The Malliavin Calculus and Hypoelliptic Differential Operators
Total Page:16
File Type:pdf, Size:1020Kb
The Malliavin Calculus and Hypoelliptic Differential Operators Denis Bell Department of Mathematics, University of North Florida 4567 St. Johns Bluff Road South,Jacksonville, FL 32224, U. S. A. email: [email protected] Dedicated to the memory of Paul Malliavin 1 Introduction The first half of the twentieth century saw the rise of stochastic analysis as an in- dependent branch of probability theory. Remarkable results were obtained. Wiener constructed a rigorous mathematical model of Brownian motion. Kolmogorov discov- ered that the transition probabilities of a diffusion process define the fundamental solution to the associated heat equation. It^odeveloped a theory of stochastic integra- tion that made it possible to represent a diffusion process with a given infinitesimal generator as the solution of a stochastic differential equation (sde). These develop- ments established a link betwen the classical area of partial differential equations and the nascent area of stochastic calculus, whereby results in the former area could be used to prove results in the latter. More specifically, let X0;X1;:::;Xn denote a collection of smooth vector fields on Rd and define the second-order differential operator1 n 1 X L ≡ X2 + X (1:1) 2 i 0 i=1 Consider the (Stratonovich) sde n X dξt = Xi(ξt) ◦ dwi + Xo(ξt)dt (1:2) i=1 where w = (w1; :::; wn) is an n-dimensional standard Wiener process. Then the so- lution ξ to (1.2) is a time-homogeneous Markov process with infinitesimal generator L, whose transition probabilities p(t; x; dy) satisfy the following PDE (known as Kol- mogorov's forward equation) in the weak sense @p = L∗p: @t y 1The presence of the factor 1/2 in (1.1) is a matter of convenience and is otherwise unimportant. 1 The Malliavin Calculus and Hypoelliptic Differential Operators 2 A differential operator G is said to be hypoelliptic if, whenever Gu is smooth for some distribution u defined on some open subset of the domain of G, then u is smooth. In 1967, H¨ormanderproved that the operator L in (1.1) is hypoelliptic if the Lie algebra generated by X0;:::;Xn has dimension d throughout the domain of L. (This hypothesis is known as H¨ormander'scondition and will be referred to in the sequel as HC.) It follows immediately that the transition probabilities p of ξ in (1.2) have @p ∗ smooth densities at all positive times if the parabolic operator @t − L satisfies HC at ξ0. In the reverse direction, if it is possible to establish by a direct probabilistic argu- ment the existence of smooth transition probabilities for ξ under HC, then one can deduce hypoellipticity of L and thereby obtain a probabilistic proof of H¨ormander's theorem. This exciting line of research was initiated in the mid-seventies by Malliavin in his seminal paper [Ma]. The idea underlying Malliavin's probabilistic approach to hypoellipticity is the following. The measure p(t; x; dy) is the image of Wiener measure under the map gt : w 7! ξt. Now the Wiener measure has a well-understood analytic structure as an infinite-dimension analogue of the Gaussian distribution in Euclidean space. If the map g were smooth, then regularity properties of p could be obtained by a process of integration by parts on the Wiener space. In fact, this is not the case, it turns out that g is most pathological from the standpoint of standard calculus. Malliavin solved this problem by constructing a calculus applicable to the class of Wiener functionals defined by sde's with smooth coefficients. Malliavin's original work was a technical tour de force. It has since been clarified and extended, and has become a powerful tool in stochastic analysis. Different ap- proaches to the subject have been introduced, in particular the functional analytic approach of Kusuoka-Stroock [KS], the variational approach of Bismut [Bi.1], the el- ementary approach of the author [Be]. Following further pioneering work by Kusuoka and Stroock, the Malliavin calculus has yielded the desired probabilistic proof of H¨ormander'stheorem and much more. Achievements of the field to date include new contributions to filtering theory by Michel [Mi] and others, a deeper understanding of the Skorohod integral and the development of an anticipating stochastic calculus by Nualart and Pardoux [NP], an extension of Clark's formula by Ocone [Oc], Bismut's probabilistic analysis of the small-time asymptotics of the heat kernel of the Dirac operator on a Riemannian manifold [Bi.2] (which itself has led to striking develop- ments in the area of index theory) and, more recently, numerous applications in the area of mathematical finance related to the computation of Greeks (cf. e.g. Malliavin & Thalmaier [MT]). While these developments fall outside the scope of the present article, the interested reader is strongly encouraged to study them. The article is organized as follows. In Section 2 we derive Malliavin's integration by parts formula, by which one obtains smoothness of the transition probabilities discussed above. In particular the Malliavin covariance matrix, which plays a central role throughout, is introduced in this section of the article. We also state and prove a result of Bismut that makes explicit the connection between Malliavin's covariance matrix and H¨ormander'scondition. In Section 3, we give the proof of the probabilistic version of H¨ormander'stheorem. This result establishes the existence of smooth densities for the diffusion process ξt under the assumption that the vector fields X0;:::;Xn satisfy a parabolic version of HC at the point ξ0. The Malliavin Calculus and Hypoelliptic Differential Operators 3 In Section 4 we prove a generalized version of H¨ormander'stheorem. This result establishes hypoellipticity of H¨ormandertype operators under conditions that allow HC to fail on hypersurfaces of codimension 1. The theorem is sharp within the category of operators with smooth coefficients. 2 Integration by parts and the regularity of mea- sures As indicated above, the Malliavin calculus has reached a state of maturity, to the point where it is now used routinely by a wide variety of researchers in stochastic analysis and other areas of mathematics. Following the influential book by Nualart [Nu], the subject is generally formulated in terms of Sobolev spaces. In this section we present a more elementary approach to the theory. Following Malliavin, we will make use of the following result from harmonic anal- ysis. Lemma 2.1. Let ν denote a finite Borel measure on Rd. Suppose that for all non- 1 negative integers k and d1; : : : ; dk, there exists a constant C such that for all test (C compact support) functions φ on Rd Z @d1+···+dk d d φ dν ≤ Cjjφ1jj: (2:1) d 1 k R @x1 : : : @xk If this condition holds for k = 1, then ν is absolutely continuous with respect to the Lebesgue measure on Rd. If the condition holds for all k ≥ 1, then ν has a C1 density. The next result is included in order to motivate the approach that follows. It is a finite-dimensional analogue of the infinite-dimensional problem that will be addressed shortly. Theorem 2.2. Suppose T : Rm 7! Rd is a C2 map. Let γ denote the standard Gaussian measure on Rm and let ν ≡ γ ◦ T −1 denote the induced measure on Rd by T (also denoted by T (γ)). Define N to be the set of points in Rm where T has a non-surjective derivative. Then the condition γ(N) = 0 (2:2) is necessary and sufficient for absolute continuity of ν. Proof. The necessity of condition (2.2) follows immediately from Sard's theorem, which asserts that the set T (N) has zero Lebesgue measure. To prove sufficiency, we argue as follows. Define a d × d matrix σ(x) = DT (x)DT (x)∗: (2:3) Let k : [0; 1] 7! [0; 1] denote a sequence of smooth bump functions such that (i) ( 1; 0 ≤ t ≤ k k(t) = 0; t ≥ k + 1 The Malliavin Calculus and Hypoelliptic Differential Operators 4 p (ii) For all p ≥ 1, supkjD k(t)j < 1. d d Define Rk : R ⊗ R 7! [0; 1] by ( −1 k(jjαjj ; α 2 GL(d) Rk(α) = 0; α2 = GL(d) 1 d d Note that Rk is C since GL(d) is an open subset of R ⊗ R . d We define sequences of measures γk on C0 and νk on R by dγ k = R (σ(x)); dγ k −1 νk = γk ◦ T : Assume (2.2) holds, then σ(x) 2 GL(d) a.s. This implies νk ! ν in variation. Hence it suffices to prove absolute continuity of νk for every k. To this end, let e1; : : : ed denote d m m the standard basis of R and define, for 1 ≤ i ≤ d; hi : R 7! R by ( ∗ −1 DT (x) σ (x)ei; σ 2 GL(d) hi(x) = 0; σ2 = GL(d) Let φ be a test function on Rd. Then Z Z @φ −1 dνk = Dφ(T (x))ei k(jjσ (x)jj) dγ(x) Rd @xi Rm Z −1 = D(φ ◦ T )(x)hi(x) k(jjσ (x)jj) dγ(x): (2:4) Rm Denote by Div the (Gaussian) divergence operator, Div(G)(x) ≡< G(x); x > −traceDG(x): Integrating by parts in (2.4) yields Z @φ Z dνk = φ ◦ T (x)Xi(x) dγ(x) (2:5) Rd @xi Rd −1 where Xi = Div[ k(jjσ jj)hi]: Since Xi is continuous with compact support, it 1 follows that Xi 2 L (γ). From (2.5), we obtain Z @φ dνk ≤ jjφjj1jjXijjL1(γ) Rd @xi and the absolute continuity of νk follows from Lemma 2.1.