• Aucun résultat trouvé

Optimal stopping with irregular reward functions

N/A
N/A
Protected

Academic year: 2021

Partager "Optimal stopping with irregular reward functions"

Copied!
34
0
0

Texte intégral

(1)

HAL Id: hal-00796701

https://hal-upec-upem.archives-ouvertes.fr/hal-00796701

Submitted on 4 Mar 2013

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.

To cite this version:

Damien Lamberton. Optimal stopping with irregular reward functions. Stochastic Processes and their Applications, Elsevier, 2009, 119 (10), pp.3253-3284. �hal-00796701�

(2)

Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party

websites are prohibited.

In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information

regarding Elsevier’s archiving and manuscript policies are encouraged to visit:

http://www.elsevier.com/copyright

(3)

Stochastic Processes and their Applications 119 (2009) 3253–3284

www.elsevier.com/locate/spa

Optimal stopping with irregular reward functions

Damien Lamberton

Universit´e Paris-Est, Laboratoire d’Analyse et de Math´ematiques Appliqu´ees, Projet Mathfi, 5 Boulevard Descartes, 77 454 Marne-la-Vall´ee CEDEX 2, France

Received 25 April 2008; received in revised form 23 April 2009; accepted 12 May 2009 Available online 20 May 2009

Abstract

We consider optimal stopping problems with finite horizon for one-dimensional diffusions. We assume that the reward function is bounded and Borel-measurable, and we prove that the value function is continuous and can be characterized as the unique solution of a variational inequality in the sense of distributions.

c 2009 Elsevier B.V. All rights reserved.

MSC:60G40; 60J60; 60H10

Keywords:Optimal stopping; One-dimensional diffusions; Irregular reward functions

1. Introduction

In a recent paper with M. Zervos (see [10]), we studied optimal stopping problems with infinite horizon for one-dimensional diffusions. In particular, we proved that, under very general conditions, the value function is the unique solution (in the sense of distributions) of a stationary variational inequality. The purpose of the present paper is to examine optimal stopping problems with finite horizon and bounded Borel-measurable reward functions. We will prove that the value function is continuous and can be characterized as the unique solution (in the sense of distributions) of a suitable variational inequality.

The connection between optimal stopping and variational inequalities goes back to the work of Bensoussan and Lions [3] and Friedman [6]. This approach is very general and applies to multi- dimensional problems, but it requires uniform ellipticity of the diffusion and some regularity of

Tel.: +33 1 60 95 75 36; fax: +33 1 60 95 75 45.

E-mail address:damien.lamberton@univ-mlv.fr.

0304-4149/$ - see front matter c2009 Elsevier B.V. All rights reserved.

doi:10.1016/j.spa.2009.05.005

(4)

the reward function. Note that the techniques of viscosity solutions do not require ellipticity, but generally impose some continuity conditions on the reward function and the coefficients of the diffusion. For our results, we will not need any regularity assumption on the reward function, and we will deal with very general one-dimensional diffusions. On the other hand, our analysis will be limited to one-dimensional situations (cf. Remark 2.3).

The paper is organized as follows. In Section 2, we present our assumptions and the main results. In particular, we give the proper formulation of the variational inequality. In Section 3, we prove the continuity of the value function. In Section 4, we essentially relate the value function to the Snell envelope. Section 5 is devoted to the analytic interpretation of the supermartingale property. The proof that the value function satisfies the variational inequality is given in Section 6.

Uniqueness of the solution is proved in Section 7. In the last section, we have gathered a number of auxiliary results, which are classical in somewhat different contexts, but which require some justification under our assumptions. In particular, we derive regularity estimates for the semi- group of one-dimensional diffusions (see Theorem 8.11 and Corollary 8.13), which we have not found in the literature.

2. Assumptions and main results

We consider an open interval I = (α, β) (with −∞ ≤ α < β ≤ +∞ ) and a stochastic differential equation

d X

t

= b(X

t

) dt + σ (X

t

) dW

t

, X

0

= xI , (1)

where W is a standard one-dimensional Brownian motion, and b, σ : I

R

are Borel- measurable functions satisfying the following condition.

A1. For all x ∈ (α, β), σ

2

(x ) > 0, and ∃ ε > 0,

Rx+ε xε

1+|b(y)|

σ2(y)

dy < ∞ .

Under assumption A1, we have existence and uniqueness in law of a weak solution of (1) up to a possible explosion time (cf. [9, Section 5.5C]). In fact, we will also assume that the diffusion is non-explosive. This assumption can be expressed in terms of the so-called scale function p(x ) and speed measure m (dx ), defined by

p(x) =

Z x

c

exp

− 2

Z y

c

b(z) σ

2

(z) dz

dy, for xI, (2)

m (dx ) = 2

σ

2

(x ) p

(x) dx , (3)

where c is an arbitrary fixed element of I . The condition for no explosion can now be written as follows, according to Feller’s test (see Theorem 5.5.29 in Karatzas and Shreve [9]).

A2. We have lim

xα

l(x) = lim

xβ

l(x ) = ∞ , where l (x ) =

Z x

c

[ p(x)p(y)] m (dy), for xI .

Throughout the paper, assumptions A1 and A2 are in force. A weak solution of (1) is defined by a triple

(

,

F

, (

Ft

)

t0

,

Px

), W , X

, where (

,

F

, (

Ft

)

t0

,

Px

) is a filtered probability space with the filtration (

Ft

)

t0

satisfying the usual conditions, W = (W

t

)

t0

is a standard (

Ft

)- Brownian motion and X is a continuous adapted process satisfying (1). Given such a weak solution, we denote by (

F0

t

)

t0

the natural right-continuous filtration of X .

(5)

We now introduce an optimal stopping problem with a discounting rate (or interest rate) r . The function r : I

R

is assumed to be non-negative, Borel-measurable and locally bounded on I . We denote by

T0

t

(resp.

T

¯

0

t

) the set of all stopping times with respect to the filtration (

F0

t

)

t0

, with values in the interval [ 0, t ) (resp. [ 0, t ] ). Given a bounded Borel-measurable function f on I , we introduce the functions u

f

and v

f

defined on (0, +∞ ) × I as follows:

u

f

(t, x ) = sup

τ∈Tt0

Ex

h

e

−Λτ

f ( X

τ

)

i

, (4)

v

f

(t , x ) = sup

τ∈ ¯Tt0

Ex

h

e

−Λτ

f (X

τ

)

i

, (5)

where

Λt

=

Z t 0

r ( X

s

) ds.

Note that, due to the fact that we consider stopping times with respect to the natural filtration, the functions u

f

and v

f

depend only on the law of X , which is uniquely defined under assumptions A1 and A2. On the other hand, let

Tt

(resp.

T

¯

t

) be the set of all stopping times with respect to the filtration (

Ft

)

t0

, with values in the interval [ 0, t ) (resp. [ 0, t ] ). If we define by u ¯

f

(resp. v ¯

f

) the value function where

T0

t

(resp.

T

¯

0

t

) is replaced with

Tt

(resp.

T

¯

t

), we have u ¯

f

= u

f

and v ¯

f

= v

f

(see Section 8, Remark 8.7).

We obviously have u

f

≤ v

f

. Our first observation is the following result, the proof of which is quite similar to the one given for infinite horizon problems (see [10], Lemma 7), and is therefore omitted.

Proposition 2.1. Let f : I

R

be a bounded Borel-measurable function on I . Denote by f the ˆ upper semicontinuous envelope of f :

f ˆ (x ) = lim sup

yx

f (y), xI.

Then we have u

f

= u

fˆ

.

Our next result concerns the joint continuity of the value function. The following theorem will be proved in Section 3.

Theorem 2.2. We have u

f

= v

f

and the function v

f

is jointly continuous on (0, +∞ ) × I . Remark 2.3. The fact that we have a one-dimensional diffusion is essential for the continuity of the value function. Indeed, consider a two-dimensional Brownian motion (W

t1

, W

t2

)

t0

and let f be the indicator function of the singleton { 0 } . Since Brownian motion starting from x 6= 0 will never hit 0 with probability one, we clearly have (with similar notations as above) u

f

= v

f

= f , so that v

f

is discontinuous. In fact, crucial to the continuity of the value function is the fact that the diffusion hits any given point close to the initial point with positive probability. Note that the regularization procedure that we develop in Section 5 also depends heavily on the one- dimensional setting.

In order to write the variational inequality satisfied by the value function, we need to introduce

the infinitesimal generator

L0

of the diffusion. For a twice continuously differentiable function

(6)

u,

L0

u is defined by

L0

u(x) = σ

2

(x )

2 u

′′

(x ) + b(x )u

(x ), xI .

As should be expected, the variational inequality will involve the operator −

t

+

L

, where the operator

L

is defined by

L

u(t , x ) =

L0

u(t, x ) − r (x )u(t, x ), (t , x ) ∈ (0, +∞ ) × I .

In fact, in order to be able to apply the operator to possibly non-smooth functions, we will rather consider the operator

A

, where

A

= 2 σ

2

p

− ∂

t +

L

.

The following proposition is the key to the extension of

A

to irregular functions, in the sense of distributions.

Proposition 2.4. If uC

1,2

((0, +∞ ) × I ), for any C

function

Φ

, with compact support in (0, +∞ ) ×

R

, we have

Z Z

A

u (t, x )

Φ

(t , x )dt dx =

Z Z

u(t , x ) ∂

Φ

t +

(t, x ) dt m (dx ).

Proof. It follows from integration by parts with respect to time and from the definition of the speed measure that

Z Z

2 σ

2

(x ) p

(x )

u

t (t , x )

Φ

(t, x )dt dx =

Z Z

u(t , x ) ∂

Φ

t (t, x ) dt m (dx ).

On the other hand, using the fact that the scale function p satisfies d

dx

1

p

= 2b σ

2

1 p

, we have

L0

u = σ

2

p

2

1 p

2

u

x

2

+ 2b σ

2

1 p

u

x

= σ

2

p

2

x

1

p

u

x

.

Hence, integrating by parts twice with respect to x,

Z Z

2

σ

2

(x) p

(x )

L0

u(t , x)

Φ

(t, x )dt dx =

Z Z

u(t, x )

L0Φ

(t , x )dt m (dx ).

The result now follows easily. ⋄

In view of Proposition 2.4, it is natural, given a locally bounded measurable function u on (0, +∞ ) × I , to define the distribution

A

u by setting, for any smooth test function

Φ

,

h

A

u,

Φ

i =

Z Z

u(t, x ) ∂

Φ

t +

(t , x ) dt m(dx ).

(7)

Remark 2.5. We will also need the distribution

A

˜ u, defined, for u locally bounded on (0, T ) × I (where T is a fixed positive number) by

h ˜

A

u,

Φi = Z Z

u(t , x )

− ∂

Φ

t +

(t , x) dt m(dx ),

for

Φ

smooth with compact support in (0, T ) × I . Note that one can prove, as in Proposition 2.4 that if uC

1,2

((0, T ) × I ),

A

˜ u =

σ22p

+

t

+

L

u.

We can now state our main result. Recall that f ˆ denotes the upper semicontinuous envelope of f . Theorem 2.6. The value function v

f

is the only continuous and bounded function on the open set (0, +∞ ) × I satisfying the following conditions.

1. On the set (0, +∞ ) × I , we have v ≥ f , and the distribution

A

v satisfies

A

v ≤ 0.

2. We have

A

v = 0 on the open set U := { (t , x ) ∈ (0, +∞ ) × I | v(t, x ) > f ˆ (x ) } . 3. For every xI , lim

t0

v(t , x ) = ˆ f (x ).

3. Continuity of the value function

This section is devoted to the proof of Theorem 2.2. At the end of the section, we also include a proposition concerning the behaviour of the value function for small time (see Proposition 3.4).

The equality u

f

= v

f

is an easy consequence of the following proposition.

Proposition 3.1. Let τ be a stopping time with values in [ 0, t ] . We have

Ex

h

e

−Λτ

f (X

τ

)

i

= lim

st,s<tEx

h

e

−Λτs

f (X

τs

)

i

.

Proof. We have

Ex

h

e

−Λτs

f ( X

τs

)

i

=

Ex

h

e

−Λτ

f (X

τ

)1

{τ <s}i

+

Ex

h

e

−Λs

f ( X

s

)1

{τs}

i

, (6)

and

Ex

h

e

−Λs

f (X

s

)1

{τs}

i

=

Ex

h

e

−Λt

f ( X

s

)1

{τ=t}

i

+

Ex

h

f (X

s

)

e

−Λs

1

{τs}

− e

−Λt

1

{τ=t}

i

.

By dominated convergence,

s

lim

t,s<tEx

h

e

−Λτ

f (X

τ

)1

{τ <s} i

=

Ex

h

e

−Λτ

f (X

τ

)1

{τ <t} i

,

and

s

lim

t,s<tEx

h

f (X

s

)

e

−Λs

1

{τs}

− e

−Λt

1

{τ=t}

= 0

i

.

We now want to prove

s

lim

t,s<tEx

| f ( X

s

) − f (X

t

) | = 0. (7)

This is clearly true if f is continuous. If f is arbitrary, we have

Ex

| f (X

s

) − f (X

t

) | ≤

Ex

| f (X

s

) − ϕ(X

s

) | +

Ex

| ϕ(X

s

) − ϕ(X

t

) |

+

Ex

| ϕ( X

t

) − f ( X

t

) | .

(8)

So, in order to prove (7), we need only prove that, given ε > 0, one can find a bounded continuous function ϕ such that

sup

t/2st

Ex

| f ( X

s

) − ϕ( X

s

) | ≤ ε.

From Corollary 8.13 we know that, given x

I

, there exists a constant C

x

> 0 such that, for all t > 0 and h ≥ 0,

P

t

h(x ) ≤ C

x

1 + 1

t

k h k

L2(m)

. (8)

Now assume that fL

2

(m ) (the extension to f bounded on I is straightforward). Given ε > 0, one can find a continuous function ϕ with compact support such that k f − ϕ k

L2(m)

< ε, so that (using (8) with h = | f − ϕ | )

sup

t/2st

Ex

| f ( X

s

) − ϕ( X

s

) | ≤ C

x

1 +

√ 2

t

!

ε,

which completes the proof of (7). ⋄

Remark 3.2. The proof of Proposition 3.1 relies on the convergence (in probability) of f (X

s

) to f (X

t

), when st , for f bounded and Borel-measurable. As proved in [4], this is related to the relative weak compactness of the laws of the random variables X

s

. The argument we give can be seen as a way of proving this property.

For the proof of the continuity of the value function, we will also need the following lemma.

Lemma 3.3. Let (

,

F

, (

Ft

)

t0

,

Px

, W, X ) be a weak solution of (1). For y ∈ I , define τ

y

= inf { t ≥ 0 | X

t

= y } .

We have lim

yxPx

y

< ∞ ) = 1 and, for all t > 0, lim

yx Px

y

t ) = 0. We also have lim

yx Ey

(e

−Λτx

) = 1 and lim

yxPy

x

t ) = 0, for all t > 0.

Proof. It is well known that the function (x, y) 7→

Ex

(e

τy

) is jointly continuous on I × I (see for instance [8]). In particular, we have lim

yx Ex

(e

τy

) = 1. We have

Px

y

t ) =

Px

(1 − e

τy

≥ 1 − e

t

) ≤

Ex

(1 − e

τy

) 1 − e

t

.

Hence, lim

yx Px

y

t) = 0. A similar argument gives lim

yx Py

x

t ) = 0. Since

Px

y

= ∞ ) ≤

Px

y

t), we have lim

yx Px

y

< ∞ ) = 1.

On the other hand, we have

Ey

(e

−Λτx

) ≥

Ey

(e

τx−Λτx

) =

Ey

(e

R0τx(1+r(Xs))ds

) and we also have that (x, y) 7→

Ex

(e

R0τy(1+r(Xs))ds

) is jointly continuous on I × I (see for instance [8]), so that lim

yxEy

(e

R0τx(1+r(Xs))ds

) = 1. Hence, lim

yx Ey

(e

−Λτx

) = 1. ⋄

Proof of Theorem 2.2. As mentioned above, the equality u

f

= v

f

follows easily from Propo- sition 3.1. On the other hand, since u

f

= u

fˆ

= v

fˆ

, we also have v

f

= v

fˆ

. It is known that if f is upper semicontinuous, so is v

f

(see [2], Proposition 17 or [5]). It remains to prove that v

f

is lower semicontinuous.

Fix (t, x ) ∈ (0, +∞ ) × I and τ ∈ ¯

Tt0

. Since τ is a stopping time of (

Ft0

), we have { τ = 0 }

F0

0

, and we deduce from the zero–one law (cf. Remark 8.4) that

Px

(τ = 0) ∈ { 0, 1 } .

(9)

Suppose

Px

(τ = 0) = 1. We then have

Ex

e

−Λτ

f (X

τ

)

= f (x ). On the other hand, we have, for any (s, y) ∈ (0, +∞ ) × I , with M = sup

yI

| f (y) | ,

v

f

(s, y)

Ey

e

−Λτx∧s

f ( X

τxs

)

Ey

e

−Λτx∧s

f (X

τxs

)1

{τx<s}

M

Py

x

s)

= f (x )

Ey

(e

−Λτx

1

{τx<s}

) − M

Py

x

s ).

Hence v

f

(s, y)f (x )

Ey

(e

−Λτx

) − 2 M

Py

x

s). Using Lemma 3.3, we have

lim

yxEy

(e

−Λτx

) = 1 and lim

yx Py

x

t /2) = 0. Hence f (x ) ≤ lim inf

(s,y)(t,x)

v

f

(s, y).

We now assume that

Px

(τ = 0) = 0. From Proposition 3.1, we know that, given ε > 0, there exists δ ∈ (0, t) such that, for all s ∈ [ t − δ, t ] ,

Ex

(e

−Λτs

f (X

τs

)) −

Ex

(e

−Λτ

f ( X

τ

))

≤ ε. (9)

Obviously, this inequality is also true for st . Now, we have, for all (s, y) ∈ (0, +∞ ) × I ,

Ex

e

−Λτs

f (X

τs

)

=

Ex

e

−Λτs

f ( X

τs

)1

{τyτs}

+

Ex

e

−Λτs

f (X

τs

)1

{τys}

.

We have

Ex

e

−Λτs

f (X

τs

)1

{τys}

M

Px

y

> τ ∧ s).

On the other hand, with the notation τ

s

for τ ∧ s , we have

Ex

e

−Λτs

f (X

τs

)1

{τyτs}

=

Ex

e

−Λτy

1

{τyτs}

e

Rsτy)+

0 r(Xθ)dθ

f ( X

τy+sτy)+

)

Ex

1

{τy<∞}

e

R

sτy)+

0 r(Xθ)dθ

f (X

τy+sτy)+

)

=

Px

y

< ∞ )

Ex

e

R

sτy)+

0 r(Xθ)dθ

f (X

τy+sτy)+

) | τ

y

< ∞

.

Note that, conditionally on { τ

y

< ∞} ,

(

,

F

, (

Fτ

y+θ

)

θ0

,

P

), (W

τy+θ

W

τy

), ( X

τy+θ

) is a weak solution of the stochastic differential equation with starting point y, and

s

− τ

y

)

+

is an (

Fτ

y+θ

)

θ0

-stopping time. Hence (using Remark 8.7)

Ex

e

−Λτs

f (X

τs

)1

{τyτs}

Px

y

< ∞ )v

f

(s, y).

Therefore, we have

Ex

e

−Λτs

f (X

τs

)

Px

y

< ∞ )v

f

(s, y) + M

Px

y

> τ ∧ s).

Now, take s ∈ [ t − δ, +∞ ). We have, using (9),

Ex

e

−Λτ

f (X

τ

)

− ε ≤

Ex

e

−Λτs

f (X

τs

)

Px

y

< ∞ )v

f

(s, y) + M

Px

y

> τ ∧ s).

(10)

It follows from Lemma 3.3 that lim

yx Px

y

< ∞ ) = 1 and lim

(s,y)(t,x)Px

y

> τ ∧ s) = 0.

Hence

Ex

e

−Λτ

f (X

τ

)

≤ lim inf

(s,y)(t,x)

v

f

(s, y).

We conclude that v

f

is lower semicontinuous. ⋄

The following Proposition clarifies the asymptotic behaviour of the value function as time goes to zero.

Proposition 3.4. We have lim

t0

v

f

(t , x ) = ˆ f (x ).

Recall that f ˆ is the upper semicontinuous envelope of f . In view of Proposition 3.4, it is natural to extend the definition of v

f

(t , x) at t = 0 by setting v

f

(0, x ) = ˆ f (x ).

Proof of Proposition 3.4. Note that t 7→ v

f

(t , x ) is clearly non-decreasing, so that the limit exists. Since v

f

= v

fˆ

, we have v

f

(t, x ) ≥ ˆ f (x ), so that lim

t0

v

f

(t, x ) ≥ ˆ f (x ). Now, let τ ∈ ¯

T0

t

. For any ε > 0 such that (x − ε, x + ε) ⊂ I , we have

Ex

e

−Λτ

f (X

τ

)

≤ sup

y(xε,x+ε)

f (y) + sup

I

f

Px

xε

∧ τ

x+ε

t ).

Hence

v

f

(t , x ) ≤ sup

y(xε,x+ε)

f (y ) + sup

I

f (

Px

xε

t ) +

Px

x+ε

t )) .

Observe that lim

t0Px

x±ε

t ). Therefore lim

t0

v

f

(t , x ) ≤ sup

y(xε,x+ε)

f ( y), and by making ε go to 0, we get lim

t0

v

f

(t , x ) ≤ ˆ f (x ). ⋄

4. The value function along the paths

The main result of this section is the following.

Theorem 4.1. Fix a positive number T and let f : I

R

be a bounded, non-negative and upper semicontinuous function. For any weak solution (

,

F

, (

Ft

)

t0

,

Px

, W , X ) of (1), the process V defined by

V

t

= e

−Λt

v

f

(T − t , X

t

), 0 ≤ tT ,

is a supermartingale. Moreover, if τ ˆ = inf { t ≥ 0 | v

f

(T − t , X

t

) = f (X

t

) } , the process (V

t∧ ˆτ

)

0tT

is a martingale.

This result is not surprising. It appears in various forms in the literature (see for instance [5]).

However, since, under our assumptions, it does not seem to follow directly from known results, we will give a complete proof.

The proof of Theorem 4.1 will be based on the following lemma.

Lemma 4.2. Let f : I

R

be bounded, non-negative and Borel-measurable. For all T > 0 and for all τ ∈ ¯

T0

T

, we have

xI ,

Ex

e

−Λτ

v

f

(T − τ, X

τ

)

≤ v

f

(T , x ).

(11)

Proof. For the proof of this lemma, we will need the strong Markov property. So, we will work with the canonical realization of the process X . More precisely, denote by (

,

F0

) the canonical space, where

is the set of all continuous functions on

R+

, with values in I , and

F0

is the σ -algebra generated by the finite-dimensional cylinder sets. We endow this space with the right- continuous natural filtration (

Ft0

) of the coordinate mapping process X defined by X

t

(ω) = ω(t ), for t ≥ 0 and ω ∈

. This space supports the family of shift operators (θ

t

, t ≥ 0), defined by θ

t

(ω) = ω(t + · ), for t ≥ 0 and ω ∈

. Given an initial condition xI , we denote by

Px

the (unique) law of a weak solution of (1). The fact that we have the strong Markov property for the family of probability measures (

Px

, xI ) on the canonical space follows from the Markov property and the fact that the semi-group preserves continuity (the weak Markov property and the fact that the semi-group preserves continuity are proved in Section 8, and the strong Markov property can be deduced by classical arguments, see [11], chapter III, Section 3).

For (t , x ) ∈ [ 0, TI , set U (t , x ) = v

f

(T − t , x ). Note that, since f is upper semicontinuous, U (T , x ) = f (x ). Given a finite subset F of the interval I , we denote by

T F

T

the set of all stopping times in

T

¯

0

T

, such that, on the set { τ < T } , X

τ

takes its values in F . We will first prove that

∀ τ ∈

T F

T

,

Ex

e

−Λτ

U (τ, X

τ

)

U (0, x ). (10)

Suppose F = { a

1

, . . . , a

n

} , with a

1

< · · · < a

n

, and let τ ∈

T F

T

. Note that, since X

T

has a density (this is an immediate consequence of Corollary 8.13), we have

Px

(X

T

F ) = 0, so that, with probability 1,

e

−Λτ

U (τ, X

τ

) = e

−ΛT

U (T , X

T

)1

{τ=T}

+

Xn

i=1

e

−Λτ

U (τ, a

i

)1

{Xτ=ai}

.

Let ρ = (ρ

0

= 0 < ρ

1

< · · · < ρ

m1

< ρ

m

= T ) be a subdivision of the interval [ 0, T ] . For i ∈ { 1, . . . , n } and j ∈ { 1, . . . , m } , define

A

i j

= { ρ

j1

≤ τ < ρ

j

} ∩ { X

τ

= a

i

} . We have (using U (T , · ) = f )

e

−Λτ

U (τ, X

τ

) = e

−ΛT

f (X

T

)1

{τ=T}

+

n

X

i=1 m

X

j=1

e

−Λτ

U (τ, a

i

)1

Ai j

=

Xn i=1

m1

X

j=1

e

−Λτ

U

j

, a

i

)1

Ai j

+ e

−Λτ

U (τ, X

τ

)1

{ρm1τ <T}

+ e

−ΛT

f ( X

T

)1

{τ=T}

+

n

X

i=1 m1

X

j=1

e

−Λτ

U (τ, a

i

) − U

j

, a

i

)

1

Ai j

. (11)

We have

U

j

, a

i

) = u

f

(T − ρ

j

, a

i

) = sup

τ∈TT0ρ

j

Eai

e

−Λτ

f (X

τ

)

.

Fix ε > 0 and denote by τ

i j

a stopping time in

T0

Tρj

, such that U

j

, a

i

) ≤

Eai

e

−Λτi j

f (X

τi j

)

+ ε.

(12)

Let

˜

τ = T 1

{ρm1τT}

+ 1

{0τ <ρm1}

τ +

n

X

i=1 m1

X

j=1

1

Ai j

τ

i j

◦ θ

τ

!

.

This clearly defines a stopping time with values in [ 0, T ] . Therefore U (0, x ) ≥

Ex

e

−Λτ˜

f (X

τ˜

)

=

Ex

e

−ΛT

f (X

T

)1

{ρm1τT}

+

n

X

i=1 m1

X

j=1

Ex

1

Ai j

e

−Λτ+τi j◦θτ

f ( X

τ+τi jθτ

) .

Using the strong Markov property and A

i j

F0

τ

, we have

Ex

1

Ai j

e

−Λτ+τi j◦θτ

f (X

τ+τi jθτ

)

=

Ex

h

1

Ai jE

e

−Λτ+τi j◦θτ

f ( X

τ+τi jθτ

) |

F0

τ

i

=

Ex

h

1

Ai j

e

−ΛτEai

e

−Λτi j

f ( X

τi j

)

i

Ex

1

Ai j

e

−Λτ

U

j

, a

i

) − ε .

Hence, using r ≥ 0, U (0, x ) ≥

Ex

e

−ΛT

f (X

T

)1

{ρm1τT}

+

Ex

n

X

i=1 m1

X

j=1

1

Ai j

e

−Λτ

U

j

, a

i

) − ε

!

Ex

e

−ΛT

f (X

T

)1

{ρm1τT}

+

Ex

n

X

i=1 m1

X

j=1

1

Ai j

e

−Λτ

U

j

, a

i

)

!

− ε.

It follows from (11) that

n

X

i=1 m1

X

j=1

1

Ai j

e

−Λτ

U

j

, a

i

) = e

−Λτ

U (τ, X

τ

) − e

−Λτ

U (τ, X

τ

)1

{ρm1τT}

Xn

i=1 m1

X

j=1

e

−Λτ

U (τ, a

i

) − U

j

, a

i

) 1

Ai j

.

Hence

U (0, x ) ≥

Ex

e

−Λτ

U (τ, X

τ

)

+

Ex

h

e

−ΛT

f (X

T

) − e

−Λτ

U (τ, X

τ

)

1

{ρm1τ <T}

i

Ex n

X

i=1 m1

X

j=1

e

−Λτ

U (τ, a

i

) − U

j

, a

i

) 1

Ai j

!

− ε.

Let | ρ | = max

1jm

| ρ

j

− ρ

j1

| . By passing to the limit as | ρ | → 0, we get, using the continuity of U on [ 0, T ) × I , U (0, x ) ≥

Ex

e

−Λτ

U (τ, X

τ

)

− ε, and, since ε is arbitrary, (10) is proved.

Now, suppose τ ∈ ¯

T0

T

, and denote by (a

n

)

n1

a dense sequence of elements in I . Set F

n

= { a

1

, . . . , a

n

} and

τ

n

= inf { t ≥ τ | X

t

F

n

} ∧ T .

We have τ

n

T Fn

, so that, according to (10),

Ex

e

−Λτn

U

n

, X

τn

)

U (0, x ). On the other

hand, the sequence (τ

n

)

n1

is non-increasing and lim

n→∞

τ

n

= τ . Indeed, if we denote the limit

Références

Documents relatifs

We prove that arbitrary actions of compact groups on finite dimensional C ∗ -algebras are equivariantly semiprojec- tive, that quasifree actions of compact groups on the Cuntz

We will use the idea of Bakry and Emery [1] to give a notion of a ”lower Ricci curvature bound” on graphs and to prove that Ricci curvature on Locally finite graphs is always

Let X be a n-dimensional compact Alexandrov space satisfying the Perel’man conjecture. Let X be a compact n-dimensional Alexan- drov space of curvature bounded from below by

First introduced by Faddeev and Kashaev [7, 9], the quantum dilogarithm G b (x) and its variants S b (x) and g b (x) play a crucial role in the study of positive representations

For example, he proved independently that two completely regular spaces X and Y have homeomorphic realcompactifications νX and νY whenever there is a biseparating linear maps from

The second mentioned author would like to thank James Cogdell for helpful conversation on their previous results when he was visiting Ohio State University.. The authors also would

In this paper, we propose a nonlocal model for linear steady Stokes equation with no-slip boundary condition.. The main idea is to use volume constraint to enforce the no-slip

Aware of the grave consequences of substance abuse, the United Nations system, including the World Health Organization, has been deeply involved in many aspects of prevention,