ebook img

On a flow of transformations of a Wiener space PDF

0.15 MB·
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview On a flow of transformations of a Wiener space

ON A FLOW OF TRANSFORMATIONS OF A WIENER SPACE J. NAJNUDEL, D. STROOCK, M. YOR 1 Abstract. In this paper, we define, via Fourier transform, an ergodic flow of transfor- 1 mations of a Wiener space which preserves the law of the Ornstein-Uhlenbeck process and 0 which interpolates the iterations of a transformation previously defined by Jeulin and Yor. 2 Then,wegiveamoreexplicitexpressionforthisflow,andweconstructfromitacontinuous n gaussian process indexed by R2, such that all its restriction obtained by fixing the first a coordinate are Ornstein-Uhlenbeck processes. J 5 2 1. Introduction ] R An abstract Wiener space is a triple (H,E, ) consisting of a separable, real Hilbert W P space H, a separable real Banach space E in which H is continuously embedded as a dense . subspace, and a Borel probability measure on E with the property that, for each x∗ E∗, h t the -distribution of the map x E Wx,x∗ R, from E to R, is a centered gau∈ssian a W ∈ 7−→ h i ∈ m random distribution with variance khx∗k2H, where hx∗ is the element of H determined by [ (h,hx∗)H = h,x∗ for all h H. See Chapter 8 of [5] for more information on this topic. h i ∈ 1 Because hx∗ : x∗ E∗ is dense in H and hx∗ H = ,x∗ L2(W), there is a unique { ∈ } k k kh· ik v isometry, known as the Paley–Wiener map, : H L2( ) such that (h) = ,x∗ 4 I 7−→ W I h· i 7 if h = hx∗. In fact, for each h ∈ H, I(h) under W is a centered Gaussian variable with 9 variance khk2H. Because, when h = hx∗, I(h) provides an extention of (·,h)H to E, for 4 intuitive purposes one can think of x [ (h)](x) as a giving meaning to the inner product . I 1 x (x,h) , although for general h this will be defined only up to a set of -measure 0. H 0 W An important property of abstract Wiener spaces is that they are invariant under orthog- 1 1 onal transformations on H. To be precise, given an orthogonal transformation on H, O : there is a -almost surely unique T : E E with the property that, for each h H, v O W −→ ∈ i (h) T = ( ⊤h) -almost surely. Notice that this is the relation which one would X O I ◦ I O W predict if one thinks of [ (h)](x) as the inner product of x with h. In general, T can be r I O a constructed by choosing {x∗m : m ≥ 1} ⊆ E∗ so the {hx∗m : m ≥ 1} is an orthonormal basis in H and then taking ∞ TOx = hx,x∗miOhx∗m, m=1 X where the series converges in E for -almost every x as well as in Lp( ;E) for every W W p [1, ). See Theorem 8.3.14 in [5] for details. In the case when admits an extension ∈ ∞ O as a continuous map on E into itself, T can be the taken equal to that extension. In any O case, it is an easy matter to check that the measure is preserved by T . Less obvious O W is a theorem, originally formulated by I.M. Segal (cf. [6]), which says that T is ergodic if O and only admits no non-trivial, finite dimensional, invariant subspace. Equivalently, T O O Date: January 27, 2011. 1 is ergodic if and only if the complexification has a continuous spectrum as a unitary c O operator on the complexification H of H. c The classical Wiener space provides a rich source of examples to which the preceding applies. Namely, take H = H1 to be the space of absolutely continuous h Θ whose 0 ∈ derivative h˙ is in L2([0,∞)), and set khkH01 = kh˙kL2([0,∞)). Then H01 with norm k · kH01 is a separable Hilbert space. Next, take E = Θ, where Θ is the space of continuous paths θ : [0, ) R such that θ(0) = 0 and ∞ −→ θ(t) | | 0 as t > 0 tends to 0 or , t12 log(e+ logt ) −→ ∞ | | and set θ(t) θ = sup | | . k kΘ t>0 t12 log(e+ logt ) | | Then Θ with norm is a separable Banach space in which H1 is continuously embedded k · kΘ 0 as a dense subspace. Finally, the renowned theorem of Wiener combined with the Brownian law of the iterated logarithm says that there is a Borel probability measure WH01 on Θ for which (H01,Θ,WH01) is an abstract Wiener space. Indeed, it is the clasRsical Wiener space on which the abstraction is modeled, and WH01 is the distribution of an -valued Brownian motion. One of the simplest examples of an orthogonal transformation on H1 for which the as- 0 sociated transformation on Θ is ergodic is the Brownian scaling map S given by S θ(t) = α α 1 α−2θ(αt) for α > 0. It is an easy matter to check that the restriction Oα of Sα to H01 is orthogonal, and so, since S is continuous on Θ, we can take T = S . Furthermore, as long α Oα α as α = 1, an elementary computation shows that lim g, nh = 0, first for smooth 6 n→∞ Oα H g, h H1 with compact support in (0, ) and thence for all g, h H1. Hence, when α = 1, ∈ 0 ∞ (cid:0) ∈(cid:1) 0 6 admits no non-trivial, finite dimensional subspace, and therefore S is ergodic; and so, α α O by the Birkoff’s Individual Ergodic Theorem, for p ∈ [1,∞) and f ∈ Lp(WH01), n−1 1 nl→im∞ n f ◦Sαn = f dWH01 m=0 Z X both WH01-almost surely and in Lp(WH01). Moreover, since {Sα : α ∈ (0,∞)} is a mul- tiplicative semigroup in the sense that S = S S , one has the continuous parameter αβ α β ◦ version 1 a dα al→im∞ loga (f ◦Sα) α = f dWH01 Z1 Z of the preceding result. A more challenging ergodic transformation of the classical Wiener space was studied by Jeulin and Yor (see [1], [2] and [4]), and, in the framework of this article, it is obtained by considering the transformation on H1, defined by O 0 t h(s) [ h](t) = h(t) ds. (1.1) O − s 2 Z0 An elementary calculation shows that is orthogonal. Moreover, admits a continuous O O extension to Θ given by replacing h H1 in (1.1) by θ Θ. That is ∈ 0 ∈ t θ(s) [T θ] = θ(t) ds for θ Θ and t 0. (1.2) O − s ∈ ≥ Z0 In addition, one can check that lim g, nh = 0 for all g, h H1, which proves that n→∞ O H01 ∈ 0 TO is ergodic for WH01. (cid:0) (cid:1) In order to study the transformation T in greater detail, it will be convenient to refor- O mulate it in terms of the Ornstein–Uhlenbeck process. That is, take HU to be the space of absolutely continuous functions h : R R such that −→ h 1h(t)2 +h˙(t)2 dt < . k kHU ≡ s R 4 ∞ Z (cid:0) (cid:1) Then HU becomes a separable Hilbert space with norm . Moreover, the map F : HU k · k H1 HU given by 0 −→ [F(g)](t) = e−2tg(et), for g H1 and t R, (1.3) ∈ 0 ∈ is an isometric surjection which extends as an isometry from Θ onto Banach space of con- U tinuous ω : R satisfying lim |ω(t)| = 0 with norm ω = sup log(e+ t ) −1 ω(t) . −→ |t|→∞ log|t| k kU t∈R | | | | Thus, (HU,U,WHU) is an abstract Wiener space, where WHU = F∗W(cid:0) H01 is the(cid:1)image of WH01 under the map F. In fact, WHU is the distribution of a standard, reverisible Ornstein– Uhlenbeck process. Notethatthe scaling transformationsfor theclassical Wiener space become translations in the Ornstein–Uhlenbeck setting. Namely, for each α > 0, F S = τ F, where τ denotes α logα s ◦ ◦ the time-translation map given by [τ ω](t) = ω(s+ t). Thus, for s = 0, the results proved s 6 about the scaling maps say that τ is an ergodic transformation for . In particular, for s HU W p [1, ) and f Lp( HU), ∈ ∞ ∈ W 1 n−1 1 T lim f τ = lim f τ ds = f d ns s HU n→∞ n ◦ T→∞ T ◦ W m=0 Z0 Z X both -almost surely and in Lp( ). HU HU W W The main goal of this article is to show that the reformulation of transformation T O coming from the Jeulin–Yor transformation in terms of the Ornstein–Uhlenbeck process allows us to embed T in a continuous-time flow of transformations on the space , each O U of which is WH01-measure preserving and all but one of which is ergodic. In Section 2, this flow is described via Fourier transforms. In Section 3, a direct and more explicit expression, involving hypergeometric functions and principal values, is computed. In Section 4, we study the two-parameter gaussian process which is induced by the flow introduced in Section 2. In particular, we compute its covariance and prove that it admits a version which is jointly continuous in its parameters. 2. Preliminary description of the flow Let and T be the transformations on H1 and Θ given by (1.1) and (1.2), and recall O O 0 the unitary map F : H1 HU in (1.3) and its continuous extension as an isometry from 0 −→ 3 Θ onto . Clearly, the inverse of F is given by U F−1(ω)(t) = √tω(logt) for t > 0. Because F is unitary and is orthogonal on H1, F F−1 is an orthognal transfor- O 0 − ◦O◦ mation on HU, and because S := F T F−1 O − ◦ ◦ is continuous extension of F F−1 to , we can identify S as T−F◦O◦F−1. − ◦O◦ U Another expression for action of S is ∞ [S(ω)](t) = ω(t)+ e−2sω(t s)ds for t R. − − ∈ Z0 Equivalently, S(ω) = ω µ, ∗ where µ is the finite, signed measure µ given by µ := δ +e−2t1 dt. 0 t≥0 − To confirm that ω µ is well-defined as a Lebesgue integral and that it maps continuously into itself, note tha∗t, for any ω and t R, U ∈ U ∈ ∞ ∞ s s e−2 ω(t s) ds ω e−2 log e+ t +s)ds U | − | ≤ k k | | Z0 Z0 (cid:0) ∞ s ω log(e+ t ) e−2(1+s)ds 9 ω log(e+ t ) U U ≤ k k | | ≤ k k | | Z0 The Fourier transform µ of µ is given by ∞ 1 1 2iλ µ(λ) = e−iλxdµ(x) = 1+ e−x(1/2+iλ)dx = 1+ = − = e−2iArctg(2λ). b − − 1/2+iλ 1+2iλ ZR Z0 Hbence, for all h HU and λ R, ∈ ∈ [ h µ(λ) = e−2iArctg(2λ)h(λ), (2.1) ∗ which, since b 1 h 2 = h(λ) 2 1+4λ2 dλ, k kHU 8π | | R Z (cid:0) (cid:1) provides another proof that S ↾ HU is isometbric. The preceeding, and especially (2.1), suggests a natural way to embed S ↾ HU into a continuous group of orthogonal transformations. Namely, for u R, let µ∗u to be the unique ∈ tempered distribution whose Fourier transform is given by µ∗u(λ) = e−2iuArctg(2λ), (2.2) and define uϕ = ϕ µ∗u for ϕ in the Schwartz test function class S of smooth functions S ∗ c which, together with all their derivatives, are rapidly decreasing. Because uϕ(λ) = e−2iuArctg(2λ)ϕˆ(λ), S 4 d it is obvious that u has a unique extension as an orthogonal transformation on HU, which we will again denoSte by u. Furthermore, it is clear that u+v = u v for all u,v R. Finally, for all g,h HU,Su R, S S ◦ S ∈ ∈ ∈ 1 (g, uh)HU = g(λ)h(λ)e−2iuArctg(2λ)(1+4λ2)dλ S 8π R Z = 1 bπ/2 gb tan(τ) h tan(τ) 1+tan2(τ) 2e−2iuτ dτ, 16π 2 2 Z−π/2 (cid:18) (cid:19) (cid:18) (cid:19) (cid:0) (cid:1) where b b 1 π/2 tan(τ) tan(τ) 1 g h 1+tan2(τ) 2 dτ = g(λ) h(λ) (1+4λ2)dλ 16π 2 2 8π | || | Z−π/2(cid:12) (cid:18) (cid:19)(cid:12) (cid:12) (cid:18) (cid:19)(cid:12) ZR (cid:12) (cid:12) (cid:12) (cid:12)(cid:0) (cid:1) 1 (cid:12)b (cid:12) (cid:12)b 1/2 (cid:12) 1/2 b b (cid:12) g(λ) 2(1+(cid:12) (cid:12)4λ2)dλ (cid:12) h(λ) 2(1+4λ2)dλ = g HU h HU < . ≤ 8π | | | | || || || || ∞ R R (cid:18)Z (cid:19) (cid:18)Z (cid:19) Hence, by Riemann–Lebesgue lemma, shows that (g, uh) tends to zero when u goes to b HU b S | | infinity. Now define the associated transformations Su := TSu on for each u R. By the U ∈ general theory summarized in the introduction and the preceding discussion, we know that Su : u R is a flow of HU-measure preserving transformations and that, for each u = 0, { ∈ } W 6 Su is ergodic. 3. A more explicit expression So far we know very little about the transformations Su for general u R. By getting a ∈ handle on the tempered distributions µ∗u, in this section we will attempt to find out a little more. We begin with the case when u is an integer n Z. Recalling that µ = δ0 +e−2t1t≥0dt, ∈ − one can use induction to check that, for n 0, ≥ µ∗n = ( 1)n δ +e−2tL′ (t)1 dt , − 0 n t≥0 where L is the nth Laguerre polynomial. Indeed, the Laguerre polynomials satisfy the n (cid:0) (cid:1) following relations: for all n 0, ≥ L (0) = 1 n and for all n 0, t R, ≥ ∈ L′ (t) = L′ (t) L (t). n+1 n − n Similarly, starting from µ∗−1 = δ0 +e2t1t≥0dt, one finds that − µ∗n = ( 1)n δ +e2tL′ ( t)1 dt − 0 n − t≤0 for n 0. In particular, µ∗n is a finite, signed measure for n Z and Snω can be identified (cid:0) (cid:1) as µ∗n≤ ω for all ω and n Z. ∈ As th∗e next resul∈t sUhows, wh∈en u / Z, µ∗u is more singular tempered distribution than a ∈ finite, signed measure. Proposition 3.1. For each u / Z, the distribution µ∗u is given by the following formula: ∈ sin(πu) µ∗u = cos(πu)δ (x)+ pv(1/x)+Φ (x), (3.1) 0 u π 5 where pv denotes the principal value, and Φ L2(R) is the function for which Φ (x) equals u u ∈ u sin(πu) ∞ (1 usgn(x)) x k Γ′ Γ′ e−|x|/2 − k| | (1+k usgn(x)) (1+k) − π k!(k +1)! Γ − − Γ k=0 (cid:20) X Γ′ sin(πu) sinπu (2+k)+log( x ) + , −Γ | | πx − πx (cid:21) (cid:19) Γ′/Γ being the logarithmic derivative of the Euler gamma function and ( ) being the k Pochhammer symbol. Proof. Define the functions ψu and θu from R∗ = R 0 to R so that θu(x) = e−x2ψu(x) and \{ } ψ (x) equals u u sin(πu) ∞ (1 usgn(x)) x k Γ′ Γ′ k − | | (1+k usgn(x)) (1+k) − π k!(k +1)! Γ − − Γ k=0 (cid:20) X Γ′ sin(πu) (2+k)+log( x ) + . −Γ | | πx (cid:21) From Lebedev [3], p. 264, equation (9.10.6), with the parameters α = 1 u or α = 1+u, n = 1, z = x or z = x, the function ψ satisfies, for all x R∗, the differ−ential equation: u − ∈ xψ′′(x)+(2 x )ψ′(x)+(u sgn(x))ψ (x) = 0, u −| | u − u and grows at most polynomially at infinity. One then deduces that θ decreases as least u exponentially at infinity, and satisfies (for x = 0) the following equation: 6 x xθ′′(x)+2θ′(x)+ u θ (x) = 0. (3.2) u u − 4 u (cid:16) (cid:17) At the same time, by writing e−|x|/2 = (e−|x|/2 1)+1 − and expanding θ (x) accordingly, we obtain: u sin(πu) usin(πu) Γ′ Γ′ Γ′ θ (x) = (1 usgn(x)) (1) (2)+log( x ) u πx − π Γ − − Γ − Γ | | (cid:20) (cid:21) sin(πu) sgn(x)+η (x), u − 2π for η (x) = xη(1)(x)+ x η(2)(x)+xlog( x )η(3)(x)+ x log( x )η(4)(x), u u | | u | | u | | | | u (1) (2) (3) (4) where η , η , η , η are all smooth functions. The derivatives of the functions x, x , u u u u | | xlog x , x log x inthesense ofthedistributions areobtainedbyinterpreting their ordinary | | | | | | derivatives as distributions. Similarly, the product by x of their second distributional deriva- tives are obtained by multiplying their ordinary second derivatives by x. Hence, both η′(x) u and xη′′(x) as distributions can be obtained by computing η′(x) and xη′′(x) as functions on u u u R∗. Now, let ν be the distribution given by the expression: u sin(πu) sin(πu) ν (x) = cos(πu)δ (x)+ pv(1/x)+ θ (x) . (3.3) u 0 u π − πx (cid:20) (cid:21) 6 Note that the term in brackets, in the definition of ν , is a locally integrable function, and u that ν coincides with the function θ in the complement of the neighborhood of zero. Let us u u now prove that ν satisfies the analog of the equation (3.2), in the sense of the distributions. u One has: sin(πu) usin(πu) Γ′ ν (x) = cos(πu)δ (x)+ pv(1/x) (1 usgn(x)) u 0 π − π Γ − (cid:20) Γ′ Γ′ sin(πu) (1) (2)+log( x ) sgn(x)+η (x). u −Γ − Γ | | − 2π (cid:21) Since Γ′ Γ′ d (Γ(1+u)Γ(1 u)) d (πu/sin(πu)) 1 (1+u) (1 u) = du − = du = πcot(πu), Γ − Γ − Γ(1+u)Γ(1 u) πu/sin(πu) u − − one obtains, after straightforward computation, sin(πu) ucos(πu) usin(πu) ν (x) = cos(π)δ (x)+ pv(1/x) sgn(x) log( x )+c(u)+η (x), u 0 u π − 2 − π | | where c(u) does not depend on x. One deduces that sin(πu) ν (x) = cos(πu)δ (x)+ pv(1/x)+χ (x), u 0 u,1 π where χ denotes a locally integrable function. Moreover, u,1 sin(πu) usin(πu) ν′(x) = cos(πu)δ′(x) fp(1/x2) ucos(πu)δ (x) pv(1/x)+η′(x), u 0 − π − 0 − π u where fp(1/x2) denotes the finite part of 1/x2, and then sin(πu) usin(πu) xν′(x) = cos(πu)δ (x) pv(1/x) +xη′(x). u − 0 − π − π u By differentiating again, one obtains: sin(πu) ν′(x)+xν′′(x) = cos(πu)δ′(x)+ fp(1/x2)+η′(x)+xη′′(x). u u − 0 π u u Therefore, x sin(πu) xν′′(x)+2ν′(x)+ u ν (x) = χ (x)+ cos(πu)δ′(x)+ fp(1/x2) u u − 4 u u,2 − 0 π (cid:18) (cid:19) (cid:16) (cid:17) sin(πu) usin(πu) + cos(πu)δ′(x) fp(1/x2) ucos(πu)δ (x) pv(1/x) 0 − π − 0 − π (cid:18) (cid:19) sin(πu) +u cos(πu)δ (x)+ pv(1/x) = χ (x), 0 u,2 π (cid:18) (cid:19) where χ is a locally integrable function. Since θ satisfies (3.2), χ is identically zero. u,2 u u,2 Hence, ν is a tempered distribution solving the differential equation: u x xν′′(x)+2ν′(x)+ u ν (x) = 0, u u − 4 u (cid:16) (cid:17) or equivalently, x d2 ν (x) (xν (x)) uν (x) = 0. 4 u − d2x u − u 7 Multiplying by 4i and taking the Fourier transform (in the sense of the distributions), one − deduces: ν ′(λ)(1+4λ2) = 4iuν (λ). u u − This linear equation admits a unique solution, up to a multiplicative factor c: b b λ 4iu ν (λ) = cexp − dt = cexp( 2iuArctg(2λ)). u 1+4t2 − (cid:18)Z0 (cid:19) Hence, ν is proportional to µ∗u. In order to determine the constant c, let us observe that u b the distribution ν given by u,0 csin(πu) ν (x) = ν (x) ccos(πu)δ (x) pv(1/x) u,0 u 0 − − π admits the Fourier transform: ν (λ) = ce−2iuArctg(2λ) ce−πiusgn(λ). u,0 − One deduces that ν is a function in L2, which implies that ν is also a function in L2, u,0 u,0 and then locally integrable.dSince the last term in (3.3) is also a locally integrable function, one deduces that c = 1, and then d µ∗u = ν , u which proves Proposition 3.1. (cid:3) The reasonably explicit expression for µ∗u found in Proposition 3.1 yields a reaonably explicit expression for the action of u. Indeed, only the term pv(1/x) is a source of concern. S However, convolution with respect of pv(1/x) is, apart from a multiplicative constant, just the Hilbert transform, whose properties are well-known. In particular, it is a translation invariant, bounded map on L2(R), and as such it is also a bounded map on HU. Thus, we can unambiguously write u(h) = h µ∗u for all h HU. On the other hand, the S ∗ ∈ interpretation of ω µ∗u for ω needs some thought. No doubt, ω µ∗u is well-defined as ∗ ∈ U ∗ an element of S ′, the space tempered distributions, but it is not immediately obvious that it is can be represented by an element of or, if it can, that the element of which represents U U it can be identified as Suω. In fact, the best that we should expect is that such statements will be true of -almost every ω . The following result justifies that expectation. HU W ∈ U Proposition 3.2. For -almost every ω , the tempered distribution ω µ∗u is repre- HU W ∈ U ∗ sented by an element of which can be can be identified as Suω. U Proof. Recall that, for ϕ S, ϕ µ∗−u is the element of S whose Fourier transform is given ∈ ∗ by ϕ\µ∗−u(λ) = ϕ(λ)e2iuArctg(2λ) for all λ R. ∗ ∈ Also, if T S ′, then T µ∗u is the tempered distribution whose action on ϕ S is given ∈ ∗ ∈ by b S ϕ,T µ∗u S′ = S ϕ µ∗−u,T S′. h ∗ i h ∗ i Now choose an orthonormal basis h : n 1 for HU all of whose members are elements n { ≥ } of S, and, for each n 1, set g = 1h +h′′. Next, think of g as the element of ∗ whose ≥ n 4 n n n U action on ω is given by ∈ U U ω,gn U∗ = S gn,ω S′. h i h i 8 It is then an easy matter to check that, in the notation of the introduction, h = h . Hence, n gn if B is the subset of ω for which ∈ U n n ω = lim S gn,ω S′hn and Suω = lim S gn,ω S′hn µ∗u, n→∞ h i n→∞ h i ∗ m=1 m=1 X X where the convergence is in , then (B) = 1. HU U W Now let ω B. Then, for each ϕ S, ∈ ∈ n S ϕ,ω µ∗u S′ = S ϕ µ∗−u,ω S′ = lim S gn,ω S′S ϕ,hn µ∗u S′ h ∗ i h ∗ i n→∞ h i h ∗ i m=1 X n = lim S gn,ω S′S ϕ, uhn S′ = S ϕ,Suω S′. n→∞ h i h S i h i m=1 X Thus, for ω B, ω µ∗u S ′ is represented by Suω . (cid:3) ∈ ∗ ∈ ∈ U 4. A two parameter gaussian process By construction, Suω(t) : (u,t) R2 is a gaussian family in L2( ). In this conclud- HU { ∈ } W ing section, we will show that this family admits a modification which is jointly continuous in (u,t). Let ϕ,ψ S and u,v R2 be given. Then, by Proposition 3.2, for HU-almost every ∈ ∈ W ω , ∈ U ϕ(s)ψ(t)(Su(ω))(s)(Sv(ω))(t)dsdt = S ϕ,ω µ∗u S′S ψ,ω µ∗v S′, h ∗ i h ∗ i ZZ R2 where the integral in the left-hand side is absolutely convergent. Because E Suω(t)2 is WHU finite and independent of (u,t) ∈ R2, by taking the expectation with respect t(cid:2)o WHU a(cid:3)nd using (2.2), one can pass from this to ϕ(s)ψ(t)EWHU (Su(ω))(s)(Sv(ω))(t) dsdt = EWHU Shϕ,ω ∗µ∗uiS′Shψ,ω ∗µ∗viS′ ZZ R2 h i (cid:2) (cid:3) 2 ∞ e2i(u−v)Arctg(2λ) 2 ei[(t−s)λ+2(u−v)Arctg(2λ)] = ϕ(λ)ψ(λ)dλ = ϕ(s)ψ(t)dsdtdλ. π 1+4λ2 π 1+4λ2 Z−∞ ZZZ R3 b b Hence, 2 ∞ ei[(t−s)λ+2(u−v)Arctg(2λ)] E [Su(ω))(s)(Sv(ω))(t)] = dλ, (4.1) WHU π 1+4λ2 Z−∞ first for almost every and then, by continuity, for all (s,t) R2. In particular, we now know that the -distribution of Su(ω))(t) : (u,t) R2 is s∈tationary. HU W { ∈ } To show that there is a continuous version of this process, we will use Kolmogorov’s continuity criterion, which, because it is stationary and gaussian, comes down to showing that 1 E [(Su(ω))(s)(Sv(ω))(t)] C (u,s) (v,t) α − WHU ≤ − 9 (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) for some C < and α > 0. But ∞ 2 ∞ dλ 1 E [(Su(ω))(s)(Sv(ω))(t)] ei[(t−s)λ+2(u−v)Arctg(2λ)] 1 − WHU ≤ π 1+4λ2 − Z−∞ (cid:12) 2 ∞ dλ (cid:12) 2 ∞ dλ (cid:12) (cid:12) (cid:12) ei(t−s)λ 1 +(cid:12) e2i(cid:12)(u−v)Arctg(2λ) 1 (cid:12) ≤ π 1+4λ2 − π 1+4λ2 − Z−∞ Z−∞ 2 ∞ dλ (cid:12) (cid:12) 4 ∞ dλ (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) ( t s λ 2)+ (u v)Arctg(2λ) , ≤ π 1+4λ2 | − || |∧ π 1+4λ2| − | Z−∞ Z−∞ and, after simple estimation, this shows that 1 1 E[(Su(ω))(s)(Sv(ω))(t)] C u v + t s 1+log 1+ , | − | ≤ | − | | − | (t s)2 (cid:20) (cid:18) (cid:18) − (cid:19)(cid:19)(cid:21) where C < . Clearly, the desired conclusion follows. ∞ Remark 4.1. A question about filtrations comes naturally when one considers the group of transformations (Su)u∈R on the space U. Indeed, for all t,u ∈ R, let Ftu be the σ-algebra generatedby the -negligiblesubsets of ofandthevariables(Su(ω))(s), fors ( ,t] HU W U ∈ −∞ (these variables are well-defined up to a negligible set). From the results of Jeulin and Yor, oune Rqu:ite easily deduces the following properties of the filtrations of the form (Ftu)t∈R for ∈ For all t,u R, u is generated by u+1 and (Su(ω))(t). • For all t,u ∈ R, Ftu+1 and (Su(ω))(tF) tare independent under . • Forall t,u ∈R, tFhet decreasing intersection of u+n for n Z iWs tHriUvial (i.e. it satisfies • ∈ Ft ∈ the zero-one law). If u R is fixed, the σ-algebra generated by u+n for t R does not depend on • n Z∈. Ft ∈ ∈ All these statements concern the sequence of filtrations ( u+n)n∈Z for fixed u R. A natural F ∈ question arises: how can these results be extended to the continuous family of filtrations ( u)u∈R? Unfortunately, for the moment, we have no answer to this question (in particular F the family does not seem to be decreasing with u). References [1] T. Jeulin, M. Yor, Filtration des ponts browniens et ´equations diff´erentielles stochastiques lin´eaires, S´eminaire de Probabilit´es, XXIV, Lecture Notes in Math., vol. 1426, Springer-Verlag, 1990, pp. 227– 265. [2] P.-A. Meyer, Sur une transformation du mouvement brownien due `a Jeulin et Yor, S´eminaire de Proba- bilit´es, XXVIII, Lecture Notes in Math., vol. 1583, Springer-Verlag,1994, pp. 98–101. [3] N.-N. Lebedev, Special functions and their applications (Translated from Russian), Dover, New York, 1972. [4] M. Yor, Some aspects of Brownian motion. Part I: Some Special Functionals, Lectures in Mathematics, Ed. ETH Zu¨rich, Birkhau¨ser, 1992. [5] D. Stroock, Probability Theory, an Analytic View, 2nd edition, Cambridge University Press, 2011. [6] D. Stroock, Some thoughts about Segal’s ergodic theorem, Colloq. Math. vol. 118 #1, pp. 89-105,2010. 10

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.