• Aucun résultat trouvé

The exit from a metastable state: concentration of the exit point distribution on the low energy saddle points, part 2

N/A
N/A
Protected

Academic year: 2021

Partager "The exit from a metastable state: concentration of the exit point distribution on the low energy saddle points, part 2"

Copied!
38
0
0

Texte intégral

(1)

HAL Id: hal-03058529

https://hal.archives-ouvertes.fr/hal-03058529

Preprint submitted on 11 Dec 2020

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.

The exit from a metastable state: concentration of the exit point distribution on the low energy saddle points,

part 2

Tony Lelièvre, Dorian Le Peutrec, Boris Nectoux

To cite this version:

Tony Lelièvre, Dorian Le Peutrec, Boris Nectoux. The exit from a metastable state: concentration of

the exit point distribution on the low energy saddle points, part 2. 2020. �hal-03058529�

(2)

The exit from a metastable state: concentration of the exit point distribution on the low energy saddle points, part 2

Tony Leli` evre , Dorian Le Peutrec and Boris Nectoux

Abstract

We consider the first exit point distribution from a bounded domain Ω of the stochastic process (X t ) t≥0 solution to the overdamped Langevin dynamics

dX t = −∇f (X t )dt + √ h dB t

starting from deterministic initial conditions in Ω, under rather general assumptions on f (for instance, f may have several critical points in Ω). This work is a continuation of the previous paper [14] where the exit point distribution from Ω is studied when X 0 is initially distributed according to the quasi-stationary distribution of (X t ) t≥0 in Ω. The proofs are based on analytical results on the dependency of the exit point distribution on the initial condition, large deviation techniques and results on the genericity of Morse functions.

1 Introduction and main results

The aim of this article is to extend the results of [14] on the concentration of the first exit point distribution from a domain to general initial conditions within the domain. For the sake of consistency, we first recall in Section 1.1 the motivation for such a study, some related works in the literature, and an informal presentation of our results. Section 1.2 then gives precise statements of the results we prove.

1.1 Motivation and informal presentation of the results We are interested in the overdamped Langevin dynamics

dX t = −∇f(X t )dt + √

h dB t , (1)

where X t ∈ R d is a vector in R d , f : R d → R is a C function, h is a positive parameter and (B t ) t≥0 is a standard d-dimensional Brownian motion. The process (1) can be used to model the evolution of molecular systems, for example. In this case, f is the potential function and h > 0 is proportional to the temperature of the system. Let us consider a domain Ω ⊂ R d and the associated exit event from Ω. More precisely, let us introduce

τ = inf{t ≥ 0|X t ∈ / Ω} (2)

CERMICS, ´ Ecole des Ponts, Universit´ e Paris-Est, INRIA, 77455 Champs-sur-Marne, France. E-mail:

tony.lelievre@enpc.fr

Institut Denis Poisson, Universit´ e d’Orl´ eans, Universit´ e de Tours, CNRS, Orl´ eans, France. E-mail:

dorian.le-peutrec@univ-orleans.fr

LMBP, Universit´ e Clermont Auvergne, Aubi` ere, France E-mail: boris.nectoux@uca.fr

(3)

the first exit time from Ω. We are interested in the limit of the law of X τ

when h → 0.

Under some assumptions which will be made precise below, this law concentrates on a subset of ∂Ω, concentration being defined in the following sense.

Definition 1. Let Y ⊂ ∂Ω and let us consider a family of random variable (Y h ) h>0 which admits a limit in distribution when h → 0. The law of Y h concentrates on Y in the limit h → 0 if for every neighborhood V Y of Y in ∂Ω

h→0 lim P [Y h ∈ V Y ] = 1, and if for all x ∈ Y and for all neighborhood V x of x in ∂Ω

h→0 lim P [Y h ∈ V x ] > 0.

In other words, the law of Y h concentrates on Y if Y is the support of the law of Y h in the limit h → 0.

In this work, we investigate the concentration of the law of X τ

on a subset of ∂Ω when X 0 = x ∈ Ω, under general assumptions on the function f : Ω → R. This is of practical interest in order to predict where the process (1) is more likely to leave Ω in the zero-noise limit. The study of the exit event in the small temperature regime has interesting theoretical and numerical counterparts, to relate continuous state space dynamics such as (1) to discrete state space dynamics (jump Markov model), and to accelerate the sampling of metastable trajectories, see [12, 31].

Review of the literature. Let us mention the main results from the mathematical liter- ature on the exit problem related to our problem. We refer to [6] for a more comprehensive review.

The concentration of the law of X τ

on arg min ∂Ω f in the small temperature regime (h → 0) has been studied in [35, 37, 43] through formal computations, see also [33]. Many of these results have been rigorously proven either by studying the underlying partial differential equations, or by using large deviations techniques. In particular, when it holds

∂ n f > 0 on ∂Ω, (3)

where ∂ n f is the normal derivative of f on ∂Ω (n is the unit outward normal vector to D), and

{x ∈ Ω, |∇f (x)| = 0} = {x 0 } with f (x 0 ) = min

f and det Hessf (x 0 ) > 0, (4) and f attains its minimum on ∂Ω at one single point y 0 , it is proved in [17, Theorem 2.1]

that the law of X τ

concentrates on y 0 in the limit h → 0, when X 0 = x ∈ Ω. This result has been extended in [7, 8, 23, 24, 41] when only (3) and (4) are satisfied: it is proved there that the the law of X τ

concentrates in this case on arg min ∂Ω f in the limit h → 0, when X 0 = x ∈ Ω.

In [17, Theorem 5.1], for Σ ⊂ ∂Ω, the limit of h ln P [X τ

∈ Σ] when h → 0 is related

to a minimization problem involving the quasipotential of the process (1). Let us mention

two limitations of [17, Theorem 5.1]. First, this theorem requires to be able to compute the

quasipotential in order to get useful information: this is trivial under the assumptions (3)

and (4) but more complicated under more general assumptions on f (in particular when

(4)

f has several critical points in Ω). Second, even when the quasipotential is analytically known, this result only gives the subset of ∂Ω where exit will not occur on an exponential scale in the limit h → 0. It does not allow to exclude exit points with probability which goes to zero polynomially in h (this may indeed occur, see Remark 6 below and examples in [11, Section 1.4]), and, when the exit point distribution concentrates on more than one point, it does not give the relative probabilities to exit through the various exit points in the limit h → 0. However, let us emphasize that the strength of large deviation theory is that it is very general: f may have several local minima in Ω, and the theory actually applies to a much wider class of dynamics (in particular for non-gradient drift and non- isotropic noise) and in a broader geometric setting [17, 40]. See for example [1, 25, 26] for recent contributions to the non reversible case. Other references where the exit problem appears as an intermediate tool to study spectral properties of the inifinitesimal generator are [2, 3, 34, 36].

Objective of this work. Our work aims at generalizing in the reversible case the results of [17, Theorem 2.1] and [7, 8, 23, 24, 41], when f has several critical points in Ω. First, we exhibit a general set of assumptions on f and an ensemble of initial conditions for which the law of X τ

concentrates on points belonging to arg min ∂Ω f and we compute the relative probabilities to leave the domain through each of them (see Theorem 1):

in this case, the limiting exit point distribution is the same as starting from the quasi- stationary distribution in Ω, and we thus rely on our previous work [14]. Second, using this first result, we identify the exit points when the process starts more generally from initial conditions contained in a potential well which touches the boundary of Ω, in a sense to be made precise (see Theorem 2 and Theorem 4).

Concerning the assumptions on f, ∂ n f is not assumed to be positive on ∂Ω and f may have several critical points in Ω with larger energies than min ∂Ω f . However, we do not consider the case when f has critical points on ∂Ω.

Here are representative examples of outputs of this work, which are new to the best of our knowledge. Let us assume that f : Ω → R and its restriction f | ∂Ω to the boundary of Ω are smooth Morse functions, and that f has no critical point on ∂Ω (see (A0) below).

• We prove that if {y ∈ Ω, f(y) < min ∂Ω f} is connected and contains all the local minima of f in Ω, and if ∂ n f > 0 on arg min ∂Ω f, then the law of X τ

concentrates on arg min ∂Ω f when X 0 = x ∈ {y ∈ Ω, f(y) < min ∂Ω f } (see Theorem 1).

• Besides, when {y ∈ Ω, f (y) < min ∂Ω f } is not connected (we denote in this case by C 1 , . . . , C N its connected components, with N ≥ 2) and if, for some j ∈ {1, . . . , N},

∂C j ∩ ∂Ω 6= ∅ and |∇f | 6= 0 on ∂C j , then, when X 0 = x ∈ C j , the law of X τ

concentrates on ∂C j ∩ ∂Ω in the limit h → 0 (see Theorem 2).

• Furthermore, when X 0 = x ∈ C j and z ∈ arg min ∂Ω f \ ∂C j , for all sufficiently small neighborhood Σ z of z in ∂Ω, P [X τ

∈ Σ z ] = O(e

hc

) for h small enough (see item 1 in Theorem 2).

Let us mention that the preprint [11] concatenates most of the results of this manuscript and [14]. A simplified version of the results of this work is also presented in [32].

On the results from [14]: metastability and the quasi-stationary distribution.

As explained above, this article generalizes to a broader class of initial conditions in Ω the

(5)

results of [14] where it is assumed that X 0 is distributed according to the quasi-stationary distribution in Ω. Let us quickly recall what is the quasi-stationary distribution, and why it is relevant to study the exit event starting from this distribution.

Let us assume that Ω ⊂ R d is smooth, open, bounded and connected.

Definition 2. A quasi-stationary distribution for the dynamics (1) in Ω is a probability measure ν h supported in Ω such that for all measurable sets A ⊂ Ω and for all t ≥ 0

ν h (A) = Z

P x [X t ∈ A, t < τ Ω ] ν h (dx) Z

P x [t < τ ] ν h (dx)

. (5)

Here and in the following, the subscript x indicates that the stochastic process starts from x ∈ R d (X 0 = x). In words, (5) means that if X 0 is distributed according to ν h , then for all t > 0, X t is still distributed according to ν h conditionally on X s ∈ Ω for all s ∈ [0, t].

The quasi-stationary distribution is related to an eigenvalue problem on the infinites- imal generator of the dynamics (1), namely the differential operator

L f,h = − h

2 ∆ + ∇f · ∇. (6)

Let us define

L 2 w (Ω) = n

u : Ω → R, Z

u 2 e

2h

f < ∞ o .

The weighted Sobolev spaces H w k (Ω) are defined similarly. The operator L f,h with homo- geneous Dirichlet boundary conditions on ∂Ω is denoted by L D f,h . Its domain is D(L D f,h ) = H w,0 1 (Ω) ∩ H w 2 (Ω), where H w,0 1 (Ω) = {u ∈ H w 1 (Ω), u = 0 on ∂Ω}. It is well know that

−L D f,h is self adjoint on L 2 w (Ω), positive and has compact resolvent. Moreover, from stan- dard results on elliptic operator (see for example [16,19]), its smallest eigenvalue λ h (a.k.a.

the principal eigenvalue) is non degenerate and its associated eigenfunction u h has a sign on Ω and is in C (Ω). Without loss of generality, one can then assume that:

u h > 0 on Ω and Z

u 2 h e

h2

f = 1. (7)

The following result (see for example [28]) relates the quasi-stationary distribution ν h to the principal eigenfunction u h .

Proposition 3. The unique quasi-stationary distribution ν h associated with the dynam- ics (1) and the domain Ω is given by:

ν h (dx) = u h (x)e

2h

f (x) Z

u h (y)e

2h

f (y) dy

dx. (8)

Moreover, whatever the law of the initial condition X 0 with support in Ω, it holds:

t→∞ lim kLaw(X t |t < τ ) − ν h k T V = 0. (9)

(6)

Here, Law(X t |t < τ ) denotes the law of X t conditional to the event {t < τ }. For a given initial distribution of the process (1), if the convergence in (9) is much quicker than the exit from Ω, the exit from the domain Ω is said to be metastable. In [14], we have investigated the concentration of the law of X τ

on arg min ∂Ω f in the limit h → 0 when X 0 ∼ ν h , namely when the exit is metastable. In this work, we extend this study to the general case: X 0 = x ∈ Ω.

1.2 Main results

In all this work, we assume that Ω ⊂ R d is smooth, open, bounded and connected, and that f : Ω → R is a C function. This section is dedicated to the statement of the main result of this work.

1.2.1 Assumptions

Let us now introduce the basic assumption which is used throughout this work:

The function f : Ω → R is a C Morse function.

For all x ∈ ∂Ω, |∇f (x)| 6= 0.

The function f : {x ∈ ∂Ω, ∂ n f(x) > 0} → R is a Morse function.

The function f has at least one local minimum in Ω.

 

 

 

 

(A0)

Remark 4. We recall that a function φ : Ω → R is a Morse function if all its critical points are non degenerate (which implies in particular that φ has a finite number of critical points since Ω is compact and a non degenerate critical point is isolated from the other critical points). Let us recall that a critical point z ∈ Ω of φ is non degenerate if the hessian matrix of φ at z, denoted by Hess φ(z), is invertible. We refer for example to [22, Definition 4.3.5]

for a definition of the hessian matrix on a manifold (see also [13, Remark 10] for explicit formulas). A non degenerate critical point z ∈ Ω of φ is said to have index p ∈ {0, . . . , d}

if Hess φ(z) has precisely p negative eigenvalues (counted with multiplicity). In the case p = 1, z is called a saddle point.

In order to introduce the remaining assumptions, we need to introduce three notations.

First, the following notation will be used for the level sets of f : for a ∈ R, {f < a} = {x ∈ Ω, f(x) < a}, {f ≤ a} = {x ∈ Ω, f (x) ≤ a}, and

{f = a} = {x ∈ Ω, f (x) = a}.

Second, for any local minimum x of f in Ω, one defines H f (x) := inf n

t∈[0,1] max f γ(t) γ ∈ C 0 ([0, 1], Ω), γ(0) = x, and γ(1) ∈ ∂Ω o

, (10)

where C 0 ([0, 1], Ω) is the set of continuous paths from [0, 1] to Ω. Intuitively, H f (x) is

the minimal energy any path connecting x to the boundary ∂Ω has to cross. This energy

is necessarily either the energy of a saddle point z in Ω (see e.g. ∂C 2 on Figure 1), or

the energy of a generalized saddle point z on ∂Ω (see e.g. ∂Ω ∩ ∂C 3 and ∂Ω ∩ ∂C max on

(7)

Figure 1; see also Equation (15) for a proper definition of a generalized saddle point).

Third, for a local minimum x of f in Ω,

C(x) is the connected component of {f < H f (x)} containing x. (11) Finally, set

C :=

C(x), x is a local minimum of f in Ω . (12) Let us mention that when (A0) holds, one has: for all local minima x of f in Ω, C(x) is an open subset of Ω (see [11, Remark 7]).

We are now in position to state the assumptions we will use in addition to (A0):

• First geometric assumption:

(A0) holds and ∃!C max ∈ C s.t. max

C∈C

n max

C

f − min

C

f o

= max

C

max

f − min

C

max

f. (A1)

• Second geometric assumption:

(A1) holds and ∂C max ∩ ∂Ω 6= ∅. (A2)

• Third geometric assumption:

(A1) holds and ∂C max ∩ ∂Ω ⊂ arg min

∂Ω

f. (A3)

Assumption (A1) implies that there is a unique deepest well, namely C max . Assump- tions (A2) and (A3) mean that the closure of this deepest well intersects ∂Ω at points where f reaches its minimum on ∂Ω. Notice that these assumptions imply that C max con- tains the global minima of f in Ω. Assumptions (A2) and (A3) ensure that when X 0 ∼ ν h

h being the quasi-stationary distribution introduced in Definition 2) or X 0 = x ∈ C max , the law of X τ

concentrates on the set ∂C max ∩ ∂Ω, (see [14, Theorem 1] and items 1 and 2 in Theorem 1 below). Finally, the last assumption is:

• Fourth geometric assumption:

(A1) holds and ∂C max ∩ Ω contains no separating saddle point of f, (A4) where the proper definition of a separating saddle point of f is introduced below in Section 1.2.2.

Assumption (A4) is equivalent to the fact that any minimal energy path connecting a point in C max to the boundary ∂Ω remains necessarily within C max . In particular, such a path leaves Ω on ∂C max ∩ ∂Ω. Notice indeed that Assumption (A4) implies

∂C max ∩ ∂Ω 6= ∅ (this is a consequence of [14, Proposition 15]). The assumptions (A4)

together with (A0), (A1), (A2), and (A3), ensure that the probability that the pro-

cess (1) starting from x ∈ C max leaves Ω through any sufficiently small neighborhood

of z ∈ ∂Ω \ ∂C max in ∂Ω is exponentially small when h → 0, see item 3 in Theorem 1

below. We refer to Remark 6 below for a discussion on the necessity of the assump-

tions (A0), (A1), (A2), (A3), and (A4) to get these results. Figure 1 gives a one-

dimensional example where (A1), (A2), (A3), and (A4) are satisfied, and Figure 2 gives

a typical example where (A1), (A2), and (A3) are satisfied but not (A4).

(8)

Ω C max

C 2

C 3

H f (x 1 ) − f (x 1 )

∂Ω ∩ ∂C max

x 1

x 2

x 3

x 5 x 4

Figure 1: A one-dimensional case where (A1), (A2), (A3) and (A4) are satisfied. On the figure, f (x 1 ) = f (x 5 ), H f (x 1 ) = H f (x 4 ) = H f (x 5 ), C = {C max , C 2 , C 3 }, ∂C 2 ∩ ∂C max = ∅ and

∂C 3 ∩ ∂C max = ∅.

C max

C 2

C 3

∂Ω z 5

z 4

x 1

x 2

z 1

z 3

z 2

x 3

∂Ω f | ∂Ω

z 3

z 1 z 2

z 4

Figure 2: Schematic representation of C (see (12)) and f | ∂Ω when the assumptions (A0), (A1), (A2) and (A3) are satisfied. In this representation, x 1 ∈ Ω is the global minimum of f in Ω and the other local minima of f in Ω are x 2 and x 3 . Moreover, min ∂Ω f = f (z 1 ) = f (z 2 ) = f (z 3 ) = H f (x 1 ) = H f (x 2 ) < H f (x 3 ) = f (z 4 ), {f < H f (x 1 )}

has two connected components: C max (see (A1)) which contains x 1 and C 2 which con-

tains x 2 . Thus, one has C = {C max , C 2 , C 3 }. In addition, {z 1 , z 2 , z 3 } = arg min ∂Ω f

(k ∂Ω 1 = 3), {z 5 } = C max ∩ C 2 is a separating saddle point of f (thus (A4) is not satisfied),

and ∂C max ∩ ∂Ω = {z 1 , z 2 } (k ∂C 1

max

= 2).

(9)

1.2.2 Notation associated with the function f

In order to state our main result, we will need a few more notation associated with the function f.

Domain of attraction A(D). The domain of attraction of a subset D of Ω for the dynamics ˙ x = −∇f(x) is defined as follows. Let x ∈ Ω and denote by ϕ t (x) the solution to the ordinary differential equation

d

dt ϕ t (x) = −∇f (ϕ t (x)) with ϕ 0 (x) = x, (13) on the interval t ∈ [0, t x ], where t x > 0 is defined by

t x = inf{t ≥ 0, ϕ t (x) ∈ / Ω}.

Let x ∈ Ω be such that t x = +∞. The ω-limit set of x, denoted by ω(x), is defined by ω(x) = {y ∈ Ω, ∃(s n ) n∈ N ∈ ( R + ) N , lim

n→∞ s n = +∞, lim

n→∞ ϕ s

n

(x) = y}.

Let us recall that the ω-limit set ω(x) is included in the set of the critical points of f in Ω. Moreover, when f has a finite number of critical points in Ω, ω(x) is either empty (if t x < ∞) or of cardinality one (if t x = ∞). Let D be a subset of Ω. The domain of attraction of a subset D of Ω is defined by

A(D) = {x ∈ Ω, t x = +∞ and ω(x) ⊂ D}. (14) Remark 5. Recall that when (A0) holds, one has: for all local minima x of f in Ω, C(x) ⊂ Ω, where C(x) is defined in (11) (see [11, Remark 7]). This implies that for all y ∈ C(x), t y = +∞ and then, C(x) ⊂ A(C(x)).

Generalized saddle points of f . We introduce in this paragraph an ensemble of points on ∂Ω, which will contain the exit points of the process (X t ) t≥0 from Ω. Let us define

U ∂Ω 1 := {z ∈ ∂Ω, z is a local minimum of f | ∂Ω but not a local minimum of f in Ω}.

Notice that an equivalent definition of U ∂Ω 1 is

U ∂Ω 1 = {z ∈ ∂Ω, z is a local minimum of f | ∂Ω and ∂ n f (z) > 0}, (15) which follows from the fact that |∇f (x)| 6= 0 for all x ∈ ∂Ω. The set U ∂Ω 1 contains the so- called generalized saddle points of f on the boundary ∂Ω [20,29]. These generalized saddle points are indeed geometrically saddle points of the function f extended by −∞ outside Ω, which is consistent with the fact that homogeneous Dirichlet boundary conditions are used to define the operator L D f,h .

Let us introduce the following notation for the k ∂Ω 1 elements of U ∂Ω 1 ∩ arg min ∂Ω f : {z 1 , . . . , z k

∂Ω

1

} = U ∂Ω 1 ∩ arg min

∂Ω

f. (16)

Let us assume that the assumptions (A1), (A2), and (A3) are satisfied. Let us recall that C max is defined in (A1). Moreover, in this case, one has k ∂Ω 1 ≥ 1 and

∂C max ∩ ∂Ω ⊂ {z 1 , . . . , z k

∂Ω 1

}.

(10)

Indeed, by assumption ∂C max ∩ ∂Ω ⊂ {f = min ∂Ω f } (see (A3)) and there is no local minimum of f in Ω on ∂C max (since C max is a connected component of a sublevel set of f).

We assume lastly that the set {z 1 , . . . , z k

∂Ω

1

} is ordered such that:

{z 1 , . . . , z k

∂Cmax 1

} = {z 1 , . . . , z k

∂Ω

1

} ∩ ∂C max . (17)

Notice that k ∂C 1

max

∈ N and k ∂C 1

max

≤ k ∂Ω 1 .

Separating saddle points. Let us finally introduce the notion of separating saddle point, following [21, Section 4.1]. To this end, let us first recall that according to [20, Section 5.2], for any non critical point z ∈ Ω, for r > 0 small enough

{f < f(z)} ∩ B(z, r) is connected,

and for any critical point z ∈ Ω of index p of the Morse function f , for r > 0 small enough, one has the three possible cases:

• either p = 0 (i.e. z is a local minimum of f ) and {f < f(z)} ∩ B(z, r) = ∅,

• or p = 1 and {f < f(z)} ∩ B(z, r) has exactly two connected components,

• or p ≥ 2 and {f < f(z)} ∩ B(z, r) is connected,

where B(z, r) := {x ∈ Ω s.t. |x − z| < r}. A separating saddle point of f in Ω is a saddle point of f in Ω such that for any r > 0 small enough, the two connected components of {f < f(z)} ∩ B(z, r) are contained in different connected components of {f < f(z)}.

Figure 2 gives an illustration of the notations introduced in this section.

1.2.3 Main results on the exit point distribution

In Theorem 1, we first make explicit a geometric setting and an ensemble of initial condi- tions for which the exit distribution is the same as when starting from the quasi-stationary distribution.

Theorem 1. Let us assume that the assumptions (A0), (A1), (A2), and (A3) are satis- fied. Let F ∈ L (∂Ω, R ) and (Σ i ) i∈{1,...,k

∂Ω

1

} be a family of pairwise disjoint open subsets of ∂Ω (i.e. such that Σ i ∩ Σ j = ∅ whenever i 6= j) such that

for all i ∈

1, . . . , k ∂Ω 1 , z i ∈ Σ i , where we recall that

z 1 , . . . , z k

∂Ω

1

= U ∂Ω 1 ∩ arg min ∂Ω f (see (16)). Let K be a compact subset of Ω such that K ⊂ A(C max ) (see (A1) and (14) for the definitions of C max and A).

Let x ∈ K. Then:

1. There exists c > 0 such that in the limit h → 0:

E x [F (X τ

)] =

k

∂Ω1

X

i=1

E x [(1 Σ

i

F ) (X τ

)] + O e

hc

(18) and

k

∂Ω1

X

i=k

∂Cmax1

+1

E x [(1 Σ

i

F ) (X τ

)] = O h

14

, (19)

where we recall that

z 1 , . . . , z k

∂Cmax

1

= ∂C max ∩ ∂Ω (see (17)).

(11)

2. When for some i ∈

1, . . . , k ∂C 1

max

the function F is C in a neighborhood of z i , one has when h → 0:

E x [(1 Σ

i

F ) (X τ

)] = F (z i ) a i + O(h

14

), (20) where

a i = ∂ n f (z i ) p det Hessf| ∂Ω (z i )

k

∂Cmax1

X

j=1

n f (z j ) p det Hessf | ∂Ω (z j )

−1

. (21)

3. When (A4) is satisfied the remainder term O(h

14

) in (19) is of the order O e

hc

for some c > 0 and the remainder term O h

14

in (20) is of the order O(h) and admits a full asymptotic expansion in h (see (22) below).

Finally, the constants involved in the remainder terms in (18), (19), and (20) are uniform with respect to x ∈ K.

Let us recall that for α > 0, (r(h)) h>0 admits a full asymptotic expansion in h α if there exists a sequence (a k ) k≥0 ∈ R N such that for any N ∈ N , it holds in the limit h → 0:

r(h) =

N

X

k=0

a k h αk + O h α(N +1)

. (22)

According to (18), when the function F belongs to C (∂Ω, R ) and x ∈ A(C max ), one has in the limit h → 0:

E x [F (X τ

)] =

k

∂Cmax1

X

i=1

a i F (z i ) + O(h

14

) =

k

∂Cmax1

P

i=1

R

Σ

i

F ∂ n f e

h2

f

k

∂Cmax1

P

i=1

R

Σ

i

n f e

2h

f

+ o h (1),

where the order in h of the remainder term o h (1) depends on the support of F and on whether or not the assumption (A4) is satisfied. This is reminiscent of previous results obtained in [7, 8, 23, 24, 41].

Theorem 1 implies that when X 0 = x ∈ A(C max ), the law of X τ

concentrates on {z 1 , . . . , z k

Cmax

1

} = ∂Ω ∩ ∂C max in the limit h → 0, with explicit formulas for the probabil- ities to exit through each of the z i ’s. Moreover, the probability to exit through a global minimum z of f | ∂Ω which satisfies ∂ n f (z) < 0 is exponentially small in the limit h → 0 (see (18)) and when assuming (A4), the probability to exit through z k

Cmax

1

+1 , . . . , z k

∂Ω 1

is also exponentially small even though all these points belong to arg min ∂Ω f . Theorem 1 is thus a generalization of [14, Theorem 1] to other initial conditions than the quasi-stationary distribution ν h .

Remark 6. Assumptions (A0) and (A1) ensure that C max appearing in Theorem 1 is

well defined. Concerning the assumption (A2), there exist functions f satisfying (A0)

and (A1) but not (A2) such that when X 0 = x ∈ C max , the law of X τ

concentrates on

points which do not belong to arg min ∂Ω f (see indeed [11, Figure 4]). The same holds for

the assumption (A3): there exist functions f which satisfy (A0), (A1), and (A2) but

not (A3) such that when X 0 = x ∈ C max , the law of X τ

concentrates on points which do

(12)

not belong to arg min ∂Ω f (see Figure 5 and [11, Figure 5]). Thus, (A2) and (A3) are necessary to ensure that the law of X τ

concentrates on points belonging to arg min ∂Ω f.

Finally, assumption (A4) is necessary to get item 3 in Theorem 1. Indeed, in [11, Sec- tion 1.4], we provide examples of functions f satisfying (A0), (A1), (A2), and (A3) but not (A4), such that for all sufficiently small neighborhood Σ z of z ∈ arg min ∂Ω f \ ∂C max

in ∂Ω, P [X τ

∈ Σ z ] = C √

h (1 + o(1)) in the limit h → 0, for some C > 0 independent of h and when X 0 = x ∈ C max .

Remark 7. When x ∈ Ω is such that t x < +∞, it is a simple consequence of the large deviations estimate (52) below that, in the limit h → 0, the process (1) almost surely exits Ω through any neighborhood of ϕ t

x

(x) in ∂Ω.

It is also possible to describe the exit point distribution when X 0 = x ∈ A(C) and C ∈ C is not necessarily C max (we recall that C is defined in (12)). This is the objective of Theorem 2, whose proof uses Theorem 1 applied to a suitable subdomain of Ω containing C.

Theorem 2. Let us assume that (A0) holds. Let C ∈ C. Let us assume that

∂C ∩ ∂Ω 6= ∅ and |∇f | 6= 0 on ∂C. (23) Recall that ∂C ∩ ∂Ω ⊂ U ∂Ω 1 (see (15) for a definition of U ∂Ω 1 ). Let F ∈ L (∂Ω, R). For all z ∈ ∂C ∩ ∂Ω, let Σ z be an open subset of ∂Ω such that z ∈ Σ z . Let K be a compact subset of Ω such that K ⊂ A(C). Then:

1. There exists c > 0 such that for h small enough, sup

x∈K E x

h

F 1 ∂Ω\ S

z∈∂C∩∂Ω

Σ

z

(X τ

) i

≤ e

hc

.

Assume moreover that the sets (Σ z ) z∈∂C∩∂Ω are pairwise disjoint (i.e. such that Σ z ∩Σ z

0

=

∅ whenever z 6= z 0 ). Let z ∈ ∂C ∩ ∂Ω.

2. If F is C in a neighborhood of z, it holds for all x ∈ K,

E x [(F 1 Σ

z

) (X τ

)] = F (z) ∂ n f (z) p det Hessf | ∂Ω (z)

 X

y∈∂C∩∂Ω

n f (y) p det Hessf | ∂Ω (y)

−1

+O(h), in the limit h → 0 and uniformly in x ∈ K.

Theorem 2 implies that when C ∈ C satisfies (23) (for instance, this is the case for C 3 on Figure 2), the law of X τ

when X 0 = x ∈ A(C) concentrates when h → 0 on ∂C ∩ ∂Ω.

Theorems 1 and 2 are actually special cases of a more general result which will be

stated and illustrated in Section 3.3 (see indeed Theorem 4). The proof of Theorem 4

is a simple extension of the proof of Theorem 2. For pedagogical purposes, we prefer

to first present Theorem 1 and 2, before stating the more general and abstract result of

Theorem 4.

(13)

1.3 Organization of the paper

The rest of this paper is dedicated to the proofs of Theorem 1, in Section 2 and of The- orems 2 and 4, in Section 3. The proof of Theorem 1 heavily relies on results from [14]

(recalled in Theorem 3 below), together with a so-called leveling result on x 7→ E x [F (X τ

)].

The proof of Theorem 2 uses Theorem 1 applied to a domain Ω C which contains C. The construction of this domain uses tools from differential topology related to the genericity of Morse functions. Finally, Section 4 gives conclusions and perspectives.

2 Proof of Theorem 1

After recalling some results from [14] in Section 2.1, we prove a so-called leveling result (as initially introduced in [4]) in C max for x 7→ E x [F (X τ

)], in Section 2.2. Then, combining the results of these two sections, one proves Theorem 1 in Section 2.3 for a smooth function F and finally for a measurable bounded F in Section 2.4.

2.1 Previous results on the principal eigenfunction of L D f,h and on ν h Let us recall the following result from [14, Theorem 4] on the spectral gap of L D f,h . We recall that λ h is the principal eigenvalue of −L D f,h (see Section 1.1).

Proposition 8. Assume that the assumptions (A0) and (A1) are satisfied. Let us denote by f max the value of f on ∂C max and x max a minimum point of f in C max :

f ≡ f max on ∂C max and x max ∈ arg min

C

max

f. (24)

Notice that f max = max C

max

f and x max ∈ C max . Then, there exists C > 1 and γ ∈ R , such that, for h small enough,

C −1 h γ e

h2

(f

max

−f(x

max

)) ≤ λ h ≤ C h γ e

h2

(f

max

−f(x

max

)) .

Moreover, there exists c > 0 such that for h small enough, min σ(L D f,h ) \ {λ h } ≥ e

hc

λ h . We now give a series of three results which are consequences of the previous proposition.

These results can all be found in [14]. We provide the proofs for these results since they are short and they are opportunities to introduce some notation which will be needed later on.

A direct corollary of Proposition 8 is the following.

Corollary 9. Let us assume that the assumptions (A0) and (A1) are satisfied. Then, there exists β 0 > 0 such that for all β ∈ (0, β 0 ), there exists h 0 > 0 such that for all h ∈ (0, h 0 ), the orthogonal projector in L 2 w (Ω)

π h := π

0,e

h2(fmax−f(xmax)−β)

(L D f,h ) has rank 1.

Here and in the following, for a Borel set E ⊂ R , π E (L D f,h ) is the spectral projector of L D f,h on E.

For α > 0, one defines

C max (α) = C max

f < f max − α . (25)

(14)

For α > 0, let χ ∈ C (Ω) be such that

χ = 0 on Ω \ C max (α) and χ = 1 on C max (2α).

Using Corollary 9, we obtain a good approximation of u h using the function χ.

Proposition 10. Let us assume that the assumptions (A0) and (A1) are satisfied. Let us define

ψ := χ

kχk L

2 w

(Ω)

.

Then, for all α > 0 small enough, there exists c > 0 such that in the limit h → 0:

u h = ψ + O(e

hc

) in L 2 w (Ω).

Proof. Let us recall that for all u ∈ H w,0 1 (Ω) and b > 0,

π [b,+∞) (L D f,h ) u

2 ≤

h 2

R

Ω |∇u| 2 e

2h

f

b .

Thus, one has:

(1 − π h )ψ

2

L

2w

(Ω) ≤ e

h2

(f

max

−f(x

max

)−β) h 2

∇ψ

2

L

2w

(Ω) . (26)

Moreover, by definition of the function ψ, it holds, ∇ψ

2 L

2w

(Ω) =

R

Ω |∇χ| 2 e

2h

f R

Ω χ 2 e

2h

f

. (27)

Since χ = 0 on Ω \ C max (α), it holds, R

Ω χ 2 e

h2

f = R

C

max

(α) χ 2 e

2h

f . Since f has finite number of global minima in C max which are all included in C max , one deduces that for all α > 0 small enough,

arg min

C

max

f ⊂ C max (2α). (28)

Consequently, using a Laplace’s method together with the fact that χ = 1 on C max (2α), one has in the limit h → 0,

Z

χ 2 e

2h

f = h

d2

π

d2

e

h2

f(x

max

) X

x∈arg min

Cmax

f

det Hessf (x) −

1

2

1 + O(h)

. (29)

Since the support of ∇χ is included in C max (α) \ C max (2α) (thus f ≥ f max − 2α on the support of ∇χ it holds for some C > 0:

Z

|∇χ| 2 e

2h

f ≤ C h 1−

d2

e

h2

(f

max

−2α) .

Plugging these estimates in (27) and using (26), one finally deduces that for k(1 − π h )ψk 2 L

2

w

(Ω) ≤ C e

2h

(β−2α) . Choosing α < β/4, this implies that for h small enough,

(1 − π h

L

2w

(Ω) ≤ e

βh

. There- fore, using Corollary 9 and the fact that the functions u h and ψ are non negative,

u h = π h ψ kπ h ψk L

2

w

(Ω)

= ψ + O(e

hc

) in L 2 w (Ω)

for some positive c. This concludes the proof of Proposition 10.

(15)

We end this section, by the following consequence of Proposition 10.

Corollary 11. Let us assume that the assumptions (A0) and (A1) are satisfied. Let us moreover assume that

min

C

max

f = min

f.

Then, when h → 0:

Z

u h e

2h

f = h

d4

π

d4

e

1h

min

f X

x∈arg min

Cmax

f

det Hessf (x) −

1

2

12

1 + O(h) .

Moreover, if O is an open subset of Ω such that O ∩ arg min C

max

f = ∅, then, there exists c > 0 such that for h small enough,

Z

O

u h e

2h

f = O e

1h

(min

f+c) .

Proof. Let O be an open subset of Ω. Using Proposition 10 and thanks to the Cauchy- Schwarz inequality, one obtains in the limit h → 0:

Z

O

u h e

h2

f = Z

O

ψ e

h2

f + O(e

hc

) s

Z

O

e

h2

f = Z

O

ψ e

h2

f + O

e

1h

(min

f+c)

.

Then, the first statement in Corollary 11 follows choosing O = Ω in the previous equal- ity and using (29) (notice that the same estimate holds replacing χ 2 by χ in (29)), the definition of ψ (see Proposition 10) together with the fact that by assumption f (x max ) = min C

max

f = min f. The second statement in Corollary 11 follows using (29) and the fact that when O ∩ arg min C

max

f = ∅, there exists c > 0 such that for h small enough, R

O χ e

2h

f = O e

2h

(min

f+c) .

We end this section by recalling the results of [14, Theorem 1] on the law of X τ

when X 0

is initially distributed according to the quasi-stationary distribution ν h of the process (1).

Theorem 3. Let us assume that the assumptions (A0), (A1), (A2), and (A3) are satis- fied. Let F ∈ L (∂Ω, R) and (Σ i ) i∈{1,...,k

∂Ω

1

} be a family of pairwise disjoint open subsets of ∂Ω such that

for all i ∈

1, . . . , k ∂Ω 1 , z i ∈ Σ i , where we recall that

z 1 , . . . , z k

∂Ω

1

= U ∂Ω 1 ∩ arg min ∂Ω f (see (16)). Then:

1. There exists c > 0 such that when h → 0:

E ν

h

[F (X τ

)] =

k

∂Ω1

X

i=1

E ν

h

[(1 Σ

i

F ) (X τ

)] + O e

ch

(30) and

k

∂Ω1

X

i=k

∂Cmax1

+1

E ν

h

[(1 Σ

i

F ) (X τ

)] = O h

14

, (31)

where we recall that

z 1 , . . . , z k

∂Cmax

1

= ∂C max ∩ ∂Ω (see (17)).

(16)

2. When for some i ∈

1, . . . , k ∂C 1

max

the function F is C in a neighborhood of z i , one has when h → 0:

E ν

h

[(1 Σ

i

F ) (X τ

)] = F (z i ) a i + O(h

14

) where a i is defined by (21). (32) 3. When (A4) is satisfied the remainder term O(h

14

) in (19) is of the order O e

hc

for some c > 0 and the remainder term O h

14

in (20) is of the order O(h) and admits a full asymptotic expansion in h.

2.2 Leveling results

To go from Theorem 3 to Theorem 1, the basic idea is to write E ν

h

[F (X τ

)] =

Z

E x [F (X τ

)]ν h (dx)

and to use the fact that the function x 7→ E x [F (X τ

)] is “more and more constant” as h → 0: this is called a leveling property.

Definition 12. Let K be a compact subset of Ω and F ∈ C (∂Ω, R ). We say that x 7→ E x [F (X τ

)] satisfies a leveling property on K if

h→0 lim

E x [F (X τ

)] − E y [F (X τ

)]

= 0 (33) and this limit holds uniformly with respect to (x, y) ∈ K × K.

The leveling property (33) has been widely studied in the literature in various geometrical settings, see for example [4, 10, 15, 17, 23, 24, 41]. We prove the following proposition which is a leveling property in our framework.

Proposition 13. Let us assume that the assumption (A0) holds. Let λ ∈ R and C be a connected component of {f < λ} such that C ⊂ Ω. Then, for any path-connected compact set K ⊂ C and for any F ∈ C (∂Ω, R ), there exist c > 0 and M > 0, such that for all (x, y) ∈ K × K,

E x [F (X τ

)] − E y [F (X τ

)]

≤ M e

ch

. (34) Proof. The proof is inspired from techniques used in [10]. The proof of Proposition 13 is divided into two steps. In the following C > 0 is a constant which can change from one occurrence to another and which does not depend on h.

Step 1. Let F ∈ C (∂Ω, R). Let us denote by v h ∈ H 1 (Ω) the unique weak solution to the elliptic boundary value problem

 h

2 ∆v h − ∇f · ∇v h = 0 on Ω v h = F on ∂Ω.

(35)

Then, we will prove in this step that v h belongs to C (Ω, R ) and that for all k ∈ N , there exist C > 0, n ∈ N and h 0 > 0 such that for all h ∈ (0, h 0 ), it holds

kv h k H

k+2

(Ω) ≤ C

h n k∇v h k L

2

(Ω) + 1

. (36)

(17)

Here L 2 (Ω) =

u : Ω → R , R

Ω u 2 < ∞ , and, for k ≥ 1, H k (Ω) =

u : Ω → R , ∀α ∈ N k such that |α| ≤ k, Z

|∂ α u| 2 < ∞ .

Moreover, the Dynkin’s formula implies that

∀x ∈ Ω, v h (x) = E x [F (X τ

)] . (37) Let us prove that v h belongs to C (Ω, R ) and (36). Since F is C , for all k ≥ 1, there exists F e ∈ H k (Ω) such that F e = F on ∂Ω and

k F e k H

k

≤ CkFk

H

k−12

(∂Ω) .

From (35), the function e v h = v h − F e ∈ H 1 (Ω) is the weak solution to

∆ e v h = 2

h ∇f · ∇v h − ∆ F e on Ω e v h = 0 on ∂Ω.

(38)

Thus, using [16, Theorem 5, Section 6.3], v e h ∈ H 2 (Ω) (and thus v h ∈ H 2 (Ω)) and there exist C > 0 and h 0 > 0 such that for all h ∈ (0, h 0 )

k e v h k H

2

(Ω) ≤ C 1

h k∇v h k L

2

(Ω) + k Fk e H

2

(Ω)

≤ C h

k∇v h k L

2

(Ω) + kF k

H

32

(∂Ω)

,

and thus

kv h k H

2

(Ω) ≤ C

h k∇v h k L

2

(Ω) + 1

. (39)

This proves (36) for k = 0. The inequality (36) is then obtained by a bootstrap argument, by induction on k. This implies by Sobolev embeddings that v h belongs to C (Ω, R ).

Let us now prove that there exist α > 0 and C > 0 such that

kv h k L

(Ω) + k∇v h k L

(Ω) ≤ Ch −α . (40) Notice that from (37), one has that for all h > 0, kv h k L

(Ω) ≤ kFk L

(∂Ω) . Using this bound, (35) and (39), there exists C > 0 such that for any ε > 0 and ε 0 > 0,

h Z

|∇v h | 2 ≤ C

h Z

∂Ω

|F ∂ n v h | dσ + kF k L

(∂Ω)

Z

|∇f · ∇v h |

≤ C h

kF k 2 L

2

(∂Ω)

4ε + h ε kv h k 2 H

2

(Ω) +

k∇fk 2 L

2

(Ω)

0 + ε 0 k∇v h k 2 L

2

(Ω)

!

≤ C h

kF k 2 L

2

(∂Ω)

4ε + C h ε C 1 h −2

k∇v h k 2 L

2

(Ω) + 1

+ C

k∇f k 2 L

2

(Ω)

0 + C ε 0 k∇v h k 2 L

2

(Ω) . Choosing ε = 4(CC h

2

1

+1) and ε 0 = 4(C+1) h we get k∇v h k L

2

(Ω) ≤ C

h .

(18)

Therefore, using (36), one obtains that for all k ≥ 0, there exist C > 0, n ∈ N and h 0 > 0 such that for all h ∈ (0, h 0 )

kv h k H

k

(Ω) ≤ C h n .

Let k ≥ 0 such that k − d 2 > 1. Then, one obtains (40) from the continuous Sobolev injection H k (Ω) ⊂ W 1,∞ (Ω).

Step 2. Let us assume that (A0) holds. Let λ ∈ R and C be a connected component of {f < λ} such that C ⊂ Ω. Let us now define the set C r by

C r = {f < λ − r} ∩ C ⊂ Ω (41) which is not empty and C for all r ∈ (0, r 1 ), for some r 1 > 0. Indeed, the boundary of C r is the set {f = λ − r} ∩ C (since C ⊂ Ω) which contains no critical points of f for r ∈ (0, r 1 ), with r 1 > 0 small enough (since there is a finite number of critical points under the assumption (A0)). In this step, we will prove that for all r 0 ∈ (0, r 1 ) there exists α 0 > 0 such that

k∇v h k L

(C

r

0

) ≤ e

αh0

. (42)

Equation (42) implies that for any compact subset K of C, there exist c > 0 and C > 0 such that

∀(x, y) ∈ K × K, |v h (x) − v h (y)| ≤ Ce

hc

,

which will then conclude the proof of (34). This follows from the fact that there exists r 0 ∈ (0, r 1 ) such that K ⊂ C r

0

.

Let us now prove (42). Let r be such that 2 n r = r 0 where n ∈ N will be fixed later (since r ≤ r 0 , r ∈ (0, r 1 )). Equation (35) rewrites

( div e

2h

f ∇v h

= 0 on Ω v h = F on ∂Ω.

Using (40), there exist C > 0 and α > 0 such that,

Z

C

r/2

|∇v h | 2 e

h2

f

=

Z

∂C

r/2

e

h2

f v h ∂ n v h dσ

≤ C

h α e

2h

(λ−

r2

) ,

where we used the Green formula (valid since C r is C for all r ∈ (0, r 1 )) and the inclusion

∂C r/2 ⊂ {f = λ − r 2 }. In addition, since C r ⊂ C r/2 it holds, e

h2

(λ−r)

Z

C

r

|∇v h | 2 ≤ Z

C

r

|∇v h | 2 e

h2

f ≤ Z

C

r/2

|∇v h | 2 e

h2

f ≤ C

h α e

h2

(λ−

r2

) .

Therefore, there exists β > 0 such that for h small enough, Z

C

r

|∇v h | 2 ≤ C

h α e

hr

≤ C e

βh

,

and from (35), we then have k∆v h k L

2

(C

r

) ≤ C e

βh

for some constant β > 0. In the

following, β > 0 is a constant which may change from one occurrence to another and

does not depend on h. Let χ 1 ∈ C c (C r ) be such that χ 1 ≡ 1 on C 2r . Since ∆(χ 1 v h ) =

(19)

χ 1 ∆v h + v h ∆χ 1 + 2∇χ 1 · ∇v h , there exists C, such that k∆(χ 1 v h )k L

2

(C

r

) ≤ C for h small enough. By elliptic regularity (see [16, Theorem 5, Section 6.3]) it comes

kv h k H

2

(C

2r

) ≤ C.

Let α ∈ (0, 1) be an irrational number such that p 1 = d−2α 2d > 0. From the Gagliardo- Nirenberg interpolation inequality (see [39, Lecture II]), the following inequality holds

k∇v h k L

p1

(C

2r

) ≤ Ckv h k α H

2

(C

2r

) k∇v h k 1−α L

2

(C

2r

) + Ck∇v h k L

2

(C

2r

) ≤ C e

βh

.

From (35), k∆v h k L

p1

(C

2r

) ≤ C e

βh

. Using a cutoff function χ 2 ∈ C c (C 2r ) such that χ 2 ≡ 1 on C 4r , we get, as previously, from the elliptic regularity kv h k W

2,p1

(C

4r

) ≤ C. Let p 2 = d−4α 2d (i.e. 1/p 2 = 1/p 1 − α/d). If p 2 < 0, then [39, Lecture II] implies

k∇v h k L

(C

4r

) ≤ Ckv h k α W

2,p1

(C

4r

) k∇v h k 1−α L

p1

(C

4r

) + Ck∇v h k L

p1

(C

4r

) ≤ C e

βh

.

Thus, (42) is proved (if one chooses n = 2, i.e. 2 2 r = r 0 ). Otherwise, we prove (42) by induction as follows. From the Gagliardo-Nirenberg interpolation inequality (see [39, Lecture II]), we get

k∇v h k L

p2

(C

4r

) ≤ Ckv h k α W

2,p1

(C

4r

) k∇v h k 1−α L

p1

(C

4r

) + Ck∇v h k L

p1

(C

4r

) ≤ C e

βh

.

We repeat this procedure n times where n is the first integer such that d−2nα < 0 and the Gagliardo-Nirenberg interpolation inequality implies that k∇v h k L

(C

2nr

) ≤ C e

βh

which ends the proof of (42). This concludes the proof of Proposition 13.

2.3 Link between the law of X τ

when X 0 ∼ ν h and X 0 = x ∈ A(C max ) Using Proposition 13, one can now compare E ν

h

[F (X τ

)] and E x [F (X τ

)] for smooth functions F : the next proposition combined with Theorem 3 already gives the result of Theorem 1 for smooth functions F .

Proposition 14. Assume that the assumptions (A0) and (A1) are satisfied and that min

C

max

f = min

f,

where we recall that C max is introduced in (A1). Let K be a compact subset of Ω such that K ⊂ A(C max ) and let F ∈ C (∂Ω, R ). Then, there exists c > 0 such that for all x ∈ K:

E ν

h

[F (X τ

)] = E x [F (X τ

)] + O e

hc

in the limit h → 0 and uniformly in x ∈ K .

Proof. Assume that the assumptions (A0) and (A1) are satisfied and that min

C

max

f = min

f.

Step 1. For α > 0 small enough, let C max (α) be as introduced in (25):

C max (α) = C max

f < f max − α .

(20)

Let F ∈ C (∂Ω, R ). In this first step, we will prove that ∃α 0 > 0, ∀α ∈ (0, α 0 ), ∃c >

0, ∀y ∈ C max (α):

E ν

h

[F (X τ

)] = E y [F (X τ

)] + O e

hc

(43) in the limit h → 0 and uniformly in y ∈ C max (α). Let us recall that from the notation of Proposition 13 (see (37)), for all x ∈ Ω:

v h (x) = E x [F (X τ

)] . From (8), one has:

E ν

h

[F (X τ

)] = Z

u h e

h2

f −1 Z

v h u h e

2h

f

= 1

Z h (Ω) Z

C

max

(α)

v h u h e

2h

f + 1 Z h (Ω)

Z

Ω\C

max

(α)

v h u h e

h2

f , (44) where

Z h (Ω) :=

Z

u h e

h2

f

and u h is the principal eigenfunction of −L D f,h which satisfies (7). Let us first deal with the second term in (44). Since (A0) and (A1) hold, and because it is assumed that min C

max

f = min f, one obtains from Corollary 11 that there exists C > 0 such that for h small enough:

1

Z h (Ω) ≤ Ch

d4

e

1 h

min

f .

Let us recall that for α > 0 small enough, one has (see (28)), arg min

C

max

f ⊂ C max (α).

Therefore, using the second statement in Corollary 11 with O = Ω \ C max (α), for all α > 0 small enough, there exists c > 0 such that when h → 0:

Z

Ω\C

max

(α)

u h e

h2

f = O e

1 h

(min

f+c) .

Thus, there exists α 0 > 0 such that for all α ∈ (0, α 0 ) there exists c > 0 such that when h → 0:

1 Z h (Ω)

Z

Ω\C

max

(α)

u h e

h2

f = O e

hc

. (45)

Then, since kv h k L

(Ω) ≤ kF k L

(∂Ω) , one obtains that 1

Z h (Ω) Z

Ω\C

max

(α)

v h u h e

2h

f = O e

ch

. (46)

Let us now deal with the first term in (44). Let us recall that C max ⊂ Ω is a connected component of {f < max C

max

f}. Moreover, for α ∈ (0, α 0 ) (α 0 > 0 small enough), the compact set C max (α) is connected and C max (α) ⊂ C max . Therefore, from Proposition 13 applied to K = C max (α) for α ∈ (0, α 0 ), one obtains that there exists δ α > 0 such that for all y ∈ C max (α),

1 Z h (Ω)

Z

C

max

(α)

v h u h e

2h

f = v h (y) Z h (Ω)

Z

C

max

(α)

u h e

2h

f + O e

δαh

Z h (Ω)

Z

C

max

(α)

u h e

h2

f (47)

(21)

in the limit h → 0 and uniformly with respect to y ∈ C max (α). Moreover, for all α ∈ (0, α 0 ) there exists c > 0 such that in the limit h → 0:

1 Z h (Ω)

Z

C

max

(α)

u h e

2h

f = 1 + O e

hc

. (48)

which follows from the fact that 1

Z h (Ω) Z

C

max

(α)

u h e

2h

f = 1 − 1 Z h (Ω)

Z

Ω\C

max

(α)

u h e

h2

f ,

together with (45). Let us now fix α ∈ (0, α 0 ). Then, using (47) and (48), ∃c > 0, ∃δ α > 0,

∀y ∈ C max (α):

1 Z h (Ω)

Z

C

max

(α)

v h u h e

2h

f = v h (y)

1 + O

e

hc

+ O

e

δαh

(49) in the limit h → 0 and uniformly with respect to y ∈ C max (α). Therefore, using (44), (46) and (49), ∃α 0 > 0, ∀α ∈ (0, α 0 ), ∃c > 0, ∀y ∈ C max (α):

E ν

h

[F (X τ

)] = E y [F (X τ

)] + O e

hc

,

in the limit h → 0 and uniformly with respect to y ∈ C max (α). This concludes the proof of (43).

Step 2. Let us now conclude the proof of Proposition 14 by considering a compact subset K of Ω such that K ⊂ A(C max ). Let us recall that (see (14)):

A(C max ) = {x ∈ Ω, t x = +∞ and ω(x) ⊂ C max }.

Since C max is open and stable by the flow ϕ t (·) (defined by (13)), the continuity of ϕ t (·) implies that there exists T K ≥ 0 such that for all x ∈ K,

ϕ T

K

(x) ∈ C max .

Moreover, since K is a compact subset of Ω and for all x ∈ K, t x = +∞ (i.e. ϕ t (x) ∈ Ω for all t ≥ 0), there exists δ > 0 such that all continuous curves γ : [0, T K ] → Ω such that

∃x ∈ K, sup

t∈[0,T

K

]

γ(t) − ϕ t (x) ≤ δ,

satisfy:

∀t ∈ [0, T K ], γ(t) ∈ Ω. (50)

Furthermore, up to choosing δ > 0 smaller, there exists α K > 0 such that

ϕ T

K

(x) + z, x ∈ K and |z| ≤ δ ⊂ C max (α K ) (51) where, we recall, C max (α K ) is defined by (25). Let us now recall the following estimate of Freidlin and Wentzell (see [17, Theorems 2.2 and 2.3 in Chapter 3, and Theorem 1.1 in Chapter 4], [5], [9, Theorem 3.5] and [18, Theorem 5.6.3]): for all x ∈ K, it holds:

lim sup

h→0

h ln P x

h sup

t∈[0,T

K

]

X t − ϕ t (x) ≥ δ i

≤ −I x,T

K

, (52)

Références

Documents relatifs

Compute the Fourier transform of the response function χ(ω).. Analyse the poles of this function e (for the

Hence in the intermediate range of temperatures the free energy of this stationary point corresponds to the replica symmetry breaking solution, which is higher than the equilibrium

We provide Large Deviation estimates for the bridge of a d -dimensional general diffusion process as the conditioning time tends to 0 and apply these results to the evaluation of

We also performed combined water entry and exit experiments with a conical mock-up in order to investigate the performance of the LED edge-lighting technique for the tracking of

Ce d´ eveloppement ce trouve dans [QZ13] dans le chapitre Th´ eor` emes limites en Probabilit´ e.. Toute s´ erie enti` ere de rayon de convergence fini admet un

In the special case when the boundary of the domain G is diffeomor- phic to a 3-dimensional sphere, the fact that the closed complex tangential curve ⊂ S ⊂ ∂ G mentioned above does

Coo case and the complex analytic or algebraic case (that we will refer to as the real or complex case), indicating when necessary the peculiarities.. of each

cumulative reaction depths in case of a random starting point, we observe an analogous surprising behaviour to what was noted above: the cumulative reaction depth over the