• Aucun résultat trouvé

The Laplace transform of the integral of the absolute value of a real Brownian motion has been computed in 1945 by M

N/A
N/A
Protected

Academic year: 2023

Partager "The Laplace transform of the integral of the absolute value of a real Brownian motion has been computed in 1945 by M"

Copied!
5
0
0

Texte intégral

(1)

THOMAS SIMON

Abstract. The Laplace transform of the integral of the absolute value of a real Brownian motion has been computed in 1945 by M. Kac [4], with the help of a long and subtle as- ymptotic analysis on Bessel functions. Up to now, there does not seem to exist a shorter proof of this well-known computation. In this semi-historical note we observe that Kac’s argument could have been greatly simplified back then, had he used the eigenfunction ex- pansion associated to a real Schr¨odinger operator with linear potential, evaluated in 1944 (and in the same linguistic part of the world) by R.-P. Bell [2].

1. Recalls on the Laplace transforms of Lp-Brownian functionals Let{Wt, t≥0} be a standard linear Brownian motion and denote byPx its law starting fromx∈R, setting P0 =P for simplicity. Consider for every p >0 the Lp-functionals

||W||p = Z 1

0

|Wt|pdt 1p

.

Introducing the function up(t, x) = Ex

exp−

Z t

0

|Ws|pds

= Ex h

e−tp/2+1||W||ppi

t≥0, x∈R,

a version of the Feynman-Kac formula which can be found e.g. in [5] Proposition 5.8, entails that it solves the PDE

∂up

∂t (t, x) = 1 2

2up

∂x2 (t, x) − |x|pup(t, x)

with initial condition up(0, x) = 1. On the other hand, it is known - see Chapter V in [8] - that the self-adjoint negative operator

Ap :f 7→ 1

2f00 − |x|pf

acting on real C2 functions has a discrete and simple spectrum, and that there exists an orthonormal basis of L2(R) made out of the corresponding eigenfunctions. In other words, there exists a sequence 0 < λp1 ≤ λp2 ≤ . . . ≤ λpn → +∞ and a sequence of fonctions {ψnp, n ≥ 1} such that for any m, n ≥ 1 one has < ψpn, ψpm >= δmn (with Kronecker’s notation) and Apψnp = −λpnψnp,and such that any f ∈ L2(R) has the decomposition

2000Mathematics Subject Classification. 60G15; 01A60.

Key words and phrases. Lp-Brownian functional - Eigenfunction expansion - Feynman-Kac formula - Series representation.

1

(2)

(1) f = X

n≥1

< f, ψnp > ψpn

inL2, where<, >stands for the usual scalar product. It is easy to see from basic properties of Brownian Motion that the function x7→ up(t, x) is C with all its derivatives in L2. By Theorem 2.7 (i) in [8], this entails that (1) applied to up(t, .) is also true in L: one has

up(t, x) = X

n≥1

ϕpn(t)ψnp(x)

for everyt ≥0 andx∈R,where theϕpn are defined by the formula ϕpn(t) =< up(t, .), ψnp > . Differentiating under the integral yields

pn dt (t) =

Z

R

∂up

∂t (t, x)ψnp(x)dx = Z

R

Apup(t, x)ψnp(x)dx

= Z

R

X

n≥1

ϕpn(t)Apψnp(x)

!

ψnp(x)dx

= −

Z

R

X

n≥1

λpnϕpn(t)ψpn(x)

!

ψnp(x)dx = −λpnϕpn(t), whence ϕpn(t) = cpne−λpnt for every t ≥ 0. Recalling that up(0, x) = 1 for every x ∈ R, we finally compute

cpn = Z

R

ψpn(y)dy which yields

up(t, x) = X

n≥1

e−λpnt Z

R

ψpn(y)dy

ψnp(x).

Takingx= 0 and changing the variable, we can now write down the formula for the Laplace transforms of the Lp-Brownian functionals:

(2) E

h

e−t||W||pp i

= X

n≥1

e−λpnt

p+22 Z

R

ψnp(y)dy

ψnp(0), t≥0.

2. Specifying to p= 1

The eigenelements {ψn, λn} of the operator f 7→ f00 − |x|f have been evaluated in [2], a paper seemingly unnoticed by Kac at the time. For the probabilistic operator A1, a change of variable and the normalization< ψn1, ψn1 >= 1 yield

λ1n = 2−1/3λn and ψn1(x) = 21/6ψn(21/3x).

In view of (2), it is enough to consider the even eigenfunctions and the corresponding eigen- values, and we will relabel them by {ψn, λn} for simplicity. By (5) in [2], the λn are the

(3)

successive zeroes of the function

x 7→ J2/3(ζ)−J−2/3(ζ),

where Jα stands for the Bessel function of the first kind with index α and ζ = 2x3/2/3. By e.g. (10.4.17) in [1], we see that the λn are the successive zeroes of x7→Ai0(−x), where

Ai(x) = 1 π

Z

0

cos(tx+t3/3)dt

is the Airy function. For the corresponding eigenfunction ψn, equations (3) and (10) in [2]

give

ψ1n(0) = 21/6ψn(0) = 2−1/3

√λn· We last compute

Z

R

ψn1(y)dy = 2−1/6 Z

R

ψn(y)dy = 25/6 Z

0

ψn(y)dy = 25/6 Z λn

0

ψn(y)dy + Z

λn

ψn(y)dy

. Again, by (3) and (10) in [2] and using the notation ζn = 2λ3/2n /3, the first integral is given by

Z λn

0

ψn(y)dy = Z λn

0

√x(J1/3(ζ) +J−1/3(ζ))

√2λn(J1/3n) +J−1/3n))

!

dx = 1

√2λnAi(−λn) Z λn

0

Ai(−x)dx, where the second equality comes from (10.4.16) in [1]. Using now (4) and (10) in [2] with the same notations, we get

Z

λn

ψn(y)dy = Z

0

√3xK1/3(ζ) π√

n(J1/3n) +J−1/3n))

! dx

= 1

√2λnAi(−λn) Z

0

xK1/3(ζ) π√

3

dx = 1

3√

nAi(−λn)

whereK1/3is the Macdonald function of index 1/3, and the last equality stems from (10.4.82) in [1]. Putting everything together with (2) entails the formula

(3) E

e−t||W||1

= X

n≥1

1 3λnAi(−λn)

1 + 3

Z λn

0

Ai(−x)dx

e−2−1/3λnt2/3 for every t≥0, which is up to some different notation exactly the one stated in [7].

Remarks. (a) For conciseness, we did not detail the proofs of the results quoted from [2]. A perusal of pp. 583-584 therein shows that those results rely on completely classical consider- ations on Bessel functions. Recall that the method of [4] was a discretization of the problem with the help of a doubly exponential random walk, and several new theorems on Bessel functions taking eight pages of calculations, overall significantly more difficult to read. For this reason, although we used classical methods we thought that our simplification concern- ing this old result might still be useful.

(4)

(b) From the historical viewpoint the Feynman-Kac formula, which is of course a crucial point in the above computation, was actually obtained shortly after [4] in 1947. Notice that the simplified version which is used here hinges only upon the semi-group property of Brownian motion - see the proof of Theorem 5.8 in [5], and does not cover the full extent of the original Feynman-Kac formula. We refer to [3] and the references therein for a modern account on the latter and its applications to the computation of Laplace transforms of several Brownian additive functionals.

(c) The eigenelements {ψn2, λ2n} were computed by Titchmarsh in April 1940 in terms of the Hermite polynomials - see Section 4.2 in [7], and with the same argument as above one can show that this computation leads to the well-known Cameron-Martin formula

(4) E

h

e−t||W||22i

= 1

pcosh√ 2t

for all t≥0.However, the eigenelement argument turns out to be overall lenghtier than the two classical methods to obtain the Cameron-Martin formula, which rely respectively on the Karhunen-Lo`eve decomposition and on Girsanov’s transformation - see Example 3 in [3] for details, and much more on this topic.

(d) It is easy to see that the function t 7→ E[e−t||W||1] is analytic over R. In [4], Kac raises the question how formula (3) should be continued to the negative real axis. To the best of our knowledge, this question remains open.

3. Series representations

With the help of the positive stable density of index 2/3, the Laplace transform (3) was inversed by Tak´acs who gave a series representation for the densityf1 of||W||1 (see Theorem 1 in [7]):

f1(x) = 22/3x−7/3

√27π X

n≥1

λne−2λ3n/27x2 3Ai(−λn)

1 + 3

Z λn

0

Ai(−x)dx

U(1/6,4/3,2λ3n/27x2) whereU is the usual confluent hypergeometric function. It is possible to inverse (2) for every p >0 exactly the same way. For every p >0, letgp be the density of the positive stable law with index 2/(p+ 2) normalized such that for every t≥0,

Z

0

gp(x)e−txdx = e−t

2 p+2,

and set fp for the density of ||W||p. By (2) and a change of variable we obtain

(5) fp(x) = X

n≥1

µpnxp−1gp(xppn)−(p/2+1)) for every x≥0, with the notation

µpn = p(λpn)−(p/2+1)ψpn(0) Z

R

ψpn(y)dy

.

(5)

Let us consider the case p = 2. Changing the variable z(x) = y(21/4x) in Equation (4.2.1) of [8], one gets

λ2n = 2n−1

√2 and ψ2n(x) = e−x2/

2Hn(21/4x) 2n218(n!)12π14

for everyn ≥1, whereHn stands for then−th Hermite polynomial with physical notations, in other words the coefficient of tn in the series expansion e2xt−t2 :

e2xt−t2 = X

n≥1

Hn(x)tn n!

for every x, t∈R. Since on the other hand the density g2 is explicit and equals g2(x) = 1

2√

πe−1/4xx−3/2, we obtain after some simple calculations the following expansion

f2(x) = x−2X

p≥0

(−1)p(2p)!(4p+ 1)

√π22p(p!)2 e−(4p+1)2/8x2,

which could of course have been obtained in a shorter way, inverting directly the Cameron- Martin formula (4) and using the series expansion of (1 + t)−1/2. However, except in the cases p= 1 and p = 2, the coefficients µpn cannot be made explicit in terms of usual special functions (see Lemma 4.2 in [3] and the discussion thereafter), although several variational formulæ for the eigenvalues λpn can be derived (see [6] for an encyclopedic account).

Aknowledgement. This paper was written under the auspices of Grant ANR-09-BLAN- 0084-01.

References

[1] M. AbramowitzandI. A. Stegun.Handbook of mathematical functions.Wiley, New-York, 1972.

[2] R. P. Bell.Eigen-values and eigen-functions for the operator d2/dx2− |x|.Philos. Magazine (7) 35, 582-588, 1944.

[3] M. Jeanblanc, J. PitmanandM. Yor.The Feynman-Kac formula and decomposition of Brownian paths.Mat. Apl. Comput.16(1), 27-52, 1997.

[4] M. Kac. On the average of a certain Wiener functional and a related limit theorem in calculus of probability.Trans. Amer. Math. Soc.59, 401-414, 1946.

[5] N. Privault.Potential theory in classical probability. In Franz et al. (ed.)The Greifswald 2007 Summer School on Quantum Potential Theory.Lecture Notes Math.1954, pp. 3-59, 2008.

[6] M. ReedandB. Simon.Methods of modern mathematical physics. IV: Analysis of operators.Academic Press, New York, 1978.

[7] L. Tak´acs. On the distribution of the integral of the absolute value of the Brownian motion. Ann.

Appl. Probab.3(1), 186-197, 1993.

[8] E. C. Titchmarsh. Eigenfunction expansions associated with second-order differential equations.

Clarendon Press, Oxford, 1946.

Laboratoire Paul Painlev´e, U. F. R. de Math´ematiques, Universit´e de Lille 1, F-59655 Villeneuve d’Ascq Cedex. Email : simon@math.univ-lille1.fr

Références

Documents relatifs

RBsumB. - Le mouvement d'une particule selon Brown dans une suspension cause un Blargisse- ment de la ligne d'absorption resonante. Au moyen de cet effet le developpement des

First, we give an elementary proof of the fact that the solution to the SDE has a positive density for all t &gt; 0 when the diffusion coefficient does not vanish, echoing in

The classical scaled path properties of branching Brownian motion (BBM) have now been well-studied: for ex- ample, see Lee [13] and Hardy and Harris [3] for large deviation results

This is to say, we get an expression for the Itô excursion measure of the reflected Langevin process as a limit of known measures.. This result resembles the classical approximation

Theorem 1 further leads to an explicit expression for the Laplace transform of the stationary distribution of reflected Brownian motion in an arbitrary convex wedge, as stated in

This convergence result will be instrumental in Section 4 to prove that the Cartan development in diffeomorphism spaces of the time rescaled kinetic Brownian motion in Hilbert spaces

We focus here on results of Berestycki, Berestycki and Schweinsberg [5] concerning the branching Brownian motion in a strip, where particles move according to a Brownian motion

The linear Boltzmann equation has been derived (globally in time) from the dynamics of a tagged particle in a low density Lorentz gas, meaning that.. • the obstacles are