Geometry
Geometry
BRUCE K. DRIVER
Contents
1. Introduction 1
2. Manifold Primer 2
3. Riemannian Geometry Primer 15
4. Flows and Cartan’s Development Map 41
5. Stochastic Calculus on Manifolds 50
6. Heat Kernel Derivative Formula 75
7. Calculus on W (M ) 78
8. Malliavin’s Methods for Hypoelliptic Operators 107
9. Appendix: Martingale and SDE Estimates 119
References 129
1. Introduction
These notes represent a much expanded and updated version of the “mini course”
that the author gave at the ETH (Zürich) and the University of Zürich in February
of 1995. The purpose of these notes is to first provide some basic background
to Riemannian geometry and stochastic calculus on manifolds and then to cover
some of the more recent developments pertaining to analysis on “curved Wiener
spaces.” Essentially no differential geometry is assumed. However, it is assumed
that the reader is comfortable with stochastic calculus and differential equations
on Euclidean spaces. Here is a brief description of what will be covered in the text
below.
Section 2 is a basic introduction to differential geometry through imbedded sub-
manifolds. Section 3 is an introduction to the Riemannian geometry that will be
needed in the sequel. Section 4 records a number of results pertaining to flows of
vector fields and “Cartan’s rolling map.” The stochastic version of these results
will be important tools in the sequel. Section 5 is a rapid introduction to stochas-
tic calculus on manifolds and related geometric constructions. Section 6 briefly
gives applications of stochastic calculus on manifolds to representation formulas for
derivatives of heat kernels. Section 7 is devoted to the study of the calculus and in-
tegral geometry associated with the path space of a Riemannian manifold equipped
with “Wiener measure.” In particular, quasi-invariance, Poincaré and logarithmic
Sobolev inequalities are developed for the Wiener measure on path spaces in this
section. Section 8 is a short introduction to Malliavin’s probabilistic methods for
This research was partially supported by NSF Grants DMS 96-12651, DMS 99-71036 and DMS
0202939. This article will appear in “Real and Stochastic Analysis: New Perspectives.”
1
2 BRUCE K. DRIVER
dealing with hypoelliptic diffusions. The appendix in section 9 records some basic
martingale and stochastic differential equation estimates which are mostly used in
section 8.
Although the majority of these notes form a survey of known results, many proofs
have been cleaned up and some proofs are new. Moreover, Section 8 is written
using the geometric language introduced in these notes which is not completely
standard in the literature. I have also tried (without complete success) to give an
overview of many of the major techniques which have been used to date in this
subject. Although numerous references are given to the literature, the list is far
from complete. I apologize in advance to anyone who feels cheated by not being
included in the references. However, I do hope the list of references is sufficiently
rich that the interested reader will be able to find additional information by looking
at the related articles and the references that they contain.
Acknowledgement: It is pleasure to thank Professor A. Sznitman and the ETH
for their hospitality and support and the opportunity to give the talks which started
these notes. I also would like to thank Professor E. Bolthausen for his hospitality
and his role in arranging the first lecture to be held at University of Zürich.
2. Manifold Primer
Conventions:
(1) If A, B are linear operators on some vector space, then [A, B] := AB − BA
is the commutator of A and B.
(2) If X is a topological space we will write A ⊂o X, A ⊏ X and A ⊏⊏ X to
mean A is an open, closed, and respectively a compact subset of X.
(3) Given two sets A and B, the notation f : A → B will mean that f is
a function from a subset D(f ) ⊂ A to B. (We will allow D(f ) to be the
empty set.) The set D(f ) ⊂ A is called the domain of f and the subset
R(f ) := f (D(f )) ⊂ B is called the range of f. If f is injective, let f −1 :
B → A denote the inverse function with domain D(f −1 ) = R(f ) and range
R(f −1 ) = D(f ). If f : A → B and g : B → C, then g ◦ f denotes the
composite function from A to C with domain D(g ◦ f ) := f −1 (D(g)) and
range R(g ◦ f ) := g ◦ f (D(g ◦ f )) = g(R(f ) ∩ D(g)).
Notation 2.1. Throughout these notes, let E and V denote finite dimensional
vector spaces. A function F : E → V is said to be smooth if D(F ) is open in
E (D(F ) = ∅ is allowed) and F : D(F ) → V is infinitely differentiable. Given a
smooth function F : E → V, let F ′ (x) denote the differential of F at x ∈ D(F ).
Explicitly, F ′ (x) = DF (x) denotes the linear map from E to V determined by
d
(2.1) DF (x) a = F ′ (x)a := |0 F (x + ta) ∀ a ∈ E.
dt
We also let
d d
(2.2) F ′′ (x) (v, w) = F ′′ (x) (v, w) := (∂v ∂w F ) (x) = |0 |0 F (x + tv + sw) .
dt ds
2.1. Imbedded Submanifolds. Rather than describe the most abstract setting
for Riemannian geometry, for simplicity we choose to restrict our attention to
imbedded submanifolds of a Euclidean space E = RN . 1 We will equip RN with the
1Because of the Whitney imbedding theorem (see for example Theorem 6-3 in Auslander and
MacKenzie [9]), this is actually not a restriction.
CURVED WIENER SPACE ANALYSIS 3
This completes the proof because; 1) every matrix can be put into upper triangular
form by a similarity transformation, and 2) “det” and “tr” are invariant under
similarity transformations.
Definition 2.8. Let E and V be two finite dimensional vector spaces and M d ⊂ E
and N k ⊂ V be two imbedded submanifolds. A function f : M → N is said to be
smooth if for all charts x ∈ A(M ) and y ∈ A(N ) the function y ◦f ◦x−1 : Rd → Rk
is smooth.
Exercise 2.9. Let M d ⊂ E and N k ⊂ V be two imbedded submanifolds as in
Definition 2.8.
6 BRUCE K. DRIVER
at m shows
′
z −1 (z(m))z ′ (m) = I.
Therefore,
d d
σ ′ (0) = |0 x−1 (z< (m) + sw) = |0 z −1 (z< (m) + sw, 0)
ds ds
′
= z −1 ((z< (m), 0))(z< ′
(m)v, 0)
−1 ′ ′ ′
= z ((z< (m), 0))(z< (m)v, z> (m)v)
′
= z −1 (z(m))z ′ (m)v = v,
The reason for the strange notation in Eq. (2.5) will be explained after Notation
2.20. By definition, every element of Tm M is of the form σ ′ (0) where σ is a smooth
path into M such that σ (0) = m. Moreover by Theorem 2.13, {∂/∂xi |m }di=1 is a
basis for Tm M.
Definition 2.19. Suppose that f : M → V is a smooth function, m ∈ D(f ) and
vm ∈ Tm M. Write
d
vm f = df (vm ) := |0 f (σ(s)),
ds
where σ is any smooth path in M such that σ ′ (0) = vm . The function df : T M → V
will be called the differential of f.
Notation 2.20. If M and N are two manifolds f : M × N → V is a smooth
function, we will write dM f (·, n) to indicate that we are computing the differential
of the function m ∈ M → f (m, n) ∈ V for fixed n ∈ N.
To understand the notation in (2.5), suppose that f = F ◦ x = F (x1 , x2 , . . . , xd )
where F : Rd → R is a smooth function and x is a chart on M. Then
∂f (m) ∂
:= |m f = (Di F )(x(m)),
∂xi ∂xi
∂
where Di denotes the ith – partial derivative of F. Also notice that dxj ∂x
i |m = δij
i d i d
so that dx |Tm M i=1 is the dual basis of {∂/∂x |m }i=1 and therefore if vm ∈ Tm M
then
d
X ∂
(2.6) vm = dxi (vm ) |m .
i=1
∂xi
This explicitly exhibits vm as a first order differential operator acting on “germs”
of smooth functions defined near m ∈ M.
Remark 2.21 (Product Rule). Suppose that f : M → V and g : M → End(V ) are
smooth functions, then
d
vm (gf ) = |0 [g(σ(s))f (σ(s))] = vm g · f (m) + g(m)vm f
ds
10 BRUCE K. DRIVER
or equivalently
d(gf )(vm ) = dg(vm )f (m) + g(m)df (vm ).
This last equation will be abbreviated as d(gf ) = dg · f + gdf.
Definition 2.22. Let f : M → N be a smooth map of imbedded submanifolds.
Define the differential, f∗ , of f by
f∗ vm = (f ◦ σ)′ (0) ∈ Tf (m) N,
where vm = σ ′ (0) ∈ Tm M, and m ∈ D(f ).
Lemma 2.23. The differentials defined in Definitions 2.19 and 2.22 are well de-
fined linear maps on Tm M for each m ∈ D(f ).
Proof. I will only prove that f∗ is well defined, since the case of df is similar.
By Proposition 2.10, there is a smooth function F : E → V, such that f = F |M .
Therefore by the chain rule
d
(2.7) f∗ vm = (f ◦ σ)′ (0) := |0 f (σ(s)) = [F ′ (m)v]f (m) ,
ds f (σ(0))
where σ is a smooth path in M such that σ ′ (0) = vm . It follows from (2.7) that
f∗ vm does not depend on the choice of the path σ. It is also clear from (2.7), that
f∗ is linear on Tm M.
Remark 2.24. Suppose that F : E → V is a smooth function and that f := F |M .
Then as in the proof of Lemma 2.23,
(2.8) df (vm ) = F ′ (m)v
for all vm ∈ Tm M , and m ∈ D(f ). Incidentally, since the left hand sides of (2.7)
and (2.8) are defined “intrinsically,” the right members of (2.7) and (2.8) are inde-
pendent of the possible choices of functions F which extend f.
Lemma 2.25 (Chain Rules). Suppose that M, N, and P are imbedded submanifolds
and V is a finite dimensional vector space. Let f : M → N, g : N → P, and
h : N → V be smooth functions. Then:
(2.9) (g ◦ f )∗ vm = g∗ (f∗ vm ), ∀ vm ∈ T M
CURVED WIENER SPACE ANALYSIS 11
and
(2.10) d(h ◦ f )(vm ) = dh(f∗ vm ), ∀ vm ∈ T M.
These equations will be written more concisely as (g◦f )∗ = g∗ f∗ and d(h◦f ) = dhf∗
respectively.
Proof. Let σ be a smooth path in M such that vm = σ ′ (0). Then, see Figure 6,
(g ◦ f )∗ vm := (g ◦ f ◦ σ)′ (0) = g∗ (f ◦ σ)′ (0)
= g∗ f∗ σ ′ (0) = g∗ f∗ vm .
Similarly,
d
d(h ◦ f )(vm ) := |0 (h ◦ f ◦ σ)(s) = dh((f ◦ σ)′ (0))
ds
= dh(f∗ σ ′ (0)) = dh(f∗ vm ).
relative to the bases {∂/∂xi |m }di=1 of Tm M and {∂/∂y j |f (m) }kj=1 of Tf (m) N is
(∂(y j ◦ f )(m)/∂xi ). Indeed, if vm = di=1 v i ∂/∂xi |m , then
P
k
X
f∗ vm = dy j (f∗ vm )∂/∂y j |f (m)
j=1
k
X
= d(y j ◦ f )(vm )∂/∂y j |f (m) (by Eq. (2.10))
j=1
k X d
X ∂(y j ◦ f )(m)
= i
· dxi (vm )∂/∂y j |f (m) (by Eq. (2.11))
j=1 i=1
∂x
k X d
X ∂(y j ◦ f )(m) i
= i
v ∂/∂y j |f (m) .
j=1 i=1
∂x
Notation 2.37. The Lie bracket of two smooth vector fields, Y and W, on M is
the vector field [Y, W ] which acts on C ∞ (M ) by the formula
(2.13) [Y, W ]f := Y (W f ) − W (Y f ), ∀ f ∈ C ∞ (M ).
(In general one might suspect that [Y, W ] is a second order differential operator,
however this is not the case, see Exercise 2.38.) Sometimes it will be convenient to
write LY W for [Y, W ].
Exercise 2.38. Show that [Y, W ] is again a first order differential operator on
C ∞ (M ) coming from a vector-field. In particular, if x is a chart on M, Y =
Pd i i
Pd i i
i=1 Y ∂/∂x and W = i=1 W ∂/∂x , then on D(x),
d
X
(2.14) [Y, W ] = (Y W i − W Y i )∂/∂xi .
i=1
Proposition 2.39. If Y (m) = (m, y(m)) and W (m) = (m, w(m)) and y, w : M →
E are smooth functions such that y(m), w(m) ∈ τm M, then we may express the Lie
bracket, [Y, W ](m), as
(2.15) [Y, W ](m) = (m, (Y w − W y)(m)) = (m, dw(Y (m)) − dy(W (m))).
Proof. Let f be a smooth function M which we may take, by Proposition 2.10,
to be the restriction of a smooth function on E. Similarly we we may assume that
y and w are smooth functions on E such that y(m), w(m) ∈ τm M for all m ∈ M.
Then
(Y W − W Y )f = Y [f ′ w] − W [f ′ y]
= f ′′ (y, w) − f ′′ (w, y) + f ′ (Y w) − f ′ (W y)
(2.16) = f ′ (Y w − W y)
wherein the last equality we have use the fact that mixed partial derivatives com-
mute to conclude
f ′′ (u, v) − f ′′ (v, u) := (∂u ∂v − ∂v ∂u ) f = 0 ∀ u, v ∈ E.
Taking f = z> in Eq. (2.16) with z = (z< , z> ) being a chart on E as in Definition
2.2, shows
′
0 = (Y W − W Y )z> (m) = z> (dw(Y (m)) − dy(W (m)))
and thus (m, dw(Y (m)) − dy(W (m))) ∈ Tm M. With this observation, we then have
f ′ (Y w − W y) = df ((m, dw(Y (m)) − dy(W (m))))
which combined with Eq. (2.16) verifies Eq. (2.15).
Exercise 2.40. Let M = SL(n, R) and A, B ∈ sl(n, R) and à and B̃ be the
associated left invariant vector fields on M as introduced in Example 2.34. Show
^
h i
Ã, B̃ = [A, B] where [A, B] := AB − BA is the matrix commutator of A and B.
2.3. More References. The reader wishing to learn about manifolds is referred
to [1, 9, 19, 41, 42, 94, 110, 111, 112, 113, 114, 162]. The texts by Kobayashi and
Nomizu are very thorough while the books by Klingenberg give an idea of why
differential geometers are interested in loop spaces. There is a vast literature on
Lie groups and there representations. Here are just two books which I have found
very useful, [24, 176].
CURVED WIENER SPACE ANALYSIS 15
where
x
(m) := h∂/∂xi |m , ∂/∂xj |m im = g ∂/∂xi |m , ∂/∂xj |m .
gi,j
x
Typically gi,j will be abbreviated by gij if no confusion is likely to arise.
Example 3.2. Let M = RN and let x = (x1 , x2 , . . . , xN ) denote the standard
chart on M, i.e. x(m) = m for all m ∈ M. The standard Riemannian metric on
RN is determined by
XN N
X
ds2 = (dxi )2 = dxi · dxi ,
i=1 i=1
and so g x is the identity matrix here. The general Riemannian metric on RN is
PN
determined by ds2 = i,j=1 gij dx dx , where g = (gij ) is a smooth gl(N, R) –
i j
valued function on RN such that g(m) is positive definite matrix for all m ∈ RN .
Let M be an imbedded submanifold of a finite dimensional inner product space
(E, h·, ·i). The manifold M inherits a metric from E determined by
ds2 (vm ) = hv, vi ∀ vm ∈ T M.
It is a well known deep fact that all finite dimensional Riemannian manifolds may
be constructed in this way, see Nash [141] and Moser [136, 137, 138]. To simplify the
exposition, in the sequel we will usually assume that (E, h·, ·i) is an inner product
space, M d ⊂ E is an imbedded submanifold, and the Riemannian metric on M is
determined in this way, i.e.
hvm , wm i = hv, wiRN , ∀ vm , wm ∈ Tm M and m ∈ M.
16 BRUCE K. DRIVER
x
In this setting the components gi,j of the metric ds2 relative to a chart x may be
computed as gi,j (m) = (φ;i (x(m)), φ;j (x(m))), where {ei }di=1 is the standard basis
x
for Rd ,
d
φ := x−1 and φ;i (a) := |0 φ(a + tei ).
dt
Example 3.3. Let M = G := SL(n, R) and Ag ∈ Tg M.
(1) Then
(3.4) ds2 (Ag ) := tr(A∗ A)
defines a Riemannian metric on G. This metric is the inherited metric from
the inner product space E = gl(n, R) with inner product hA, Bi := tr(A∗ B).
(2) A more “natural” choice of a metric on G is
(3.5) ds2 (Ag ) := tr((g −1 A)∗ g −1 A).
This metric is invariant under left translations, i.e. ds2 (Lk∗ Ag ) = ds2 (Ag ),
for all k ∈ G and Ag ∈ T G. According to the imbedding theorem of Nash
and Moser, it would be possible to find another imbedding of G into a
Euclidean space, E, so that the metric in Eq. (3.5) is inherited from an
inner product on E.
Example 3.4. Let M = R3 be equipped with the standard Riemannian metric
and (r, ϕ, θ) be spherical coordinates on M , see Figure 7. Here r, ϕ, and θ are
R The problem with this notion of integration is that (as the notation indicates)
M
f dx depends on the choice of chart x. To remedy this, consider a small cube
C(δ) of side δ contained in R(x), see Figure 8. We wish to estimate “the volume”
of φ(C(δ)) where φ := x−1 : R(x) → D(x). Heuristically, we expect the volume of
φ(C(δ)) to be approximately equal to the volume of the parallelepiped, C̃(δ), in
the tangent space Tm M determined by
( d )
X
(3.10) C̃(δ) := si δ · φ;i (x (m))|0 ≤ si ≤ 1, for i = 1, 2, . . . , d ,
i=1
where we are using the notation proceeding Example 3.3, see Figure 8. Since Tm M
is an inner product space, the volume of C̃(δ) is well defined. Forexamplechoose
an isometry θ : Tm M → Rd and define the volume of C̃(δ) to be m θ(C̃(δ)) where
m is Lebesgue measure on Rd . The next elementary lemma will be used to give a
formula for the volume of C̃ (δ) .
dim V
Lemma 3.7. If V is a finite dimensional inner product space, {vi }i=1 is any
basis for V and A : V → V is a linear transformation, then
det [hAvi , vj i]
(3.11) det (A) = ,
det [hvi , vj i]
18 BRUCE K. DRIVER
where det [hAvi , vj i] is the determinant of the matrix with i-j th – entry being
hAvi , vj i. Moreover if
( d )
X
C̃(δ) := δsi · vi : 0 ≤ si ≤ 1, for i = 1, 2, . . . , d
i=1
Using
p the second assertion in Lemma 3.7, the volume of C̃(δ) in Eq. (3.10)
is δ d det g x (m), where gij
x
(m) = hφ;i (x(m)), φ;j (x(m))im . Because of the above
CURVED WIENER SPACE ANALYSIS 19
and hence we have proved the existence of λM . The uniqueness assertion is easy
and will be left to the reader.
20 BRUCE K. DRIVER
Example 3.11. Let M = R3 with the standard Riemannian metric, and let x
denote the standard coordinates on M determined by x(m) = m for all m ∈ M.
Then λR3 is Lebesgue measure which in spherical coordinates may be written as
dλR3 = r2 sin ϕdrdϕdθ
p
because g (r,ϕ,θ) = r2 sin ϕ by Eq. (3.7). Similarly using Eq. (3.9),
dλM = ρ2 sin ϕdϕdθ
when M ⊂ R3 is the sphere of radius ρ centered at 0 ∈ R3 .
Exercise 3.12. Compute the “volume element,” dλR3 , for R3 in cylindrical coor-
dinates.
Theorem 3.13 (Change of Variables Formula). Let (M, h·, ·iM ) and (N, h·, ·iN )
be two Riemannian manifolds, ψ : M → N be a diffeomorphism and ρ ∈
C ∞ (M, (0, ∞)) be determined by the equation
p
tr ψ
ρ (m) = det [ψ∗m ∗m ] for all m ∈ M,
tr
where ψ∗m denotes the adjoint of ψ∗m relative to Riemannian inner products on
Tm M and Tψ(m) N. If f : N → R+ is a positive Borel measurable function, then
Z Z
f dλN = ρ · (f ◦ ψ) dλM .
N M
This implies
Z Z q
f dλN = f ◦ (ψ ◦ φ) (t) det [h∂i (ψ ◦ φ) (t) , ∂j (ψ ◦ φ) (t)iN ]dt
N R(x)
Z q
= (f ◦ ψ) ◦ φ(t) · ρ (φ (t)) det [h∂i φ (t) , ∂j φ (t)iM ]dt
R(x)
√
Z Z
= (f ◦ ψ) · ρ · g x dx = ρ · f ◦ ψ dλM .
D(x) M
CURVED WIENER SPACE ANALYSIS 21
Example 3.14. Let M = SL(n, R) as in Example 3.3 and let h·, ·iM be the metric
given by Eq. (3.5). Because Lg : M → M is an isometry, Theorem 3.13 implies
Z Z
f (gx) dλG (x) = f (x) dλG (x) for all g ∈ G.
SL(n,R) SL(n,R)
for all g ∈ G.
3.3. Gradients, Divergence, and Laplacians. In the sequel, let M be a
Riemannian manifold, x be a chart on M, gij := h∂/∂xi , ∂/∂xj i, and ds2 =
Pd i j
i,j=1 gij dx dx .
Definition 3.15. Let g ij denote the i-j th – matrix element for the inverse matrix
to the matrix, (gij ).
Given f ∈ C ∞ (M ) and m ∈ M, dfm := df |Tm M is a linear functional on Tm M.
Hence there is a unique vector vm ∈ Tm M such that dfm = hvm , ·im .
Definition 3.16. The vector vm above is called the gradient of f at m and will
~ (m) .
be denoted by either grad f (m) or ∇f
Exercise 3.17. If x is a chart on M and m ∈ D(x) then
d
~ (m) = gradf (m) =
X ∂f (m) ∂
(3.18) ∇f g ij (m) |m ,
i,j=1
∂xi ∂xj
x −1 ~ is a
where as usual, gij = gij and g ij = (gij ) . Notice from Eq. (3.18) that ∇f
smooth vector field on M.
Exercise 3.18. Suppose M ⊂ RN is an imbedded submanifold with the induced
Riemannian structure. Let F : RN → R be a smooth function and set f := F |M .
~ (m))m , where ∇F
Then grad f (m) = (P (m)∇F ~ (m) denotes the usual gradient on
R , and P (m) denotes orthogonal projection of RN onto τm M.
N
d d √
√ ∂( g Y i )
Z Z X Z X
Y f dλM = Y i ∂f /∂xi · gdx = − f dx
M M i=1 M i=1 ∂xi
d √
1 ∂( gY i )
Z X
=− f √ dλM ,
M i=1
g ∂xi
where the second equality follows by an integration by parts. This shows that if
div Y exists it must be given on D(x) by Eq. (3.20). This proves the uniqueness
assertion. Using what we have already proved, it is easy to conclude that the
formula for div Y is chart independent. Hence we may define smooth function
div Y on M using Eq. (3.20) in each coordinate chart x on M. It is then possible to
show (again using a smooth partition of unity argument) that this function satisfies
Eq. (3.19).
Remark 3.20. We may write Eq. (3.19) as
Z Z
(3.21) hY, grad f i dλM = − divY · f dλM , ∀ f ∈ Cc∞ (M ),
M M
so that div is the negative of the formal adjoint of grad .
Exercise 3.21 (Product Rule). If f ∈ C ∞ (M ) and Y ∈ Γ (T M ) then
~ · (f Y ) = h∇f,
∇ ~ Yi+f ∇ ~ · Y.
Lemma 3.22 (Integration by Parts). Suppose that Y ∈ Γ(T M ), f ∈ Cc∞ (M ), and
h ∈ C ∞ (M ), then
Z Z
Y f · h dλM = f {−Y h − h · divY } dλM .
M M
Proof. By the definition of div Y and the product rule,
Z Z Z
f h divY dλM = − Y (f h) dλM = − {hY f + f Y h} dλM .
M M M
1
divY = {∂r (r2 sin ϕYr ) + ∂ϕ (r2 sin ϕYϕ ) + r2 sin ϕ∂θ Yθ }
r2
sin ϕ
1 1
= 2 ∂r (r2 Yr ) + ∂ϕ (sin ϕYϕ ) + ∂θ Yθ ,
r sin ϕ
and
1 1 1
∆f = 2
∂r (r2 ∂r f ) + 2 ∂ϕ (sin ϕ∂ϕ f ) + 2 2 ∂θ2 f.
r r sin ϕ r sin ϕ
Example 3.27. Let M = G = O (n) with Riemannian metric determined by Eq.
(3.5) and for A ∈ g := Te G let à ∈ Γ (T G) be the left invariant vector field,
d
à (x) := Lx∗ A = |0 xetA
dt
as was done for SL(n, R) in Example 2.34. Using the invariance of dλG under right
translations established in Example 3.14, we find for f, h ∈ C 1 (G) that
d
Z Z
|0 f xetA · h (x) dλG (x)
Ãf (x) · h (x) dλG (x) =
G G dt
d
Z
f xetA · h (x) dλG (x)
= |0
dt
ZG
d
f (x) · h xe−tA dλG (x)
= |0
dt G
d
Z
f (x) · |0 h xe−tA dλG (x)
=
G dt
Z
=− f (x) · Ãh (x) dλG (x) .
G
24 BRUCE K. DRIVER
Taking h ≡ 1 implies
Z Z D E
0= Ãf (x) dλG (x) = ~ (x) dλG (x)
à (x) , ∇f
G G
Z
=− ∇~ · Ã (x) · f (x) dλG (x)
G
~ · Ã = 0,
and, by the product rule and ∇
h i X D E
~ · ∇f
~ = ~ · Ãf à = ~ Ãf, à =
X X
∆f = ∇ ∇ ∇ Ã2 f.
A∈S0 A∈S0 A∈S0
Proposition 3.32 (Properties of ∇/ds). Let W (s) = (σ(s), w(s)) and V (s) =
(σ(s), v(s)) be two smooth vector fields along a path σ in M. Then:
(1) ∇W (s)/ds may be computed as:
∇W (s) d
(3.25) := (σ(s), w(s) + (dQ(σ ′ (s)))w(s)).
ds ds
(2) ∇ is metric compatible, i.e.
d ∇W (s) ∇V (s)
(3.26) hW (s), V (s)i = h , V (s)i + hW (s), i.
ds ds ds
Now suppose that (s, t) → σ(s, t) is a smooth function into M, W (s, t) =
d
(σ(s, t), w(s, t)) is a smooth function into T M, σ ′ (s, t) := (σ(s, t), ds σ(s, t))
d
and σ̇(s, t) = (σ(s, t), dt σ(s, t)). (Notice by assumption that w(s, t) ∈
Tσ(s,t) M for all (s, t).)
(3) ∇ has zero torsion, i.e.
∇σ ′ ∇σ̇
(3.27) = .
dt ds
(4) If R is the curvature tensor of ∇ defined by
(3.28) R(um , vm )wm = (m, [dQ(um ), dQ(vm )]w),
then
∇ ∇
∇∇ ∇∇
(3.29) , W := ( − )W = R(σ̇, σ ′ )W.
dt ds dt ds ds dt
26 BRUCE K. DRIVER
X
hRic u, zi = hR(u, a)a, zi
a∈S
X
= [hdQ(a)a, dQ(u)zi − hdQ(u)a, dQ(a)zi]
a∈S
X
= [ha, dQ(a)dQ(u)zi − hdQ(u)a, dQ(z)ai]
a∈S
X
= [ha, dQ(dQ(u)z)ai − hdQ(z)dQ(u)a, ai]
a∈S
= tr(dQ(dQ(u)z) − dQ(z)dQ(u))
which proves Eq. (3.32). The assertion that Ricm : Tm M → Tm M is a symmetric
operator follows easily from this formula and item 3.
Notation 3.37. To each v ∈ RN , let ∂v denote the vector field on RN defined by
d
∂v (at x) = vx = |0 (x + tv).
dt
So if F ∈ C ∞ (RN ), then
d
(∂v F )(x) := |0 F (x + tv) = F ′ (x) v
dt
and
(∂v ∂w F ) (x) = F ′′ (x) (v, w) ,
see Notation 2.1.
Notice that if w : RN → RN is a function and v ∈ RN , then
(∂v ∂w F ) (x) = ∂v [F ′ (·) w (·)] (x) = F ′ (x) ∂v w (x) + F ′′ (x) (v, w (x)) .
The following variant of item 4. of Proposition 3.36 will be useful in proving the
key Bochner-Weitenböck identity in Theorem 3.49 below.
Proposition 3.38. Suppose that Z ∈ Γ (T M ) , v, w ∈ Tm M and let X, Y ∈ Γ (T M )
such that X (m) = v and Y (m) = w. Then
(1) ∇2v⊗w Z defined by
(3.35) ∇2v⊗w Z := (∇X ∇Y Z − ∇∇X Y Z) (m)
is well defined, independent of the possible choices for X and Y.
(2) If Z(m) = (m, z(m)) with z : RN → RN a smooth function such z (m) ∈
τm M for all m ∈ M, then
(3.36)
∇2v⊗w Z = dQ (v) dQ (w) z (m) + P (m) z ′′ (m) (v, w) − P (m) z ′ (m) [dQ (v) w] .
(3) The curvature tensor R (v, w) may be computed as
(3.37) ∇2v⊗w Z − ∇2w⊗v Z = R (v, w) Z (m) .
(4) If V is a smooth vector field along a path σ (s) in M, then the following
product rule holds,
∇
(3.38) ∇V (s) Z = ∇ ∇ V (s) Z + ∇2σ′ (s)⊗V (s) Z.
ds ds
30 BRUCE K. DRIVER
Proof. We will prove items 1. and 2. by showing the right sides of Eq. (3.35)
and Eq. (3.36) are equal. To do this write X(m) = (m, x(m)), Y (m) = (m, y(m)),
and Z(m) = (m, z(m)) where x, y, z : RN → RN are smooth functions such that
x(m), y(m), and z(m) are in τm M for all m ∈ M. Then, suppressing m from the
notation,
∇X ∇Y Z − ∇∇X Y Z = P ∂x [P ∂y z] − P ∂P ∂x y z
= P (∂x P ) ∂y z + P ∂x ∂y z − P ∂P ∂x y z
= P (∂x P ) ∂y z + P z ′′ (x, y) + P z ′ [∂x y − P ∂x y]
= (∂x P ) Q∂y z + P z ′′ (x, y) + P z ′ [Q∂x y] .
d
∇ X
Vi′ (s) Ei (σ (s)) + Vi (s) ∇σ′ (s) Ei
(3.40) V (s) =
ds i=1
and
d
!
∇ ∇ X
∇V (s) Z = Vi (s) (∇Ei Z) (σ (s))
ds ds i=1
d
X d
X
= Vi′ (s) (∇Ei Z) (σ (s)) + Vi (s) ∇σ′ (s) (∇Ei Z) .
i=1 i=1
and using this in the previous equation along with Eq. (3.40) shows
d
∇ X
Vi (s) ∇2σ′ (s)⊗Ei (σ(s)) Z
∇V (s) Z = ∇P d {V ′ (s)Ei (σ(s))+Vi (s)∇ ′ Ei } Z +
ds i=1 i σ (s)
i=1
2
= ∇ ∇ V (s) Z + ∇σ′ (s)⊗V (s) Z.
ds
CURVED WIENER SPACE ANALYSIS 31
For the second proof, write V (s) = (σ (s) , v (s)) = v (s)σ(s) and p (s) :=
P (σ (s)) , then
∇ d
(∇V Z) − ∇ ∇ V Z = p (pz ′ (v)) − pz ′ (pv ′ )
ds ds ds
= p [p′ z ′ (v) + pz ′′ (σ ′ , v) + pz ′ (v ′ )] − pz ′ (pv ′ )
= pp′ z ′ (v) + pz ′′ (σ ′ , v) + pz ′ (qv ′ )
= p′ qz ′ (v) + pz ′′ (σ ′ , v) − pz ′ (q ′ v)
= ∇2σ′ (s)⊗V (s) Z
wherein the last equation we have made use of Eq. (3.39).
3.5. Formulas for the Divergence and the Laplacian.
Theorem 3.39. Let Y be a vector field on M, then
(3.41) div Y = tr(∇Y ).
(Note: (vm → ∇vm Y ) ∈ End(Tm M ) for each m ∈ M, so it makes sense to take the
trace.) Consequently, if f is a smooth function on M, then
(3.42) ∆f = tr(∇ grad f ).
Proof. Let x be a chart on M , ∂i := ∂/∂xi , ∇i := ∇∂i , and Y i := dxi (Y ). Then
by the product rule and the fact that ∇ is Torsion free (item 2. of the Proposition
3.36),
Xd d
X
∇i Y = ∇i (Y j ∂j ) = (∂i Y j ∂j + Y j ∇i ∂j ),
j=1 j=1
and ∇i ∂j = ∇j ∂i . Hence,
d
X d
X d
X
i i
tr(∇Y ) = dx (∇i Y ) = ∂i Y + dxi (Y j ∇i ∂j )
i=1 i=1 i,j=1
d
X d
X
= ∂i Y i + dxi (Y j ∇j ∂i ).
i=1 i,j=1
Therefore, according to Eq. (3.20), to finish the proof it suffices to show that
d
X √
dxi (∇j ∂i ) = ∂j log g.
i=1
From Lemma 2.7,
d
√ 1 1 1 X kl
∂j log g = ∂j log(det g) = tr(g −1 ∂j g) = g ∂j gkl ,
2 2 2
k,l=1
Before continuing, let us record the following useful corollary of the previous
proof.
CURVED WIENER SPACE ANALYSIS 33
Proof. Let fi (m) := θ(m)P (m)ei and gi (m) = xi (m) = hm, ei iRN where
N
{ei }i=1 is the standard basis for RN and P (m) is orthogonal projection of RN onto
τm M for each m ∈ M.
Definition 3.43. For f ∈ C ∞ (M ) and vm , wm in Tm M , let
∇df (vm , wm ) := (∇vm df )(wm ),
so that
∇df : ∪m∈M (Tm M × Tm M ) → R.
We call ∇df the Hessian of f.
Lemma 3.44. Let f ∈ C ∞ (M ), F ∈ C ∞ (RN ) such that f = F |M , X, Y ∈ Γ(T M )
and vm , wm ∈ Tm M. Then:
(1) ∇df (X, Y ) = XY f − df (∇X Y ).
(2) ∇df (vm , wm ) = F ′′ (m)(v, w) − F ′ (m)dQ(vm )w.
(3) ∇df (vm , wm ) = ∇df (wm , vm ) – another manifestation of zero torsion.
Proof. Using the product rule (see Eq. (3.44)):
XY f = X(df (Y )) = (∇X df )(Y ) + df (∇X Y ),
and hence
∇df (X, Y ) = (∇X df )(Y ) = XY f − df (∇X Y ).
This proves item 1. From this last equation and Proposition 3.36 (∇ has zero
torsion), it follows that
∇df (X, Y ) − ∇df (Y, X) = [X, Y ]f − df (∇X Y − ∇Y X) = 0.
This proves the third item upon choosing X and Y such that X(m) = vm and
Y (m) = wm . Item 2 follows easily from Lemma 3.41 applied with θ := F ′ .
Definition 3.45. Given a point m ∈ M, a local orthonormal frame {Ei }di=1 at
d
m is a collection of local vector fields defined near m such that {Ei (p)}i=1 is an
orthonormal basis for Tp M for all p near m.
Corollary 3.46. Suppose that F ∈ C ∞ (RN ), f := F |M , and m ∈ M. Let {ei }di=1
be an orthonormal basis for τm M and let {Ei }di=1 be an orthonormal frame near
m ∈ M. Then
Xd
(3.47) ∆f (m) = ∇df (Ei (m), Ei (m)),
i=1
d
X
(3.48) ∆f (m) = {Ei Ei f )(m) − df (∇Ei (m) Ei )},
i=1
and
d
X
(3.49) ∆f (m) = F ′′ (m)(ei , ei ) − F ′ (m)(dQ(Ei (m))ei )
i=1
34 BRUCE K. DRIVER
which proves Eq. (3.47). Equations (3.48) and (3.49) follows from Eq. (3.47) and
Lemma 3.44.
N
Notation 3.47. Let {ei }i=1 be the standard basis on RN and define Xi (m) :=
P (m) ei for all m ∈ M and i = 1, 2, . . . , N.
In the next proposition we will express the gradient, divergence and the Laplacian
in terms of the vector fields, {Xi }N.
i=1 . These formula will prove very useful when
we start discussing Brownian motion on M.
Proposition 3.48. Let f ∈ C ∞ (M ) and Y ∈ Γ (T M ) then
PN
(1) vm = i=1 hvm , Xi (m)iXi (m) for all vm ∈ Tm M.
~ = grad f = N Xi f · Xi
P
(2) ∇f i=1
~ · Y = div(Y ) = PN h∇Xi Y, Xi i
(3) ∇
PN i=1
(4) i=1 ∇Xi Xi = 0
PN
(5) ∆f = i=1 Xi2 f.
Proof. 1. The main point is to show
N
X d
X
(3.50) Xi (m) ⊗ Xi (m) = ui ⊗ ui
i=1 i=1
d
where {ui }i=1 is an orthonormal basis for Tm M. But this is easily proved since
N
X N
X
Xi (m) ⊗ Xi (m) = P (m) ei ⊗ P (m) ei
i=1 i=1
N
and the latter expression is independent of the choice of orthonormal basis {ei }i=1
for RN . Hence if we choose {ei }N
i=1 so that ei = ui for i = 1, . . . , d, then
N
X d
X
P (m) ei ⊗ P (m) ei = ui ⊗ ui
i=1 i=1
PN
as desired. Since i=1 hvm , Xi (m)iXi (m) is quadratic in Xi , it now follows that
N
X d
X
hvm , Xi (m)iXi (m) = hvm , ui iui = vm .
i=1 i=1
PN
3. Again i=1 h∇Xi Y, Xi i (m) is quadratic in Xi and so by Eq. (3.50) and
Theorem 3.39,
N
X d
X
h∇Xi Y, Xi i (m) = h∇ui Y, ui i (m) = div(Y ).
i=1 i=1
4. By definition of Xi and ∇ and using Lemma 3.30,
N
X N
X N
X
(3.51) (∇Xi Xi ) (m) = P (m) dP (Xi (m)) ei = dP (P (m) ei ) Q (m) ei .
i=1 i=1 i=1
N
The latter expression is independent of the choice of orthonormal basis {ei }i=1 for
N
RN . So again we may choose {ei }i=1 so that ei = ui for i = 1, . . . , d, in which case
P (m) ej = 0 for j > d and so each summand in the right member of Eq. (3.51) is
zero.
~ and the product rule
5. To compute ∆f, use items 2.– 4., the definition of ∇f
to find
N
~ · (∇f
~ )= ~ Xi i
X
∆f = ∇ h∇Xi ∇f,
i=1
N N N
~ Xi i − ~ ∇Xi Xi i =
X X X
= Xi h∇f, h∇f, Xi Xi f.
i=1 i=1 i=1
The following commutation formulas are at the heart of many of the results to
appear in the latter sections of these note.
Theorem 3.49 (The Bochner-Weitenböck Identity). Let f ∈ C ∞ (M ) and a, b, c ∈
Tm M, then
(3.52) ~ ci = h∇2 ∇f,
h∇2 ∇f, ~ bi
a⊗b a⊗c
and if S ⊂ Tm M is an orthonormal basis, then
~ = (grad ∆f ) (m) + Ric ∇f
~ (m) .
X
(3.53) ∇2a⊗a ∇f
a∈S
This result is the first indication that the Ricci tensor is going to play an im-
portant role in later developments. The proof will be given after the next technical
lemma which will be helpful in simplifying the proof of the theorem.
Lemma 3.50. Given m ∈ M and v ∈ Tm M there exists V ∈ Γ (T M ) such that
d
V (m) = v and ∇w V = 0 for all w ∈ Tm M. Moreover if {ei }i=1 is an orthonormal
d
basis for Tm M, there exists a local orthonormal frame {Ei }i=1 near m such that
∇w Ei = 0 for all w ∈ Tm M.
Proof. In the proof to follow it is assume that V, Q and P have all been extended
off M to smooth function on the ambient space. If V is to exist, we must have
0 = ∇w V = V ′ (m) w + ∂w Q (m) v,
i.e.
V ′ (m) w = −∂w Q (m) v for all w ∈ Tm M.
This helps to motivate defining V by
V (x) := P (x) (v − (∂x−m Q) (m) v) ∈ Tx M for all x ∈ M.
36 BRUCE K. DRIVER
as desired.
d
For the second assertion, choose a local frame {Vi }i=1 such that Vi (m) = ei
and ∇w Vi = 0 for all i and w ∈ Tm M. The desired frame {Ei }di=1 is now con-
d
structed by performing Gram-Schmidt orthogonalization on {Vi }i=1 . The resulting
d
orthonormal frame, {Ei }i=1 , still satisfies ∇w Ei = 0 for all w ∈ Tm M. For example,
E1 = hV1 , V1 i−1/2 V1 and since
it follows that
∇w E1 = w hV1 , V1 i−1/2 · V1 (m) + hV1 , V1 i−1/2 (m) ∇w V1 (m) = 0.
wherein the last equality we have used (∇A B) (m) = 0. Interchanging B and C in
this equation and subtracting then implies
~ ci − h∇2 ∇f,
(A [B, C] f ) (m) = h∇2a⊗b ∇f, ~ bi + Ah∇f,
~ ∇B C − ∇C Bi (m)
a⊗c
~ ci − h∇2a⊗c ∇f,
= h∇2a⊗b ∇f, ~ bi + Ah∇f,
~ [B, C]i (m)
~ ci − h∇2 ∇f,
= h∇2a⊗b ∇f, ~ bi + (A[B, C]f ) (m)
a⊗c
Since
d d d
~ ei i = ~ Ei i (m) = ~ Ei i (m)
X X X
h∇2c⊗ei ∇f, h∇C ∇Ei ∇f, Ch∇Ei ∇f,
i=1 i=1 i=1
~
= (C∆f ) (m) = h ∇∆f (m) , ci
and therefore,
∇ ∇
∇//s v Y = [h//s v, Xi (σ (s))i · (∇Xi Y ) (σ (s))]
ds ds
(3.60) = h//s v, Xi (σ (s))i · ∇σ′ (s) (∇Xi Y ) + h//s v, ∇σ′ (s) Xi i · (∇Xi Y ) (σ (s)) .
Now
∇σ′ (s) (∇Xi Y ) = ∇2σ′ (s)⊗Xi Y + ∇σ′ (s)Xi Y
and so again using Proposition 3.48,
(3.61)
h//s v, Xi (σ (s))i · ∇σ′ (s) (∇Xi Y ) = ∇2σ′ (s)⊗//s v Y + h//s v, Xi (σ (s))i · ∇σ′ (s)Xi Y.
Taking ∇/ds of Eq. (3.59) shows
0 = h//s v, ∇σ′ (s) Xi iXi (σ (s)) + h//s v, Xi (σ (s))i∇σ′ (s) Xi .
and so
(3.62) h//s v, Xi (σ (s))i · ∇σ′ (s)Xi Y = −h//s v, ∇σ′ (s) Xi i · (∇Xi Y ) (σ) (s) .
Assembling Eqs. (3.59), (3.61) and (3.62) proves Eq. (3.58).
d
Second proof. Let {Ei }i=1 be an orthonormal frame near σ (s) , then
∇ ∇
∇//s v Y = [h//s v, Ei (σ (s))i · (∇Ei Y ) (σ (s))]
ds ds
(3.63) = h//s v, ∇σ′ (s) Ei i · (∇Ei Y ) (σ (s)) + h//s v, Ei (σ (s))i · ∇σ′ (s) ∇Ei Y.
Working as in the first proof,
h//s v, Ei (σ (s))i · ∇σ′ (s) ∇Ei Y = h//s v, Ei (σ (s))i · ∇2σ′ (s)⊗Ei Y + ∇∇σ′ (s) Ei Y
= ∇2σ′ (s)⊗//s v Y + ∇h//s v,Ei (σ(s))i∇ Ei
Y
σ′ (s)
and using
∇
0= //s v = h//s v, ∇σ′ (s) Ei i · Ei (σ (s)) + h//s v, Ei (σ (s))i · ∇σ′ (s) Ei
ds
we learn
h//s v, Ei (σ (s))i · ∇σ′ (s) ∇Ei Y = ∇2σ′ (s)⊗//s v Y − h//s v, ∇σ′ (s) Ei i · (∇Ei Y ) (σ (s)) .
This equation combined with Eq. (3.63) again proves Eq. (3.58).
The remainder of this section discusses a covariant derivative on M × RN which
“extends” ∇ defined above. This will be needed in Section 5, where it will be
convenient to have a covariant derivative on the normal bundle:
N (M ) := ∪m∈M ({m} × τm M ⊥ ) ⊂ M × RN .
Analogous to the definition of ∇ on T M, it is reasonable to extend ∇ to the
normal bundle N (M ) by setting
∇V (s)
= (σ(s), Q(σ(s))v ′ (s)) = (σ(s), v ′ (s) + dP (σ ′ (s))v(s)),
ds
for all smooth paths s → V (s) = (σ(s), v(s)) in N (M ). Then this covariant deriva-
tive on the normal bundle satisfies analogous properties to ∇ on the tangent bundle
T M. The covariant derivatives on T M and N (M ) can be put together to make a
40 BRUCE K. DRIVER
Theorem 4.2 (Flow Theorem). Suppose that Xt is a smooth time dependent vector
field on M. Then for each m ∈ M, there exists a maximal open interval Jm ⊂ R
such that 0 ∈ Jm and t → TtX (m) exists for t ∈ Jm . Moreover the set D (X) :=
∪m (Jm × {m}) ⊂ R × M is open and the map (t, m) ∈ D (X) → TtX (m) ∈ M is a
smooth map.
Proof. Let Yt be a smooth extension of Xt to a vector field on E where E is the
Euclidean space in which M is imbedded. The stated results with X replaced by
Y follows from the standard theory of ordinary differential equations on Euclidean
spaces. Let TtY denote the flow of Y on E. We will construct T X by setting
42 BRUCE K. DRIVER
TtX (m) := TtY (m) for all m ∈ M and t ∈ Jm . In order for this to work we must
show that TtY (m) ∈ M whenever m ∈ M.
To verify this last assertion, let x be a chart on M such that m ∈ D (x) , then
σ (t) solves σ̇ (t) = Xt (σ (t)) with σ (0) = m iff
d
[x ◦ σ (t)] = dx (σ̇ (t)) = dx (Xt (σ (t))) = dx Xt ◦ x−1 (x ◦ σ (t))
dt
with x◦ σ (0) = m. Since this is a differential equation for x◦ σ (t) ∈ R (z) and R (z)
is an open subset Rd , the standard local existence theorem for ordinary differential
equations implies x ◦ σ (t) exists for small time. This then implies σ (t) ∈ M exists
for small t and satisfies
σ̇ (t) = Xt (σ (t)) = Yt (σ (t)) with σ (0) = m.
By uniqueness of solutions to ordinary differential equations, we must have
TtY (m) = σ (t) for small t and in particular TtY (m) ∈ M for small t. Let
τ := sup t ∈ Jm : TsY (m) ∈ M for 0 ≤ s ≤ t
then
d d
0 = id = [Tt ◦ St ] = Xt ◦ Tt ◦ St + Tt∗ Ṡt .
dt dt
So it follows that St solves
−1
Ṡt = −Tt∗ Xt ◦ Tt ◦ St = − AdT −1 Xt ◦ St
t
2Actually, for those in the know, any torsion zero covariant derivative could be used here.
44 BRUCE K. DRIVER
Remark 4.6. As a warm up for writing the stochastic version of Eq. (4.3) in Itô
∇
form let us pause to compute dt (∇Tt∗ v Y ) for Y ∈ Γ (T M ) . Using Eqs. (3.38),
(3.37) and (3.35) of Proposition 3.38,
∇
∇T v Y = ∇2Ṫt (m)⊗Tt∗ v Y + ∇ ∇ Tt∗ v Y = ∇2Xt (Tt (m))⊗Tt∗ v Y + ∇∇Tt∗ v Xt Y
dt t∗ dt
= ∇Tt∗ v⊗Xt (Tt (m)) Y + R∇ (Xt (Tt (m)) , Tt∗ v) Y (Tt (m)) + ∇∇Tt∗ v Xt Y
2
Theorem 4.7 (Differentiating TtX in X). Suppose (t, m) → Xt (m) and (t, m) →
Yt (m) are smooth time dependent vector fields on M and let
d
(4.5) ∂Y TtX := |0 T X+sY .
ds t
Then
Z t −1
Z t
(4.6) ∂Y TtX = Tt∗
X
TτX∗ Yτ ◦ TτX dτ = Tt∗
X
Ad−1
T X Yτ dτ.
τ
0 0
d d
X
Since dt , ds |0 = 0, the previous two displayed equations imply Tt∗ V̇t f =
Yt ◦ TtX f and because this holds for all f ∈ C ∞ (M ),
X
(4.8) Tt∗ V̇t = Yt ◦ TtX .
where
Adg A = Rg−1 ∗ Lg∗ A for all g ∈ G and A ∈ g.
The next theorem expresses [Xt , Y ] using the flow T X . The stochastic analog of
this theorem is a key ingredient in the “Malliavin calculus,” see Proposition 8.14
below.
Theorem 4.9. If Xt and TtX are as above and Y ∈ Γ (T M ) , then
d h X −1 i
X −1
Y ◦ TtX = Tt∗ [Xt , Y ] ◦ TtX
(4.9) Tt∗
dt
or equivalently put
d
(4.10) Ad−1X = Ad−1 L
TtX Xt
dt Tt
where LX Y := [X, Y ] .
46 BRUCE K. DRIVER
X −1
Y ◦ TtX which is equivalent to Tt∗
X
Vt = Y ◦ TtX , or
Proof. Let Vt := Tt∗
more explicitly to
Y f ◦ TtX = Y ◦ TtX f = Tt∗X
Vt f = Vt f ◦ TtX for all f ∈ C ∞ (M ).
4.3. Cartan’s Development Map. For this section assume that M is compact3
Riemannian manifold and let W ∞ (T0 M ) be the collection of piecewise smooth
paths, b : [0, 1] → To M such that b (0) = 0o ∈ To M and let Wo∞ (M ) be the
collection of piecewise smooth paths, σ : [0, 1] → M such that σ (0) = o ∈ M.
Theorem 4.10 (Development Map). To each b ∈ W ∞ (T0 M ) there is a unique
σ ∈ Wo∞ (M ) such that
(4.11) σ ′ (s) := (σ(s), dσ(s)/ds) = //s (σ)b′ (s) and σ(0) = o,
where //s (σ) denotes parallel translation along σ.
Proof. Suppose that σ is a solution to Eq. (4.11) and //s (σ)vo = (o, u(s)v),
where u(s) : τo M → RN . Then u satisfies the differential equation
(4.12) u′ (s) + dQ(σ ′ (s))u(s) = 0 with u(0) = u0 ,
where u0 v := v for all v ∈ τo M, see Remark 3.54. Hence Eq. (4.11) is equivalent
to the following pair of coupled ordinary differential equations:
(4.13) σ ′ (s) = u(s)b′ (s) with σ(0) = o,
and
(4.14) u′ (s) + dQ((σ(s), u(s)b′ (s))u(s) = 0 with u(0) = u0 .
Therefore the uniqueness assertion follows from standard uniqueness theorems for
ordinary differential equations. The slickest prove of existence to Eq. (4.11) is to
first introduce the orthogonal frame bundle, O (M ) , on M defined by O (M ) :=
∪m∈M Om (M ) where Om (M ) is the set of all isometries, u : To M → Tm M. It is then
possible to show that O (M ) is an imbedded submanifold in RN × Hom τo M, RN
and that coupled pair of ordinary differential equations (4.13) and (4.14) may be
viewed as a flow equation on O(M ). Hence the existence of solutions may be deduced
3It would actually be sufficient to assume that M is a “complete” Riemannian manifold for
this section.
CURVED WIENER SPACE ANALYSIS 47
from the Theorem 4.2, see, for example, [47] for details of this method. Here I will
sketch a proof which does not require us to develop the frame bundle formalism in
detail.
Looking at the proof of Lemma 2.30, Q has an extension to a neighborhood
in RN of m ∈ M in such a way that Q(x) is still an orthogonal projection onto
Nul(F ′ (x)), where F (x) = z> (x) is as in Lemma 2.30. Hence for small s, we may
define σ and u to be the unique solutions to Eq. (4.13) and Eq. (4.14) with values in
RN and Hom(τo M, RN ) respectively. The key point now is to show that σ(s) ∈ M
and that the range of u(s) is τσ(s) M.
Using the same proof as in Theorem 3.52, w(s) := Q(σ(s))u(s) satisfies,
w′ = dQ (σ ′ ) u + Q (σ) u′ = dQ (σ ′ ) u − Q (σ) dQ(σ ′ )u
= P (σ) dQ (σ ′ ) u = dQ (σ ′ ) Q (σ) u = dQ (σ ′ ) w,
where Lemma 3.30 was used in the last equality. Since w(0) = 0, it follows by
uniqueness of solutions to linear ordinary differential equations that w ≡ 0 and
hence
Ran [u(s)] ⊂ Nul [Q(σ(s))] = Nul [F ′ (σ(s))] .
Consequently
dF (σ(s))/ds = F ′ (σ(s))dσ(s)/ds = F ′ (σ(s))u(s)b′ (s) = 0
for small s and since F (σ(0)) = F (o) = 0, it follows that F (σ(s)) = 0, i.e. σ(s) ∈ M.
So we have shown that there is a solution (σ, u) to (4.13) and (4.14) for small
s such that σ stays in M and u(s) is parallel translation along s. By standard
ordinary differential equation methods, there is a maximal solution (σ, u) with these
properties. Notice that (σ, u) is a path in M × Iso(To M, RN ), where Iso(To M, RN )
is the set of isometries from To M to RN . Since M × Iso(To M, RN ) is a compact
space, (σ, u) can not explode. Therefore (σ, u) is defined on the same interval where
b is defined.
The geometric interpretation of Cartan’s map is to roll the manifold M along a
freshly painted curve b in To M to produce a curve σ on M, see Figure 11.
Notation 4.11. Let φ : W ∞ (T0 M ) → Wo∞ (M ) be the map b → σ, where σ is
the solution to (4.11). It is easy to construct the inverse map Ψ := φ−1 . Namely,
Ψ(σ) = b, where
Z s
Ψs (σ) = b(s) := //r (σ)−1 σ ′ (r)dr.
0
We now conclude this section by computing the differentials of Ψ and φ. For more
details on computations of this nature the reader is referred to [46, 47] and the
references therein.
Theorem 4.12 (Differential of Ψ). Let (t, s) → Σ(t, s) be a smooth map into M
such that Σ(t, ·) ∈ Wo∞ (M ) for all t. Let
H(s) := Σ̇(0, s) := (Σ(0, s), dΣ(t, s)/dt|t=0 ),
so that H is a vector-field along σ := Σ(0, ·). One should view H as an element of
the “tangent space” to Wo∞ (M ) at σ, see Figure 12. Let u(s) := //s (σ), h(s) :=
//s (σ)−1 H(s) b := Ψs (σ) and, for all a, c ∈ To M, let
(4.15) (Ru (a, c))(s) := u(s)−1 R(u(s)a, u(s)c)u(s).
48 BRUCE K. DRIVER
Then
Z Z
(4.16) dΨ(H) = dΨ(Σ(t, ·))/dt|t=0 = h + Ru (h, δb) δb,
0 0
′
R
where δb(s) is short hand notation for b (s)ds, and 0
f δb denotes the function
Rs
s → 0 f (r)b′ (r)dr when f is a path of matrices.
d d
Proof. To simplify notation let “ · ”= dt |0 , “ ′ ”= ds , B(t, s) := Ψ(Σ(t, ·))(s),
U (t, s) := //s (Σ(t, ·)), u(s) := //s (σ) = U (0, s) and
ḃ(s) := (dΨ(H))(s) := dB(t, s)/dt|t=0 .
I will also suppress (t, s) from the notation when possible. With this notation
(4.17) Σ′ = U B ′ , Σ̇ = H = uh,
CURVED WIENER SPACE ANALYSIS 49
and
∇U
(4.18) = 0.
ds
∇U ∇U
In Eq. (4.18), ds : To M → TΣ M is defined by ds = P (Σ) U ′ or equivalently by
∇U ∇ (U a)
a := for all a ∈ To M.
ds ds
Taking ∇/dt of (4.17) at t = 0 gives, with the aid of Proposition 3.32,
∇U
|t=0 b′ + uḃ′ = ∇Σ′ /dt|t=0 = ∇Σ̇/ds = uh′ .
dt
Therefore,
(4.19) ḃ′ = h′ + Ab′ ,
where A := −U −1 ∇U
dt |t=0 , i.e.
∇U
(0, ·) = −uA.
dt
Taking ∇/ds of this last equation and using ∇u/ds = 0 along with Proposition
3.32 gives
∇ ∇
′ ∇ ∇
= R(σ ′ , H)u
−uA = U = , U
ds dt t=0 ds dt t=0
and hence A′ = Ru (h, b′ ). By integrating this identity using A(0) = 0
(∇U (t, 0)/dt = 0 since U (t, 0) := //0 (Σ(t, ·)) = I is independent of t) shows
Z
(4.20) A = Ru (h, δb)
0
The theorem now follows by integrating (4.19) relative to s making use of Eq. (4.20)
and the fact that ḃ(0) = 0.
Theorem 4.13 (Differential of φ). Let b, k ∈ W ∞ (T0 M ) and (t, s) → B(t, s)
be a smooth map into To M such that B(t, ·) ∈ W ∞ (T0 M ) , B(0, s) = b(s), and
Ḃ(0, s) = k(s). (For example take B(t, s) = b(s) + tk(s).) Then
d
φ∗ (kb ) := |0 φ(B(t, ·)) = //· (σ)h,
dt
where σ := φ(b) and h is the first component in the solution (h, A) to the pair of
coupled differential equations:
(4.21) k ′ = h′ + Ab′ , with h(0) = 0
and
(4.22) A′ = Ru (h, b′ ) with A(0) = 0.
Proof. This theorem has an analogous proof to that of Theorem 4.12. We can
also deduce the result from Theorem 4.12 by defining Σ by Σ(t, s) := φs (B(t, ·)).
We now assume the same notation used in Theorem 4.12 and its proof. Then
B(t, ·) = Ψ(Σ(t, ·)) and hence by Theorem 4.13
d
Z Z
k = |0 Ψ(Σ(t, ·)) = dΨ(H) = h + ( Ru (h, δb))δb.
dt 0 0
R
Therefore, defining A := 0 Ru (h, δb) and differentiating this last equation relative
to s, it follows that A solves (4.22) and that h solves (4.21).
50 BRUCE K. DRIVER
The following theorem is a mild extension of Theorem 4.12 to include the possi-
/ Wo∞ (M ) when t 6= 0, i.e. the base point may change.
bility that Σ(t, ·) ∈
Theorem 4.14. Let (t, s) → Σ(t, s) be a smooth map into M such that σ :=
Σ(0, ·) ∈ Wo∞ (M ). Define H(s) := dΣ(t, s)/dt|t=0 , σ := Σ(0, ·), and h(s) :=
//s (σ)−1 H(s). (Note: H(0) and h(0) are no longer necessarily equal to zero.) Let
U (t, s) := //s (Σ(t, ·))//t (Σ(·, 0)) : To M → TΣ(t,s) M,
Rs
so that ∇U (t, 0)/dt = 0 and ∇U (t, s)/ds ≡ 0. Set B(t, s) := 0 U (t, r)−1 Σ′ (t, r)dr,
then
Z s Z
d
(4.23) ḃ(s) := |0 B(t, s) = hs + Ru (h, δb) δb,
dt 0 0
is the mutual variation of Z and Σ. (All limits may be taken in the sense of
uniform convergence on compact subsets of R+ in probability.)
5.1. Stochastic Differential Equations on Manifolds.
n
Notation 5.4. Suppose that {Xi }i=0 ⊂ Γ (T M ) are vector fields on M. For a ∈ Rn
let
Xn
Xa (m) := X (m) a := ai Xi (m)
i=1
i.e. if
n Z
X s Z s
f (Σs ) = f (Σ0 ) + (Xi f ) (Σr ) δβri + X0 f (Σr ) dr.
i=1 0 0
Lemma 5.6 (Itô Form of Eq. P(5.1)). Suppose that β = B is an Rn – valued Brow-
1 n 2
nian motion and let L := 2 i=1 Xi + X0 . Then an M – valued semi-martingale
Σs solves Eq. (5.1) iff
Xn Z s Z s
(5.2) f (Σs ) = f (Σ0 ) + (Xi f ) (Σr ) dBri + Lf (Σr ) dr
i=1 0 0
for all f ∈ C ∞ (M ).
52 BRUCE K. DRIVER
Xn Z s Z s
i
= f (Σ0 ) + (Xi f ) (Σr ) δBr + X0 f (Σr ) dr.
i=1 0 0
f : Rn × RN → Hom(Rn , RN ) and f0 : Rn × RN → RN
(5.3) ξπ′ (s) = f (Bπ (s), ξπ (s))Bπ′ (s) + f0 (Bπ (s), ξπ (s)), ξπ (0) = a
Then, for any γ ∈ (0, 21 ) and p ∈ [1, ∞), there is a constant C(p, γ) < ∞ such that
(5.5) lim E sup |ξπ (s) − ξs | ≤ C(p, γ)|π|γp .
p
|π|→0 s≤T
This theorem is a special case of Theorem 5.7.3 and Example 5.7.4 in Kunita
[115]. Theorems of this type have a long history starting with Wong and Zakai
[178, 179]. The reader may also find this and related results in the following partial
list of references: [7, 10, 11, 20, 22, 44, 67, 93, 102, 106, 107, 117, 116, 125, 128, 131,
133, 134, 139, 140, 149, 164, 172, 165, 173, 175]. Also see [8, 53] and the references
therein for more of the geometry associated to the Wong and Zakai approximation
scheme.
Proof. Existence. If for the moment we assumed that the Brownian motion
Bs were differentiable in s, Eq. (5.1) could be written as
where
n
X ′
Xs (m) := Xi (m) B i (s) + X0 (m)
i=1
and the existence of Σs could be deduced from Theorem 4.2. We will make this
rigorous with an application of Theorem 5.8.
54 BRUCE K. DRIVER
n
Let {Yi }i=0 be smooth vector fields on E with compact support such that Yi = Xi
on M for each i and let Bπ (s) be as in Notation 5.7 and define
n
X ′
Xsπ (m) := Xi (m) Bπi (s) + X0 (m) and
i=1
n
X ′
Ysπ (m) := Yi (m) Bπi (s) + Y0 (m) .
i=1
implies
n
X
(5.7) d [F (Σs )] = Yi F (Σs ) δBsi + Y0 F (Σs ) ds.
i=1
i.e. Z s
Σs − Σ0 = P (Σr )δΣr .
0
Proof. We will first assume that M is the level set of a function F as in Theorem
2.5. Then we may assume that
Q(x) = φ(x)F ′ (x)∗ (F ′ (x)F ′ (x)∗ )−1 F ′ (x),
where φ is smooth function on RN such that φ := 1 in a neighborhood of M and the
support of φ is contained in the set: {x ∈ RN |F ′ (x) is surjective}. By Itô’s lemma
0 = d0 = d(F (Σ)) = F ′ (Σ)δΣ.
The lemma follows in this special case by multiplying the above equation through
by φ(Σ)F ′ (Σ)∗ (F ′ (Σ)F ′ (Σ)∗ )−1 , see the proof of Lemma 2.30.
For the general case, choose two open covers {Vi } and {Ui } of M such that each
V̄i is compactly contained in Ui , there is a smooth function Fi ∈ Cc∞ (Ui → RN −d )
such that Vi ∩ M = Vi ∩ {Fi−1 ({0})} and Fi has a surjective differential onPVi ∩ M.
Choose φi ∈ Cc∞ (RN ) such that the support of φi is contained in Vi and φi = 1
on M, with the sum being locally finite. (For the existence of such covers and
functions, see the discussion of partitions of unity in any reasonable book about
manifolds.) Notice that φi · Fi ≡ 0 and that Fi · φ′i ≡ 0 on M so that
0 = d{φi (Σ)Fi (Σ)} = (φ′i (Σ)δΣ)Fi (Σ) + φi (Σ)Fi′ (Σ)δΣ
= φi (Σ)Fi′ (Σ)δΣ.
Multiplying this equation by Ψi (Σ)Fi′ (Σ)∗ (Fi′ (Σ)Fi′ (Σ)∗ )−1 , where each Ψi is a
smooth function on RN such that Ψi ≡ 1 on the support of φi and the support of
Ψi is contained in the set where Fi′ is surjective, we learn that
(5.8) 0 = φi (Σ)Fi′ (Σ)∗ (Fi′ (Σ)Fi′ (Σ)∗ )−1 Fi′ (Σ)δΣ = φi (Σ)Q(Σ)δΣ
for all i. By a stopping time argument we may assume that Σ never leaves a compact
P and therefore we may choose a finite subset I of the indices {i} such that
set,
i∈I φi (Σ)Q(Σ) = Q(Σ). Hence summing over i ∈ I in equation (5.8) shows that
0 = Q(Σ)δΣ. Since Q + P = I, it follows that
dΣ = IδΣ = [Q(Σ) + P (Σ)] δΣ = P (Σ) δΣ.
The following notation will be needed to define line integrals along a semi-
martingale Σ.
Notation 5.12. Let P (m) be orthogonal projection of RN onto τm M as above.
(1) Given a one-form α on M let α̃ : M → (RN )∗ be defined by
(5.9) α̃(m)v := α((P (m)v)m )
for all m ∈ M and v ∈ RN .
(2) Let Γ(T ∗ M ⊗ T ∗ M ) denote the set of functions ρ : ∪m∈M Tm M ⊗ Tm M →
R such that ρm := ρ|Tm M⊗Tm M is linear, and m → ρ(X(m) ⊗ Y (m))
is a smooth function on M for all smooth vector-fields X, Y ∈ Γ(T M ).
(Riemannian metrics and Hessians of smooth functions are examples of
elements of Γ(T ∗ M ⊗ T ∗ M ).)
56 BRUCE K. DRIVER
where the stochastic integrals on the right hand sides of Eqs. (5.11) and (5.12) are
¯ := P (Σ)dΣ. We also
Fisk-Stratonovich and Itô integrals respectively. Formally, dΣ
define quadratic integral:
Z · Z · N Z ·
X
(5.13) ρ(dΣ ⊗ dΣ) := ρ̃(Σ)(dΣ ⊗ dΣ) := ρ̃(Σ)(ei ⊗ ej )d[Σi , Σj ],
0 0 i,j=1 0
where {ei }Ni=1 is an orthonormal basis for R , Σ := hei , Σi, and d[Σ , Σ ] is the
N i i j
i j
differential of the mutual quadratic variation of Σ and Σ .
So as not to confuse [Σi , Σj ] with a commutator or a Lie bracket, in the sequel
we will write dΣi dΣj for d[Σi , Σj ].
Remark 5.14. The above definitions may be generalized as follows. Suppose that
α is now a T ∗ M – valued semi-martingale and Σ is the M valued semi-martingale
such that αs ∈ TΣ∗s M for all s. Then we may define
α̃s v := αs ((P (Σs )v)Σs ),
Z · Z ·
(5.14) α(δΣ) := α̃δΣ,
0 0
and
Z · Z ·
(5.15) ¯ :=
α(dΣ) α̃dΣ.
0 0
where
ρ̃s (v ⊗ w) := ρs ((P (Σs )v)Σs ⊗ (P (Σs )v)Σs )
and
N
X
(5.17) dΣ ⊗ dΣ = ei ⊗ ej dΣi dΣj
i,j=1
as in Eq. (5.13).
CURVED WIENER SPACE ANALYSIS 57
Since
ρ̃(m) = f (m) · (H ′ (m)P (m)) ⊗ (G′ (m)P (m)),
it follows from Eq. (5.13) and the two above displayed equations that
Z · Z ·X
f (Σ)d[h(Σ), g(Σ)] := f (Σ)(H ′ (Σ)P (Σ)ek )(G′ (Σ)P (Σ)el )dΣk dΣl
0 0 k,l
Z · Z ·
= ρ̃(Σ)(dΣ ⊗ dΣ) =: ρ(dΣ ⊗ dΣ).
0 0
which has a unique solution by Theorem 5.10. Using Lemma 5.6, this equation may
be rewritten in Itô form as
N N
X 1X 2
d [f (Σ)] = Xi f (Σ)dB i + X f (Σ) ds for all f ∈ C ∞ (M ).
i=1
2 i=1 i
PN
This completes the proof since Xi2 = ∆ by Proposition 3.48.
i=1
Pd
Lemma 5.21 (Lévy’s Criteria). For each m ∈ M, let I(m) := i=1 Ei ⊗ Ei , where
{Ei }di=1 is an orthonormal basis for Tm M. An M – valued semi-martingale, Σ, is
a Brownian motion iff Σ is a martingale and
(5.24) dΣ ⊗ dΣ = I(Σ)dλ.
More precisely, this last condition is to be interpreted as:
Z · Z ·
(5.25) ρ(dΣ ⊗ dΣ) = ρ(I(Σ))dλ ∀ ρ ∈ Γ(T ∗ M ⊗ T ∗ M ).
0 0
Proof. (⇒) Suppose that Σ is a Brownian motion on M (so Eq. (5.22) holds) and
f, g ∈ C ∞ (M ). Then on one hand
d(f (Σ)g(Σ)) = d [f (Σ)] · g(Σ) + f (Σ)d [g(Σ)] + d[f (Σ), g(Σ)]
∼ 1
= {∆f (Σ)g(Σ) + f (Σ)∆g(Σ)}dλ + d[f (Σ), g(Σ)],
2
where “ ∼
=” denotes equality up to the differential of a martingale. On the other
hand,
1
d(f (Σ)g(Σ)) ∼
= ∆(f g)(Σ)dλ
2
1
= {∆f (Σ)g(Σ) + f (Σ)∆g(Σ) + 2hgrad f, gradgi(Σ)}dλ.
2
Comparing the above two equations implies that
d[f (Σ), g(Σ)] = hgrad f, gradgi(Σ)dλ = df ⊗ dg(I(Σ)idλ.
CURVED WIENER SPACE ANALYSIS 61
3.41, δ(θ(Σ)P (Σ))v = ∇α(δΣ ⊗ V ), where ∇α(vm ⊗ wm ) := (∇vm α)(wm ) for all
vm , wm ∈ T M. Therefore:
δ(α(V )) = δ(θ(Σ)v) = δ(θ(Σ)P (Σ)v) = (d(θP )(δΣ))v + θ(Σ)P (Σ)δv
= (d(θP )(δΣ))v + α̃(Σ)δv = ∇α(δΣ ⊗ V ) + α(δ ∇ V ).
where Lemma 3.30 was used in the third equality. The proof of Eq. (5.34) is
completely analogous. The skeptical reader is referred to Section 3 of Driver [47]
for more details.
Definition 5.28 (Stochastic Parallel Translation). Given v ∈ RN and an M –
valued semi-martingale Σ, let //s (Σ)vΣ0 = (Σs , us v), where u solves (5.32). (Note:
Vs = //s (Σ)V0 .)
In the remainder of these notes, I will often abuse notation and write us instead
of //s := //s (Σ) and vs rather than Vs = (Σs , vs ). For example, the reader should
sometimes interpret us v as //s (Σ)vΣ0 depending on the context. Essentially, we
will be identifying τm M with Tm M when no particular confusion will arise.
Convention. Let us now fix a base point o ∈ M and unless otherwise noted,
we will assume that all M – valued semi-martingales, Σ, start of o ∈ M, i.e. Σ0 = o
a.e.
To each M – valued semi-martingale, Σ, let Ψ(Σ) := b where
Z · Z · Z ·
−1 −1
b := // δΣ = u δΣ = utr δΣ.
0 0 0
In what follows, we will assume that bs , us (or equivalently //s (Σ)), and Σs are
related by Equations (5.35) and (5.32), i.e. Σ = φ (b) and u = // = // (Σ) . Recall
¯ = P (Σ) dΣ is the Itô differential of Σ, see Definition 5.13.
that dΣ
Proposition 5.31. Let Σ = φ (b) , then
(5.36) ¯ = P (Σ)dΣ = udb.
dΣ
64 BRUCE K. DRIVER
Also
d
X
(5.37) dΣ ⊗ dΣ = udb ⊗ udb := uei ⊗ uej dbi dbj ,
i,j=1
Pd
where {ei }di=1 is an orthonormal basis for To M and b = i=1 bi ei . More precisely
Z · Z · Xd
ρ(dΣ ⊗ dΣ) = ρ(uei ⊗ uej )dbi dbj ,
0 0 i,j=1
The rolling construction of Brownian motion seems to have first been discovered
by Eells and Elworthy [62] who used ideas of Gangolli [86]. The relationship of the
stochastic development map to stochastic differential equations on the orthogonal
frame bundle O(M ) of M is pointed out in Elworthy [65, 66, 67]. The frame
bundle point of view has also been extensively developed by Malliavin, see for
example [129, 128, 130]. For a more detailed history of the stochastic development
map, see pp. 156–157 in Elworthy [67]. The reader may also wish to consult
[73, 102, 115, 131, 169, 100].
Corollary 5.34. If Σ is a Brownian motion on M,
π = {0 = s0 < s1 < · · · < sn = T }
is a partition of [0, T ] and f ∈ C ∞ (M n ) , then
Z n
Y
(5.40) Ef (Σs1 , . . . , Σsn ) = f (x1 , x2 , . . . , xn ) p∆i s (xi−1 , xi ) dλ (xi )
Mn i=1
where ∆i s := si − si−1 , x0 := o and λ := λM . In particular Σ is a Markov process
relative to the filtration, {Fs } where Fs is the σ – algebra generated by {Στ : τ ≤ s} .
Proof. By standard measure theoretic arguments, it suffices toQ prove Eq. (5.40)
n
when f is a product function of the form f (x1 , x2 , . . . , xn ) = i=1 fi (xi ) with
¯
fi ∈ C ∞ (M ). By Theorem 5.33, Ms := e(T −s)∆/2 fn (Σs ) is a martingale for s ≤ T
and therefore
"n−1 # "n−1 #
Y Y
E [f (Σs1 , . . . , Σsn )] = E fi (Σsi ) · MT = E fi (Σsi ) · Msn−1
i=1 i=1
"n−1 #
Y
=E
(5.41) fi (Σsi ) · (P∆n s fn ) Σsn−1 .
i=1
In particular if n = 1, it follows that
h i Z
¯
T ∆/2
E [f1 (ΣT )] = E e f1 (Σ0 ) = pT (o, x1 )f1 (x1 ) dλ (x1 ) .
M
Now assume we have proved Eq. (5.40) with n replaced by n − 1 and to simplify
Qn−1
notation let g (x1 , x2 , . . . , xn−1 ) := i=1 fi (xi ) . It would then follow from Eq.
(5.41) that
E [f (Σs1 , . . . , Σsn )]
Z sn −sn−1 n−1
¯ Y
= g (x1 , x2 , . . . , xn−1 ) e 2 ∆ fn (xn−1 ) p∆i s (xi−1 , xi ) dλ (xi )
M n−1 i=1
Z Z
= g (x1 , x2 , . . . , xn−1 ) fn (xn ) p∆n s (xn−1 , xn ) dλ (xn ) ×
M n−1 M
n−1
Y
× p∆i s (xi−1 , xi ) dλ (xi )
i=1
Z n
Y
= f (x1 , x2 , . . . , xn ) p∆i s (xi−1 , xi ) dλ (xi ) .
Mn i=1
This completes the induction step and hence also the proof of the theorem.
CURVED WIENER SPACE ANALYSIS 67
N
Proof. Suppose {vi }i=1 is another orthonormal basis for RN . Using the bilin-
earity of the joint quadratic variation,
X
[hei , Bi, hej , Bi] = [hei , vk ihvk , Bi, hej , vl ihvl , Bi]
k,l
X
= hei , vk ihej , vl i[hvk , Bi, hvl , Bi].
k,l
Therefore,
N
X
P (Σ)ei ⊗ Q(Σ)ej · d B i , B j
i,j=1
N
X
= [P (Σ)ei ⊗ Q(Σ)ej ] hei , vk ihej , vl id[hvk , Bi, hvl , Bi]
i,j,k,l=1
N
X
= [P (Σ)vk ⊗ Q(Σ)vl ] d[hvk , Bi, hvl , Bi]
k,l=1
wherein we have used P (Σ)u = uP (o) and Q(Σ)u = uQ(o), see Theorem 5.27. This
last expression is easily seen to be zero by choosing {ei } such that P (o)ei = ei for
i = 1, 2, . . . , d and Q (o) ej = ej for j = d + 1, . . . , N.
The next proposition is a stochastic analogue of Lemma 3.55 and the proof is
very similar to that of Lemma 3.55.
//s δs //−1 ∇
(5.43) s Vs = δs Vs =: P (Σs ) δVs
//−1 ∇
−1
s δs ∇//s w Ys = δs //s ∇//s w Ys
d
(5.45) = //−1 2
s ∇δΣs ⊗//s w Ys + //−1
s ∇//s w Ys .
ds
Furthermore if Σs is a Brownian motion, then
d
//s−1 Ys (Σs ) =//−1 −1
d s ∇//s dbs Ys + //s Ys (Σs ) ds
ds
d
1 X −1 2
(5.46) + // ∇//s ei ⊗//s ei Ys ds
2 i=1 s
Proof. We will use the convention of summing on repeated indices and write
us for //s , i.e. stochastic parallel translation along Σ on T M. Recall that us solves
Then
δ (ūs us ) = −ūs dQ (δΣs ) us + ūs dQ (δΣs ) us = 0
from which it follows that ūs us = I for all s and hence ūs = u−1
s . This proves Eq.
(5.43) since
us δs u−1
−1 −1
s Vs = us us dQ (δΣs ) Vs + us δVs
= dQ (δΣs ) Vs + δVs = δ ∇ Vs ,
and
//s w = h//s w, Xi (Σs )iXi (Σs ) = hw, //−1
s Xi (Σs )iXi (Σs )
or equivalently,
(5.48) w = hw, //−1 −1
s Xi (Σs )i//s Xi (Σs ) .
Taking the covariant differential of Eq. (5.47), making use of Eq. (5.44), gives
δs∇ ∇//s w Ys
Proof. By Eq. (5.51) and Theorem 5.33, Σ is a martingale and from Eq. (5.1),
n n
j
Xki (Σ) Xkj (Σ) ds
X X
i j i k l
dΣ dΣ = Xk (Σ) Xl (Σ) dB dB =
k,l=1 k=1
N
where is the standard basis for R , Σ := hΣ, ei i and Xki (Σ) = hXk (Σ) , ei i.
{ei }i=1 N i
Notation 5.42. The pull back, Ric//s , of the Ricci tensor by parallel translation
is defined by
(5.60) Ric//s := //−1
s RicΣs //s .
Theorem 5.43 (Itô form of Eq. (5.59)). The Itô form of Eq. (5.59) is
dzs v = //−1
(5.61) s ∇//s zs v X dBs + αs ds
where
(5.62)
n n
" ! #
X 1X ∇
αs := //−1
s ∇//s zs v ∇Xi Xi + X0 − R (//s zs v, Xi (Σs )) Xi (Σs ) ds.
i=1
2 i=1
If we further assume that n = N and Xi (m) = P (m) ei (so that Eq. (5.1) is
equivalent to Eq. (5.42) if X0 ≡ 0), then αs = − 21 Ric//s zs vds, i.e. Eq. (5.59) is
equivalent to
−1 −1 1
(5.63) dzs v = //s P (Σs ) dP (//s zs v) dBs + //s ∇//s zs v X0 − Ric//s zs v ds.
2
Proof. In this proof there will always be an implied sum on repeated indices.
Using Proposition 5.36,
h i
d //s−1 ∇//s zs v X dBs = //−1 ∇2X(Σs )dBs ⊗//s zs v X + ∇//s dzs v X dBs
s
h i
= //s−1 ∇2X(Σs )dBs ⊗//s zs v X + ∇(∇// z v X)dBs X dBs
s s
h i
−1 2
(5.64) = //s ∇Xi (Σs )⊗//s zs v Xi + ∇(∇// z v Xi ) Xi ds.
s s
Eq. (5.61) is now a follows directly from this equation and Eq. (5.59).
If we further assume n = N, Xi (m) = P (m) ei and X0 (m) = 0, then
∇//s zs v X dBs = //−1
(5.66) s P (Σs ) dP (//s zs v) dBs .
Moreover, from the definition of the Ricci tensor in Eq. (3.31) and making use of
Eq. (3.50) in the proof of Proposition 3.48 we have
(5.67) R∇ (//s zs v, Xi (Σs )) Xi (Σs ) = Ric//s //s zs v.
Combining Eqs. (5.66) and (5.67) along with ∇Xi Xi = 0 (from Proposition 3.48)
with Eqs. (5.61) and (5.62) implies Eq. (5.63).
In the next result, we will filter out the “redundant noise” in Eq. (5.63). This is
useful for deducing intrinsic formula from their extrinsic cousins, see, for example,
Corollary 6.4 and Theorem 7.39 below.
74 BRUCE K. DRIVER
Theorem 5.44 (Filtering out the Redundant Noise). Keep the same setup in The-
orem 5.43 with n = N and Xi (m) = P (m) ei . Further let M be the σ – algebra
generated by the solution Σ = {Σs : s ≥ 0} . Then there is a version, z̄s , of E [zs |M]
such that s → z̄s is continuous and z̄ satisfies,
Z s
1
(5.68) z̄s v = v + //−1
r ∇ X
//r z̄r v 0 − Ric z̄
//r r v dr.
0 2
In particular if X0 = 0, then
d 1
(5.69) z̄s = − Ric//s z̄s with z̄0 = id,
ds 2
Proof. In this proof, we let bs be the martingale part of the anti-development
map, Ψs (Σ) , i.e.
Z s Z s
bs := //−1
r P (Σ r ) δB r = //−1
r P (Σr ) dBr .
0 0
Since (Σs , us ) solves the stochastic differential equation,
δΣs = us δbs + X0 (Σs ) ds with Σ0 = o
δu = −Γ (δΣ) u = −Γ (uδb) u with u0 = I ∈ O(N )
it follows that (Σ, u) may be expressed as a function of the Brownian motion, b.
Therefore by the martingale representation property, see Corollary 7.20 below, any
measurable function, f (Σ) , of Σ may be expressed as
Z 1 Z 1
f (Σ) = f0 + har , dbr i = f0 + har , //−1
r [P (Σr ) dBr ]i.
0 0
Hence, using P dP = dP Q, the previous equation and the isometry property of the
Itô integral,
Z s
E [P (Σr ) dP (//r zr v) dBr ] f (Σ)
0
Z s Z 1
=E [dP (//r zr v) Q (Σr ) dBr ] hP (Σr ) //r ar , dBr i
Z0 s 0
=E [dP (//r zr v) Q (Σr ) P (Σr ) //r ar ] dr = 0.
0
This shows that Z s
E P (Σr ) dP (//r zr v) dBr |M = 0
0
and hence taking the conditional expectation, E [·|M] , of the integrated version of
Eq. (5.63) implies Eq. (5.68). In performing this operation we have used the fact
that (Σ, //) is M – measurable and that zs appears linearly in Eq. (5.63). I have
also glossed over the technicality of passing the conditional expectation past the
integrals involving a ds term. For this detail and a much more general presentation
of these ideas the reader is referred to Elworthy, Li and Le Jan [70].
5.7. More References. For more details on the sorts of results in this section,
the books by Elworthy [68], Emery [73], and Ikeda and Watanabe [103], Malliavin
[131], Stroock [169], and Hsu [100] are highly recommended. The following articles
and books are also relevant, [14, 20, 21, 40, 63, 62, 64, 109, 128, 135, 142, 152, 153,
154, 177].
CURVED WIENER SPACE ANALYSIS 75
1 h
~
~
i
= //−1 ∇ 2
∇F (s, ·) − ∇∆F (s, ·) (Σ s ) ds
2 s //s ei ⊗//s ei
+ //−1 ~ i
s ∇//s ei ∇F (s, ·)dbs
1 ~ (s, Σs )ds + //−1 ∇// e ∇F ~ (s, ·)dbi
= //−1 Ric ∇F
2 s s s i s
1 ~
= Ric//s Ws ds + //−1 s ∇//s ei ∇F (s, ·)dbs
i
2
where {ei }di=1 is an orthonormal basis for To M and there is an implied sum on
repeated indices. Hence if Q solves Eq. (6.1), then
1 1 −1 ~ i
d [Qs Ws ] = − Qs Ric//s Ws ds + Qs Ric//s Ws ds + //s ∇//s ei ∇F (s, ·)dbs
2 2
= Qs //−1 ~ i
s ∇// e ∇F (s, ·)dbs
s i
wherein the the third equality we have used (by Lemma 6.1) that s →
Qs //−1 ~
s ∇(F (s, ·))(Σs ) is a martingale. Hence
Z t0
~ t∆/2 f )(o) = 1 E
∇(e Qs dbs (e(t−t0 )∆/2 f )(Σt0 )
t0 0
from which Eq. (6.4) follows using either the Markov property of Σs or the fact
that s → e(t−s)∆/2 f (Σs ) is a martingale.
The following theorem is an non-intrinsic form of Theorem 6.2. In this theo-
rem we will be using the notation introduced before Theorem 5.41. Namely, let
n
{Xi }i=0 ⊂ Γ (T M ) be as in Notation 5.4, Bs be an Rn – valued Brownian motion,
and Ts (m) = Σs where Σs is the solution to Eq. (5.1) with Σs = m ∈ M and
β = B.
# −1
: T m M → Rn ,
(6.5) X (m) = X (m) |Nul(X(m))⊥
where the orthogonal complement is taken relative to the standard inner product
#
on Rn . (See Lemma 7.38 below for more on X (m) .) Then for all v ∈ To M,
0 < to < t < ∞ and f ∈ C (M ) we have
Z t0
1 #
(6.6) v etL/2 f = E f (Σt ) hX (Σs ) Zs v, dBs i
t0 0
Proof. Let L = ni=1 Xi2 + 2X0 be the generator of the diffusion, {Ts (m)}s≥0 .
P
C ∞ (M ) . So, using results similar to those in Fact 5.32, it makes sense to define
Fs (m) := e(t−s)L/2 f (m) and Nsm = Fs (Ts (m)) . Then
1
∂s Fs + LFs = 0 with Ft = f
2
and by Itô’s lemma,
n
X
(6.7) dNsm = d [Fs (Ts (m))] = (Xi Fs ) (Ts (m))dBsi .
i=1
This shows Nsm is a martingale for all m ∈ M and, upon integrating Eq. (6.7) on
s, that
n Z
X t
tL/2
f (Tt (m)) = e f (m) + (Xi Fs ) (Ts (m))dBsi .
i=1 0
Rt
Hence if as ∈ R is a predictable process such that E 0 |as |2 ds < ∞, then by the
n
Since Nsm = Fs (Ts (m)) is a martingale for all m, we may deduce that
(6.10) v (m → Nsm ) = dM Fs (Ts∗o v) = dM Fs (Zs v)
is a martingale as well for any v ∈ To M. In particular, s ∈ [0, t] → E [(dM Fs ) (Zs v)]
is constant and evaluating this expression at s = 0 and s = t implies
(6.11) E [(dM Fs ) (Zs v)] = v etL/2 f = E [(dM f ) (Zt v)] .
#
where we have used L = ∆ (see Proposition 3.48) and X (m) = P (m) in this
setting. By Theorem 5.40,
Z t0 Z t0
h//s zs v, dBs i = h//s zs v, P (Σs ) dBs i
0 0
Z t0 Z t0
−1
= hzs v, //s P (Σs ) dBs i = hzs v, dbs i
0 0
and therefore Eq. (6.12) may be written as
Z t0
1
v et∆/2 f = E f (Σt ) hzs v, dbs i .
t0 0
Using Theorem 5.44, this may also be expressed as
Z t0 Z t0
1 1
(6.13) v et∆/2 f = E f (Σt ) hz̄s v, dbs i = E f (Σt ) hv, z̄str dbs i
t0 0 t0 0
where z̄s solves Eq. (5.69). By taking transposes of Eq. (5.69) it follows that z̄str
satisfies Eq. (6.1) and hence z̄str = Qs . Since v ∈ To M was arbitrary, Equation (6.4)
is now an easy consequence of Eq. (6.13) and the definition of ∇(e ~ t∆/2 f )(o).
7. Calculus on W (M )
In this section, (M, o) is assumed to be either a compact Riemannian manifold
equipped with a fixed point o ∈ M or M = Rd with o = 0.
Notation 7.1. We will be interested in the following path spaces:
W (To M ) := {ω ∈ C([0, 1] → To M )|ω(0) = 0o ∈ To M },
Z 1
H (To M ) := {h ∈ W (To M ) : h(0) = 0, & hh, hiH := |h′ (s)|2To M ds < ∞}
0
and
W (M ) := {σ ∈ C([0, 1] → M ) : σ (0) = 0 ∈ M } .
(By convention hh, hiH = ∞ if h ∈ W (To M ) is not absolutely continuous.) We refer
to W (To M
) as Wiener space, W (M ) as curved Wiener space and H (To M )
or H Rd as the Cameron-Martin Hilbert space.
Definition 7.2. Let µ and µW (M) denote the Wiener measures on W (To M ) and
W (M ) respectively, i.e. µ = Law (b) and µW (M) = Law (Σ) where b and Σ are
Brownian motions on To M and M starting at 0 ∈ To M and o ∈ M respectively.
Notation 7.3. The probability space in this section will often be W (M ) , F , µW (M) ,
where F is the completion of the σ – algebra generated by the projection maps,
Σs : W (M ) → M defined by Σs (σ) = σs for s ∈ [0, 1]. We make this into a filtered
probability space by taking Fs to be the σ – algebra generated by {Σr : r ≤ s} and
the null sets in Fs . Also let //s be stochastic parallel translation along Σ.
Definition 7.4. A function F : W (M ) → R is called a C k – cylinder function
if there exists a partition
(7.1) π := {0 = s0 < s1 < s2 · · · < sn = 1}
k n
of [0, 1] and f ∈ C (M ) such that
(7.2) F (σ) = f (σs1 , . . . , σsn ) for all σ ∈ W (M ) .
CURVED WIENER SPACE ANALYSIS 79
Notation 7.11. Given a Cameron-Martin vector field X on W (M ) , µW (M) and
a cylinder function F ∈ F C 1 (W (M )) as in Eq. (7.2), let XF denote the random
variable
Xn
(7.7) XF (σ) := (gradi F (σ), Xsi (σ)),
i=1
where
(7.8) gradi F (σ) := (gradi f ) (σs1 , . . . , σsn )
and (gradi f ) denotes the gradient of f relative to the ith variable.
Notation 7.12. The gradient, DF, of a smooth cylinder functions, F, on W (M )
is the unique Cameron-Martin process such that G (DF, X) = XF for all X ∈ X .
The explicit formula for D, as the reader should verify, is
n
!
X
−1
(7.9) (DF )s = //s s ∧ si //si gradi F (σ) .
i=1
The formula in Eq. (7.9) defines a densely defined operator, D : L2 (µ) → X with
D (D) = F C 1 (W (M )) as its domain.
7.1. Classical Wiener Space Calculus. In this subsection (which is a warm up
M = R , o = 0 ∈ R . To
d d
for the sequel) we will specialize to the case where
simplify notation let W := W (R ), H := H R , µ = µW (Rd ) , bs (ω) = ωs for
d d
all s ∈ [0, 1] and ω ∈ W. Recall that {Fs : s ∈ [0, 1]} is the filtration on W as
explained in Notation 7.3 where we are now writing b for Σ. Cameron and Martin
[25, 26, 28, 27] and Cameron [28] began the study of calculus on this classical
Wiener space. They proved the following two results, see Theorem 2, p. 387 of [26]
and Theorem II, p. 919 of [28] respectively. (There have been many extensions of
these results partly initiated by Gross’ work in [89, 90].)
Theorem 7.13 (Cameron & Martin 1944). Let (W, F , µ) be the classical Wiener
space described above and for h ∈ W, define Th : W → W by Th (ω) = ω + h for all
ω ∈ W. If h is C 1 , then µTh−1 is absolutely continuous relative to µ.
This theorem was extended by Maruyama [132] and Girsanov [87] to allow the
same conclusion for h ∈ H and more general Cameron-Martin processes. Moreover
it is now well known µTh−1 ≪ µ iff h ∈ H. From the Cameron and Martin theorem
one may prove Cameron’s integration by parts formula.
Theorem 7.14 (Cameron 1951). Let h ∈ H and F, G ∈ L∞− (µ) := ∩1≤p<∞ Lp (µ)
d d
such that ∂h F := dε F ◦ Tεh |ε=0 and ∂h G := dε G ◦ Tεh |ε=0 where the derivatives are
7 p
supposed to exist in L (µ) for all 1 ≤ p < ∞. Then
Z Z
∂h F · G dµ = F ∂h∗ G dµ,
W W
R1
where ∂h∗ G = −∂h G + zh G and zh := 0
′
hh (s) , dbs iRd .
7The notion of derivative stated here is weaker than the notion given in [28]. Nevertheless
Cameron’s proof covers this case without any essential change.
CURVED WIENER SPACE ANALYSIS 81
In this flat setting parallel translation is trivial, i.e. //s = id for all s. Hence the
gradient operator D in Eq. (7.9) reduces to the equation,
n
!
X
(DF )s (ω) = s ∧ si gradi F (ωs ) .
i=1
Proof. We start by proving the theorem under the additional assumption that
(7.10) sup |h′s | ≤ C,
s∈[0,1]
Corollary 7.16. The operator D∗ is densely defined and hence D is closable. (Let
D̄ denote the closure of D.)
Proof. Let h ∈ H and F and K be smooth cylinder functions. Then, by the
product rule,
hDF, KhiX = E[hKDF, hiH ] = E[hD (KF ) − F DK, hiH ]
= E[F · KD∗ h − F hDK, hiH ].
Therefore Kh ∈ D(D∗ ) (D(D∗ ) is the domain of D∗ ) and
D∗ (Kh) = KD∗ h − hDK, hiH .
Since the subspace,
{Kh|h ∈ H and K is a smooth cylinder function},
is a dense subspace of X , D∗ is densely defined.
and hence
F (b) = (e(sn −sn−1 )∆n /2 f )(bs1 , . . . , bsn−1 , bsn−1 )
Z sn
+ hgradn e(sn −s)∆n /2 f )(bs1 , . . . , bsn−1 , bs , dbs i
sn−1
Z sn
(sn −sn−1 )∆n /2
(7.13) = (e f )(bs1 , . . . , bsn−1 , bsn−1 ) + hαs , dbs i,
sn−1
Z 1
d
F = EF + E
(7.16) D̄F s (b) Fs , dbs .
0 ds
In particular if F = f (bs1 , . . . , bsn ) is a smooth cylinder function on W (M ) then
Z 1 * "X n
# +
(7.17) F = EF + E 1s≤si gradi f (bs1 , . . . , bsn ) Fs , dbs .
0 i=1
RProof.
1 ′ 2
Let h be a predictable Cameron-Martin valued process such that
E 0 |hs | ds < ∞. Then using Theorem 7.15 and the Itô isometry property,
Z 1
EhD̄F, hiH = E [F D∗ h] = E F hh′s , dbs i
0
Z 1 Z 1 Z 1
′ ′
(7.18) = E EF + ha, dbi hhs , dbs i = E has , hs ids
0 0 0
and hence
Ms := E [Hε |Fs ] = E F 2 + ε|Fs ≥ ε
where M0 = EHε .
Let φ (x) = x ln x so that φ′ (x) = ln x + 1 and φ′′ (x) = x−1 . Then by Itô’s
formula,
1 2
d [φ (Ms )] = φ (M0 ) + φ′ (Ms ) dMs + φ′′ (Ms ) |as | ds
2
1 1 2
= φ (M0 ) + φ′ (Ms ) dMs + |as | ds.
2 Ms
Integrating this equation on s and then taking expectations shows
Z 1
1 1 2
(7.23) E [φ (M1 )] = φ (EM1 ) + E |as | ds .
2 0 Ms
We may now let ε ↓ 0 in this inequality to find Eq. (7.22) is valid for F ∈ F C 1 (W ) .
Since F C 1 (W ) is a core for D̄, standard limiting arguments show that Eq. (7.22)
is valid in general.
The main objective for the rest of this section is to generalize the previous
theorems to the setting of general compact Riemannian manifolds. Before doing
this we need to record the stochastic analogues of the differentiation formula in
Theorems 4.7, 4.12, and 4.13.
CURVED WIENER SPACE ANALYSIS 87
//−1
s Zs . (See Theorem 5.41 for more on the processes Z and z.) Recall from Nota-
tion 5.4 that
n
X
Xa (m) := ai Xi (m) = X (m) a.
i=1
the pull back of X under the stochastic development map should be the process Y
defined by
Z s Z r
(7.28) Ys = hs + R//ρ (hρ , δbρ ) δbr
0 0
where
(7.29) R//s (hs , δbs ) = //s−1 R(//s hs , //s δbs )//s
like in Eq. (4.15). Since
Z r Z r
1
R//ρ (hρ , δbρ ) δbr = R//ρ (hρ , δbρ ) dbr + R//ρ (hρ , dbρ )dbρ
0 0 2
Z r d
1X
= R//ρ (hρ , δbρ ) dbr + R//ρ (hρ , ei )ei dρ
0 2 i=1
where {ei }di=1 is an orthonormal basis for To M, Eq. (7.28) may be written in Itô
form as
Z · Z ·
(7.30) Y· = Cs dbs + rs ds,
0 0
where
s
1
Z
(7.31) Cs := R//σ (hσ , δbσ ), rs = h′s + Ric//s hs and
0 2
(7.32) Ric//s a := //−1
s Ric //s a ∀ a ∈ To M.
Then there exists a version of φs (bt ) which is continuous in (s, t), differentiable in
d
t and dt |0 φ (bt ) = X h .
Proof. For the proof of this theorem and its generalization to more general h,
the reader is referred to Section 3.1 of [45] and to [47]. Let me just point out here
that formally the proof is very analogous to the deterministic version in Theorems
4.12 and 4.13.
This theorem first appeared in Driver [47] for h ∈ H (To M )∩C 1 ([0, 1], ToM ) and
was soon extended to all h ∈ H (To M ) by E. Hsu [95, 96]. Other proofs may also
be found in [75, 126, 144]. The proof of this theorem is rather involved and will not
be given here. A sketch of the argument and more information on the technicalities
involved may be found in [49].
Example 7.30. When M = Rd , //s (σ)vo = vσs for all v ∈ Rd and σ ∈ W (Rd ).
h
Thus Xsh (σ) = (hs )σs and etX (σ) = σ + th and so Theorem 7.29 becomes the
classical Cameron-Martin Theorem 7.13.
Corollary 7.31 (Integration by Parts for µW (M) ). For h ∈ H(To M ) and F ∈
F C 1 (W (M )) as in Eq. (7.2), let
d h
(X h F )(σ) = |0 F (etX (σ)) = G DF, X h
dt
as in Notation 7.11. Then
Z Z
h
X F dµW (M) = F z h dµW (M)
W(M) W(M)
where
1
1
Z
z h := Ric//s h′s , dbs i,
hh′s +
0 2
Z s
bs (σ) := Ψs (σ) = //−1
r δσr
0
and Ric//s ∈ End(To M ) is as in Eq. (5.60).
90 BRUCE K. DRIVER
Proof. A special case of this Corollary 7.31 with F (σ) = f (σs ) for some f ∈
C ∞ (M ) first appeared in Bismut [21]. The result stated here was proved in [47] as
an infinitesimal form of the flow Theorem 7.29. Other proofs of this corollary may
be found in [2, 5, 50, 71, 72, 69, 75, 77, 95, 96, 121, 122, 126, 144]. This corollary
is a special case of Theorem 7.32 below.
This proves the theorem in the special case that h′ is uniformly bounded.
−1
Let X be a general adapted R s ′Cameron-Martin vector-field and h := // X.n For
each n ∈ N, let hn (s) := 0 h (r) · 1|h′ (r)|≤n dr be as in Eq. (7.11). Set X :=
//hn , then by the special case above we know that X n ∈ D(D∗ ) and D∗ X n =
R1
0 hB(hn ), dbi. It is easy to check that
hX − X n , X − X n iX = Ehh − hn , h − hn iH → 0 as n → ∞.
Furthermore,
Z 1
∗ m n 2
E |D (X − X )| = E |B(hm − hn )|2 ds ≤ CEhhm − hn , hm − hn iH ,
0
Since
span{KX h|h ∈ H and K ∈ F C ∞ } ⊂ D(D∗ )
is is a dense subspace of X , D∗ is densely defined.
Theorem 7.32 may be extended to allow for vector-fields on the paths of M which
are not based. This theorem and it Corollary 7.37 will not be used in the sequel
and may safely be skipped.
Theorem 7.36. Let h be an adapted To M – valued process such that h(0) is non-
random and h − h(0) is a Cameron-Martin process, X := X h := //h, Ex denote
the path space expectation for a Brownian motion starting at x ∈ M, F : C([0, 1] →
M ) → R be a cylinder function as in Definition 7.4 and X h F be defined as in Eq.
(7.7). Then (writing hdf, vi for df (v))
(7.40) Eo [X h F ] = Eo [F D∗ X h ] + hd(E(·) F ), h(0)o i,
where
1 1
1
Z Z
∗ h
D X := hh′s + Ric//s hs , dbs i := hB(h), dbi,
0 2 0
as in Eq. (7.35) and B(h) is defined in Eq. (7.36).
Proof. Start by choosing a smooth path α in M such that α̇(0) = h(0)o . Let
Z
C := R// (h, δb),
1
r = h′ + Ric// (h),
Z s 2 Z s
t tC
bs = e db + t rdλ and
0 0
Z 1
1 2 1
Z
tC
Zt = exp − thr, e dbi + t hr, rids
0 2 0
be defined by the same formulas as in the proof of Theorem 7.32. Let u0 (t) denote
parallel translation along α, that is
du0 (t)/dt + Γ(α̇(t))u0 (t) = 0 with u0 (0) = id.
For t ∈ R, define Σ(t, ·) by
Σ(t, δs) = u(t, s)δbts with Σ(t, 0) = α(t)
and
u(t, δs) + Γ(u(t, s)δs bts )u(t, s) = 0 with u(t, 0) = uo (t).
Appealing to a stochastic version of Theorem 4.14 (after choosing a good version
d
of Σ) it is possible to show that Σ̇(0, ·) = X, so the XF = dt |0 F [Σ(t, ·)] . As in
the proof of Theorem 7.32, b is a Brownian motion relative to the expectation Et
t
CURVED WIENER SPACE ANALYSIS 93
and therefore by integration by parts on the flat Wiener space (Theorem 7.32) with
M = Rn ) implies
" Z T #
E dW (M) F (Σ) (Z· k· ) = E [∂h [F (Σ)]] = E F (Σ) hh′s , dBs i
0
" #
Z T
tr
= E F (Σ) hX (Σs ) Zs ks′ , dBs i .
0
By factoring out the redundant noise in Theorem 7.39, we get yet another proof
of Corollary 7.35 which also easily gives another proof of Theorem 7.32.
Theorem 7.40 (Factoring out the redundant noise). Assume X (m) = P (m) and
X0 = 0, ks is a Cameron-Martin valued process adapted to the filtration, FsΣ :=
σ (Σr : r ≤ s) , then
" Z T #
E dW (M) F //Qtr = E F (Σ) h//s Qtr ′
t k s ks , dbs i
0
Proof. Using
" #
Z T
E dW (M) F (//zk) = E F (Σ) h//s zs ks′ , P
(Σs ) dBs i
0
" #
Z T
= E F (Σ) h//s zs ks′ , dbs i
0
7.6. Fang’s Spectral Gap Theorem and Proof. As in the flat case we let
L = D∗ D̄ – an unbounded operator on L2 W (M ) , µW (M) which is a “curved”
analogue of the Ornstein-Uhlenbeck operator used in Theorem 7.23. It has been
shown in Driver and Röckner [55] that this operator generates a diffusion on W (M ).
This last result also holds for pinned paths on M and free loops on RN , see [6].
In this section, we will give a proof of S. Fang’s [78] spectral gap inequality for
L. Hsu’s stronger logarithmic Sobolev inequality will be covered later in Theorem
7.52 below.
Theorem 7.41 (Fang). Let D̄ be the closure of D and L be the self-adjoint operator
on L2 µW (M) defined by L = D∗ D̄. (Note, if M = Rd then L would be an infinite
dimensional Ornstein-Uhlenbeck operator.) Then the null-space of L consists of the
constant functions on W (M ) and L has a spectral gap, i.e. there is a constant
c > 0 such that hLF, F iL2 (µW (M ) ) ≥ chF, F iL2 (µW (M ) ) for all F ∈ D(L) which are
perpendicular to the constant functions.
This theorem is the W (M ) analogue of Theorem 7.23. The proof of this theorem
will be given at the end of this subsection. We first will need to represent F in
terms of DF. (Also see Section 7.7 below.)
Lemma 7.42. For each F ∈ L2 W (M ) , µW (M) , there is a unique adapted
and
Z 1
(7.45) F = EF + has , dbs i.
0
96 BRUCE K. DRIVER
= hY, B̄ ∗ B̄(X)iX ,
where in going from the first to the second line we have used E [D∗ Y ] = 0. From
the above displayed equation it follows that QD̄F = B̄ ∗ B̄(X) and hence X =
(B̄ ∗ B̄)−1 QD̄F = T (D̄F ).
where C is the operator norm of B̄T. In particular if F ∈ D(L), then hD̄F, D̄F iX =
E[LF · F ], and hence
hLF, F iL2 (µW (M ) ) ≥ C −1 hF − EF, F − EF iL2 (µW (M ) ) .
E //−1
si gradi f Σs1 , . . . , Σsn , Σsn+1
Fsn
¯
(7.49) = //−1
si gradi (e
(sn+1 −sn )∆n+1 /2
f ) (Σs1 , . . . , Σsn , Σsn ) .
98 BRUCE K. DRIVER
Proof. Let us begin with the special case where f = g⊗h for some g ∈ C ∞ (M n )
and h ∈ C ∞ (M ) where g ⊗ h (x1 , . . . , xn+1 ) := g (x1 , . . . , xn ) h (xn+1 ) . In this case
//−1 −1
si gradi f Σs1 , . . . , Σsn , Σsn+1 = //si gradi g (Σs1 , . . . , Σsn ) · h Σsn+1
where //−1
si gradi g (Σs1 , . . . , Σsn ) is Fsn – measurable. Hence by the Markov prop-
erty we have
E //−1
si gradi f Σs1 , . . . , Σsn , Σsn+1
Fsn
= //−1s gradi g (Σs1 , . . . , Σsn ) E h Σsn+1
Fsn
i
¯
= //−1
si gradi g (Σs1 , . . . , Σsn ) (e(sn+1 −sn )∆/2 h) (Σsn )
¯
= //−1
si gradi (e
(sn+1 −sn )∆n+1 /2
f ) (Σs1 , . . . , Σsn , Σsn ) .
¯
Alternatively, as we have already seen, Ms := (e(sn+1 −s)∆/2 h) (Σs ) is a martingale
for s ≤ sn+1 , and therefore,
¯
E h Σsn+1 Fsn = E Msn+1 Fsn = Msn = (e(sn+1 −sn )∆/2 h) (Σsn ) .
Since Eq. (7.49) is linear in f, this proves Eq. (7.49) when f is a linear combination
of functions of the form g ⊗ h as above.
Using a partition unity argument along with the standard convolution ap-
proximation methods; to any f ∈ C ∞ M n+1 there exists a sequence of fk ∈
C ∞ M n+1 with each fk being a linear combination of functions of the form g ⊗ h
such that fk along with all of its derivatives converges uniformly to f. Passing to
the limit in Eq. (7.49) with f being replaced by fk , shows that Eq. (7.49) holds
for all f ∈ C ∞ M n+1 .
Recall that Qs is the End (To M ) – valued process determined in Eq. (6.1) and
since
d −1 d
Qs = −Q−1 s Q s Qs ,
−1
ds ds
Q−1
s solves the equation,
d −1 1 −1
(7.50) Q = Ric//s Q−1 s with Q0 = I.
ds s 2
Theorem 7.47 (Representation Formula). Suppose that F is a smooth cylinder
function of the form F (σ) = f (σs1 , . . . , σsn ) , then
Z 1
(7.51) F (Σ) = EF + has , dbs i
0
where as is a bounded predictable process, as is zero if s ≥ sn and s → as is
continuous off the partition set, {s1 , . . . , sn }. Moreover as may be expressed as
" n #
X
−1 −1
(7.52) as := Qs E 1s≤si Qsi //si gradi f (Σs1 , . . . , Σsn ) Fs .
i=1
¯
By Lemma 6.1, Qs //−1
s grad e
(t−s)∆/2
f (Σs ) is a martingale, and hence
¯
Qs //−1 (t−s)∆/2
f (Σs ) = E Qt //−1
s grad e t grad f (Σt ) Fs
from which it follows that
¯ −1
as = 10≤s≤t //−1 (t−s)∆/2
f (Σs ) = 10≤s≤t Q−1
s E Qt //t grad f (Σt ) Fs .
s grad e
This shows that Eq. (7.52) is valid for n = 1.
To carry out the inductive step, suppose the result holds for level n and now
suppose that
F (Σ) = f Σs1 , . . . , Σsn+1
with 0 < s1 < s2 · · · < sn+1 ≤ 1. Let
(∆n+1 f )(x1 , x2 , . . . , xn+1 ) = (∆g)(xn+1 )
where g(x) := f (x1 , x2 , . . . , xn , x). Similarly, let gradn+1 denote the gradient acting
th
on the (n + 1) – variable of a function f ∈ C ∞ (M n+1 ). Set
¯
H(s, Σ) := (e(sn+1 −s)∆n+1 /2 f )(Σs1 , . . . , Σsn , Σs )
for sn ≤ s ≤ sn+1 . By Itô’s Lemma, (see Corollary 5.18 and also Eq. (5.38),
¯
d [H(s, Σs )] = hgradn+1 e(sn+1 −s)∆n+1 /2 f )(Σs1 , . . . , Σsn , Σs , //s dbs i
for sn ≤ s ≤ sn+1 . Integrating this last expression from sn to sn+1 yields:
¯
F (Σ) = (e(sn+1 −sn )∆n+1 /2 f )(Σs1 , . . . , Σsn , Σsn )
Z sn+1
¯ n+1 /2
(7.54) + h//−1
s gradn+1 e
(sn+1 −s)∆
f ) (Σs1 , . . . , Σsn , Σs ) , dbs i
sn
Z sn+1
¯
(7.55) = (e(sn+1 −sn )∆n+1 /2 f )(Σs1 , . . . , Σsn , Σsn ) + hαs , dbs i,
sn
¯ n+1 /2
where αs := //−1
s (gradn+1 e
(sn+1 −s)∆
f )(Σs1 , . . . , Σsn , Σs ). By the induction
hypothesis, the smooth cylinder function,
¯
(e(sn+1 −sn )∆n+1 /2 f )(Σs1 , . . . , Σsn , Σsn ),
R1
may be written as a constant plus 0 hãs , dbs i, where ãs is bounded and piecewise
continuous and ãs ≡ 0 if s ≥ sn . Thus if we let as := ãs + 1sn <s≤sn+1 αs , we have
shown Z sn+1
F (Σ) = C + has , dbs i
0
for some constant C. Taking expectations of both sides of this equation then shows
C = E [F (Σ)] and the proof of Eq. (7.51) is complete. So to finish the proof it only
remains to verify Eq. (7.52).
Again by Lemma 6.1,
¯
s → Ms := Qs //−1
s (gradn+1 e
(sn+1 −s)∆n+1 /2
f )(Σs1 , . . . , Σsn , Σs )
is a martingale for s ∈ [sn , sn+1 ] and therefore,
¯
Ms = Qs //−1
s (gradn+1 e
(sn+1 −s)∆n+1 /2
f )(Σs1 , . . . , Σsn , Σs )
(7.56)
h i
= E Msn+1 Fs = E Qsn+1 //−1
grad f (Σ , . . . , Σ , Σ sn+1 Fs ,
)
sn+1 n+1 s1 sn
100 BRUCE K. DRIVER
i.e.
¯
//−1
s (gradn+1 e
(sn+1 −s)∆n+1 /2
f )(Σs1 , . . . , Σsn , Σs )
h i
= Q−1 −1
s E Qsn+1 //sn+1 gradn+1 f (Σs1 , . . . , Σsn , Σsn+1 ) Fs .
(7.57)
where
¯
g (x1 , . . . , xn ) := (e(sn+1 −sn )∆n+1 /2 f ) (x1 , . . . , xn , xn ) .
By the induction hypothesis,
g(Σs1 , . . . , Σsn )
Z 1* " n # +
X
(7.59) =C+ Q−1
s E 1s≤si Qsi //s−1 gradi g (Σs1 , . . . , Σsn ) Fs , dbs
i
0 i=1
h i
¯ n+1 /2
E Qsn //−1 grad (e (sn+1 −sn )∆
f ) (Σ , . . . , Σ , Σ sn Fs
)
sn n s1 sn
h h i i
¯ n+1 /2
= E Qsn E //−1 sn gradn (e
(sn+1 −sn )∆
f ) (Σs1 , . . . , Σsn , Σsn ) Fsn Fs
= E Qsn //−1
Fs
sn gradn f Σs1 , . . . , Σsn , Σsn+1
from Eq. (7.57) with s = sn . Combining the previous three displayed equations
shows,
E Qsn //−1
sn gradn g (Σs1 , . . . , Σsn ) Fs
= E Qsn //−1
sn gradn f Σs1 , . . . , Σsn , Σsn+1
Fs
h i
+ E Qsn+1 //−1
(7.62) grad f (Σ , . . . , Σ , Σ n+1 Fs
)
sn+1 n+1 s1 sn s
F (Σ) = E [F (Σ)]
Z 1* " n+1 # +
−1
X
−1
+ Qs E 1s≤si Qsi //si gradi f Σs1 , . . . , Σsn , Σsn+1 Fs , dbs .
0 i=1
and let
n
X n
X
αs := 1s≤si Q−1 −1
s Qsi //si gradi f (Σs1 , . . . , Σsn ) = 1s≤si Q−1
s Qsi vi .
i=1 i=1
and therefore
Z 1 Z 1
αs = −Q−1
s Qr dξr = − Q−1
s Qr dξr .
s s
102 BRUCE K. DRIVER
Conversely we may give a proof of Theorem 7.47 which is based on the integration
by parts Theorem 7.32.
Theorem 7.51 (Representation Formula). Suppose F is a cylinder function on
d
W (M ) as in Eq. (7.2) and ξs := //−1
s ds (DF )s , then
Z 1
1 1 −1
Z
F = EF + E ξs −
(7.64) Q Qr Ric//r ξr dr Fs , dbs .
0 2 s s
where Qs is the solution to Eq. (6.1).
R1 h ∈
Proof. Let
2
Xa be a predictable adapted Cameron-Martin valued process
such that E 0 |h′s | ds < ∞. By the martingale representation property in Corollary
7.20,
Z 1
(7.65) F = EF + ha, dbi
0
R1 2
for some predictable process a such that E 0 |as | ds < ∞. Then from Corollary
7.35 and the Itô isometry property,
h i h ∗ i Z 1
tr tr
E X Q hF = E F · X Q h 1 = E F · hQtr h′ , dbi
0
Z 1 Z 1
(7.66) =E hQtr ′
s hs , as ids =E hh′s , Qs as ids .
0 0
h tr
i
On the other hand we may compute E X Q F as: h
Z 1
h tr
i d
E X Q h F = E hDF, //Qtr hiH = E Qtr h s ids
hξs ,
0 ds
Z 1
1
(7.67) =E ξs , Qtr
s sh ′
− Ric //s Q tr
s h s ds
0 2
where we have used Eq. (7.39) in the last equality. We will now rewrite the right
side of Eq. (7.67) so that it has the same form as Eq. (7.66) To do this let
ρs := 21 Ric//s and notice that
Z 1 Z 1 Z s
tr ∗ ′
hξs , ρs Qs hs ids = Qs ρs ξs , hr dr ds
0 0 0
Z Z 1 Z 1
= drds10≤r≤s≤1 hQs ρ∗s ξs , h′r i = Qr ρ∗r ξr dr, h′s ds
0 s
wherein the last equality we have interchanged the role of r and s. Using this result
back in Eq. (7.67) implies
h i Z 1 Z 1
tr
(7.68) E X Q hF = E Qs ξs − Qr ρ∗r ξr dr, h′s ds.
0 s
and comparing this with Eq. (7.66) shows
Z 1 Z 1
∗ ′
(7.69) E Qs as − Qs ξs + Qr ρr ξr dr, hs ds = 0
0 s
for all h ∈ Xa .
Up to now we have only used F ∈ D (D) and not the fact that F is a cylinder
function. We will use this hypothesis now. From the easy part of Theorem 7.47
104 BRUCE K. DRIVER
Z 1 Z 1 2
// (DF )′ − 1 ′
−1 −1 −1
(7.70) + 2E s s Q s Q r Ric //r // r (DF )r dr ds,
0 2 s
′ d
where (DF )s := ds (DF )s . Moreover, there is a constant C = C (Ric) such that
E F log F 2 ≤ CE hDF, DF iH(To M) + EF 2 · log EF 2 .
2
(7.71)
Proof. The proof we give here follows the paper of Capitaine, Hsu and
Ledoux [29]. We begin in the same way as the proof of Theorem 7.24. Let
F ∈ F C 1 (W (M )) , ε > 0, Hε := F 2 + ε ∈ D D̄ and
1 1 −1
Z
as := E ξs −
Qs Qr Ric//r ξr dr Fs
2 s
where
d d
ξs = //−1
s (DHε )s = 2F · //−1
s (DF )s .
ds ds
Then by Theorem 7.47, Z 1
Hε = EHε + ha, dbi.
0
CURVED WIENER SPACE ANALYSIS 105
Combining the last two equations along with Eq. (7.24) implies
Eφ (Hε ) ≤φ (EHε )
Z 1 " Z 1 2 #
−1 ′ 1 −1 −1 ′
E //s (DF )s −
+ 2E Qs Qr Ric//r //r (DF )r dr Fs ds
0 2 s
=φ (EHε )
Z 1 2
1 1 −1
Z
−1 ′ −1 ′
+ 2E //s (DF )s − 2 Qs Qr Ric//r //r (DF )r dr ds.
0 s
We may now let ε ↓ 0 in this inequality to learn Eq. (7.70) holds for all F ∈
F C 1 (W ) . By compactness of M, Ricm is bounded on M and so by simple Gronwall
type estimates on Q and Q−1 , there is a non-random constant K < ∞ such that
−1
Qs Qr Ric//
≤ K for all r, s.
r op
Therefore,
Z 1 2
// (DF )′ − 1 ′
−1 −1 −1
s s Qs Qr Ric//r //r (DF )r dr
2 s
Z 1 2
′ 1 ′
≤ (DF )s + K
(DF )s ds
2 0
Z 1 2
′ 2 1 2 ′
≤ 2 (DF )s + K
(DF )s ds
2 0
Z 1
′ 2 1 (DF )′ 2 ds
≤ 2 (DF )s + K 2
s
2 0
and hence
1 Z 1 2
(DF )′ − 1
Z
−1 ′
2E s Q s Q r Ric //r (DF )r dr ds
0
2 s
Z 1
≤ 4 + K2 (DF )′ 2 ds.
s
0
106 BRUCE K. DRIVER
Combining this estimate with Eq. (7.70) implies Eq. (7.71) holds with C =
4 + K 2 . Again, since F C 1 (W ) is a core for D̄, standard limiting arguments
show that Eq. (7.70) and Eq. (7.71) are valid for all F ∈ D D̄ .
Theorem 7.52 was first proved by Hsu [97] with an independent proof given
shortly thereafter by Aida and Elworthy [4]. Hsu’s original proof relied on a Markov
dependence version of a standard additivity property for logarithmic Sobolev in-
equalities and makes key use of Corollary 7.37. On the other hand Aida and Elwor-
thy show, using the projection construction of Brownian motion, the logarithmic
Sobolev inequality on W (M ) is a consequence of Gross’ [91] original logarithmic
Sobolev inequality on the classical Wiener space W (RN ), see Theorem 7.24. In
Aida’s and Elworthy’s proof, Theorem 5.43 plays an important role.
7.9. More References. Many people have now proved some version of integration
by parts for path and loop spaces in one context or another, see for example [21, 28,
32, 26, 28, 27, 47, 48, 49, 75, 74, 77, 84, 121, 127, 144, 159, 157, 158, 161, 101]. We
have followed Bismut in these notes who proved integration by parts formulas for
cylinder functions depending on one time. However, as is pointed out by Leandre
and Malliavin and Fang, Bismut’s technique works with out any essential change
for arbitrary cylinder functions. In [47, 48], the flow associated to a general class of
vector fields on paths and loop spaces of a manifold were constructed. The reader
is also referred to the texts [70, 99, 169] and the related articles [80, 79, 35, 76, 81,
82, 83, 34, 37, 33, 38, 36, 39, 124].
Many of the results in this section extend to pinned Wiener measure on loop
spaces, see [48] for example. Loop spaces are more interesting than path spaces
since they have nontrivial topology, The issue of the spectral gap and logarithmic
Sobolev inequalities for general loop spaces is still an open problem. In [92], Gross
has prove a logarithmic Sobolev inequality on Loop groups with an added “potential
term” for a special geometry on loop groups. Here Gross uses pinned Wiener
measure as the reference measure. In Driver and Lohrenz [54], it is shown that a
logarithmic Sobolev inequality without a potential term does hold on the Loop
group provided one replace pinned Wiener measure by a “heat kernel” measure.
The quasi-invarariance properties of the heat kernel measure on loop groups was
first established in [50, 51]. For more results on heat kernel measures on the loop
groups see for example, [56, 3, 30, 31, 81, 82, 105].
The question as to when or if the potential is needed in Gross’s setting for
logarithmic Sobolev inequalities is still an open question, but see Gong, Röckner
and Wu [88] for a positive result in this direction. Eberle [58, 59, 60, 61] has
provided examples of Riemannian manifolds where the spectral gap inequality fails
in the loop space setting. The reader is referred to [52, 53] and the references
therein for some more perspective on the stochastic analysis on loop spaces.
CURVED WIENER SPACE ANALYSIS 107
(3) Let
∆(ω) := det C(ω) = det(DF (ω)DF (ω)∗ )
and assume ∆−1 ∈ L∞− (µ) .
Then the law (µF = F∗ µ = µ ◦ F −1 ) of F is absolutely continuous relative to
Lebesgue measure, λ, on Rd and the Radon-Nikodym derivative, ρ := dµF /dλ, is
smooth.
Proof. For each vector field Y ∈ Γ T Rd , define
induction,
(Y1 Y2 · · · Yk φ)(F (ω)) = (Y1 Y2 · · · Yk (φ ◦ F ))(ω)
and therefore,
Z Z
(Y1 Y2 · · · Yk φ)dµF = (Y1 Y2 · · · Yk φ)(F (ω)) dµ(ω)
Rd
ZW
= (Y1 Y2 · · · Yk (φ ◦ F ))(ω) dµ(ω)
ZW
(8.3) = φ(F (ω)) · (Y∗k Y∗k−1 · · · Y∗1 1)(ω) dµ(ω).
W
By the remarks in the previous paragraph, (Y∗k Y∗k−1 · · · Y∗1 1) ∈ L∞− (µ) which
along with Eq. (8.3) shows
Z
d (Y1 Y2 · · · Yk φ)dµF ≤ C kφkL∞ (Rd ) ,
R
where C =
Y∗k Y∗k−1 · · · Y∗1 1
L1 (µ) < ∞. It now follows from Sobolev imbedding
theorems or simple Fourier analysis that µF ≪ λ and that ρ := dµF /dλ is a smooth
function.
The remainder of Section 8 will be devoted to an infinite dimensional analogue
of Theorem 8.2 (see Theorem 8.9) where Rd is replaced by a manifold M d ,
W := {ω ∈ C ([0, ∞), Rn ) : ω (0) = 0} ,
CURVED WIENER SPACE ANALYSIS 109
and DΣt : H → TΣt M be defined by (DΣt ) h := ∂h TtB (o) as defined Theorem 7.26.
Recall from Theorem 7.26 that
Z t Z t
−1
(8.4) (DΣt ) h := Zt Zτ X (Στ ) ḣτ dτ = //t zt zτ−1 //−1
τ X (Στ ) ḣτ dτ,
0 0
d B
where ḣτ := dτ hτ , Zt := Tt ∗o : To M → TΣt M, //t is stochastic parallel transla-
tion along Σ and zt := //−1t Zt . In the sequel, adjoints will be denote by either “
∗
tr
” or “ ” with the former being used if an infinite dimensional space is involved
and the latter if all spaces involved are finite dimensional.
Definition 8.4 (Reduced Malliavin Covariance). The End (To M ) – valued random
variable,
Z t
tr tr
(8.5) C̄t := Zτ−1 X (Στ ) X (Στ ) Zτ−1 dτ
0
Z t
tr −1 tr
zτ−1 //−1
(8.6) = τ X (Στ ) X (Στ ) //τ zτ dτ,
0
which implies Eq. (8.7). Combining Eqs. (8.4) and (8.7), using
tr
Zτtr = (//τ zτ ) = zτtr //tr tr −1
τ = zτ //τ ,
shows
Z t tr
∗ tr
DΣt (DΣt ) //t v = Zt Zτ−1 X (Στ ) X (Στ ) //τ zt zτ−1 vdτ
0
Z t tr
tr
= Zt Zτ−1 X (Στ ) X (Στ ) Zτ−1 Zttr //t vdτ.
0
Therefore,
Ct = Zt C̄t Zttr = //t zt C̄t zttr //−1
t
from which Eq. (8.8) follows.
The next crucial theorem is at the heart of Malliavin’s method and constitutes
the deepest part of the theory. The proof of this theorem will be postponed until
Section 8.4 below.
¯ t := det C̄t . If Hörmander’s re-
Theorem 8.6 (Non-degeneracy of C̄t ). Let ∆
stricted bracket condition at o ∈ M holds then ∆ ¯ t > 0 a.e. (i.e. C̄t is invertible
¯ −1
a.e.) and moreover ∆t ∈ L ∞−
(µ) .
Following the general strategy outlined in Theorem 8.2, given a vector field
Y ∈ Γ (T M ) we wish to lift it via the map Σt : W → M to a vector field Yt on
W := W (Rn ) . According to the prescription used in Eq. (8.1) in Theorem 8.2,
∗ ∗ −1 ∗
(8.10) Yt := (DΣt ) DΣt (DΣt ) Y (Σt ) = (DΣt ) Ct−1 Y (Σt ) ∈ H.
From Eq. (8.8)
−1 −1 −1 −1
Ct−1 = //t zttr C̄t zt //t
and combining this with Eq. (8.10), using Eq. (8.7), implies
d t d h −1 −1 −1 −1 i
Yτ = 1τ ≤t (DΣt ) //t zttr C̄t zt //t Y (Σt )
dτ dτ τ
tr tr −1 −1 −1 −1
= 1τ ≤t X (Στ )tr //τ zt zτ−1 zt C̄t zt //t Y (Σt )
tr
= 1τ ≤t X (Στ )tr //τ zτ−1 C̄t−1 Zt−1 Y (Σt )
tr tr
= 1τ ≤t X (Στ ) Zτ−1 C̄t−1 Zt−1 Y (Σt ) .
Hence, the formula for Yt in Eq. (8.10) may be explicitly written as
Z s∧t
tr
(8.11) t
Ys = Zτ X (Στ ) dτ C̄t−1 Zt−1 Y (Σt ) .
−1
0
CURVED WIENER SPACE ANALYSIS 111
The reader should observe that the process s → Yts is non-adapted since C̄t−1 Zt−1 Y (Σt )
depends on the entire path of Σ up to time t.
Proof. We only sketch the proof here and refer the reader to [145, 12, 146] with
d
regard to some of the technical details which are omitted below. Let {ei }i=1 be an
orthonormal basis for To M, then
d Z s d
X
−1 −1
tr X
Yts Zτ−1 X (Στ ) ei dτ ai his
(8.13) = ei , C̄t Zt Y (Σt ) =
i=1 0 i=1
where
Z s∧t tr
ai := ei , C̄t−1 Zt−1 Y (Σt ) and his := Zτ−1 X (Στ ) ei dτ.
0
Z s∧t tr
tr ~ (Σt ) dτ.
Zτ−1 Zttr ∇f
D̄ [f (Σt )] s
= X (Στ )
0
112 BRUCE K. DRIVER
= (Y f ) (Σt ) .
= E F · GD∗ Yt
∗
from which it follows that G ∈ D (Yt ) and
∗
(8.15) Yt G = −Yt G + GD∗ Yt .
From the general theory (see [146] for example), D∗ U is Malliavin smooth if U
∗
is Malliavin smooth. In particular (Yt ) G is Malliavin smooth if G is Malliavin
smooth.
Theorem 8.9 (Smoothness of Densities). Assume the restricted Hörmander con-
dition holds at o ∈ M (see Definition 8.1) and suppose f ∈ C ∞ (M ) and
k
{Yi }i=1 ⊂ Γ (T M ) . Then
E [(Y1 . . . Yk f ) (Σt )] = E Yt1 . . . Ytk [f (Σt )]
h ∗ ∗ i
(8.16) = E [f (Σt )] Ytk . . . Yt1 1 .
Moreover, the law of Σt is smooth.
Proof. By an induction argument using Eq. (8.14),
Yt1 . . . Ytk [f (Σt )] = (Y1 . . . Yk f ) (Σt )
from which Eq. (8.16) is a simple consequence. As has already been observed,
∗ ∗ ∗ ∗
(Ytk ) . . . (Yt1 ) 1 is Malliavin smooth and in particular (Ytk ) . . . (Yt1 ) 1 ∈ L1 (µ) .
Therefore it follows from Eq. (8.16) that
∗ ∗
(8.17) |E [(Y1 . . . Yk f ) (Σt )]| ≤
Ytk . . . Yt1 1
kf k∞ .
1 L (µ)
Since the argument used in the proof of Theorem 8.2 after Eq. (8.16) is local in
nature, it follows from Eq. (8.17) that the Law(Σt ) has a smooth density relative
to any smooth measure on M and in particular the Riemannian volume measure.
CURVED WIENER SPACE ANALYSIS 113
8.3. The Invertibility of C̄t in the Elliptic Case. As a warm-up to the proof
of the full version of Theorem 8.6 let us first consider the special case where X (m) :
Rn → Tm M is surjective for all m ∈ M. Since M is compact this will imply there
exists and ε > 0 such that
X(m)Xtr (m) ≥ εITm M for all m ∈ M.
Notation 8.10. We will write f (ε) = O (ε∞− ) if, for all p < ∞,
|f (ε)|
lim = 0.
ε↓0 εp
Proposition 8.11 (Elliptic Case). Suppose there is an ε > 0 such that
X(m)Xtr (m) ≥ εITm M
−1
∈ L∞− (µ) .
for all m ∈ M, then det C̄t
Proof. Let δ ∈ (0, 1) and
(8.18) Tδ := inf {t > 0 : |zt − ITo M | > δ}
where, as usual,
zt := //−1 −1
TtB
t Zt = //t ∗o
.
Since for all a ∈ To M,
−1
hZτ−1 X(Στ )Xtr (Στ ) Zτtr a, ai
D −1 −1 E
= X(Στ )Xtr (Στ ) Zτtr a, Zτtr a
D E D −1 E
−1 −1
≥ ε Zτtr a, Zτtr a = ε a, Zτtr Zτtr
a ,
we have
−1
Zτ−1 X(Στ )Xtr (Στ ) Zτtr
−1 tr −1
−1 −1
≥ εZτtr Zτtr = εzttr //tr zttr = εzttr zttr
t //t .
Hence
Z t −1
C̄t = Zτ−1 X(Στ )Xtr (Στ ) Zτtr dτ
0
Z t −1
Z t∧Tδ −1
≥ε Zτ−1 Zτtr dτ ≥ ε zτtr zτtr dτ
0 0
and therefore,
!
Z t∧Tδ −1
¯ t = det C̄t ≥ εn det zτtr zτtr
∆ dτ .
0
which will be finite for all p > 1 iff µ (t ∧ Tδ ≤ τ ) = µ (Tδ ≤ τ ) = O(τ k ) as τ ↓ 0 for
all k > 0.
By Chebyschev’s inequalities and Eq. (9.10) of Proposition 9.5 below,
(8.19) µ (Tδ ≤ τ ) = µ sup |zs − I| > δ ≤ δ −p E sup |zs − I|p = O(τ p/2 ).
s≤τ s≤τ
To verify this claim, notice that λ0 := inf v∈S hC̄t v, vi is the smallest eigenvalue of C̄t .
Since det ¯ t := det C̄t ≥ λn0
C̄t is the product
n
of the eigenvalues of C̄t it follows that ∆
and so det C̄t < ε ⊂ {λ0 < ε} and hence
µ det C̄t < εn ≤ µ (λ0 < ε) = O(ε∞− ).
By replacing ε by ε1/n above this implies µ ∆ ¯ t < ε = O(ε∞− ). From this estimate
According to Lemma 8.13 and Proposition 8.15 below, Eq. (8.20) holds with
(8.21) T = Tδ := inf {t > 0 : max {|zt − ITo M | , dist(Σt , Σ0 )} > δ}
provided δ > 0 is chosen sufficiently small.
CURVED WIENER SPACE ANALYSIS 115
The rest of this section is now devoted to the proof of Lemma 8.13 and Propo-
sition 8.15 below. In what follows we will make repeated use of the identity,
n Z T
X
−1 2
(8.22) hC̄T v, vi = Zτ Xi (Στ ), v dτ.
i=1 0
n
To prove this, let {ei }i=1 be the standard basis for Rn . Then
n
−1 X −1
Zτ−1 X(Στ )Xtr (Στ ) Zτtr v= Zτ−1 X(Στ )ei hei , Xtr (Στ ) Zτtr vi
i=1
Xn
= hZτ−1 Xi (Στ ), vi Zτ−1 Xi (Στ )
i=1
so that
D E Xn
−1 2
Zτ−1 X(Στ )Xtr (Στ ) Zτtr
−1
v, v = Zτ Xi (Στ ), v
i=1
which upon integrating on τ gives Eq. (8.22).
In the proofs below, there will always be an implied sum on repeated indices.
Lemma 8.13 (Compactness Argument). Let Tδ be as in Eq. (8.21) and suppose
for all v ∈ S there exists i ∈ {1, . . . , n} and an open neighborhood N ⊂o S of v
such that
Z Tδ !
−1 2
Zτ Xi (Στ ), u dτ < ε = O ε∞− ,
(8.23) sup µ
u∈N 0
For w ∈ To M, let ∂w denote the directional derivative acting on functions f (v) with
v ∈ To M. Because for all v, w ∈ Rn with |v| ≤ 1 and |w| ≤ 1 (using Eq. (8.22)),
Xn Z Tδ
−1
Zτ Xi (Στ ), v Zτ−1 Xi (Στ ), w dτ
∂w C̄T v, v ≤2
δ
i=1 0
n Z Tδ
Zτ Xi (Στ )2
X −1
≤2 Hom(Rn ,To M)
dτ
i=1 0
n Z Tδ
z // Xi (Στ )2
X −1 −1
=2 τ τ Hom(Rn ,To M)
dτ,
i=1 0
by choosing δ > 0 in Eq. (8.21) sufficiently small we may assume there is a non-
random constant θ < ∞ such that
sup ∂w C̄Tδ v, v ≤ θ < ∞.
|v|,|w|≤1
There exists D < ∞ satisfying: for any ε > 0, there is an open cover of S with at
n
most D · (θ/ε) balls of the form B(vj , ε/θ). From Eq. (8.25), for any v ∈ S there
exists j such that v ∈ B(vj , ε/θ) ∩ S and
C̄Tδ v, v − C̄Tδ vj , vj < ε.
So if inf v∈S C̄Tδ v, v < ε then minj C̄Tδ vj , vj < 2ε, i.e.
[
inf C̄Tδ v, v < ε ⊂ min C̄Tδ vj , vj < 2ε ⊂ C̄Tδ vj , vj < 2ε .
v∈S j
j
Therefore,
X
µ inf C̄Tδ v, v < ε ≤ µ C̄Tδ vj , vj < 2ε
v∈S
j
n
≤ D · (θ/ε) · sup µ C̄Tδ v, v < 2ε
v∈S
n ∞−
≤ D · (θ/ε) O(ε ) = O(ε∞− ).
n
( )
1 X −1 2
Zs−1 [X0 , W ](Σs )
(8.27) + + Z LXi W (Σs ) ds,
2 i=1 s
where LX W := [X, W ] as in Theorem 4.9.
Proof. Write W (Σs ) = Zs ws , i.e. let ws := Zs−1 W (Σs ). By Proposition 5.36
and Theorem 5.41,
∇δΣs W = δ ∇ [W (Σs )] = δ ∇ [Zs ws ] = δ ∇ Zs ws + Zs δws
Proof. The proof given here will follow Norris [145]. Hörmander’s condition
implies there exist l ∈ N and β > 0 such that
1 X
K(o)K(o)tr ≥ 3βI
|Kl |
K∈Kl
K∈Kl τ ≤Tδ
τ ≤Tδ
τ ≤Tδ
Write K = LXir . . . LXi2 Xi1 with r ≤ l. If it happens that r = 1 then Eq. (8.28)
becomes
Z Tδ !
−1 2
Zτ Xi1 (Στ ), u dt ≤ ε = O ε∞−
sup µ C̄Tδ u, u ≤ ε ≤ sup µ
u∈U u∈U 0
where
(Z )
Tδ
−1 2
Ω1 (u) := Zt Kj−1 (Σt ), u dt < εq ,
0
n
(Z )
Tδ X
−1 2
Ω2 (u) := Zτ [Xi , Kj−1 ](Στ ), u dτ ≥ ε
0 i=1
8.5. More References. The literature on the “Malliavin calculus” is very exten-
sive and I will not make any attempt at summarizing it here. Let me just add to
references already mentioned the articles in [174, 104, 151] which carry out Malli-
avin’s method in the geometric context of these notes. Also see [148] for another
method which works if Hörmander’s bracket condition holds at level 2, namely when
span({K(m) : K ∈ K2 }) = Tm M for all m ∈ M,
see Definition 8.1. The reader should also be aware of the deep results of Ben Arous
and Leandre in [17, 18, 16, 15, 123].
where
n
2
X X
|A| = tr (AA∗ ) = (AA∗ )ii = Aij Aij = tr (A∗ A) .
i=1 i,j
120 BRUCE K. DRIVER
Proof. We may assume the right side of Eq. (9.2) is finite for otherwise there
is nothing to prove. For the moment further assume α ≡ 0. By a standard limiting
argument involving stopping times we may further assume there is a non-random
constant C < ∞ such that
Z T
∗ 2
YT + |Aτ | dτ ≤ C.
0
p
Let f (y) = |y| and ŷ := y/ |y| for y ∈ Rd . Then, for a, b ∈ Rd ,
∂a f (y) = p |y|p−1 ŷ · a = p |y|p−2 y · a
and
∂b ∂a f (y) = p (p − 2) |y|p−4 (y · a) (y · b) + p |y|p−2 b · a
p−2
= p |y| [(p − 2) (ŷ · a) (ŷ · b) + b · a] .
So by Itô’s formula
p
d |Yt | = d [f (Yt )]
p−1 p p−2
h i
= p |Yt | Ŷt · dYt +
|Yt | (p − 2) Ŷt · dYt Ŷt · dYt + dYt · dYt .
2
Taking expectations of this formula (using Y is a martingale) then gives
p t p−2 h
Z i
p
(9.3) E |Yt | = E |Y | (p − 2) Ŷ · dY Ŷ · dY + dY · dY .
2 0
Using dY = AdB,
2
dY · dY = Aei · Aej dB i dB j = ei · A∗ Aei dt = tr(A∗ A)dt = |A| dt
and
2
Ŷ · dY = Ŷ · Aei Ŷ · Aej dB i dB j = A∗ Ŷ · ei A∗ Ŷ · ei dt
2
= A∗ Ŷ · A∗ Ŷ dt = AA∗ Ŷ · Ŷ dt ≤ |A| dt.
th
and taking the p – power of this equation proves Eq. (9.2).
Remark 9.3. A slightly different application of Hölder’s inequality to the right side
of Eq. (9.4) gives
Z t h i Z t
p−2
∗ p ∗ p−2 2 ∗ p p p 2/p
E |Yt | ≤ C E |Yt | |Aτ | dτ ≤ C [E |Yt | ] [E |Aτ | ] dτ
0 0
p−2
Z t
2/p
= [E |Yt∗ |p ] p
C [E |Aτ |p ] dτ
0
which leads to the estimate
Z t p/2
p p 2/p
E |Yt∗ | ≤ C [E |Aτ | ] dτ .
0
Proof. Since
1
Xi (Σt )δB i (t) = Xi (Σt )dB i (t) + d [Xi (Σt )] · dB i (t)
2
i 1
= Xi (Σt )dB (t) + ∂Xi (Σt ) Xi (Σt )dt,
2
the Itô form of Eq. (5.1) is
1
∂Xi (Σt ) Xi (Σt ) dt + Xi (Σt )dB i (t) with Σ0 = x,
δΣt = X0 (Σt ) +
2
122 BRUCE K. DRIVER
or equivalently,
t Z t
1
Z
Xi (Στ )dBτi
Σt = x + + X0 (Στ ) + ∂Xi (Στ ) Xi (Στ ) dτ.
0 0 2
By Proposition 9.2,
Z t p/2
p p p 2
E |Σt | ≤ E (Σ∗t ) ≤ Cp |x| + Cp E |X(Στ )| dτ
0
Z t p
+ Cp E X0 (Στ ) + 1 ∂X (Σ ) Xi (Στ ) dτ .
(9.7) i τ
0
2
Using the bounds on the derivatives of X we learn
2 2
|X(Στ )| ≤ C 1 + |Στ | and
X0 (Στ ) + 1 ∂X (Σ ) Xi (Στ ) ≤ C (1 + |Στ |)
i τ
2
which combined with Eq. (9.7) gives the estimate
p p
E |Σt | ≤ E (Σ∗t )
Z t p/2 Z t p
p 2
≤ Cp |x| + Cp E C 1 + |Στ | dτ + Cp E C (1 + |Στ |) dτ .
0 0
Now assuming t ≤ T < ∞, we have by Jensen’s (or Hölder’s) inequality that
p p
E |Σt | ≤E (Σ∗t )
Z t p/2 dτ
p 2
≤C |x| + Ctp/2 E 1 + |Στ |
0 t
Z t
p dτ
+ Ctp E (1 + |Στ |)
0 t
Z t p/2
≤C |x|p + CT (p/2−1) E 1 + |Στ |2 dτ
0
Z t
+ CT (p−1) E (1 + |Στ |)p dτ
0
from which it follows
Z t
(9.8) E |Σt |p ≤ E (Σ∗t )p ≤ C |x|p + C(T ) (1 + E |Στ |p ) dτ.
0
p
An application of Gronwall’s inequality now shows supt≤T E |Σt | < ∞ for all p < ∞
and feeding this back into Eq. (9.8) with t = T proves Eq. (9.6).
Proposition 9.5. Suppose {Xi }ni=0 is a collection of smooth vector fields on M,
Σt solves Eq. (5.1) with Σ0 = o ∈ M and β = B, zt is the solution to Eq. (5.59)
(i.e. zt := //−1 B 9
t Tt∗o ) and further assume there is a constant K < ∞ such that
kA (m)kop ≤ K < ∞ for all m ∈ M, where A (m) ∈ End (Tm M ) is defined by
n n
" ! #
1 X X
A (m) v := ∇v ∇Xi Xi + X0 − R∇ (v, Xi (m)) Xi (m)
2 i=1 i=1
and
n
X
|∇v Xi | ≤ K |v| for all v ∈ T M.
i=1
Then for all p < ∞ and T < ∞,
p
(9.9) E sup |zt | < ∞
t≤T
and
∗p
E (z· − I)t = O tp/2 as t ↓ 0.
(9.10)
Proof. In what follows C will denote a constant depending on K, T and p. From
Theorem 5.43, we know that the integrated Itô form of Eq. (5.59) is
Z t
1
//−1
(9.11) zt = ITo M + τ ∇//τ zτ (·) X dBτ + A//τ zτ vdτ
0 2
where A//t := //t−1 A (Σt ) //t . By Proposition 9.2 and the assumed bounds on A
and ∇· X,
Z tXn
!p/2
∗ p p −1 2
E (zt ) ≤C |I| + CE //τ ∇// z (·) Xi dτ
τ τ
0 i=1
Z t p
+ CE A// zτ dτ
τ
0
Z t p/2 Z t p
2
≤C + CE |zτ | dτ + CE |zτ | dτ
0 0
Z t
p
≤C + C E |zτ | dτ
0
and
Z t p/2 Z t p
∗p 2
E (z· −
I)t ≤ CE |zτ | dτ + CE |zτ | dτ
0 0
p
(9.12) ≤ C · E |zt∗ | · tp/2 + tp
where we have made use of Hölder’s (or Jensen’s) inequality. Since
Z t
p ∗ p p
(9.13) E |zt | ≤ E (zt ) ≤ C + C E |zτ | dτ,
0
Gronwall’s inequality implies
p
sup E [|zt | ] ≤ CeCT < ∞.
t≤T
Feeding the last inequality back into Eq. (9.13) shows Eq. (9.9). Eq. (9.10) now
follows from Eq. (9.9). and Eq. (9.12).
Exercise 9.6. Show under the same hypothesis of Proposition 9.5 that
p
E sup zt−1 < ∞
t≤T
9.2. Martingale Estimates. This section follows the presentation in Norris [145].
Lemma 9.7 (Reflection Principle). Let βt be a 1 - dimensional Brownian motion
starting at 0, a > 0 and Ta = inf {t > 0 : βt = a} – be first time βt hits height a,
see Figure 15. Then
Z ∞
2 2
P (Ta < t) = 2P (βt > a) = √ e−x /2t dx
2πt a
and hence that β̃t is another Brownian motion. Since β̃t hits level a for the first
time exactly when βt hits level a,
n o
Ta = T̃a := inf t > 0 : β̃t = a
n o
and T̃a < t = {Ta < t} . Furthermore (see Figure 16),
n o n o
{Ta < t & βt < a} = T̃a < t & β̃t > a = β̃t > a .
CURVED WIENER SPACE ANALYSIS 125
Figure 16. The Brownian motion βt and its reflection β̃t about
the line y = a. Note that after time Ta , the labellings of the βt and
the β̃t could be interchanged and the picture would still be possible.
This should help alleviate the readers fears that Brownian motion
has some funny asymmetry after the first hitting of level a.
Therefore,
P (Ta < t & βt < a) = P (β̃t > a) = P (βt > a)
which completes the proof.
Remark 9.8. An alternate way to get a handle on the stopping time Ta is to compute
its Laplace transform. This can be done by considering the martingale
1 2
Mt := eλβt − 2 λ t .
Since Mt is bounded by eλa for t ∈ [0, Ta ] the optional sampling theorem may be
applied to show
h 1 2 i h 1 2
i
eλa E e− 2 λ Ta = E eλa− 2 λ Ta = EMTa = EM0 = 1,
h 1 2 i
i.e. this implies that E e− 2 λ Ta = e−λa . This is equivalent to
√
E e−λTa = e−a 2λ .
From this point of view one would now have to invert the Laplace transform to get
the density of the law of Ta .
Corollary 9.9. Suppose now that T = inf {t > 0 : |βt | = a} , i.e. the first time βt
leaves the strip (−a, a). Then
Z ∞
4 2
P (T < t) ≤ 4P (βt > a) = √ e−x /2t dx
2πt a
r !
8t −a2 /2t
(9.15) ≤ min e ,1 .
πa2
Notice that P (T < t) = P (βt∗ ≥ a) where βt∗ = max {|βτ | : τ ≤ t} . So Eq. (9.15)
may be rewritten as
r !
∗ 8t −a2 /2t 2
(9.16) P (βt ≥ a) ≤ 4P (βt > a) ≤ min 2
e , 1 ≤ 2e−a /2t .
πa
126 BRUCE K. DRIVER
Proof. By definition T = Ta ∧ T−a so that {T < t} = {Ta < t} ∪ {T−a < t} and
therefore
P (T < t) ≤ P (Ta < t) + P (T−a < t)
Z ∞
4 2
= 2P (Ta < t) = 4P (βt > a) = √ e−x /2t dx
2πt a
Z ∞ ∞ r
4 x −x /2t
2 4 t −x2 /2t 8t −a2 /2t
≤ √ e dx = √ − e = e .
2πt a a 2πt a
a πa2
This proves everything but the very last inequality in Eq. (9.16). To prove this
inequality first observe the elementary calculus inequality:
4 −y 2 /2 2
(9.17) min √ e , 1 ≤ 2e−y /2 .
2πy
4
√
Indeed Eq. (9.17) holds √2πy ≤ 2, i.e. if y ≥ y0 := 2/ 2π. The fact that Eq. (9.17)
holds for y ≤ y0 follows from the following trivial inequality
1 2
1 ≤ 1.4552 ∼
= 2e− π = e−y0 /2 .
√
Finally letting y = a/ t in Eq. (9.17) gives the last inequality in Eq. (9.16).
Theorem 9.10. Let N be a continuous martingale such that N0 = 0 and T be a
stopping time. Then for all ε, δ > 0,
2
P (hN iT < ε & NT∗ ≥ δ) ≤ P (βε∗ ≥ δ) ≤ 2e−δ /2ε
.
Proof. By the Dambis, Dubins & Schwarz’s theorem (see p.174 of [108]) we
may write Nt = βhN it where β is a Brownian motion (on a possibly “augmented”
probability space). Therefore
{hN iT < ε & NT∗ ≥ δ} ⊂ {βε∗ ≥ δ}
and hence from Eq. (9.16),
2
P (hN iT < ε & NT∗ ≥ δ) ≤ P (βε∗ ≥ δ) ≤ 2e−δ /2ε
.
we learn
P (Yt∗ ≥ a) ≤ P (Mt∗ ≥ a/2) ≤ P (βct ∗
≥ a/2)
s s
8ct −(a/2)2 /2ct 8ct −(a/2)2 /2ct
≤ 2e = 2e
π (a/2) π (a/2)
s
a2
8c (a/2c) −(a/2)2 /2ct 4
≤ 2 e = √ exp −
π (a/2) πa 8ct
wherein the last inequality we have used the restriction t < a/2c.
Lemma 9.12. If f : [0, ∞) → R is a locally absolutely continuous function such
that f (0) = 0, then
r
|f (t)| ≤ 2
f˙
kf kL1 ([0,t]) ∀ t ≥ 0.
L∞ ([0,t])
We are now ready for a key result needed in the probabilistic proof of
Hörmander’s theorem. Loosely speaking it states that if Y is a Brownian semi-
martinagale, then it can happen only with small probability that the L2 – norm
of Y is small while the quadratic variation of Y is relatively large.
Proposition 9.13 (A key martingale inequality). Let T be a stopping time bounded
by t0 < ∞, Y = y+M +A where M is a continuous martingale and A is a process of
bounded variation such that M0 = A0 = 0. Further assume, on the set {t ≤ T } , that
hM it and |A|t are absolutely continuous functions and there exists finite positive
constants, c1 and c2 , such that
dhM it d |A|t
≤ c1 and ≤ c2 .
dt dt
Then for all ν > 0 and q > ν + 4 there exists constants c = c(t0 , q, ν, c1 , c2 ) > 0 and
ε0 = ε0 (t0 , q, ν, c1 , c2 ) > 0 such that
Z T !
2 q 1
= O ε−∞
(9.18) P Yt dt < ε , hY iT = hM iT ≥ ε ≤ 2 exp − ν
0 2c1 ε
for all ε ∈ (0, ε0 ].
q−ν
R·
Proof. Let q0 = 2 (so that q0 ∈ (2, q/2)), N := 0 Y dM and
q
(9.19) Cε := {hN iT ≤ c1 ε , NT∗ q0
≥ ε }.
We will show shortly that for ε sufficiently small,
(Z )
T
(9.20) Bε := Yt2 dt < εq , hY iT ≥ ε ⊂ Cε .
0
From Lemma 9.12 with f (t) = hY it and the assumption that dhY it /dt ≤ c1 ,
s
r
Z T
˙
(9.23) hY iT ≤ 2
f
kf kL1 ([0,T ]) ≤ 2c1 hY it dt.
∞ L ([0,T ]) 0
References
[1] Ralph Abraham and Jerrold E. Marsden, Foundations of mechanics, Benjamin/Cummings
Publishing Co. Inc. Advanced Book Program, Reading, Mass., 1978, Second edition, revised
and enlarged, With the assistance of Tudor Raţiu and Richard Cushman. MR 81e:58025
[2] S. Aida, On the irreducibility of certain dirichlet forms on loop spaces over compact homoge-
neous spaces, New Trends in Stochastic Analysis (New Jersey) (K. D. Elworthy, S. Kusuoka,
and I. Shigekawa, eds.), Proceedings of the 1994 Taniguchi Symposium, World Scientific,
1997, pp. 3–42.
[3] Shigeki Aida and Bruce K. Driver, Equivalence of heat kernel measure and pinned Wiener
measure on loop groups, C. R. Acad. Sci. Paris Sér. I Math. 331 (2000), no. 9, 709–712.
MR 1 797 756
[4] Shigeki Aida and David Elworthy, Differential calculus on path and loop spaces. I. Loga-
rithmic Sobolev inequalities on path spaces, C. R. Acad. Sci. Paris Sér. I Math. 321 (1995),
no. 1, 97–102.
[5] Helene Airault and Paul Malliavin, Integration by parts formulas and dilatation vector fields
on elliptic probability spaces, Probab. Theory Related Fields 106 (1996), no. 4, 447–494.
[6] Sergio Albeverio, Rémi Léandre, and Michael Röckner, Construction of a rotational invari-
ant diffusion on the free loop space, C. R. Acad. Sci. Paris Sér. I Math. 316 (1993), no. 3,
287–292.
[7] Y. Amit, A multiflow approximation to diffusions, Stochastic Process. Appl. 37 (1991),
no. 2, 213–237.
[8] Lars Andersson and Bruce K. Driver, Finite-dimensional approximations to Wiener measure
and path integral formulas on manifolds, J. Funct. Anal. 165 (1999), no. 2, 430–498. MR
2000j:58059
[9] Louis Auslander and Robert E. MacKenzie, Introduction to differentiable manifolds, Dover
Publications Inc., New York, 1977, Corrected reprinting. MR 57 #10717
[10] Vlad Bally, Approximation for the solutions of stochastic differential equations. I. Lp -
convergence, Stochastics Stochastics Rep. 28 (1989), no. 3, 209–246.
[11] , Approximation for the solutions of stochastic differential equations. II. Strong con-
vergence, Stochastics Stochastics Rep. 28 (1989), no. 4, 357–385.
[12] Denis R. Bell, Degenerate stochastic differential equations and hypoellipticity, Pitman Mono-
graphs and Surveys in Pure and Applied Mathematics, vol. 79, Longman, Harlow, 1995. MR
99e:60124
[13] Denis R. Bell and Salah Eldin A. Mohammed, The Malliavin calculus and stochastic delay
equations, J. Funct. Anal. 99 (1991), no. 1, 75–99. MR 92k:60124
[14] Ya. I. Belopol′ skaya and Yu. L. Dalecky, Stochastic equations and differential geometry,
Mathematics and its Applications (Soviet Series), vol. 30, Kluwer Academic Publishers
Group, Dordrecht, 1990, Translated from the Russian. MR 91b:58271
[15] G. Ben Arous and R. Léandre, Décroissance exponentielle du noyau de la chaleur sur la
diagonale. I, Probab. Theory Related Fields 90 (1991), no. 2, 175–202. MR 93b:60136a
[16] , Décroissance exponentielle du noyau de la chaleur sur la diagonale. II, Probab.
Theory Related Fields 90 (1991), no. 3, 377–402. MR 93b:60136b
[17] Gérard Ben Arous, Développement asymptotique du noyau de la chaleur hypoelliptique sur
la diagonale, Ann. Inst. Fourier (Grenoble) 39 (1989), no. 1, 73–99. MR 91b:58272
[18] , Flots et séries de Taylor stochastiques, Probab. Theory Related Fields 81 (1989),
no. 1, 29–77. MR 90a:60106
[19] Richard L. Bishop and Richard J. Crittenden, Geometry of manifolds, AMS Chelsea Pub-
lishing, Providence, RI, 2001, Reprint of the 1964 original. MR 2002d:53001
[20] Jean-Michel Bismut, Mecanique aleatoire. (french) [random mechanics], Springer-Verlag,
Berlin-New York, 1981, Lecture Notes in Mathematics, Vol. 866.
[21] , Large deviations and the Malliavin calculus, Progress in Mathematics, vol. 45,
Birkhäuser Boston Inc., Boston, Mass., 1984.
[22] G. Blum, A note on the central limit theorem for geodesic random walks, Bull. Austral.
Math. Soc. 30 (1984), no. 2, 169–173.
[23] Nicolas Bouleau and Francis Hirsch, Dirichlet forms and analysis on Wiener space, de
Gruyter Studies in Mathematics, vol. 14, Walter de Gruyter & Co., Berlin, 1991. MR
93e:60107
130 BRUCE K. DRIVER
[24] Theodor Bröcker and Tammo tom Dieck, Representations of compact Lie groups, Gradu-
ate Texts in Mathematics, vol. 98, Springer-Verlag, New York, 1995, Translated from the
German manuscript, Corrected reprint of the 1985 translation. MR 97i:22005
[25] R. H. Cameron and W. T. Martin, Transformations of Wiener integrals under translations,
Ann. of Math. (2) 45 (1944), 386–396. MR 6,5f
[26] , The transformation of Wiener integrals by nonlinear transformations, Trans. Amer.
Math. Soc. 66 (1949), 253–283. MR 11,116b
[27] , Non-linear integral equations, Ann. of Math. (2) 51 (1950), 629–642. MR 11,728d
[28] Robert H. Cameron, The first variation of an indefinite Wiener integral, Proc. Amer. Math.
Soc. 2 (1951), 914–924. MR 13,659b
[29] Mireille Capitaine, Elton P. Hsu, and Michel Ledoux, Martingale representation and a
simple proof of logarithmic Sobolev inequalities on path spaces, Electron. Comm. Probab. 2
(1997), 71–81 (electronic).
[30] Trevor R. Carson, Logarithmic sobolev inequalities for free loop groups, Uni-
versity of California at San Diego Ph.D. thesis. This may be retrieved at
http://math.ucsd.edu/∼driver/driver/thesis.htm, 1997.
[31] , A logarithmic Sobolev inequality for the free loop group, C. R. Acad. Sci. Paris Sér.
I Math. 326 (1998), no. 2, 223–228. MR 99g:60108
[32] Carolyn Cross, Differentials of measure-preserving flows on path space, Uni-
versity of California at San Diego Ph.D. thesis. This may be retrieved at
http://math.ucsd.edu/∼driver/driver/thesis.htm, 1996.
[33] A. B. Cruzeiro, S. Fang, and P. Malliavin, A probabilistic Weitzenböck formula on Rie-
mannian path space, J. Anal. Math. 80 (2000), 87–100. MR 2002d:58045
[34] A. B. Cruzeiro and P. Malliavin, Riesz transforms, commutators, and stochastic integrals,
Harmonic analysis and partial differential equations (Chicago, IL, 1996), Chicago Lectures
in Math., Univ. Chicago Press, Chicago, IL, 1999, pp. 151–162. MR 2001d:60062
[35] Ana Bela Cruzeiro and Shizan Fang, An L2 estimate for Riemannian anticipative stochastic
integrals, J. Funct. Anal. 143 (1997), no. 2, 400–414. MR 98d:60106
[36] Ana-Bela Cruzeiro and Shizan Fang, A Weitzenböck formula for the damped Ornstein-
Uhlenbeck operator in adapted differential geometry, C. R. Acad. Sci. Paris Sér. I Math.
332 (2001), no. 5, 447–452. MR 2002a:60090
[37] Ana-Bela Cruzeiro and Paul Malliavin, Frame bundle of Riemannian path space and Ricci
tensor in adapted differential geometry, J. Funct. Anal. 177 (2000), no. 1, 219–253. MR
2001h:60103
[38] , A class of anticipative tangent processes on the Wiener space, C. R. Acad. Sci.
Paris Sér. I Math. 333 (2001), no. 4, 353–358. MR 2002g:60084
[39] , Stochastic calculus of variations and Harnack inequality on Riemannian path
spaces, C. R. Math. Acad. Sci. Paris 335 (2002), no. 10, 817–820. MR 2003k:58058
[40] R. W. R. Darling, Martingales in manifolds—definition, examples, and behaviour under
maps, Seminar on Probability, XVI, Supplement, Lecture Notes in Math., vol. 921, Springer,
Berlin, 1982, pp. 217–236. MR 84j:58133
[41] E. B. Davies, Heat kernels and spectral theory, Cambridge Tracts in Mathematics, vol. 92,
Cambridge University Press, Cambridge, 1990. MR 92a:35035
[42] Manfredo Perdigão do Carmo, Riemannian geometry, Mathematics: Theory & Applications,
Birkhäuser Boston Inc., Boston, MA, 1992, Translated from the second Portuguese edition
by Francis Flaherty. MR 92i:53001
[43] Jozef Dodziuk, Maximum principle for parabolic inequalities and the heat flow on open
manifolds, Indiana Univ. Math. J. 32 (1983), no. 5, 703–716. MR 85e:58140
[44] H. Doss, Connections between stochastic and ordinary integral equations, Biological Growth
and Spread (Proc. Conf., Heidelberg, 1979) (Berlin-New York), Springer, 1979, Lecture
Notes in Biomath., 38, pp. 443–448.
[45] B. K. Driver, The Lie bracket of adapted vector fields on Wiener spaces, Appl. Math. Optim.
39 (1999), no. 2, 179–210. MR 2000b:58063
[46] Bruce K. Driver, Classifications of bundle connection pairs by parallel translation and lassos,
J. Funct. Anal. 83 (1989), no. 1, 185–231.
[47] , A Cameron-Martin type quasi-invariance theorem for Brownian motion on a com-
pact Riemannian manifold, J. Funct. Anal. 110 (1992), no. 2, 272–376.
CURVED WIENER SPACE ANALYSIS 131
[72] K. D. Elworthy and Xue-Mei Li, A class of integration by parts formulae in stochastic
analysis. I, Itô’s stochastic calculus and probability theory, Springer, Tokyo, 1996, pp. 15–
30.
[73] Michel Emery, Stochastic calculus in manifolds, Universitext, Springer-Verlag, Berlin, 1989,
With an appendix by P.-A. Meyer.
[74] O. Enchev and D. Stroock, Integration by parts for pinned Brownian motion, Math. Res.
Lett. 2 (1995), no. 2, 161–169.
[75] O. Enchev and D. W. Stroock, Towards a Riemannian geometry on the path space over a
Riemannian manifold, J. Funct. Anal. 134 (1995), no. 2, 392–416.
[76] S. Fang, Stochastic anticipative calculus on the path space over a compact Riemannian
manifold, J. Math. Pures Appl. (9) 77 (1998), no. 3, 249–282. MR 99i:60110
[77] S. Z. Fang and P. Malliavin, Stochastic analysis on the path space of a Riemannian manifold.
I. Markovian stochastic calculus, J. Funct. Anal. 118 (1993), no. 1, 249–274.
[78] Shi Zan Fang, Inégalité du type de Poincaré sur l’espace des chemins riemanniens, C. R.
Acad. Sci. Paris Sér. I Math. 318 (1994), no. 3, 257–260.
[79] Shizan Fang, Rotations et quasi-invariance sur l’espace des chemins, Potential Anal. 4
(1995), no. 1, 67–77. MR 96d:60080
[80] , Stochastic anticipative integrals on a Riemannian manifold, J. Funct. Anal. 131
(1995), no. 1, 228–253. MR 96i:58178
[81] , Integration by parts for heat measures over loop groups, J. Math. Pures Appl. (9)
78 (1999), no. 9, 877–894. MR 1 725 745
[82] , Integration by parts formula and logarithmic Sobolev inequality on the path space
over loop groups, Ann. Probab. 27 (1999), no. 2, 664–683. MR 1 698 951
[83] , Ricci tensors on some infinite-dimensional Lie algebras, J. Funct. Anal. 161 (1999),
no. 1, 132–151. MR 2000f:58013
[84] I. B. Frenkel, Orbital theory for affine Lie algebras, Invent. Math. 77 (1984), no. 2, 301–352.
MR 86d:17014
[85] Sylvestre Gallot, Dominique Hulin, and Jacques Lafontaine, Riemannian geometry, second
ed., Universitext, Springer-Verlag, Berlin, 1990. MR 91j:53001
[86] R. Gangolli, On the construction of certain diffusions on a differenitiable manifold, Z.
Wahrscheinlichkeitstheorie und Verw. Gebiete 2 (1964), 406–419.
[87] I. V. Girsanov, On transforming a class of stochastic processes by absolutely continuous
substitution of measures, Teor. Verojatnost. i Primenen. 5 (1960), 314–330. MR 24 #A2986
[88] Fuzhou Gong, Michael Röckner, and Liming Wu, Poincaré inequality for weighted first order
Sobolev spaces on loop spaces, J. Funct. Anal. 185 (2001), no. 2, 527–563. MR 2002j:47074
[89] Leonard Gross, Abstract Wiener spaces, Proc. Fifth Berkeley Sympos. Math. Statist. and
Probability (Berkeley, Calif., 1965/66), Vol. II: Contributions to Probability Theory, Part
1, Univ. California Press, Berkeley, Calif., 1967, pp. 31–42. MR 35 #3027
[90] , Potential theory on Hilbert space, J. Functional Analysis 1 (1967), 123–181. MR
37 #3331
[91] , Logarithmic Sobolev inequalities, Amer. J. Math. 97 (1975), no. 4, 1061–1083. MR
54 #8263
[92] , Logarithmic Sobolev inequalities on loop groups, J. Funct. Anal. 102 (1991), no. 2,
268–313. MR 93b:22037
[93] S. J. Guo, On the mollifier approximation for solutions of stochastic differential equations,
J. Math. Kyoto Univ. 22 , no. 2 (1982), 243–254.
[94] Noel J. Hicks, Notes on differential geometry, Van Nostrand Mathematical Studies, No. 3,
D. Van Nostrand Co., Inc., Princeton, N.J.-Toronto-London, 1965. MR 31 #3936
[95] E. P. Hsu, Flows and quasi-invariance of the Wiener measure on path spaces, Stochastic
analysis (Ithaca, NY, 1993), Proc. Sympos. Pure Math., vol. 57, Amer. Math. Soc., Provi-
dence, RI, 1995, pp. 265–279.
[96] , Quasi-invariance of the Wiener measure on the path space over a compact Rie-
mannian manifold, J. Funct. Anal. 134 (1995), no. 2, 417–450.
[97] Elton P. Hsu, Inégalités de Sobolev logarithmiques sur un espace de chemins, C. R. Acad.
Sci. Paris Sér. I Math. 320 (1995), no. 8, 1009–1012.
[98] , Estimates of derivatives of the heat kernel on a compact Riemannian manifold,
Proc. Amer. Math. Soc. 127 (1999), no. 12, 3739–3744. MR 2000c:58047
CURVED WIENER SPACE ANALYSIS 133
[124] Xiang Dong Li, Existence and uniqueness of geodesics on path spaces, J. Funct. Anal. 173
(2000), no. 1, 182–202. MR 2001f:58074
[125] T. J. Lyons and Z. M. Qian, Calculus for multiplicative functionals, Itô’s formula and
differential equations, Itô’s stochastic calculus and probability theory, Springer, Tokyo, 1996,
pp. 233–250.
[126] , Stochastic Jacobi fields and vector fields induced by varying area on path spaces,
Imperial College of Science, 1996.
[127] Marie-Paule Malliavin and Paul Malliavin, An infinitesimally quasi-invariant measure on
the group of diffeomorphisms of the circle, Special functions (Okayama, 1990), ICM-90
Satell. Conf. Proc., Springer, Tokyo, 1991, pp. 234–244. MR 93h:58027
[128] Paul Malliavin, Geometrie differentielle stochastique, Séminaire de Mathématiques
Supérieures, Presses de l’Université de Montréal, Montreal, Que, 1978, Notes prepared by
Danièle Dehen and Dominique Michel.
[129] , Stochastic calculus of variation and hypoelliptic operators, Proceedings of the In-
ternational Symposium on Stochastic Differential Equations (Res. Inst. Math. Sci., Kyoto
Univ., Kyoto, 1976) (New York-Chichester-Brisbane), Wiley, 1978, pp. 195—263,.
[130] , Stochastic jacobi fields, Partial Differential Equations and Geometry (Proc. Conf.,
Park City, Utah, 1977 (New York), Dekker, 1979, Lecture Notes in Pure and Appl. Math.,
48, pp. 203–235.
[131] , Stochastic analysis, Grundlehren der Mathematischen Wissenschaften [Fundamen-
tal Principles of Mathematical Sciences], vol. 313, Springer-Verlag, Berlin, 1997.
[132] Gisiro Maruyama, Notes on Wiener integrals, Kōdai Math. Sem. Rep. 1950 (1950), 41–44.
MR 12,343d
[133] E. J. McShane, Stochastic differential equations and models of random processes, Proceed-
ings of the Sixth Berkeley Symposium on Mathematical Statistics and Probability (Univ.
California, Berkeley, Calif., 1970/1971), Vol. III: Probability theory, Univ. California Press,
Berkeley, Calif., 1972, pp. 263–294.
[134] , Stochastic calculus and stochastic models, Academic Press, New York, 1974, Prob-
ability and Mathematical Statistics, Vol. 25.
[135] P.-A. Meyer, A differential geometric formalism for the Itô calculus, Stochastic integrals
(Proc. Sympos., Univ. Durham, Durham, 1980), Lecture Notes in Math., vol. 851, Springer,
Berlin, 1981, pp. 256–270. MR 84e:60084
[136] Jürgen Moser, A new technique for the construction of solutions of nonlinear differential
equations, Proc. Nat. Acad. Sci. U.S.A. 47 (1961), 1824–1831. MR 24 #A2695
[137] , A rapidly convergent iteration method and non-linear differential equations. II,
Ann. Scuola Norm. Sup. Pisa (3) 20 (1966), 499–535. MR 34 #6280
[138] , A rapidly convergent iteration method and non-linear partial differential equations.
I, Ann. Scuola Norm. Sup. Pisa (3) 20 (1966), 265–315. MR 33 #7667
[139] J.-M. Moulinier, Théorème limite pour les équations différentielles stochastiques, Bull. Sci.
Math. (2) 112 (1988), no. 2, 185–209.
[140] S. Nakao and Y. Yamato, Approximation theorem on stochastic differential equations,
Proceedings of the International Symposium on Stochastic Differential Equations (Res.
Inst. Math. Sci., Kyoto Univ., Kyoto, 1976) (New York-Chichester-Brisbane), Wiley, 1978,
pp. 283–296.
[141] John Nash, The imbedding problem for Riemannian manifolds, Ann. of Math. (2) 63 (1956),
20–63. MR 17,782b
[142] J. R. Norris, A complete differential formalism for stochastic calculus in manifolds,
Séminaire de Probabilités, XXVI, Lecture Notes in Math., vol. 1526, Springer, Berlin, 1992,
pp. 189–209. MR 94g:58254
[143] , Path integral formulae for heat kernels and their derivatives, Probab. Theory Re-
lated Fields 94 (1993), no. 4, 525–541.
[144] , Twisted sheets, J. Funct. Anal. 132 (1995), no. 2, 273–334. MR 96f:60094
[145] James Norris, Simplified Malliavin calculus, Séminaire de Probabilités, XX, 1984/85, Lec-
ture Notes in Math., vol. 1204, Springer, Berlin, 1986, pp. 101–130.
[146] David Nualart, The Malliavin calculus and related topics, Probability and its Applications
(New York), Springer-Verlag, New York, 1995. MR 96k:60130
CURVED WIENER SPACE ANALYSIS 135
[147] Barrett O’Neill, Semi-Riemannian geometry, Pure and Applied Mathematics, vol. 103, Aca-
demic Press Inc. [Harcourt Brace Jovanovich Publishers], New York, 1983, With applications
to relativity. MR 85f:53002
[148] Jean Picard, Gradient estimates for some diffusion semigroups, Probab. Theory Related
Fields 122 (2002), no. 4, 593–612. MR 2003d:58056
[149] Mark A. Pinsky, Stochastic Riemannian geometry, Probabilistic analysis and related topics,
Vol. 1 (New York), Academic Press, 1978, pp. 199–236.
[150] M. M. Rao, Stochastic processes: general theory, Mathematics and its Applications, vol.
342, Kluwer Academic Publishers, Dordrecht, 1995. MR 97c:60092
[151] Jang Schiltz, Time dependent Malliavin calculus on manifolds and application to nonlinear
filtering, Probab. Math. Statist. 18 (1998), no. 2, Acta Univ. Wratislav. No. 2111, 319–334.
MR 2000b:60144
[152] Laurent Schwartz, Semi-martingales sur des variétés, et martingales conformes sur des
variétés analytiques complexes, Lecture Notes in Mathematics, vol. 780, Springer, Berlin,
1980. MR 82m:60051
[153] , Géométrie différentielle du 2ème ordre, semi-martingales et équations
différentielles stochastiques sur une variété différentielle, Seminar on Probability, XVI, Sup-
plement, Lecture Notes in Math., vol. 921, Springer, Berlin, 1982, pp. 1–148. MR 83k:60064
[154] , Semimartingales and their stochastic calculus on manifolds, Collection de la Chaire
Aisenstadt. [Aisenstadt Chair Collection], Presses de l’Université de Montréal, Montreal,
QC, 1984, Edited and with a preface by Ian Iscoe. MR 86b:60085
[155] Ichiro Shigekawa, Absolute continuity of probability laws of Wiener functionals, Proc. Japan
Acad. Ser. A Math. Sci. 54 (1978), no. 8, 230–233. MR 81m:60097
[156] , Derivatives of Wiener functionals and absolute continuity of induced measures, J.
Math. Kyoto Univ. 20 (1980), no. 2, 263–289. MR 83g:60051
[157] , On stochastic horizontal lifts, Z. Wahrsch. Verw. Gebiete 59 (1982), no. 2, 211–221.
MR 83i:58102
[158] , Transformations of the Brownian motion on a Riemannian symmetric space, Z.
Wahrsch. Verw. Gebiete 65 (1984), no. 4, 493–522.
[159] , Transformations of the Brownian motion on the Lie group, Stochastic analysis
(Katata/Kyoto, 1982), North-Holland Math. Library, vol. 32, North-Holland, Amsterdam,
1984, pp. 409–422.
[160] Ichirō Shigekawa, de Rham-Hodge-Kodaira’s decomposition on an abstract Wiener space, J.
Math. Kyoto Univ. 26 (1986), no. 2, 191–202. MR 88h:58009
[161] Ichiro Shigekawa, Differential calculus on a based loop group, New trends in stochastic
analysis (Charingworth, 1994), World Sci. Publishing, River Edge, NJ, 1997, pp. 375–398.
MR 99k:60146
[162] Michael Spivak, A comprehensive introduction to differential geometry. Vol. I, second ed.,
Publish or Perish Inc., Wilmington, Del., 1979. MR 82g:53003a
[163] Robert S. Strichartz, Analysis of the Laplacian on the complete Riemannian manifold, J.
Funct. Anal. 52 (1983), no. 1, 48–79. MR 84m:58138
[164] D. Stroock and S. Taniguchi, Diffusions as integral curves, or Stratonovich without Itô,
The Dynkin Festschrift, Progr. Probab., vol. 34, Birkhäuser Boston, Boston, MA, 1994,
pp. 333–369.
[165] D. W. Stroock and S. R. S. Varadhan, On the support of diffusion processes with applica-
tions to the strong maximum principle, Proceedings of the Sixth Berkeley Symposium on
Mathematical Statistics and Probability (Univ. California, Berkeley, Calif., 1970/1971), Vol.
III: Probability theory, Univ. California Press, Berkeley, Calif., 1972, pp. 333–359.
[166] Daniel W. Stroock, The Malliavin calculus, a functional analytic approach, J. Funct. Anal.
44 (1981), no. 2, 212–257. MR 83h:60076
[167] , The Malliavin calculus and its application to second order parabolic differential
equations. I, Math. Systems Theory 14 (1981), no. 1, 25–65. MR 84d:60092a
[168] , The Malliavin calculus and its application to second order parabolic differential
equations. II, Math. Systems Theory 14 (1981), no. 2, 141–171. MR 84d:60092b
[169] , An introduction to the analysis of paths on a Riemannian manifold, Mathematical
Surveys and Monographs, vol. 74, American Mathematical Society, Providence, RI, 2000.
MR 2001m:60187
136 BRUCE K. DRIVER
[170] Daniel W. Stroock and James Turetsky, Short time behavior of logarithmic derivatives of
the heat kernel, Asian J. Math. 1 (1997), no. 1, 17–33. MR 99b:58225
[171] , Upper bounds on derivatives of the logarithm of the heat kernel, Comm. Anal.
Geom. 6 (1998), no. 4, 669–685. MR 99k:58174
[172] Daniel W. Stroock and S. R. S. Varadhan, Diffusion processes with continuous coefficients.
II, Comm. Pure Appl. Math. 22 (1969), 479–530.
[173] H. J. Sussmann, Limits of the Wong-Zakai type with a modified drift term, Stochastic
analysis, Academic Press, Boston, MA, 1991, pp. 475–493.
[174] Setsuo Taniguchi, Malliavin’s stochastic calculus of variations for manifold-valued Wiener
functionals and its applications, Z. Wahrsch. Verw. Gebiete 65 (1983), no. 2, 269–290. MR
85d:58088
[175] Krystyna Twardowska, Approximation theorems of Wong-Zakai type for stochastic differ-
ential equations in infinite dimensions, Dissertationes Math. (Rozprawy Mat.) 325 (1993),
54. MR 94d:60092
[176] Nolan R. Wallach, Harmonic analysis on homogeneous spaces, Marcel Dekker Inc., New
York, 1973, Pure and Applied Mathematics, No. 19. MR 58 #16978
[177] S. Watanabe, Lectures on stochastic differential equations and Malliavin calculus, Tata
Institute of Fundamental Research Lectures on Mathematics and Physics, vol. 73, Published
for the Tata Institute of Fundamental Research, Bombay, 1984, Notes by M. Gopalan Nair
and B. Rajeev. MR 86b:60113
[178] E. Wong and M. Zakai, On the relation between ordinary and stochastic differential equa-
tions, Internat. J. Engrg. Sci. 3 (1965), 213–229.
[179] , On the relation between ordinary and stochastic differential equations and appli-
cations to stochastic problems in control theory, Automatic and remote control III (Proc.
Third Congr. Internat. Fed. Automat. Control (IFAC), London, 1966), Vol. 1, p. 5, Paper
3B, Inst. Mech. Engrs., London, 1967, p. 8.