• 検索結果がありません。

MichelBenaïm OlivierRaimond Self-interactingdiffusionsIV:Rateofconvergence ∗

N/A
N/A
Protected

Academic year: 2022

シェア "MichelBenaïm OlivierRaimond Self-interactingdiffusionsIV:Rateofconvergence ∗"

Copied!
29
0
0

読み込み中.... (全文を見る)

全文

(1)

El e c t ro nic

Journ a l of

Pr

ob a b il i t y

Vol. 16 (2011), Paper no. 66, pages 1815–1843.

Journal URL

http://www.math.washington.edu/~ejpecp/

Self-interacting diffusions IV: Rate of convergence

Michel Benaïm

Institut de Mathématiques Université de Neuchâtel,

Rue Emile-Argand 11, CH-2000 Neuchâtel, michel.benaim@unine.ch

Olivier Raimond

Laboratoire Modal’X, Université Paris Ouest, 200 avenue de la République 92000 Nanterre, France

olivier.raimond@u-paris10.fr

Abstract

Self-interacting diffusions are processes living on a compact Riemannian manifold defined by a stochastic differential equation with a drift term depending on the past empirical measureµt of the process. The asymptotics ofµt is governed by a deterministic dynamical system and under certain conditions(µt)converges almost surely towards a deterministic measureµ(see Benaïm, Ledoux, Raimond (2002) and Benaïm, Raimond (2005)). We are interested here in the rate of convergence ofµt towardsµ. A central limit theorem is proved. In particular, this shows that greater is the interaction repelling faster is the convergence.

Key words:Self-interacting random processes, reinforced processes.

AMS 2010 Subject Classification:Primary 60K35; Secondary: 60H10, 62L20, 60F05.

Submitted to EJP on July 29, 2009, final version accepted July 18, 2011.

We acknowledge financial support from the Swiss National Science Foundation Grant 200021-103625/1

(2)

1 Introduction

Self-interacting diffusions

Let M be a smooth compact Riemannian manifold and V : M × M → R a sufficiently smooth mapping1. For all finite Borel measureµ, letVµ:M→Rbe the smooth function defined by

Vµ(x) = Z

M

V(x,y)µ(d y). Let(eα) be a finite family of vector fields onM such thatP

αeα(eαf)(x) = ∆f(x), where∆is the Laplace operator onM andeα(f)stands for the Lie derivative of f alongeα. Let(Bα)be a family of independent Brownian motions.

A self-interacting diffusion on M associated to V can be defined as the solution to the stochastic differential equation (SDE)

d Xt=X

α

eα(Xt)◦d Bαt − ∇(Vµt)(Xt)d t

whereµt=1tRt

0δXsdsis the empirical occupation measure of(Xt).

In absence of drift (i.e V = 0), (Xt) is just a Brownian motion on M but in general it defines a non Markovian process whose behavior at time t depends on its past trajectory through µt. This type of process was introduced in Benaim, Ledoux and Raimond (2002) ([3]) and further analyzed in a series of papers by Benaim and Raimond (2003, 2005, 2007) ([4], [5] and [6]). We refer the reader to these papers for more details and especially to [3]for a detailed construction of the process and its elementary properties. For a general overview of processes with reinforcement we refer the reader to the recent survey paper by Pemantle (2007) ([16]).

Notation and Background

We letM(M)denote the space of finite Borel measures onM,P(M)⊂ M(M)the space of proba- bility measures. IfIis a metric space (typically,I=M,R+×Mor[0,TM) we letC(I)denote the space of real valued continuous functions onI equipped with the topology of uniform convergence on compact sets. The normalized Riemann measure on M will be denoted byλ.

Let µ ∈ P(M) and f : M → Ra nonnegative or µ−integrable Borel function. We write µf for R f dµ, and for the measure defined as fµ(A) = R

Af dµ. We let L2(µ) denote the space of functions for which µ|f|2 < ∞, equipped with the inner product 〈f,gµ = µ(f g) and the norm kfkµ=p

µf2. We simply write L2 forL2(λ).

Of fundamental importance in the analysis of the asymptotics of(µt)is the mappingΠ:M(M)→ P(M)defined by

Π(µ) =ξ(Vµ)λ (1)

1The mappingVx:MRdefined byVx(y) =V(x,y)isC2and its derivatives are continuous in(x,y).

(3)

whereξ:C(M)→C(M)is the function defined by ξ(f)(x) = e−f(x)

R

Me−f(y)λ(d y). (2)

In[3], it is shown that the asymptotics ofµtcan be precisely related to the long term behavior of a certain semiflow onP(M)induced by the ordinary differential equation (ODE) onM(M):

µ˙=−µ+ Π(µ). (3)

Depending on the nature ofV, the dynamics of (3) can either be convergent or nonconvergent lead- ing to similar behaviors for{µt}(see[3]). WhenV is symmetric, (3) happens to be aquasigradient and the following convergence result holds.

Theorem 1.1 ([5]). Assume that V is symmetric, i.e. V(x,y) =V(y,x). Then the limit set oft} (for the topology of weak* convergence) is almost surely a compact connected subset of

Fix(Π) ={µ∈ P(M):µ= Π(µ)}.

In particular, ifFix(Π) is finite then(µt) converges almost surely toward a fixed point ofΠ. This holds for a generic functionV (see[5]). Sufficient conditions ensuring thatFix(Π)has cardinal one are as follows:

Theorem 1.2([5],[6]). Assume that V is symmetric and that one of the two following conditions hold (i) Up to an additive constant V is a Mercer kernel: For some constant C, V(x,y) =K(x,y) +C,

and for all fL2,

Z

K(x,y)f(x)f(y)λ(d x)λ(d y)≥0.

(ii) For all xM,yM,uTxM,vTyM

Ricx(u,u) +Ricy(v,v) +Hessx,yV((u,v),(u,v))K(kuk2+kvk2)

where K is some positive constant. HereRicx stands for the Ricci tensor at x andHessx,y is the Hessian of V at(x,y).

ThenFix(Π)reduces to a singleton}andµtµwith probability one.

As observed in [6] the condition (i) in Theorem 1.2 seems well suited to describe self-repelling diffusions. On the other hand, it is not clearly related to the geometry of M. Condition(ii) has a more geometrical flavor and is robust to smooth perturbations (of M and V). It can be seen as a Bakry-Emery type condition for self interacting diffusions.

In[5], it is also proved that every stable (for the ODE (3)) fixed point ofΠhas a positive probability to be a limit point forµt; and any unstable fixed point cannot be a limit point forµt.

(4)

Organisation of the paper Letµ∈Fix(Π). We will assume that

Hypothesis 1.3. µtconverges a.s. towardsµ.

In this paper we intend to study the rate of this convergence. Let

t=et/2etµ).

It will be shown that, under some conditions to be specified later, for all g= (g1, . . . ,gn)∈C(M)n the process

sg1, . . . ,∆sgn,Vs

st converges in law, as t → ∞, toward a certain stationary Ornstein-Uhlenbeck process(Zg,Z)onRn×C(M). This process is defined in Section 2. The main result is stated in section 3 and some examples are developed. It is in particular observed that a strong repelling interaction gives a faster convergence. The section 4 is a proof section.

In the following K (respectively C) denotes a positive constant (respectively a positive random constant). These constants may change from line to line.

2 The Ornstein-Uhlenbeck process ( Z

g

, Z ) .

For a more precise definition of Ornstein-Uhlenbeck processes on C(M) and their basic properties, we refer the reader to the appendix (section 5). Throughout all this section we letµ∈ P(M)and g= (g1, ...,gn)∈C(M)n. ForxM we setVx :M →Rdefined byVx(y) =V(x,y).

2.1 The operatorGµ

LethC(M)and letGµ,h:R×C(M)→Rbe the linear operator defined by

Gµ,h(u,f) =u/2+Covµ(h,f), (4) where Covµis the covariance on L2(µ), that is the bilinear form acting onL2×L2defined by

Covµ(f,h) =µ(f h)−(µf)(µh). We define the linear operatorGµ:C(M)→C(M)by

Gµf(x) =Gµ,Vx(f(x),f) = f(x)/2+Covµ(Vx,f). (5) It is easily seen that kGµfk≤(2kVk+1/2)kfk. In particular,Gµ is a bounded operator. Let {e−t Gµ}denote the semigroup acting on C(M) with generator−Gµ. From now on we will assume the following:

Hypothesis 2.1. There exists κ > 0 such that µ << λ with kdλk <, and such that for all fL2(λ),〈Gµf,fλκkfk2λ.

Let

λ(−Gµ) = lim

t→∞

log(ket Gµk)

t .

This limit exists by subadditivity. Then

(5)

Lemma 2.2. Hypothesis 2.1 implies thatλ(−Gµ)≤ −κ <0.

Proof : For all fL2(λ), d

d tke−t Gµfk2λ=−2〈Gµe−t Gµf,e−t Gµfλ≤ −2κke−t Gµfkλ.

This implies thatke−t Gµfkλe−κtkfkλ. Denote by gt the solution of the differential equation d

d tgt(x) =Covµ(Vx,gt)

with g0 = fC(M). Note that et Gµf = et/2gt. It is straightforward to check that (using the fact that kk < ∞) d

d tkgtkλKkgtkλ with K a constant depending only on V and µ. Thus supt∈[0,1]kgtkλKkfkλ. Now, since for allxM andt∈[0, 1]

d d tgt(x)

KkgtkλKkfkλ, we havekg1kKkfkλ. This implies thatke−GµfkKkfkλ. Now for allt>1, and fC(M),

ket Gµfk = keGµe−(t1)GµfkKke−(t1)Gµfkλ

Ke−κ(t−1)kfkλKe−κtkfk. This implies thatket Gµk ≤Ke−κt, which proves the lemma. QED TheadjointofGµ is the operator onM(M)defined by the relation

m(Gµf) = (Gµm)f for allm∈ M(M)and fC(M). It is not hard to verify that

Gµm= 1

2m+ (V m)µ−(µ(V m))µ. (6)

2.2 The generatorAµ and its inverseQµ

LetH2be the Sobolev space of real valued functions onM, associated with the normkfk2H=kfk2λ+ k∇fk2λ. Since Π(µ) and λ are equivalent measures with continuous Radon-Nykodim derivative, L2(Π(µ)) =L2(λ). We denote byKµthe projection operator, acting onL2(Π(µ)), defined by

Kµf = f −Π(µ)f. We denote byAµthe operator acting onH2 defined by

Aµf = 1

2∆f − 〈∇,∇f〉.

(6)

Note that for f andhinH2 (denoting〈·,·〉the Riemannian inner product onM)

Aµf,hΠ(µ)=−1 2 Z

〈∇f,∇h〉(x)Π(µ)(d x). For all fC(M)there existsQµfH2such thatΠ(µ)(Qµf) =0 and

f −Π(µ)f =Kµf =−AµQµf. (7)

It is shown in[3]thatQµf isC1 and that there exists a constantK such that for all fC(M)and µ∈ P(M),

kQµfk+k∇QµfkKkfk. (8) Finally, note that for f andhin L2,

Z

〈∇Qµf,∇Qµh〉(x)Π(µ)(d x) =−2〈AµQµf,QµhΠ(µ)=2〈f,QµhΠ(µ). (9)

2.3 The covarianceCµg

We letCbµdenote the bilinear continuous formCbµ:C(M)×C(M)→Rdefined by Cbµ(f,h) =2〈f,QµhΠ(µ).

This form is symmetric (see its expression given by (9)). Note also that for some constantKdepend- ing onµ,|Cbµ(f,h)| ≤Kkfk× khk.

We let Cµ denote the mapping Cµ : M × M → R defined by Cµ(x,y) = Cbµ(Vx,Vy). Let ˜M = {1, . . . ,n} ∪M andCµg: ˜M×M˜ →Rbe the function defined by

Cµg(x,y) =

Cbµ(gx,gy) for x,y ∈ {1, . . . ,n}, Cµ(x,y) for x,yM,

Cbµ(Vx,gy) for xM, y ∈ {1, . . . ,n}. ThenCµ andCµgare covariance functions (as defined in subsection 5.2).

In the following, when n= 0, ˜M = M and Cµg = Cµ. When n ≥ 1, C(M˜) can be identified with Rn×C(M).

Lemma 2.3. There exists a Brownian motion onRn×C(M)with covariance Cµg. Proof : Since the argument are the same forn≥1, we just do it forn=0. Let

dCµ(x,y) := p

Cµ(x,x)−2Cµ(x,y) +Cµ(y,y)

= k∇Qµ(VxVy)kΠ(µ)KkVxVyk

where the last inequality follows from (8). ThendCµ(x,y)≤K d(x,y). Thus dCµ satisfies (30) and we can apply Theorem 5.4 of the appendix (section 5). QED

(7)

2.4 The process(Zg,Z)

LetGµg:Rn×C(M)→Rn×C(M)be the operator defined by Gµg=

‚ In/2 Agµ 0 Gµ

Œ

(10) where In is the identity matrix onRn andAgµ :C(M)→Rn is the linear map defined byAgµ(f) =

Covµ(f,g1), . . . , Covµ(f,gn) .

SinceGµgis a bounded operator, for any lawν onRn×C(M), there exists ˜Z= (Zg,Z)an Ornstein- Uhlenbeck process of covarianceCµg and drift−Gµg, with initial distribution given byν (using Theo- rem 5.6). More precisely, ˜Z is the unique solution of

¨ d Zt = dWtGµZtd t d Ztgi = dWtgi −€

Ztgi/2+Covµ(Zt,gi

d t, i=1, . . . ,n (11) where ˜Z0is aRn×C(M)-valued random variable of lawν and ˜W= (Wg,W)is aRn×C(M)-valued Brownian motion of covariance Cµg independent of ˜Z. In particular, Z is an Ornstein-Uhlenbeck process of covarianceCµ and drift−Gµ. Denote byPg,t µthe semigroup associated to ˜Z. Then Proposition 2.4. Assume hypothesis 2.1. Then there existsπg,µthe law of a centered Gaussian variable inRn×C(M), with varianceVar(πg,µ)where for(u,m)∈Rn× M(M),

Varg,µ)(u,m) := E€(mZ+〈u,Zg〉)2Š

= Z

0

Cbµ(ft,ft)d t with ft =e−t/2P

iuigi+V mt,and where mtis defined by mtf =m0(et Gµf) +

Xn

i=1

ui Z t

0

es/2Covµ(gi,e−(ts)Gµf)ds. (12)

Moerover,

(i) πg,µ is the unique invariant probability measure ofPt.

(ii) For all bounded continuous function ϕ on Rn × C(M) and all (u,f) ∈ Rn × C(M), limt→∞Pg,µt ϕ(u,f) =πg,µϕ.

Proof : This is a consequence of Theorem 5.7. To apply it one can remark thatGµg is an operator like the ones given in example 5.11.

The varianceVar(πg,µ)is given byVar(πg,µ)(ν) =R

0 〈ν,e−sGµgCµges(Gµg)νdsforν= (u,m)∈Rn× M(M) =C(M˜). ThusVarg,µ)(u,m) =R

0 Cbµ(ft,ft)d t with ft =P

iut(i)gi+V mt and where (ut,mt) =e−t(Gµg)(u,m). Now

(Gµg)=

‚ I/2 0 (Agµ) (Gµ)

Œ

(8)

and(Agµ)u=P

iui(giµgi)µ. Thusut=et/2uandmt is the solution withm0=mof d mt

d t =−et/2 X

i

ui(giµgi)

!

µ−(Gµ)mt. (13) Note that (13) is equivalent to

d

d t(mtf) =−et/2Covµ X

i

uigi,f

!

mt(Gµf) for all fC(M), andm0=m. From which we deduce that

mt=et Gµm0− Z t

0

es/2e−(ts)Gµ X

i

ui(giµgi

! ds which implies the formula formt given by (12). QED

An Ornstein-Uhlenbeck process of covarianceCµg and drift −Gµg will be called stationarywhen its initial distribution isπg,µ.

3 A central limit theorem for µ

t

We state here the main results of this article. We assumeµ∈Fix(Π)satisfies hypotheses 1.3 and 2.1. Set∆t=et/2etµ), Dt=Vt andDt+·= (Dt+s)s≥0. Then

Theorem 3.1. Dt+· converges in law, as t → ∞, towards a stationary Ornstein-Uhlenbeck process of covariance Cµ and driftGµ.

ForgC(M)n, we setDtg= (∆tg,Dt)andDt+·g = (Dgt+s)s≥0. Then

Theorem 3.2. Dt+·g converges in law towards a stationary Ornstein-Uhlenbeck process of covariance Cµg and driftGµg.

DefineCb:C(MC(M)→Rthe symmetric bilinear form defined by

Cb(f,h) = Z

0

Cbµ(ft,ht)d t, (14) with (ht is defined by the same formula, withhin place of f)

ft(x) =e−t/2f(x)− Z t

0

e−s/2Covµ(f,e−(t−s)GµVx)ds. (15) Corollary 3.3.tg converges in law towards a centered Gaussian variable Zg of covariance

E[ZgiZgj] =Cb(gi,gj).

Proof : Follows from theorem 3.2 and the calculus ofVar(πg,µ)(u, 0). QED

(9)

3.1 Examples

3.1.1 Diffusions

Suppose V(x,y) = V(x), so that (Xt) is just a standard diffusion on M with invariant measure µ= λe x p(−exp(−VV).

Let fC(M). Sinceet Gµ1=et/21, ft defined by (15) is equal toet/2f. Thus

Cb(f,g) =(fQµg). (16) Corollary 3.3 says that

Theorem 3.4. For all gC(M)n,gt converges in law toward a centered Gaussian variable (Zg1, . . . ,Zgn), with covariance given by

E(ZgiZgj) =2µ(giQµgj).

Remark 3.5. This central limit theorem for Brownian motions on compact manifolds has already been considered by Baxter and Brosamler in[1]and[2]; and by Bhattacharya in[7]for ergodic diffusions.

3.1.2 The caseµ=λandV symmetric.

Suppose here thatµ = λ and that V is symmetric. We assume (without loss of generality since Π(λ) =λimplies thatis a constant function) that=0.

Since V is compact and symmetric, there exists an orthonormal basis (eα)α≥0 in L2(λ) and a se- quence of reals(λα)α≥0 such thate0is a constant function and

V =X

α≥1

λαeαeα.

Assume that for allα, 1/2+λα> 0. Then hypothesis 2.1 is satisfied, and the convergence of µt

towardsλholds with positive probability (see[6]).

Let fC(M) and ft defined by (15), denoting fα = 〈f,eαλ and ftα = 〈ft,eαλ, we have ft0 = e−t/2f0and forα≥1,

ftα = e−t/2fαλαe−(1/2+λα)t

‚eλαt−1 λα

Œ fα

= e−(1/2α)tfα.

Using the fact thatCbλ(f,g) =2λ(fQλg), this implies that

Cb(f,g) =2X

α≥1

X

β≥1

1

1+λα+λβf,eαλg,eβλλ(eαQλeβ). This, with corollary 3.3, proves

(10)

Theorem 3.6. Assume hypothesis 1.3 and that 1/2+λα > 0 for all α. Then for all gC(M)n,

gt converges in law toward a centered Gaussian variable (Zg1, . . . ,Zgn), with covariance given by E(ZgiZgj) =Cb(gi,gj).

In particular,

E(ZeαZeβ) = 2

1+λα+λβλ(eαQλeβ).

When allλαare positive, which corresponds to what is named a self-repelling interaction in[6], the rate of convergence ofµt towardsλis bigger than when there is no interaction, and the bigger is the interaction (that is largerλα’s) faster is the convergence.

4 Proof of the main results

We assume hypothesis 1.3 and µ satisfies hypothesis 2.1. For convenience, we choose for the constantκin hypothesis 2.1 a constant less than 1/2. In all this section, we fix g = (g1, ...,gn) ∈ C(M)n.

4.1 A lemma satisfied byQµ

We denote byX(M)the space of continuous vector fields on M, and equip the spacesP(M)and X(M)respectively with the weak convergence topology and with the uniform convergence topology.

Lemma 4.1. For all fC(M), the mapping µ 7→ ∇Qµf is a continuous mapping fromP(M) in X(M).

Proof : Letµandν be inM(M), and fC(M). Seth=Qµf. Then f =−Aµh+ Π(µ)f and k∇Qµf − ∇Qνfk = k − ∇QµAµh+∇QνAµhk

= k∇h+∇QνAµhk

≤ k∇(h+QνAνh)k+k∇Qν(AµAν)hk. Since∇(h+QνAνh) =0 and(AµAν)h=〈∇Vµ−ν,∇h〉, we get

k∇Qµf − ∇QνfkKk〈∇Vµ−ν,∇h〉k. (17) Using the fact that (x,y) 7→ ∇Vx(y) is uniformly continuous, the right hand term of (17) con- verges towards 0, whend(µ,ν)converges towards 0, d being a distance compatible with the weak convergence. QED

4.2 The process

Setht=t andh=. Recall∆t=et/2etµ)andDt(x) =Vt(x) = ∆tVx. Then(Dt)is a continuous process taking its values inC(M)andDt=et/2(heth).

(11)

To simplify the notation, we set Ks = Kµs, Qs =Qµs andAs =Aµs. Let(Mtf)t≥1 be the martingale defined byMtf =P

α

Rt

1 eα(Qsf)(Xs)d Bαs. The quadratic covariation ofMf andMh(with f andhin C(M)) is given by

Mf,Mht= Z t

1

〈∇Qsf,∇Qsh〉(Xs)ds.

Then for allt≥1 (with ˙Qt= d tdQt) ,

Qtf(Xt)−Q1f(X1) =Mtf + Z t

1

sf(Xs)ds− Z t

1

Ksf(Xs)ds.

Thus

µtf = 1 t

Z t

1

Ksf(Xs)ds+1 t

Z t

1

Π(µs)f ds+1 t

Z 1

0

f(Xs)ds

= −1 t

‚

Qtf(Xt)−Q1f(X1)− Z t

1

sf(Xs)ds

Œ

+ Mtf t +1

t Z t

1

〈ξ(hs),fλds+1 t

Z 1

0

f(Xs)ds.

For fC(M)(using the fact thatµf =〈ξ(h),fλ),∆tf =P5

i=1itf with

1tf = et/2 −Qetf(Xet) +Q1f(X1) + Z et

1

sf(Xs)ds

!

2tf = e−t/2Meft

3tf = et/2 Z et

1

〈ξ(hs)−ξ(h)−Dξ(h)(hsh),fλds

4tf = e−t/2 Z et

1

Dξ(h)(hsh),fλds

5tf = e−t/2 Z 1

0

f(Xs)dsµf

! .

ThenDt=P5

i=1Dit, whereDit=Vit. Finally, note that

Dξ(h)(h−h),fλ=−Covµ(h−h,f). (18) 4.3 First estimates

We recall the following estimate from[3]: There exists a constantK such that for allfC(M)and t>0,

kQ˙tfkK tkfk.

(12)

This estimate, combined with (8), implies that for f andhinC(M),

MfMhtKkfhk×t and that

Lemma 4.2. There exists a constant K depending onkVksuch that for all t≥1, and all f ∈C(M) k∆1tfk+k∆5tfkK×(1+t)et/2kfk, (19) which implies that((∆1+ ∆5)t+s)s0 and((D1+D5)t+s)s0 both converge towards0(respectively in M(M)and in C(R+×M)).

We also have

Lemma 4.3. There exists a constant K such that for all t≥0and all fC(M), E[(∆2tf)2] ≤ Kkfk2,

|∆3tf| ≤ Kkfkλ×e−t/2 Z t

0

kDsk2λds,

|∆4tf| ≤ Kkfkλ×e−t/2 Z t

0

es/2kDskλds.

Proof : The first estimate follows from

E[(∆2tf)2] =e−tE[(Meft)2] =e−tE[〈Mfet] ≤ Kkfk2. The second estimate follows from the fact that

kξ(h)−ξ(h)−Dξ(h)(h−h)kλ=O(khhk2λ).

The last estimate follows easily after having remarked that

|〈Dξ(h)(hsh),f〉|=|Covµ(hsh,f)| ≤Kkfkλ× khshkλ. This proves this lemma. QED

4.4 The processes0 and D0

Set∆0= ∆2+ ∆3+ ∆4andD0=D2+D3+D4. For fC(M), set εtf =et/2〈ξ(het)−ξ(h)−Dξ(h)(heth),fλ. Then

d0tf =−∆0tf

2 d t+d Ntf +εtfd t+〈Dξ(h)(Dt),fλd t where for all fC(M),Nf is a martingale. Moreover, for f andhinC(M),

Nf,Nht= Z t

0

〈∇Qesf(Xes),∇Qesh(Xes)〉ds.

(13)

Then, for all x,

d D0t(x) =−D0t(x)

2 d t+d Mt(x) +εt(x)d t+〈Dξ(h)(Dt),Vxλd t

whereM is the martingale inC(M)defined byM(x) =NVx andεt(x) =εVtx. We also have Gµ(D0)t(x) = D0t(x)

2 − 〈Dξ(h)(D0t),Vxλ. DenotingLµ= L−Gµ (defined by equation (32) in the appendix (section 5)),

d Lµ(D0)t(x) = d D0t(x) +Gµ(D0)t(x)d t and we have

Lµ(D0)t(x) =Mt(x) + Z t

0

ε0s(x)ds withε0s(x) =ε0sVx where for all fC(M),

ε0sf =εsf +〈Dξ(h)((D1+D5)s),fλ. Using lemma 5.5,

D0t=Lµ1(M)t+ Z t

0

e−(ts)Gµε0sds. (20) Denote ∆tg = (∆tg1, . . . ,∆tgn), ∆0tg = (∆0tg1, . . . ,∆0tgn) , Ng = (Ng1, . . . ,Ngn) and ε0tg = (ε0tg1, . . . ,ε0tgn). Then, denoting Lµg=LGg

µ (withGµg defined by (10)) we have Lµg(∆0g,D0)t= (Ntg,Mt) +

Z t

0

0sg,ε0s)ds

so that (using lemma 5.5 and integrating by parts) (∆0tg,D0t) = (Lµg)1(Ng,M)t+

Z t

0

e−(ts)G

g

µ0sg,ε0s)ds. (21) Moreover

(Lµg)1(Ng,M)t=

Nbtg1, . . . ,Nbtgn,Lµ1(M)t

, where

Nbtgi =Ntgi− Z t

0

‚Nsgi

2 +Cbµ(Lµ1(M)s,gi)

Πds.

参照

関連したドキュメント

Keywords: divergence-measure fields, normal traces, Gauss-Green theorem, product rules, Radon measures, conservation laws, Euler equations, gas dynamics, entropy solu-

In this state space model, the stochastic system model is represented by the stochastic Equations (4) and (5) and the probability distributions given in Section (2.3); the

We give a new and self-contained proof of the existence and unicity of the flow for an arbitrary (not necessarily homogeneous) smooth vector field on a real supermanifold, and

(v) It is worth mentioning here that the Banach contraction principle [2] and its generalizations give the existence of a unique …xed point for a self map (for instance, Baradol et

Littlewood decomposition on partitions Multiplication- addition theorem for SC, even case Signed refinements The odd case.. Multiplication-addition theorems for

(ii) (conversely to Bercovici and Li’s result): If we know, for all self-adjoints A and B in an arbitrary finite von Neumann algebra M, calling their eigenvalue functions u and v

As application of our coarea inequality we answer this question in the case of real valued Lipschitz maps on the Heisenberg group (Theorem 3.11), considering the Q − 1

The first group contains the so-called phase times, firstly mentioned in 82, 83 and applied to tunnelling in 84, 85, the times of the motion of wave packet spatial centroids,