable of the harmonic oscillator involves algebraic singularity of order√

h+M0. This explains why the rate of decay of the integral with respect to the time is mainly driven by the behavior of f andϕnear(0,0) which corresponds of a local behavior of f(h)ϕ (h)in(h+M0)p+q2 ,yielding the main contribution for the decay in the previous Theorem.

Theorem2.2will be a straightforward consequence of Propositions7.7and7.14, proven in Sect.7.

2.2 Linear Damping

As explained above, our main result is expressed as a scattering result with the strong con- vergence ofg=rψtand by using the previous Theorem, the weak convergence ofr. The

(8)

next proposition gives the equation satisfied bygand fixes some notations used later. This Proposition will be proved at the beginning of Sect.3.

Proposition 2.3 Let r(t,x, v)be the solution of the linearized Eq.(1.6). Then the function g(t,x, v)=r(t, ψt(x, v))=rψt(x, v) (2.6) satisfies the equation

tg=C(t){η,cos(X◦ψt)} +S(t){η,sin(X◦ψt)}. (2.7) whereX:T×R→Tdenotes the projectionX(x, v)=x,and where

C(t)=C[r(t)] =C[gψ−t] =

T×Rcos(X(y, w))g(t, ψ−t(y, w))dydw (2.8) and

S(t)=S[r(t)] =S[gψ−t] =

T×Rsin(X(y, w))g(t, ψ−t(y, w))dydw. (2.9) Moreover, the coefficientsC(t)andS(t)satisfy the following Volterra integral equations

C(t)=FC(t)+ t

0

C(s)KC(ts)ds and S(t)=FS(t)+ t

0

S(s)KS(ts)ds(2.10), with

FC(t)=

T×Rcos(X◦ψt(y, w))r0(y, w)dydw, FS(t)=

T×Rsin(X◦ψt(y, w))r0(y, w)dydw, KC(t)=1{t≥0}

T×R{η,cos(X)}cos(X◦ψt), and KS(t)= −1{t≥0}

T×R{η,sin(X)}sin(X◦ψt).

(2.11)

For a functionF(t), we define its Fourier transform by F(ξ)ˆ =

RF(t)ei tξdt.

Theorem 2.4 (Linear damping) Letη(x, v) = G(h0(x, v))with G a decreasing function that satisfies the assumption

maxn≤10

yμG(n)(y)

L(R)Cμ, (2.12)

withμ >2,and assume that there existsκ >0such that

Im(ξ)≤0min |1− ˆKC(ξ)| ≥κ and min

Im(ξ)≤0|1− ˆKS(ξ)| ≥κ. (2.13) Let us assume that the initial perturbation r0satisfies

|α|≤maxmvνxα,vr0(x, v)LCm,ν,

(9)

for someν >2,and where

m≥5+3p 2 , with

p=max

k≥1, ∂x,vα r0(0,0)=0, ∀1 ≤ |α| ≤k

. (2.14)

Then, if r0satisfies the orthogonality condition

∗∈{±,◦}

J

C0(a)(r0)0(a)da=0, (2.15) there exists C>0such that for all t≥0

|C(t)| ≤ C

tmax(3,p+52 ) and |S(t)| ≤ C t2.

We shall call assumption (2.13) the Penrose criterion, by analogy with the stability con- ditions of the same name in the homogeneous case.

Let us remark that the orthogonality condition (2.15) is propagated by the flow of the linear Eq. (2.7) and therefore natural to impose. Indeed by using the action-angle variables given by Theorem2.1, we have thatη=G(h)is in fact a function ofhand hence ofaonly.

Hence the Eq. (2.7) in symplectic variables(θ,a)can be written

tg=C(t){G(a),cos(x(θ+tω(a),a))} +S(t){G(a),sin(x(θ+tω(a),a))}

SinceG(a)depends only ona, we get from the above equation that

tg0(t,a)=t

Tg(t, θ,a)dθ =0, aI, ∗ ∈ {±,◦}.

Asg0(t,a)=r0(t,a), this shows that that the orthogonality condition (2.15) is propagated along the flow of (1.6).

However, note that the condition (2.14) is in general not propagated by the flow unless some conditions onη(and henceG) are assumed to hold. Indeed, the decay is mainly driven by the Eq. (2.10) and the time decay of the kernelsKC(t)andKS(t)which depends on the order of the zero of∇ηat the origin. In the generic caseG <0 studied in this paper, the bounds given in Proposition3.1are essentially optimal, but some symmetry might yield some improvement: for example ifrandηare fully symmetric with respect to(x, v)→ −(x, v) then we can verify thatS(t)=0 for all times. A precise study of the rate of decay ofC(t)and S(t)as functions of the degree of “flatness" at the origin of the initial values and the stationary state will not be done here, but might play an important role in a nonlinear situation.

As a corollary of Theorem2.4, we get a scattering result for the solutiongof (2.7) and the weak convergence ofr(t,x, v)=g(t, ψ−t(x, v))the solution of (1.6) towards an asymptotic stater(x, v)that depends only onh0(x, v).

Corollary 2.5 Under the assumptions of Theorem2.4with p=0, we obtain that:

There exists g(x, v)and a constant C such that when t→ +∞, we have g(t)gL1x,vC

t. (2.16)

(10)

There exists r(x, v)that depends only on h, that is to say r(x, v) = r(h) for (x, v)Uand∗ ∈ {±,◦}, such that for every test functionφ, we have that

T×Rr(t,x, v)φ(x, v)dxdvt→+∞

T×Rr(x, v)φ(x, v)dxdv.

2.3 About the Penrose Criterion

Written in this form, the Penrose criterion (2.13) is difficult to check, but we can relate it to a more classical condition that was found in [31] or [3] to ensure orbital stability of inhomogeneous stationary states in the nonlinear equation.

First we shall prove that the verification of Penrose criterion (2.13) at the frequencyξ=0 is sufficient, by proving the following Theorem.

Theorem 2.6 Letηbe a state defined by(1.4), and assume that G satisfies the regularity assumption(2.12). Assume moreover that G<0and

1− ˆKC(0) >0 and 1− ˆKS(0) >0. (2.17) Then the Penrose criterion(2.13)holds true.

For regular and decreasing profileG, we first show that this condition can be recast in a more explicit expression:

Proposition 2.7 Letη be a state defined by (1.4). Assume that G satisfies the hypothesis (2.12), and that G<0.Then(2.17)holds true if and only ifη(x, v)=G(h0(x, v))satisfies

1+

R×TG(h0(x, v))cos2(x)dxdv

∗∈{±,◦}

J

G(h0(a))C0(a)2da>0. (2.18) Note that (2.18) coincides with the stability condition already mentioned in [3,31], ensuring the orbital stability of the stationary states by variational techniques.

Finally, we exhibit examples of stable stationary states given by Maxwell-Boltzmann distribution, under some condition on the coefficients of the Gaussian:

Proposition 2.8 Letα >0andβ >0such thatα2β < π2, then there exists M0 >0satisfying (1.5)such that

η(x, v)=αe−β

v2

2−M0cos(x)

, is a stable stationary states in the sense of Proposition2.7.

2.4 Organization

In Sect.3, we collect and prove some results concerning Volterra integral equations, and use them to prove the linear damping Theorem2.4by assuming Theorem 2.2giving the dispersive effect of the flow of the Pendulum. In Sect.4, we prove the scattering result corollary2.5. Section5is dedicated to the Penrose criterion, and we prove there Theorem2.6 and Proposition2.7. In Sect.6we exhibit examples of inhomogeneous stationary states which are stable in the sense of Proposition2.7and prove Proposition2.8. Finally, Sect.7contains all the technical results that we shall need concerning angle-action variables, and we prove there the dispersion Theorem2.2.

(11)

3 Proof of the Linear Damping Theorem2.4

We begin with the derivation of the Volterra Eq. (2.10). The proof then consists in showing that the kernelsKC andKShave sufficient decay in time, which, with the Penrose criterion and a Paley-Wiener argument will yield a control of the decay in time ofC(t)andS(t)by the one of the source termsFC(t)andFS(t),and the latter will be guaranteed by Theorem2.2.

In all the remainder of the paper, we will often use the notation A Bto denote an inequality of the formAC B for some constantC depending only on the assumptions made in the section of the proof but not onAorB.

Proof of Proposition2.3 Let us first prove (2.7). Ifrsolves (1.6), the functiongdefined in (2.6) satisfies

tg(t,x, v)= {r,h0}(t, ψt(x, v))+ {η, φ[r]}(t, ψt(x, v))− {r,h0}(t, ψt(x, v))

= {η, φ[g◦ψt]}(t, ψt(x, v)).

Hence asηis invariant by the flowψt,gsolves

tg(t,x, v)= {η, φ[g◦ψt] ◦ψt}(t,x, v).

Sinceψtpreserves the volume, φ[gψ−t] ◦ψt(x, v)=

T×Rcos(X◦ψt(x, v)y)g(t, ψ−t(y, w))dydw

=

T×Rcos(X◦ψt(x, v)−X(y, w))g(t, ψ−t(y, w))dydw

=cos(X◦ψt(x, v))C(t)+sin(X◦ψt(x, v))S(t),

withC(t)=C[g◦ψ−t] =C[r(t)]andS(t)=S[g◦ψ−t] =S[r(t)], which proves (2.7).

We deduce that

g(t,x, v)=r0(x, v)+ t

0

C(s){η,cos(X◦ψs)} +S(s){η,sin(X◦ψs)}ds. Using this formula and the fact thatψt preserves the Poisson bracket, we calculate that

C(t)=

T×Rcos(X(y, w))g(t, ψ−t(y, w))dydw

=

T×Rcos(X(y, w))r0−t(y, w))dydw +

t

0

C(s)

T×Rcos(X){η,cos(X◦ψs−t)}ds +

t

0

S(s)

T×Rcos(X){η,sin(X◦ψst)}ds.

Note that the flowψtis reversible with respect to the transformationν(x, v)=(x,−v),that is we haveψtν= −ν◦ψ−t. But as the Hamiltonian is even inx, the flow is also reversible with respect to(x, v)(−x, v). Hence the transformationμ(x, v):=(−x,−v)satisfies ψtμ=μψt, and this transformation preserves the Poisson bracket and is an isometry.

Let us apply this to the last term in the previous equation. We thus have for anyσ∈R

T×Rcos(X){η,sin(X◦ψσ)} =

T×Rcos(X◦μ){η,sin(X◦ψσ)} ◦μ

(12)

=

T×Rcos(X){η,sin(X◦μψσ)}

= −

T×Rcos(X){η,sin(X◦ψσ)} =0, as X◦μ= −X. For the same reason, we have

T×Rsin(X){η,cos(X◦ψσ)} =0.

Now using the identitiesην=ηand X◦ν=X,and the evenness of the cosine function, we have

T×Rcos(X){η,cos(X◦ψs−t)} = −

T×Rcos(X◦ν){η,cos(X◦ψs−tν)}

= −

T×Rcos(X){η,cos(X◦(−ν)ψt−s)}

= −

T×Rcos(X){η,cos(X◦ψt−s)}.

Integrating by parts that last integral yields then

T×Rcos(X){η,cos(X◦ψs−t)} =

T×Rcos(Xψt−s){η,cos(X)}.

Using the oddness of the sine function, we have by similar manipulations

T×Rsin(X){η,sin(X◦ψst)} = −

T×Rsin(X◦ψts){η,sin(X)}.

This ends the proof.

As a preliminary, we shall first use Theorem2.2in order to get the decay rates of the kernels. We shall prove the following result.

Proposition 3.1 Letη(x, v)= G(h0(x, v))with G a decreasing function that satisfies the assumption(2.12)withμ >2. Then there exist a constant C such that

|KC(t)| ≤ C

t3 and |KS(t)| ≤ C

t2. (3.1)

Proof In view of the expression (2.11) orKC(t)we apply Theorem2.2with the functions f(x, v)= {η,cos(X)}(x, v)andϕ=cos(X(x, v))for which we havep=1 andq=1. As we have that for all∗ ∈ {◦,±},

f0(h)= 1 2π

π

−π f(x(h, θ), v(h, θ))dθ =ω(h)G(h) 2π

π

−πθ(cos(x(h, θ)))dθ=0. (3.2) Theorem2.2then yields|KC(t)| t13. ConcerningKS(t),it suffices to apply Theorem2.2 with the functions{η,sin(X)}(x, v)and sin(X(x, v)).We have this time p=q =0,and S0(h)=0 for all∗ ∈ {◦,±}(see (7.25) and (7.39)). Hence the application of Theorem2.2

yields|KS(t)| t12.

(13)

To study the coefficientsC(t)andS(t),we shall use general results on Volterra integral equations written under the form

y(t)=Ky(t)+F(t), t∈R (3.3)

whereK,y,F vanish fort ≤ 0. Let us first recall the following Paley-Wiener result on Volterra integral equations (Theorem 4.1 of [25], see also [21,34]).

Lemma 3.2 (Paley–Wiener) Assume that KL1(R)is such that

Im(ξ)≤0min |1− ˆK(ξ)| ≥κ.

Then there exists a unique resolvent kernel RL1(R+)which vanishes for t≤0such that R(t)= −K(t)+KR(t). (3.4) Note that usingR, the solution of (3.3) can be written as

y(t)=F(t)RF(t). (3.5) We shall then use the following corollary.

Corollary 3.3 Under the assumptions of Lemma3.2, the following holds:

(i) There exists C >0such that

yLCFL. (3.6)

(ii) Ift2KLandt2FL, then there exists C>0such that

|t2yLCt2FL. (iii) Ift3KLandtαFLforα∈ [2,3], then

tαyLCtαFL. Proof To get (i) it suffices to use (3.5) and the Young inequality.

Let us prove (ii). We first observe that t12y(t)=K(t12y)+

t

0

(t12s12)K(ts)y(s)ds+t12F.

By using (i), we obtain that t12yLyLsup

t

t

0 (t12s12)|K(ts)|ds+ t12FLt12FL. (3.7) Next, we can write that

t y(t)= K(t y)+ t

0(t12s12)K(ts)s12y(s)ds+t12 t

0(t12s12)K(ts)y(s)ds+t F. Consequently, by using (i) and the assumptions onK, we obtain that

t yLsup

t

t

0

(ts)12 t−s2 ds

t12yL

+sup

t

t12

t

0

(ts)12 t−s2

1 s12ds

t12yL+ tFL

(14)

and by (3.7),

t yLtFL. (3.8)

Note that we have used that t12

t

0

(ts)12 ts2

1

s21ds 1 t32

t

t 2

1 s12 ds+

t

2 0

1

ts32 ds1.

We then estimatet2y,and for that we write

t2y=Kt2y+F2

where by similar manipulations as above, the source termF2may be estimated as follows

|F2|t12

·12|K|

(·|y|) +

·12|K|

·32|y| +t

·12|K|

·12|y|

+t32

·12|K|

(|y|)+t2|F|.

By using again (i) and (3.7), and similar arguments as above, we obtain that t2yLt32yL+ t2FL.

To conclude, we can use first the interpolation inequality t32yLtyL12t2yL12. Then we apply the Young inequality: for anyδ >0,

tyL12t2yL12≤tyL

2δ +δt2yL

2 .

Choosingδsmall enough, we conclude that

t2yLtyL+ t2FL

and the result follows by using (3.8).

To prove (iii), we can use the same arguments. We first write t y=K(t y)+F1

with

F1(t)=t F+(t K)y.

Sincet KL1, we get by using (3.6) that

t yLF1LtFL. Next, we write

t2y=Kt2y+F2, F2=(t K)t y+t F1 and by Young’s inequality

F2Lt KL1t yL+ t2FL+ t((t K)y)L

t FL+ t2FL+ t((t K)∗y)L.

(15)

It remains to see that

|(t K)y|

t

0

1 t−s2

1

styLds 1

tt yL, such that

F2Lt2FL. We conclude by using again (3.6) that

t2yLt2FL.

t3yis estimated in the same way as above.

We shall then apply the Corollary to the two Volterra equations (2.10) to prove Theo- rem2.4, starting with the one satisfied byC(t).Note that by using Proposition3.1, and the Penrose criterion (2.13), we get that the kernelKCmatches the assumptions of Corollary3.3 (iii). To estimateFC(t)given by (2.11), we can apply Theorem2.2(using the orthogonality condition (2.15)) with the functionsϕ=cos(X(x, v))and f =r0(x, v), for which we have p ≥ 0 andq = 1. Now without further assumptions this implies that FC(t) t1α with α= p+52 . Therefore the application of Corollary3.3to the first Volterra equation of (2.10) yields the estimate onC(t)claimed in Theorem2.4.

In the case of the second Volterra equation of (2.10), satisfied byS(t),we estimateFS(t) given in (2.11) by using Theorem2.2with the functionsϕ=sin(X(x, v))and f =r0(x, v) for which we havep≥0 andq=0. This yields the estimate|FS(t)| t1α withα=2+p2. As the kernelKSfalls under the scope of Corollary3.3(ii), the estimate onS(t)claimed in Theorem2.4follows.

4 Proof of the Scattering Result Corollary2.5

Let us first study the asymptotic behavior ofg,and defineg(x, v)by g(x, v)=r0(x, v)+

+∞

0

C(s){η,cos(X)} ◦ψs(x, v)+S(s){η,sin(X)} ◦ψs(x, v) ds. (4.1) Note that the above integral is convergent in L1x,v. Indeed, by using that ψs is measure preserving and Theorem2.4giving decay estimates forC(s)andS(s)we get that

C(s){η,cos(X)} ◦ψs(x, v)+S(s){η,sin(X)} ◦ψs(x, v)L1x,v 1 s2. As

g(t,x, v)=r0(x, v)+ t

0

C(s){η,cos(X)} ◦ψs(x, v)+S(s){η,sin(X)} ◦ψs(x, v) ds, this also yields that

g(t)−gL1x,v +∞

t

1

s2ds 1

t, (4.2)

(16)

which proves the first part of the statement. Now, let us study the weak convergence of r(t,x, v). Let us observe that for every test functionφ(x, v), we have by volume preservation that

T×R

r(t,x, v)φ(x, v)dxdv=

T×R

g(t,x, v)φ(ψt(x, v))dxdv

=

T×Rg(x, v)φ(ψt(x, v))dxdv+O 1

t

=:I(t)+ +O 1

t

. By using the expression (4.1), and the fact thatψsis invertible and preserves the volume, we obtain that

I(t)=

T×Rr0(x, v)φ(ψt(x, v))dxdv+ +∞

0

C(s)

T×R{η,cos(X)}φ(ψts(x, v))dxdv +S(s)

T×R{η,sin(X)}φ(ψts(x, v))dxdv

ds. (4.3)

Now, thanks to Theorem2.2, we obtain that

T×Rr0(x, v)φ(ψt(x, v))dxdvt→+∞

∗∈{±,◦}

J

(r0)0(a)φ0(a)da

=

T×Rr(x, v)φ(x, v)dxdv, withr(x, v)the angle average ofr0(θ,a),

r(x, v)=(r0)0(h)= 1 2π

(−π,π)r0(x(θ,h), v(θ,h))dθ, hI,∗ ∈ {±,◦}.

Next, we observe that{η,cos(X)}0 = {η,sin(X)}0 = 0. Consequently, by using again Theorem2.2, we obtain that

T×R{η,cos(X)}φ(ψts(x, v))dxdv +

T×R{η,sin(X)}φ(ψts(x, v))dxdv 1

t−s2. Consequently, we find that

+∞

0

C(s)

T×R{η,cos(X)}φ(ψt−s)dxdv+S(s)

T×R{η,sin(X)}φ(ψt−s)dxdv

ds

+∞

0

1 s2

1

ts2ds 1 t2.

and using (4.3) this concludes the proof of corollary2.5.

5 Penrose Condition: Proofs of Theorem2.6and Proposition2.7 5.1 Proof of Theorem2.6

Let us start with the study ofKC (see (2.11)). With the assumption on the profile function G, Proposition3.1shows thatKCL1(R+)L2(R+). We have

KˆC(ξ)= 1 2π

−∞KC(t)ei tξdt= 1 2π

T×R

0

ei tξ{η,cos(X)}cos(X◦ψt)dxdvdt,

(17)

which defines a continuous function on the set{ξ ∈C|Im(ξ)≤0}holomorphic on the set {Im(ξ) <0}.

(i) As dtd cos(X◦ψt)= {cos(X),h0} ◦ψt, forξ =0, we have after integration by part and using estimate (3.1)

KˆC(ξ)= − 1

2iξπKC(0)− 1

1 2π

0

ei tξ

T×R{η,cos(X)}{cosX,h0} ◦ψtdxdvdt.

To analyze the second term, we can use Theorem2.2with the functions f = {η,cos(X)}and ϕ= {cosX,h0} = −sin(x)vfor which we have p=q =1. As noted in (3.2) the average f0(a)vanishes and hence the integrand isO(t13)by using Theorem2.2. This shows that for {Im(ξ)≤0}andξ=0 we have

| ˆKC(ξ)| 1

|ξ|.

Hence there existsB>0 such that for|ξ| ≥B,|1− ˆKC(ξ)|> 12.

(ii) Moreover, asη=G(h0), and ash0is invariant by the flow, we have that KC(t)=1{t0}

T×RG(h0)cos(X){h0,cos(X)} ◦ψtdxdv

=1{t0}d dt

T×RG(h0)cos(X)cos(X◦ψt)dxdv=1{t0}d dtQC(t), with

QC(t)=

T×RG(h0)cos(X◦ψt)cos(X)dxdvQ0. where

Q0=

∗∈{±,◦}

J

G(h0(a))|C0(a)|2da,

whereC0(a) ∈ Ris given in (2.5) (see (7.24) and (7.38) for explicit expressions). By applying Theorem2.2with the functions f =G(h0)cosXandϕ=cosX, we obtain with this definition ofQ0that

|QC(t)| 1 t3. Hence we can write

KˆC(ξ)= +∞

0

e−i tξ d

dtQC(t)dt= −QC(0)+ +∞

0

e−i tξQC(t)dt,

where by the previous estimate the time integral is well defined and uniformly bounded in {Im(ξ)≤0}. The assumption (2.17) can actually be restated as

1− ˆKC(0)=1+QC(0)=κ0>0.

Hence, by continuity, there existsA>0 such that for|ξ| ≤A, we will have|1− ˆKC(ξ)|> κ20. (iii) Now let us expressQC(t)in action-angle variables. We have

QC(t)= −Q0+ 1 2π∗∈{◦,±}

J×(−π,π)G(h0(a))cos(X◦ψt(θ,a))cos(x(θ,a))dθda

Referenzen

ÄHNLICHE DOKUMENTE

For the dynamics of a size-structured population we prove the existence and uniqueness of a stationary state maximizing the profit on population exploitation under the assumption

The model will be presented in several stages. Capacity utilization is determined by t h e ratio of indicated production to production capacity, a measure of

Importantly, the GM and FS models are centered around the micro structure that generates deviations from the UIP condition; however, to a …rst order approximation, these end

Prime Minister Mariano Rajoy offered political support for the embattled Greek Prime Minister, Antonis Samaras, by visiting Athens before the 25 January snap general election

We consider the case with a Schwarzschild black hole in the centre and we show that there exist static massless shells of Vlasov matter with compact support and finite mass

Fach D 188, 78457 Konstanz, Germany Email: preprints@informatik.uni–konstanz.de WWW: http://www.fmi.uni–konstanz.de/Schriften. Konstanzer Online-Publikations-System (KOPS)

The ruling of the Second Senate issued on 30 June 2009 in relation to the Treaty of Lis- bon, states that “the unification of Europe in the shape of a treaty-based union of sovereign

While in the limit of very small and infinite correlation lengths ξ of the random disor- der, the fluctuating gap model (FGM) admits for an exact analytic calculation of the density